Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Oct 14.
Published in final edited form as: J Am Chem Soc. 2020 Sep 30;142(41):17294–17300. doi: 10.1021/jacs.0c08262

C2-Symmetric Dinickel Catalysts for Enantioselective [4 + 1]-Cycloadditions

Michael J Behlen 1, Christopher Uyeda 1,*
PMCID: PMC7698671  NIHMSID: NIHMS1648254  PMID: 32972140

Abstract

Dinickel naphthyridine–bis(oxazoline) catalysts promote enantioselective intermolecular [4 + 1]-cycloadditions of vinylidene equivalents and 1,3-dienes. The products of this reaction are methylenecyclopentenes, and the exocyclic alkene is generally obtained with high Z selectivity. E and Z dienes react in a stereoconvergent fashion, providing cycloadducts with the same sense of absolute stereochemistry and nearly identical ee values. This feature allows dienes that are commercially available as E/Z mixtures to be used as substrates for the cycloaddition. A DFT model for the origin of asymmetric induction is provided.

Graphical Abstract

graphic file with name nihms-1648254-f0001.jpg


Several of the most universal ligands in asymmetric catalysis employ substituted oxazolines as the source of chirality.1 Prominent examples include C2-symmetric bis(oxazoline) and pyridine–bis(oxazoline) ligands, which have been used in transformations as mechanistically diverse as cycloadditions, carbonyl additions, alkene additions, cross-couplings, and C–H functionalizations.2 Being derived from α-amino acids, oxazolines can be readily configured to meet different requirements for steric discrimination. Additionally, by constraining amino acids into five-membered rings, rotational flexibility is minimized, and the stereogenic substituent is positioned close to the substrate binding site. Whereas most oxazoline-containing ligands are designed to support single metal ions, it would be attractive to incorporate such frameworks into new chiral multinucleating ligands.

Dinuclear complexes containing metal–metal bonds can exhibit catalytic properties that are challenging to replicate using mononuclear complexes.3 For example, (NDI)Ni2 catalysts (NDI = naphthyridine–diimine) promote [4 + 1]-cycloadditions4,5,6 of vinylidenes and 1,3-dienes.7 Based on experimental observations and DFT modeling studies, we hypothesized that the preference for [4 + 1]-cycloaddition over competing cyclopropanation arises due to the stepwise nature of the vinylidene addition. In our proposed mechanism, migratory insertion of the diene into a Ni2(C=CR2) forms a dinickel metallacycle. Reductive elimination could then proceed in either a 1,3 or a 1,5 fashion, the latter of which appears to be favored.

Seeking to render this process enantioselective, we prepared several chiral NDI ligands using imines derived from α-stereogenic primary amines.7 Promising levels of asymmetric induction were observed in intramolecular [4 + 1]-cycloadditions (Figure 1A). However, ee values greater than 90% could not be obtained, and these ligands proved to be ineffective for intermolecular cycloadditions. Cramer recently described alternative chiral NDI variants based on C2-symmetric 2,6-di-(1-arylethyl)anilines.8 These ligands were used to carry out enantioselective methylenecyclopropanation reactions.

Figure 1.

Figure 1.

Asymmetric [4 + 1]-cycloaddition reactions of vinylidenes and 1,3-dienes using C2-symmetric dinickel catalysts.

We reasoned that replacing the imine donors of NDI with oxazolines might impart added rigidity and provide opportunities to sample a broader range of steric environments in the catalyst active site. To this end, we report here that (napbox)Ni2 complexes (napbox = naphthyridine–bis(oxazoline)) catalyze enantioselective intermolecular [4 + 1]-cycloaddition reactions of terminal monosubstituted and 1,2-disusbstituted 1,3-dienes (Figure 1B).9 Notably, several dienes that did not undergo cycloaddition using (NDI)Ni2 catalysts—for example, those containing Ar groups at the 1-position—are viable substrates using (napbox)Ni2 catalysts, indicating that the two classes of catalysts have complementary substrate scopes.

(napbox)Ni2OAc Complexes.

