logo
My Activity
Recently Viewed
You have not visited any articles yet, Please visit some articles to see contents here.
CONTENT TYPES

Computational Study of the Ni-Catalyzed C–H Oxidative Cycloaddition of Aromatic Amides with Alkynes

  • Humair M. Omer
    Humair M. Omer
    Department of Chemistry, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States
  • Peng Liu*
    Peng Liu
    Department of Chemistry, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States
    Department of Chemical and Petroleum Engineering, University of Pittsburgh, Pittsburgh, Pennsylvania 15261, United States
    *E-mail: [email protected].
    More by Peng Liu
Cite this: ACS Omega 2019, 4, 3, 5209-5220
Publication Date (Web):March 13, 2019
https://doi.org/10.1021/acsomega.9b00030
Copyright © 2019 American Chemical Society
Authors ChoiceACS AuthorChoice
Article Views
1255
Altmetric
-
Citations
LEARN ABOUT THESE METRICS

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

PDF (2 MB)
Supporting Info (1)»

Abstract

The mechanism of Ni-catalyzed ortho C(sp2)–H oxidative cycloaddition of aromatic amides with internal alkynes containing 2-pyridinylmethylamine directing group was investigated using density functional theory (DFT) calculations. The C–H cleavage step proceeds via σ-complex-assisted metathesis (σ-CAM) with an alkenyl-Ni(II) complex. This is in contrast to the more common carboxylate/carbonate-assisted concerted metalation–deprotonation mechanism in related Ni-catalyzed C–H bond functionalization reactions with N,N-bidentate directing groups. In this reaction, the alkyne not only serves as the coupling partner, but also facilitates the σ-CAM C–H metalation both kinetically and thermodynamically. The subsequent functionalization of the five-membered nickelacycle proceeds via alkyne insertion into the Ni–C bond to form a seven-membered nickelacycle. This process proceeds with high levels of regioselectivity to form a C–C bond with sterically more encumbered alkyne terminus. This unusual regioselectivity is due to steric repulsions with the directing group that is coplanar with the alkyne in the migratory insertion transition state. The C–N bond reductive elimination to form the isoquinolone cycloadduct is promoted by PPh3 complexation to the Ni center and the use of flexible 2-pyridinylmethylamine directing group. The origin of the cis–trans isomerism of alkene byproduct was also explained by computations.

1. Introduction

ARTICLE SECTIONS
Jump To

Functionalization of C–H bonds to construct new C–C and C–heteroatom bonds using transition-metal catalysts is a powerful, step- and atom-economical strategy in organic synthesis.(1) First introduced by Daugulis et al.,(2)N,N-bidentate directing groups are now widely used to facilitate site-selective C–H bond functionalizations using a variety of transition metals, including Pd, Ni, Ru, Rh, Co, and Cu, as catalysts (Scheme 1).(3,4) The C–H cleavage step in these reactions is most commonly proposed to proceed via a base-promoted concerted metalation–deprotonation (CMD) mechanism.(5) Although other C–H bond metalation pathways, such as oxidative addition (OA),(6) σ-complex-assisted metathesis (σ-CAM),(7) and ligand-to-ligand hydrogen transfer,(8) are well precedented in the literature, these alternative mechanisms have not been thoroughly investigated in C–H functionalization reactions involving N,N-bidentate directing groups.(9) Because previous computational mechanistic studies of this type of reactions all focused on the CMD mechanism,(10,11) it is not clear what conditions would promote an alternative C–H bond cleavage pathway. In addition, factors that control the reactivity of these alternative C–H cleavage pathways have not been investigated computationally.

Scheme 1

Scheme 1. Mechanisms of Transition-Metal-Catalyzed C–H Functionalization of Amides Using N,N-Bidentate Directing Groups
In this manuscript, we report a computational study of the Ni-catalyzed oxidative cycloaddition of aromatic amides and alkynes via C(sp2)–H functionalization assisted by a 2-pyridinylmethylamine directing group (Scheme 2a).(12,13) In contrast to the C–H functionalization of similar aromatic and aliphatic amides with other coupling partners,(14−18) this oxidative cycloaddition is mechanistically distinct in several aspects. First, unlike the majority of Ni-catalyzed C–H functionalization reactions, which often employs a Ni(II) precatalyst and under conditions with bases (e.g., acetates, carbonates, etc.), this reaction involves a Ni(0) catalyst in the absence of a base. Chatani et al. proposed a unique C–H metalation mechanism via an alkenyl-Ni(II) complex formed from insertion of an alkyne to a nickel(II)-hydride.(12,19) It is surmised that the electron-deficient Ni(II) center would coordinate to the ortho C–H bond to form a σ-complex, which then undergoes σ-complex-assisted metathesis (σ-CAM) of the ortho C–H bond with the Ni–alkenyl bond via a four-membered cyclic transition state (path b, Scheme 1). Second, this oxidative cycloaddition forms the isoquinolone cycloadduct, while reactions of similar substrates with alkynes under different reaction conditions (e.g., in the presence of a base or with O2 as oxidant) lead to acyclic alkenylation or alkynation products (Scheme 2b,c).(20) These reactions are expected to involve similar Ni(II) metalacycle intermediates regardless of the C–H metalation mechanism (Scheme 1). Factors that determine the chemoselectivity of cyclic versus acyclic products are still not clear. Finally, this oxidative cycloaddition uses alkyne as a mild oxidant,(9,21) which is unusual in oxidative C–H functionalization reactions. In most of the oxidative C–H functionalization reactions,(22−25) the oxidant is not directly involved in the C–H metalation step. In contrast, it has been proposed that the alkyne promotes the reactivity of the C–H metalation step of this reaction.(12) Therefore, the role of the alkyne oxidant warrants an in-depth investigation.