Synthesizing well-defined (napbox)Ni2 complexes necessitated developing a new metalation procedure. Reacting Me2napbox 1 with Ni(cod)2 (2.0 equiv) or with a combination of Ni(cod)2 (1.0 equiv) and Ni(dme)Br2 (1.0 equiv), routes previously used to synthesize (NDI)Ni2 species,10 yielded monometallated products with concomitant precipitation of Ni(0). Templating the formation of the Ni–Ni bond using a carboxylate counterion proved to be a more fruitful approach. Accordingly, comproportionation of Ni(cod)2 (1.5 equiv) and Ni(OAc)2 (0.5 equiv) in the presence of Me2napbox 1 at elevated temperatures (75 °C, 24 h, THF) provided (Me2napbox)Ni2(OAc) (2) as a crystalline violet solid in 60% yield (Figure 2A).

Figure 2.

Figure 2.

(a) Synthesis of (Me2napbox)Ni2(OAc) (2). (b) Comparison of structural parameters for Me2napbox (1) and (Me2napbox)Ni2(OAc) (2) showing bond distances relevant to ligand-based redox activity. (c) Calculated SOMO for (Me2napbox)Ni2(OAc) (2) (BP86/6–311G(d,p) level of DFT). (d) X-band EPR spectra for (Me2napbox)Ni2(OAc) (2) (5.5 K, g = [2.0318, 2.0093]) and (i-PrNDI)Ni2(OAc) (3) (5.5 K, g = [2.0634, 2.0186]). (e) Cyclic voltammograms for (Me2napbox)Ni2(OAc) (2) and (i-PrNDI)Ni2(OAc) (3) (N2 atmosphere, glassy carbon working electrode, 0.3 M [n-Bu4N]PF6 in THF, 100 mV/s scan rate).

The XRD structure of 2 reveals that acetate coordinates in a μ-O,O’ fashion coplanar with the napbox ligand (Figure 2B). The local geometry at Ni is square planar, and the Ni–Ni distance (2.372(1) Å) is within the range of other compounds containing Ni(I)–Ni(I) single bonds.11 The bond metrics associated with the napbox π-system exhibit characteristic features of ligand-centered reduction, such as contracted C1–C2/C3–C4 distances and elongated N1–C1/N3–C3 and N2–C2/N4–C4 distances.12 The EPR spectrum for 2 is indicative of an S = 1/2 ground state, with the unpaired electron predominantly residing on the napbox ligand (Figure 2D). The giso value at room temperature (2.0305), is near that of the free electron, and the g anisotropy in frozen solution (Me-THF, 5.5 K) is relatively small (gmaxgmin = 0.0225). Thus, the structural and spectroscopic data for 2 are consistent with it being best described as a Ni(I)–Ni(I) complex bearing a napbox radical anion.13 According to DFT models, 61% of the total spin density for 2 is localized on the napbox ligand (Figure 2C).

For the Ni2 catalyzed reductive [4 + 1]-cycloaddition, it is critical for the redox potentials of the complex to be matched with the Zn/ZnCl2 couple in order to achieve turnover. Therefore, we were interested in assessing the electronic perturbation associated with replacing imine donors with oxazolines. By cyclic voltammetry, (Me2napbox)Ni2(OAc) (2) possesses a quasi-reversible oxidation at E1/2 = –0.99 V vs. Cp2Fe/Cp2Fe+ (–0.61 V vs. SCE) and a fully reversible reduction at E1/2 = −1.81 V (–1.43 V vs. SCE) (Figure 2E).14 These redox events are nearly coincident with those observed for (i-PrNDI)Ni2(OAc) (3). The modest anodic shift in the potentials for the Me2napbox complex 2 may be a consequence of the oxygen atoms being inductively electron-withdrawing.

Enantioselective [4 + 1]-Cycloadditions.

Several chiral derivatives of napbox were then synthesized and tested in the model intermolecular [4 + 1]-cycloaddition of 1,1-dichloroalkene 4 and 1,3-diene 5 (see Supporting Information for reaction optimization studies). The (t-Bunapbox)Ni2OAc complex (7) provided 6 in 34% yield (18:1 Z/E ratio) and 89% ee for the major Z stereoisomer (Figure 1B). Significant improvements in yield and selectivity were obtained using the tetraarylated (t-BuPh2napbox)Ni2OAc catalyst 8. Under fully optimized conditions, 6 was obtained in 81% yield as a >20:1 ratio of Z/E stereoisomers. The major Z isomer was produced in 95% ee. Interestingly, (NDI)Ni2 catalysts proved to be ineffective at promoting [4 + 1]-cycloadditions with this diene. For example, (i-PrNDI)Ni2(OAc) catalyst 3 generated 6 in only 4% yield and a relatively low Z/E selectivity of 5:1.

Terminal monosubstituted and 1,2-disubstituted dienes were effective substrates in the enantioselective [4 + 1]-cycloaddition reaction, providing ee values up to 98% (Figure 3). Dienes possessing other substitution patterns, such as 1,3- and 1,4-disubstitution, generally afforded low yields. A variety of common polar functional groups are tolerated, including ethers, thioethers, esters, nitriles, boronate esters, electron-rich heterocycles, and aryl chlorides. Cycloadducts 6, 22, and 32 were obtained as crystalline solids, allowing for the unambiguous assignment of absolute stereochemistry and alkene geometry by XRD analysis.

Figure 3.

Figure 3.

Substrate Scope Studies. Isolated yields were obtained following purification by column chromatography. Reaction conditions: 1,1-dichloroalkene (0.3 mmol, 1.0 equiv), 1,3-diene (2.0 equiv), catalyst 8 (10 mol%), Zn (3.0 equiv), 5:1 MTBE:DMA (1.0 mL), 35 °C, 48 h. aReactions were run with 15 mol% catalyst loading. bReactions were run with tBunapbox (5 mol%) and Ni(dme)Cl2 (15 mol%).

A double stereodifferentiation experiment was carried out using (+)-nopadiene (29),15 which is available in enantiopure form from (–)-myrtenal. Using the S,S enantiomer of catalyst 7, cycloadduct 30 was obtained with a high dr (32:1). The R,R enantiomer of 7 reverses this selectivity, favoring the other diastereomer (31) with a more moderate dr of 4:1.

The E and Z stereoisomers of 1-phenyl-1,3-butadiene were independently reacted under the [4 + 1]-cycloaddition conditions and found to converge on the same major enantiomer of cycloadduct 32 with nearly identical ee values (97%) (Figure 4A). For the reaction with (Z)-1-phenyl-1,3-butadiene, the unreacted diene was recovered at the end of the reaction with minimal isomerization to the E form, ruling out an off-cycle Z-to-E isomerization. The stereoconvergent nature of the reaction proved to be synthetically useful in cases where 1,3-dienes are most readily available as E/Z mixtures (Figure 4B). For example, a commercial sample of 1,3-hexadiene (1.3:1 E/Z ratio) provides cycloadduct 33 in 47% yield and 93% ee. β-Ocimene (2:1 E/Z ratio) was reacted with 1,1-dichloroalkene 4 to generate 34 in 64% yield and 88% ee.

Figure 4.

Figure 4.

Stereoconvergent cycloadditions with E and Z 1,3-dienes. (a) (E)- and (Z)-1-phenyl-1,3-butadiene undergo [4 + 1]-cycloaddition to provide the same enantiomer of 32. (b) Highly enantioselective cycloadditions with dienes that are commercially available as E/Z mixtures.

Finally, we developed a computational model (BP86/6–311G(d,p) level of DFT) to rationalize the sense of stereoinduction. In our previous studies, we had identified an energetically viable stepwise pathway for the [4 + 1]-cycloaddition,7 involving migratory insertion of the diene into a Ni2(C=CR2) species followed by C–C reductive elimination. To examine the asymmetric process, we selected 1-phenyl-1,3-butadiene and 1,1-dichloroalkene 4 as model substrates. There are four possible stereoisomers of the product that can be generated in the cycloaddition: R,Z, R,E, S,Z, and S,E. Each of these products corresponds to one of four possible substrate orientations in the catalyst active site: two orientations of the diene and two orientations of the vinylidene. Additionally, the initial migratory insertion can take place on either end of the diene. Thus, eight total pathways were investigated. These results are summarized in Figure 5A, with only the lowest-energy pathway leading to each product stereoisomer shown.