Scheme 2

Scheme 2. Ni-Catalyzed C–H Functionalization of Amides with Alkynes Using N,N-Bidentate Directing Groups
To address these mechanistic ambiguities, we performed density functional theory calculations to investigate the reaction mechanisms of the Ni-catalyzed ortho C(sp2)–H oxidative cycloaddition reaction with alkynes. We provide a detailed analysis of the mechanisms of the C–H metalation step and the role of alkyne in facilitating the C–H metalation. The effects of phosphine additives and 2-pyridinylmethylamine directing group on reactivity and chemoselectivity for the isoquinolone cycloaddition products are carefully analyzed. In addition, the origins of the experimentally observed regioselectivity with unsymmetrically internal aryl alkynes and the formation of trans-alkene byproducts are also rationalized.

2. Results and Discussion

ARTICLE SECTIONS
Jump To

2.1. Proposed Reaction Mechanisms

The proposed mechanisms of the Ni-catalyzed ortho C(sp2)–H oxidative cycloaddition of aromatic amide 1 with model substrate 2-butyne 2 to afford the isoquinolone cycloaddition product 3 are provided in Scheme 3. With the low-valence Ni(0) precatalyst, coordination of the N,N-bidentate directing group most likely occurs via an oxidative addition of the amide N–H bond to form a Ni(II)-hydride (10) rather than through a base-promoted N–H deprotonation.(26,27) The ortho C–H bond in 10 is expected to yield an agostic interaction with the Ni due to its proximity to the electron-deficient metal center. From 10, two different C–H metalation pathways are possible. The σ-complex-assisted metathesis (σ-CAM) of the ortho C–H bond with the Ni–H bond in 10 will afford the H2-bound nickelacycle 11.(28) Alternatively, alkyne insertion into the Ni–H bond in 10 will form an alkenyl-Ni(II) complex 8,(29) which then undergoes a σ-CAM with the Ni–alkenyl bond to give the alkene-bound nickelacycle 9. Ligand exchange of the H2 or alkene in intermediate 11 or 9 with an alkyne yields complex 13. Subsequent alkyne migratory insertion into the Ni–C(sp2) bond in 13 forms a seven-membered nickelacycle, which upon C–N bond reductive elimination gives the isoquinolone product 3 and regenerates the Ni(0) catalyst. As discussed in Introduction, because such σ-CAM pathways have not been investigated computationally, a few important mechanistic questions still remain. These include: (a) the preference of the two competing σ-CAM pathways from either the nickel hydride 10 or the alkenyl nickel complex 8; (b) factors that promote the C–N bond reductive elimination to form the cycloaddition product; (c) factors that promote alkyne insertion into the nickelacycle 13 over the direct alkene migratory insertion from 9; (d) origins of regioselectivity with unsymmetrical internal alkynes; and (e) the mechanism to form the trans-alkene byproduct. These questions are discussed in detail in the following sections.

Scheme 3

Scheme 3. Proposed Mechanisms of the Ni-Catalyzed Ortho C(sp2)–H Oxidative Cycloaddition Reaction

2.2. Mechanisms of the Ortho C(sp2)–H Metalation Step and the Role of Alkyne as a Hydrogen Acceptor

The computed reaction energy profiles for steps leading to the C–H metalated nickelacycle 9 are shown in Figure 1a. Optimized geometries of select transition states and intermediates are shown in Figure 1b. The catalytic cycle begins with ligand exchange to replace the cod ligands in the Ni(cod)2 precatalyst with PPh3 and amide 1 to form complex 4. Under the experimental conditions of 10 mol % Ni(cod)2, 40 mol % PPh3 ligand, and three or more equivalents of internal alkyne, cod, PPh3, or the internal alkyne can potentially bind to the Ni center prior to the amide N–H oxidative addition.(30) The N–H oxidative addition pathways involving these ancillary ligands were considered computationally (see the Supporting Information for details). Our calculations indicate that the most favorable amide N–H oxidative addition pathway involves binding of two PPh3 ligands (TS1). Facilitated by the strong donor ligands (PPh3 and pyridine), this N–H oxidative addition process has an activation barrier of 28.0 kcal/mol with respect to 1 and Ni(cod)2. In the absence of PPh3, the N–H oxidative addition requires a barrier that is about 5 kcal/mol higher than that of TS1 (see the Supporting Information for details).