Figure 5.

Figure 5.

DFT model for asymmetric induction (BP86/6–311G(d,p) level of DFT): (t-Bunapbox)Ni2Cl, 4, and 1-phenyl-1,3-butadiene. (a) Calculated lowest energy [4 + 1]-cycloaddition pathways leading to the four stereoisomers of cycloadducts 32. Relative free energies for the major pathway are shown in units of kcal/mol. (b) Migratory insertion and reductive elimination transition structures highlighting key steric interactions in pathways leading to minor stereoisomers.

The migratory insertion and C–C reductive elimination transition states are sufficiently close in energy that either step in principle could be selectivity determining. Gratifyingly, the pathway leading to the R,Z cycloadduct contained the lowest activation energies, consistent with the major stereoisomer observed experimentally. The other pathways suffer from either a high-barrier migratory insertion or reductive elimination step (Figure 5B). For example, in the pro-S,Z migratory insertion transition state, prohibitive steric interactions with the diene Ph substituent causes the oxazoline N to dissociate from Ni. In the pro-S,E and pro-R,E pathways, it is the reductive elimination barriers that are high due to interactions between the oxazoline t-Bu group and either the diene Ph or vinylidene 4-OMePh substituent.

In summary, a new class of chiral (napbox)Ni2 catalysts enabled the development of highly enantioselective intermolecular [4 + 1]-cycloadditions. Interestingly, E- and Z-dienes provide the same major enantiomer of the product and similar ee values, allowing readily available E/Z mixtures to be used as substrates. Computational models are used to rationalize the sense of enantioinduction and high preference for Z stereochemistry in the cycloadducts. The formation of the minor stereoisomers are suppressed due to unfavorable steric interactions with the catalyst t-Bu substituents in either the migratory insertion or reductive elimination transition states. Ongoing efforts are directed at developing enantioselective variants of other dinickel catalyzed processes.

Supplementary Material

Supporting Information

ACKNOWLEDGMENT

This research was supported by the NIH (R35 GM124791). X-ray diffraction data were collected using an instrument funded by the NSF (CHE-1625543). We thank Matthias Zeller for assistance with X-ray crystallography and Dr. Houng Kang for experimental assistance. C.U. acknowledges support from a Camille Dreyfus Teacher-Scholar award and a Lilly Grantee award.

Footnotes

Supporting Information

The Supporting Information is available free of charge on the ACS Publications website.

Experimental details, characterization data, and spectra (PDF)

X-ray crystallography data for 1 (CIF)

X-ray crystallography data for 2 (CIF)

X-ray crystallography data for 3 (CIF)

X-ray crystallography data for 6 (CIF)

X-ray crystallography data for 22 (CIF)

X-ray crystallography data for 32 (CIF)

X-ray crystallography data for Cynapbox (CIF)

X-ray crystallography data for (i-Prnapbox)Ni2(OAc) (CIF)