Figure 1

Figure 1. Mechanisms of the C–H metalation steps in the reaction of amide 1 with 2-butyne. (a) Computed reaction energy profiles of Ni-catalyzed ortho C(sp2)–H metalation. (b) Optimized structures of σ-C–H complexes and σ-CAM transition states with select bond distances shown in Å. All energies are with respect to the separate reactants and Ni(cod)2.

From the Ni(II)-hydride intermediate 5, two different σ-CAM pathways are possible. Dissociation of the PPh3 ligand forms σ-complex 10, which has a strong agostic interaction between the ortho C–H bond and the Ni.(31) This agostic interaction slightly elongates the ortho C–H bond to 1.11 Å compared to 1.09 Å for the same bond in 5. Formation of the σ-complex promotes the σ-bond metathesis of the ortho C–H bond with the Ni–H bond via a four-membered cyclic σ-CAM transition state TS4.(32) The four-membered cycle in TS4 is completely planar, which makes the benzene ring coplanar with the 2-pyridinylmethylamine directing group, resembling the planar geometry of the forming nickelacycle intermediate 11. TS4 is only 9.2 kcal/mol higher in energy than the σ-complex 10. However, because the formation of the Ni(II)-hydride σ-complex 10 is highly endergonic (27.1 kcal/mol with respect to amide 1 and the Ni(cod)2 catalyst), the overall activation free energy of this σ-CAM pathway is relatively high (ΔG = 36.3 kcal/mol). An alternative σ-CAM pathway from complex 5 involves a ligand exchange to replace the PPh3 ligand with an alkyne to form π-alkyne complex 6, which then undergoes facile alkyne migratory insertion (TS2) into the Ni–H bond and forms alkenyl-Ni(II) complex 8.(33) An agostic interaction with the ortho C–H bond was also observed in σ-complex 8, although the distance between the C–H bond and the Ni is slightly longer than that in 10 due to the larger size of the alkenyl group compared to the hydride ligand. From 8, the C–H metalation occurs via σ-CAM transition state TS3, which requires a low activation barrier of 10.8 kcal/mol with respect to 8 to form the alkene-bound five-membered nickelacycle 9.(34) Similar to the planar geometry of TS4, the four-membered cycle in TS3 is also coplanar with the bidentate directing group. The alkenyl group is perpendicular to the plane. Therefore, no significant steric repulsions are observed between the alkenyl group and the directing group in the σ-CAM transition state. Overall, TS3 is much more stable than TS4 because of the greater stability of the alkenyl-Ni(II) complex 8 compared to the Ni(II)-hydride 10. As such, the σ-CAM occurs via the alkenyl-Ni(II) complex 8 rather than from the Ni(II)-hydride 10, consistent with the mechanism proposed by Chatani.(12) Here, the alkyne plays an important role in promoting the C–H metalation via σ-CAM. Although the initial N–H oxidative addition to form the Ni(II)-hydride is kinetically feasible, this process is thermodynamically uphill. In the presence of alkyne, the Ni(II)-hydride intermediate is converted to a thermodynamically more stable alkenyl-Ni(II) complex via alkyne migratory insertion. Due to the thermodynamic stability of the alkenyl-Ni(II) complex, this σ-CAM pathway now requires a much lower overall activation barrier. As the H2 acceptor, the alkyne also provides thermodynamic driving force for the C–H metalation. While the formation of the cis-2-butene-bound nickelacycle 9 is exergonic, the corresponding C–H metalation process in the absence of alkyne to form the H2-bound nickelacycle 11 is endergonic by 24.8 kcal/mol with respect to the reactants (1, alkyne, and Ni(cod)2). Taken together, the alkyne serves multiple roles in promoting the C–H metalation both kinetically and thermodynamically.(35)

2.3. Mechanisms of Ni–C Insertion and C–N Bond Formation Steps and the Effects of Phosphine Additives and the 2-Pyridinylmethylamine Directing Group