REFERENCES

  • (1).Yoon TP; Jacobsen EN “Privileged chiral catalysts” Science 2003, 299, 1691–1693. [DOI] [PubMed] [Google Scholar]
  • (2).(a) Ghosh AK; Mathivanan P; Cappiello J “C2-symmetric chiral bis(oxazoline)–metal complexes in catalytic asymmetric synthesis” Tetrahedron: Asymmetry 1998, 9, 1–45; [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Johnson JS; Evans DA “Chiral bis(oxazoline) copper(II) complexes: Versatile catalysts for enantioselective cycloaddition, aldol, michael, and carbonyl ene reactions” Acc. Chem. Res 2000, 33, 325–335; [DOI] [PubMed] [Google Scholar]; (c) Fache F; Schulz E; Tommasino ML; Lemaire M “Nitrogen-containing ligands for asymmetric homogeneous and heterogeneous catalysis” Chem. Rev 2000, 100, 2159–2232; [DOI] [PubMed] [Google Scholar]; (d) Desimoni G; Faita G; Quadrelli P “Pyridine-2,6-bis(oxazolines), helpful ligands for asymmetric catalysts” Chem. Rev 2003, 103, 3119–3154; [DOI] [PubMed] [Google Scholar]; (e) Desimoni G; Faita G; Jørgensen KA “C2-symmetric chiral bis(oxazoline) ligands in asymmetric catalysis” Chem. Rev 2006, 106, 3561–3651; [DOI] [PubMed] [Google Scholar]; (f) Rasappan R; Laventine D; Reiser O “Metal-bis(oxazoline) complexes: From coordination chemistry to asymmetric catalysis” Coord. Chem. Rev 2008, 252, 702–714; [Google Scholar]; (g) Babu SA; Krishnan KK; Ujwaldev SM; Anilkumar G “Applications of pybox complexes in asymmetric catalysis” Asian J. Org. Chem 2018, 7, 1033–1053. [Google Scholar]
  • (3).(a) Doyle MP “Perspective on dirhodium carboxamidates as catalysts” J. Org. Chem 2006, 71, 9253–9260; [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Davies HML; Manning JR “Catalytic C-H functionalization by metal carbenoid and nitrenoid insertion” Nature 2008, 451, 417–424; [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Davies HML; Morton D “Guiding principles for site selective and stereoselective intermolecular C-H functionalization by donor/acceptor rhodium carbenes” Chem. Soc. Rev 2011, 40, 1857–1869; [DOI] [PubMed] [Google Scholar]; (d) Paton RS; Brown JM “Dinuclear palladium complexes—precursors or catalysts?” Angew. Chem., Int. Ed 2012, 51, 10448–10450; [DOI] [PubMed] [Google Scholar]; (e) Kornecki KP; Berry JF; Powers DC; Ritter T In Prog. Inorg. Chem; Karlin KD, Ed.; John Wiley & Sons, Inc.: Hoboken, New Jersey, 2014; Vol. 58, p 225–302; [Google Scholar]; (f) Mankad NP “Selectivity effects in bimetallic catalysis” Chem.-Eur. J 2016, 22, 5822–5829; [DOI] [PubMed] [Google Scholar]; (g) Powers IG; Uyeda C “Metal-metal bonds in catalysis” ACS Catalysis 2017, 7, 936–958; [Google Scholar]; (h) Farley CM; Uyeda C “Organic reactions enabled by catalytically active metal-metal bonds” Trends in Chem. 2019, 1, 497–509. [Google Scholar]
  • (4).(a) Turkenburg LAM; de Worlf WH; Bickelhaupt F “1,4-addition of dichlorocarbene to 1,2-bismethylenecycloheptane” Tetrahedron Lett. 1982, 23, 769–770; [Google Scholar]; (b) Sugita H; Mizuno K; Mori T; Isagawa K; Otsuji Y “Unusual mode of addition of 1,2-alkadienylidene carbenes to 1,3-dienes: 1,4 addition to rigid and flexible 1,3-dienes” Angew. Chem., Int. Ed 1991, 30, 984–986; [Google Scholar]; (c) Boisvert L; Beaumier F; Spino C “Evidence for a concerted [4 + 1]- cycloaddition between electron-rich carbenes and electron-deficient dienes” Org. Lett 2007, 9, 5361–5363. [DOI] [PubMed] [Google Scholar]