We next investigated the mechanisms of the reaction of the nickelacycle intermediate 9 with alkyne to form the experimentally observed isoquinolone product and a few competing pathways to the experimentally unobserved products (Figure 2). From the alkene-bound nickelacycle 9, ligand exchange with another molecule of alkyne forms π-alkyne complex 13, which is 7.8 kcal/mol more stable than 9. The alkyne migratory insertion to the Ni–C bond in 13 forms a seven-membered nickelacycle 14 via transition state TS5. This process is more favorable than the alkene migratory insertion from 9 (via TS8). Here, TS5 is stabilized by the backdonation of the Ni d electrons to the π* of the alkyne, which is not present in TS8.(36) From 14, the C–N bond reductive elimination is promoted by coordination of a PPh3 ligand to form a four-coordinated Ni(II) complex 15. From 15, the C–N bond reductive elimination (TS6) requires only 13.7 kcal/mol to form the isoquinolone-bound Ni(0) complex 16. On the other hand, reductive elimination from complex 14 without phosphine coordination requires a much higher activation barrier of 24.2 kcal/mol (TS7) with respect to 14.

Figure 2

Figure 2. Mechanisms of the C–C and C–N bond formation steps from the nickelacycle intermediate 9. (a) Computed reaction energy profiles of C–C and C–N bond formation mechanisms. (b) Optimized structures of transition states with select bond distances shown in Å. All energies are with respect to the separate reactants and Ni(cod)2.

Experimentally, the Ni-catalyzed ortho C(sp2)–H oxidative cycloaddition reaction is the most effective with 2-pyridinylmethylamine directing group.(12) Although several different N,N-bidentate directing groups, such as 2-pyridinylmethylamine, 8-aminoquinoline, and 2-(pyridin-2-yl)isopropyl amine, have been used experimentally for related transformations, a thorough understanding of potential directing group effects in these reactions is still lacking.(10a,b) We surmised that the flexible directing group in 1 may facilitate the C–N bond reductive elimination to form the cyclic isoquinolone product. We performed calculations with model substrate 19 containing a more rigid 8-aminoquinoline moiety to explore the flexibility effect of the directing group. In reaction with 19, the C–N bond reductive elimination (TS9) requires an activation barrier of 18.9 kcal/mol with respect to the seven-membered nickelacycle 20 (Figure 3), which is more than 6 kcal/mol higher than the corresponding C–N bond reductive elimination using the 2-pyridinylmethylamine directing group (TS6). In TS6, the five-membered N,N-chelate adapts an envelope conformation in which the sp3 carbon (C1) is puckered out of plane. This allows the forming C–N bond to be coplanar with the pyridine N and the PPh3 ligand such that the Ni(II) center can adopt a less-distorted square planar geometry. On the other hand, the rigid 8-aminoquinoline directing group leads to greater distortion of the fused rings in TS9 that makes the C–N bond reductive elimination less effective.(37)

Figure 3

Figure 3. Effects of directing group on the C–N bond reductive elimination. (a) C–N reductive elimination with 2-pyridinylmethylamine directing group. (b) C–N reductive elimination with 8-aminoquinoline directing group. (c) Optimized structures of C–N reductive elimination transition states with select bond distances in angstrom (Å) and bond angles shown in degree. All energies are with respect to the phosphine-bound seven-membered nickelacycles 15 and 20.

In summary, the most favorable mechanism in the Ni-catalyzed ortho C(sp2)–H oxidative cycloaddition of aromatic amide 1 and internal alkyne 2 proceeds by oxidative addition into the amide N–H bond to form Ni(II)-hydride 5, followed by alkyne insertion to form an alkenyl-Ni(II) complex 8. The agostic interaction with the ortho C–H bond in the σ-complex 8 promotes C–H metalation via a σ-CAM mechanism to afford alkene-bound five-membered nickelacycle. Insertion of another alkyne molecule and phosphine-promoted C–N reductive elimination afford the isoquinolone product and regenerate the Ni(0) catalyst.

2.4. Origin of Regioselectivity with Unsymmetrical Internal Aryl Alkynes

We next investigated the origin of the high levels of regioselectivity in reactions with unsymmetrical internal alkynes. When phenylalkylacetylenes are used as coupling partners, this oxidative cycloaddition reaction tolerates bulky alkyl substituents, such as tert-butyl, on the alkyne (Scheme 4). Interestingly, the major regioisomeric pathway involves formation of a new C–C bond with the more sterically demanding alkyne terminus. In addition, a greater regioselectivity was observed when the size of the alkyl group increased from methyl to tert-butyl. To investigate the origin of this “counter-steric” regioselectivity, we calculated the regioselectivity-determining alkyne insertion pathways with phenylalkylacetylenes 23 and 24 (Table 1).

Scheme 4

Scheme 4. Experimentally Observed Regioselectivity with Internal Aryl Alkynes(12)
Table 1. Regioselectivity in Reactions with Unsymmetrical Alkynesa
a

Gibbs free energy and enthalpy of activation in the alkyne insertion step. All energies are in kcal/mol with respect to the alkyne-bound nickelacycle C.

In the reaction with 1-phenyl-1-propyne (23, R = Me), the alkyne insertion transition state (TS-10A) leading to the major regioisomer A is preferred by 1.6 kcal/mol, in good agreement with experimental regioselectivity (entry 1). The origin of this preference is attributed to the stabilization of the partial negative charge on the α-carbon of the forming Ni–C bond by the terminal phenyl group in TS-10A.(38) In the reaction with phenyl-t-butylacetylene (24, R = t-Bu), the major regioisomeric transition state TS-11A is stabilized by a similar electronic effect. TS-11A and TS-10A have almost identical activation barriers with respect to the corresponding π-alkyne complexes. Therefore, the reactivity of this migratory insertion is not sensitive to the steric bulk of the terminal alkyne substituent (R) adjacent to the forming C–C bond. The four-membered cyclic alkyne migration transition states TS-11A and TS-10A are not planar; the alkyl group (R) on the alkyne points out of the plane of the nickelacycle (Figure 4). As such, the steric repulsions between R and the nickelacycle in both transition states are diminished. On the other hand, in the minor regioisomeric transition state TS-11B, the bulky tert-Bu substituent is placed coplanar with the nickelacycle to achieve a square planar geometry of Ni(II). As such, substantial steric repulsions between the tert-Bu and the 2-pyridinylmethylamine directing group are observed in TS-11B. This steric effect makes the tert-Bu-substituted TS-11B 1.5 kcal/mol less stable than the Me-substituted TS-10B and thus leads to an increased regioselectivity (ΔΔG = 3.8 kcal/mol) when phenyl-t-butylacetylene (24) was used as the coupling partner.

Figure 4

Figure 4. Regioselectivity-determining insertion transition states with (a) alkyne 23 and (b) alkyne 24.

2.5. Mechanism of Cis–Trans Isomerization of the Alkene Byproduct

In the Ni-catalyzed C–H bond functionalization of amide 1 with diphenylacetylene, trans-stilbene was produced in 81% yield.(12) The alkyne-promoted σ-CAM process discussed above forms cis-alkenes rather than the trans-isomers. As such, a cis-to-trans alkene isomerization must be operational. Because nickel hydride complexes are known to catalyze alkene isomerization reactions,(27a) we surmised that Ni(II)-hydride intermediate 5 in the main catalytic cycle may serve as a catalyst to promote the cis–trans isomerization. The reaction energy profile of this pathway was calculated (Figure 5).

Figure 5

Figure 5. Reaction energy profile of the Ni(II)-hydride-catalyzed isomerization of cis-2-butene to trans-2-butene. All energies are with respect to the Ni(II)-hydride complex 5 and cis-2-butene.

Ligand exchange of PPh3 in Ni(II)-hydride 5 with cis-2-butene forms complex 25, which then undergoes alkene migratory insertion to form β-agostic alkyl-Ni(II) complex 27.(39) From 27, β-hydride elimination with a different C–H bond forms the trans-2-butene-bound Ni(II)-hydride 28, which upon ligand exchange with PPh3 extrudes the trans-2-butene byproduct. The cis–trans isomerization process in this off-cycle pathway is kinetically feasible and thermodynamically exergonic by 0.9 kcal/mol. The presence of PPh3 ligand does not significantly inhibit the reaction because the ligand exchange of PPh3 with alkene is only uphill by about 11 kcal/mol. In addition, coordination to a PPh3 ligand to form 26 does not stabilize the alkyl-Ni(II) intermediate 27. Taken together, the computed reaction energy profile indicates that the Ni(II)-hydride intermediate 5 is a competent catalyst for the isomerization of cis-alkenes to the experimentally observed trans-alkene byproducts. These results further support the formation of Ni(II)-hydride complex in the main catalytic cycle.