  • (5).Examples of tandem cyclopropanation/1,3-rearrangement reactions:; (a) Danheiser RL; Martinez-Davila C; Auchus RJ; Kadonaga JT. “Stereoselective synthesis of cyclopentene derivatives from 1,3-dienes” J. Am. Chem. Soc 1981, 103, 2443–2446; [Google Scholar]; (b) Hudlicky T; Koszyk FF; Kutchan TM; Sheth JP “Cyclopentene annulation via intramolecular addition of diazoketones to 1,3-dienes. Applications to the synthesis of cyclopentanoid terpenes” J. Org. Chem 1980, 45, 5020–5027; [Google Scholar]; (c) Danheiser RL; Bronson JJ; Okano K “Carbanion-accelerated vinylcyclopropane rearrangement. Application in a general, stereocontrolled annulation approach to cyclopentene derivatives” J. Am. Chem. Soc 1985, 107, 4579–4581; [Google Scholar]; (d) Coscia RW; Lambert TH “Development of a formal [4 + 1] cycloaddition: Pd(OAc)2-catalyzed intramolecular cyclopropanation of 1,3-dienyl β-keto esters and MgI2-promoted vinylcyclopropane−cyclopentene rearrangement” J. Am. Chem. Soc 2009, 131, 2496–2498; [DOI] [PubMed] [Google Scholar]; (e) Beaumier F; Dupuis M; Spino C; Legault CY “Formal intramolecular (4 + 1)-cycloaddition of dialkoxycarbenes: Control of the stereoselectivity and a mechanistic portrait” J. Am. Chem. Soc 2012, 134, 5938–5953. [DOI] [PubMed] [Google Scholar]
  • (6).Examples of diene carbonylation reactions:; (a) Eaton BE; Rollman B; Kaduk JA. “The first catalytic iron-mediated [4 + 1] cyclopentenone assembly: Stereoselective synthesis of 2,5-dialkylidenecyclo-3-pentenones” J. Am. Chem. Soc 1992, 114, 6245–6246; [Google Scholar]; (b) Sigman MS; Eaton BE “Catalytic iron-mediated [4 + 1] cycloaddition of diallenes with carbon monoxide” J. Am. Chem. Soc 1996, 118, 11783–11788; [Google Scholar]; (c) Murakami M; Itami K; Ito Y “A study on rhodium – vinylallene complexes leading to a new reaction, rhodium-catalyzed carbonylative [4 + 1]cycloaddition” Angew. Chem., Int. Ed 1996, 34, 2691–2694; [Google Scholar]; (d) Murakami M; Itami K; Ito Y “Rhodium-catalyzed asymmetric [4 + 1] cycloaddition” J. Am. Chem. Soc 1997, 119, 2950–2951; [Google Scholar]; (e) Murakami M; Itami K; Ito Y “Catalytic asymmetric [4 + 1] cycloaddition of vinylallenes with carbon monoxide: Reversal of the induced chirality by the choice of metal” J. Am. Chem. Soc 1999, 121, 4130–4135. [Google Scholar]
  • (7).Zhou Y-Y; Uyeda C “Catalytic reductive [4 + 1]-cycloadditions of vinylidenes and dienes” Science 2019, 363, 857–862. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (8).Cramer N; Braconi E “A chiral naphthyridine diimine ligand platform enables nickel-catalyzed asymmetric alkylidenecyclopropanations” Angew. Chem., Int. Ed, 2020, 59, 1–6. [DOI] [PubMed] [Google Scholar]
  • (9).(a) Fahrni CJ “Allylic oxidation catalyzed by chiral dinuclear copper complexes” Tetrahedron 1998, 54, 5465–5470; [Google Scholar]; (b) Fahrni CJ; Pfaltz A “Synthesis of chiral C2-symmetric binucleating ligands” Helv. Chim. Acta 1998, 81, 491–506; [Google Scholar]; (c) Fahrni CJ; Pfaltz A; Neuburger M; Zehnder M “Structure and properties of transition-metal complexes with chiral C2-symmetric binucleating ligands” Helv. Chim. Acta 1998, 81, 507–524. [Google Scholar]
  • (10).Zhou Y-Y; Hartline DR; Steiman TJ; Fanwick PE; Uyeda C “Dinuclear nickel complexes in five states of oxidation using a redox-active ligand” Inorg. Chem 2014, 53, 11770–11777. [DOI] [PubMed] [Google Scholar]