3. Conclusions

ARTICLE SECTIONS
Jump To

The reaction mechanism of Ni-catalyzed ortho C(sp2)–H oxidative cycloaddition of aromatic amides with internal alkynes containing 2-pyridinylmethylamine directing group was investigated using DFT calculations. The catalytic cycle begins by oxidative addition of the amide N–H bond to form a Ni(II)-hydride complex. The subsequent C–H metalation process occurs via a unique σ-complex-assisted metathesis (σ-CAM) mechanism where the internal alkyne acts as a hydrogen acceptor. This contrasts with the CMD mechanism that is usually involved in the Ni-catalyzed C–H metalation in the presence of carboxylate or carbonate bases. The alkyne plays significant roles in promoting the σ-CAM pathway both thermodynamically as a H2 acceptor and kinetically. Because the Ni(II)-hydride intermediate is thermodynamically unstable, σ-CAM from the Ni(II)-hydride requires a high overall barrier. On the other hand, in the presence of the alkyne, the Ni(II)-hydride is converted to a more stable alkenyl-Ni(II) species, which then undergoes more facile σ-CAM.
The subsequent reaction with the alkene-bound nickelacycle proceeds via an exergonic ligand exchange with another molecule of alkyne, followed by alkyne insertion to form a seven-membered nickelacycle. The insertion of the alkene is less favorable. The alkyne migratory insertion occurs via a nonplanar four-membered cyclic transition state, in which the steric repulsion about the forming C–C bond is diminished. As such, this reaction tolerates alkynes with very bulky terminus and offers high regioselectivity to form the sterically more encumbered C–C bond. The C–N bond reductive elimination of the seven-membered nickelacycle is a key step to form the cyclic isoquinolone products. This C–N bond reductive elimination is promoted by a PPh3 ligand and the flexible 2-pyridinylmethylamine directing group, which reduces the strain of the fused cyclic system in the reductive elimination transition state. The cis–trans isomerism of the alkene byproduct was also explored computationally. This process is catalyzed by a Ni(II)-hydride intermediate in the main catalytic cycle.
We expect that the mechanistic insights from this study, in particular, the unique roles of alkynes to promote the σ-CAM pathway, will aid in the development of other transition-metal-catalyzed C–H functionalization reactions with alkynes.

4. Computational Methods

ARTICLE SECTIONS
Jump To

All calculations were performed using Gaussian 09.(40) Images of the 3D structures of molecules were generated using CYLview.(41) Geometry optimizations and vibrational frequency calculations were performed using the B3LYP(42) functional in gas phase with the LANL2DZ effective core potential basis set for nickel and the 6-31G(d) basis set for other atoms. The nature of all stationary points was confirmed by the number of imaginary frequencies. All minima have zero imaginary frequency, and all transition states have only one imaginary frequency. IRC calculations were carried out for alkene and alkyne insertion transition states and for σ-CAM transition states to confirm that the transition state structures connected to the appropriate intermediates. Single-point energy calculations were carried out using the M06(43) functional and the SDD basis set for Ni and 6-311+G(d,p) for other atoms. The SMD(44) solvation model was used in the single-point energy calculations to incorporate solvent effects with toluene as the solvent. Thermal corrections to the Gibbs free energies and enthalpies were calculated using the harmonic oscillator approximation at 298.15 K.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b00030.

  • Reaction energy profiles of amide N–H oxidative addition with different ligands (Figure S1), alternative C−H metalation mechanisms with cis-2-butene (Figure S2), via σ-bond metathesis of phosphine-bound Ni(II)-hydride (Figure S3), via oxidative addition to Ni(0) (Figure S4), via deprotonation by the amide N (Figure S5), and C−N reductive elimination with model substrates (Figure S6); Cartesian coordinates of optimized geometries (PDF)

The authors declare no competing financial interest.

Terms & Conditions

Electronic Supporting Information files are available without a subscription to ACS Web Editions. The American Chemical Society holds a copyright ownership interest in any copyrightable Supporting Information. Files available from the ACS website may be downloaded for personal use only. Users are not otherwise permitted to reproduce, republish, redistribute, or sell any Supporting Information from the ACS website, either in whole or in part, in either machine-readable form or any other form without permission from the American Chemical Society. For permission to reproduce, republish and redistribute this material, requesters must process their own requests via the RightsLink permission system. Information about how to use the RightsLink permission system can be found at http://pubs.acs.org/page/copyright/permissions.html.

Acknowledgments

ARTICLE SECTIONS
Jump To

The authors acknowledge the NSF (CHE-1654122) for funding. Calculations were performed at the Center for Research Computing at the University of Pittsburgh and the Extreme Science and Engineering Discovery Environment (XSEDE) supported by NSF.

References

ARTICLE SECTIONS
Jump To

This article references 44 other publications.