  • (11).(a) Dible BR; Sigman MS; Arif AM “Oxygen-induced ligand dehydrogenation of a planar bis-μ-chloronickel(I) dimer featuring an NHC ligand” Inorg. Chem 2005, 44, 3774–3776; [DOI] [PubMed] [Google Scholar]; (b) Keen AL; Johnson SA “Nickel(0)-catalyzed isomerization of an aryne complex: Formation of a dinuclear Ni(I) complex via C−H rather than C−F bond activation” J. Am. Chem. Soc 2006, 128, 1806–1807; [DOI] [PubMed] [Google Scholar]; (c) Velian A; Lin S; Miller AJM; Day MW; Agapie T “Synthesis and C−C coupling reactivity of a dinuclear NiI−NiI complex supported by a terphenyl diphosphine” J. Am. Chem. Soc 2010, 132, 6296–6297; [DOI] [PubMed] [Google Scholar]; (d) Beck R; Johnson SA “Dinuclear Ni(I)—Ni(I) complexes with syn-facial bridging ligands from Ni(I) precursors or Ni(II)/Ni(0) comproportionation” Organometallics 2013, 32, 2944–2951; [Google Scholar]; (e) Cornella J; Gómez-Bengoa E; Martin R “Combined experimental and theoretical study on the reductive cleavage of inert C–O bonds with silanes: Ruling out a classical Ni(0)/Ni(II) catalytic couple and evidence for Ni(I) intermediates” J. Am. Chem. Soc 2013, 135, 1997–2009; [DOI] [PubMed] [Google Scholar]; (f) Wu J; Nova A; Balcells D; Brudvig GW; Dai W; Guard LM; Hazari N; Lin PH; Pokhrel R; Takase MK “Nickel (I) monomers and dimers with cyclopentadienyl and indenyl ligands” Chem.-Eur. J 2014, 20, 5327–5337. [DOI] [PubMed] [Google Scholar]
  • (a).(a) Knijnenburg Q; Hetterscheid D; Kooistra TM; Budzelaar Peter H. M. “The electronic structure of (diiminopyridine)cobalt(I) complexes” Eur. J. Inorg. Chem 2004, 2004, 1204–1211; [Google Scholar]; (b) Jones GD; Martin JL; McFarland C; Allen OR; Hall RE; Haley AD; Brandon RJ; Konovalova T; Desrochers PJ; Pulay P; Vicic DA “Ligand redox effects in the synthesis, electronic structure, and reactivity of an alkyl−alkyl cross-coupling catalyst” J. Am. Chem. Soc 2006, 128, 13175–13183; [DOI] [PubMed] [Google Scholar]; (c) Bart SC; Chłopek K; Bill E; Bouwkamp MW; Lobkovsky E; Neese F; Wieghardt K; Chirik PJ “Electronic structure of bis(imino)pyridine iron dichloride, monochloride, and neutral ligand complexes: A combined structural, spectroscopic, and computational study” J. Am. Chem. Soc 2006, 128, 13901–13912; [DOI] [PubMed] [Google Scholar]; (d) Bowman AC; Milsmann C; Hojilla Atienza CC; Lobkovsky E; Wieghardt K; Chirik PJ “Synthesis and molecular and electronic structures of reduced bis(imino)pyridine cobalt dinitrogen complexes: Ligand versus metal reduction” J. Am. Chem. Soc 2010, 132, 1676–1684. [DOI] [PubMed] [Google Scholar]
  • (13).Schley ND; Fu GC “Nickel-catalyzed negishi arylations of propargylic bromides: A mechanistic investigation” J. Am. Chem. Soc 2014, 136, 16588–16593. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (14).Potentials vs. SCE are approximated using reference potentials in MeCN:; Pavlishchuk VV; Addison AW. “Conversion constants for redox potentials measured versus different reference electrodes in acetonitrile solutions at 25°C” Inorg. Chim. Acta 2000, 298, 97–102. [Google Scholar]
  • (15).(a) Minuti L; Taticchi A; Marrocchi A; Broggi A; Gacs-Baitz E “High pressure Diels–Alder reactions of (+)-nopadiene with cycloalkenones” Tetrahedron: Asymmetry 2004, 15, 1187–1192; [Google Scholar]; (b) Minuti L; Taticchi A; Marrocchi A; Broggi A; Gacs-Baitz E “Synthesis of optically active polycyclic compounds by Diels–Alder reactions of (+)-nopadiene” Tetrahedron: Asymmetry 2004, 15, 3245–3248. [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