  1. 1
    (a) Godula, K.; Sames, D. C-H Bond Functionalization in Complex Organic Synthesis. Science 2006, 312, 6772, DOI: 10.1126/science.1114731 .
    (b) Albrecht, M. Cyclometalation Using d-Block Transition Metals: Fundamental Aspects and Recent Trends. Chem. Rev. 2010, 110, 576623, DOI: 10.1021/cr900279a .
    (c) Balcells, D.; Clot, E.; Eisenstein, O. C–H Bond Activation in Transition Metal Species from a Computational Perspective. Chem. Rev. 2010, 110, 749823, DOI: 10.1021/cr900315k .
    (d) Yeung, C. S.; Dong, V. M. Catalytic Dehydrogenative Cross-Coupling: Forming Carbon–Carbon Bonds by Oxidizing Two Carbon–Hydrogen Bonds. Chem. Rev. 2011, 111, 12151292, DOI: 10.1021/cr100280d .
    (e) Brückl, T.; Baxter, R. D.; Ishihara, Y.; Baran, P. S. Innate and Guided C–H Functionalization Logic. Acc. Chem. Res. 2012, 45, 826839, DOI: 10.1021/ar200194b .
    (f) Pototschnig, G.; Maulide, N.; Schnürch, M. Direct Functionalization of C–H Bonds by Iron, Nickel, and Cobalt Catalysis. Chem. Eur. J. 2017, 23, 92069232, DOI: 10.1002/chem.201605657 .
    (g) Chen, Z.; Wang, B.; Zhang, J.; Yu, W.; Liu, Z.; Zhang, Y. Transition metal-catalyzed C–H bond functionalizations by the use of diverse directing groups. Org. Chem. Front. 2015, 2, 11071295, DOI: 10.1039/C5QO00004A .
    (h) Wencel-Delord, J.; Dröge, T.; Liu, F.; Glorius, F. Towards mild metal-catalyzed C–H bond activation. Chem. Soc. Rev. 2011, 40, 47404761, DOI: 10.1039/c1cs15083a .
    (i) Davies, H. M. L.; Morton, D. Recent Advances in C–H Functionalization. J. Org. Chem. 2016, 81, 343350, DOI: 10.1021/acs.joc.5b02818
  2. 2
    (a) Zaitsev, V. G.; Shabashov, D.; Daugulis, O. Highly Regioselective Arylation of sp3 C–H Bonds Catalyzed by Palladium Acetate. J. Am. Chem. Soc. 2005, 127, 1315413155, DOI: 10.1021/ja054549f .
    (b) Nadres, E. T.; Santos, G. I. F.; Shabashov, D.; Daugulis, O. Scope and Limitations of Auxiliary-Assisted, Palladium-Catalyzed Arylation and Alkylation of sp2 and sp3 C–H Bonds. J. Org. Chem. 2013, 78, 96899714, DOI: 10.1021/jo4013628 .
    (c) Daugulis, O.; Roane, J.; Tran, L. D. Bidentate, Monoanionic Auxiliary-Directed Functionalization of Carbon–Hydrogen Bonds. Acc. Chem. Res. 2015, 48, 10531064, DOI: 10.1021/ar5004626
  3. 3

    For reviews on chelation-assisted C–H functionalization, see:

    (a) Rouquet, G.; Chatani, N. Catalytic Functionalization of C(sp2)–H and C(sp3)–H Bonds by Using Bidentate Directing Groups. Angew. Chem., Int. Ed. 2013, 52, 1172611743, DOI: 10.1002/anie.201301451 .
    (b) Rit, R. K.; Yadav, M. R.; Ghosh, K.; Sahoo, A. K. Reusable directing groups [8-aminoquinoline, picolinamide, sulfoximine] in C(sp3)–H bond activation: present and future. Tetrahedron 2015, 71, 44504459, DOI: 10.1016/j.tet.2015.03.085 .
    (c) Yang, X.; Shan, G.; Wang, L.; Rao, Y. Recent advances in transition metal (Pd, Ni)-catalyzed C(sp3)–H bond activation with bidentate directing groups. Tetrahedron Lett. 2016, 57, 819836, DOI: 10.1016/j.tetlet.2016.01.009 .
    (d) He, R.; Huang, Z.-T.; Zheng, Q.-Y.; Wang, C. Isoquinoline skeleton synthesis via chelation-assisted C–H activation. Tetrahedron Lett. 2014, 55, 57055713, DOI: 10.1016/j.tetlet.2014.08.077
  4. 4

    For reviews on Ni-catalyzed C–H functionalization, see:

    (a) Tasker, S. Z.; Standley, E. A.; Jamison, T. F. Recent advances in homogeneous nickel catalysis. Nature 2014, 509, 299309, DOI: 10.1038/nature13274 .
    (b) Chatani, N. Nickel-Catalyzed C–H Bond Functionalization Utilizing an N,N′-Bidentate Directing Group. Top. Organomet. Chem. 2016, 56, 1946.
    (c) Aihara, Y.; Wuelbern, J.; Chatani, N. The Nickel(II)-Catalyzed Direct Benzylation, Allylation, Alkylation, and Methylation of CH Bonds in Aromatic Amides Containing an 8-Aminoquinoline Moiety as the Directing Group. Bull. Chem. Soc. Jpn. 2015, 88, 438446, DOI: 10.1246/bcsj.20140387 .
    (d) Castro, L. C. M.; Chatani, N. Nickel Catalysts/N,N′-Bidentate Directing Groups: An Excellent Partnership in Directed C–H Activation Reactions. Chem. Lett. 2015, 44, 410421, DOI: 10.1246/cl.150024 .
    (e) Cai, X.; Xie, B. Recent advances in nickel-catalyzed C–H bond functionalized reactions. ARKIVOC 2015, i, 184, DOI: 10.3998/ark.5550190.p008.915 .
    (f) Yamaguchi, J.; Muto, K.; Itami, K. Nickel-Catalyzed Aromatic C-H Functionalization. Top. Curr. Chem. 2016, 374, 55, DOI: 10.1007/s41061-016-0053-z .
    (g) Yamaguchi, J.; Muto, K.; Itami, K. Recent Progress in Nickel-Catalyzed Biaryl Coupling. Eur. J. Org. Chem. 2013, 1930, DOI: 10.1002/ejoc.201200914 .
    (h) Khan, M. S.; Haque, A.; Al-Suti, M. K.; Raithby, P. R. Recent advances in the application of group-10 transition metal based catalysts in C-H activation and functionalization. J. Organomet. Chem. 2015, 793, 114133, DOI: 10.1016/j.jorganchem.2015.03.023 .
    (i) Johnson, S. A. Nickel complexes for catalytic C–H bond functionalization. Dalton Trans. 2015, 44, 1090510913, DOI: 10.1039/C5DT00032G
  5. 5
    (a) Davies, D. L.; Donald, S. M. A.; Macgregor, S. A. Computational Study of the Mechanism of Cyclometalation by Palladium Acetate. J. Am. Chem. Soc. 2005, 127, 1375413755, DOI: 10.1021/ja052047w .
    (b) Lapointe, D.; Fagnou, K. Computational Study of the Mechanism of Cyclometalation by Palladium Acetate. Chem. Lett. 2010, 39, 11181126, DOI: 10.1246/cl.2010.1118 .
    (c) Lafrance, M.; Gorelsky, S. I.; Fagnou, K. High-Yielding Palladium-Catalyzed Intramolecular Alkane Arylation: Reaction Development and Mechanistic Studies. J. Am. Chem. Soc. 2007, 129, 1457014571, DOI: 10.1021/ja076588s .
    (d) Gorelsky, S. I.; Lapointe, D.; Fagnou, K. Analysis of the Concerted Metalation-Deprotonation Mechanism in Palladium-Catalyzed Direct Arylation Across a Broad Range of Aromatic Substrates. J. Am. Chem. Soc. 2008, 130, 1084810849, DOI: 10.1021/ja802533u .
    (e) Guihaumé, J.; Clot, E.; Eisenstein, O.; Perutz, R. N. Importance of palladium–carbon bond energies in direct arylation of polyfluorinated benzenes. Dalton Trans. 2010, 39, 1051010519, DOI: 10.1039/c0dt00296h .
    (f) Gorelsky, S. I.; Lapointe, D.; Fagnou, K. Analysis of the Palladium-Catalyzed (Aromatic)C–H Bond Metalation–Deprotonation Mechanism Spanning the Entire Spectrum of Arenes. J. Org. Chem. 2012, 77, 658668, DOI: 10.1021/jo202342q .
    (g) Petit, A.; Flygare, J.; Miller, A. T.; Winkel, G.; Ess, D. H. Transition-State Metal Aryl Bond Stability Determines Regioselectivity in Palladium Acetate Mediated C–H Bond Activation of Heteroarenes. Org. Lett. 2012, 14, 36803683, DOI: 10.1021/ol301521n .
    (h) Dang, Y.; Qu, S.; Nelson, J. W.; Pham, H. D.; Wang, Z.-X.; Wang, X. The Mechanism of a Ligand-Promoted C(sp3)–H Activation and Arylation Reaction via Palladium Catalysis: Theoretical Demonstration of a Pd(II)/Pd(IV) Redox Manifold. J. Am. Chem. Soc. 2015, 137, 20062014, DOI: 10.1021/ja512374g .
    (i) Dang, Y.; Deng, X.; Guo, J.; Song, C.; Hu, W.; Wang, Z.-X. Unveiling Secrets of Overcoming the “Heteroatom Problem” in Palladium-Catalyzed Aerobic C–H Functionalization of Heterocycles: A DFT Mechanistic Study. J. Am. Chem. Soc. 2016, 138, 27122723, DOI: 10.1021/jacs.5b12112 .
    (j) Davies, D. L.; Macgregor, S. A.; McMullin, C. L. Computational Studies of Carboxylate-Assisted C–H Activation and Functionalization at Group 8–10 Transition Metal Centers. Chem. Rev. 2017, 117, 86498709, DOI: 10.1021/acs.chemrev.6b00839 .
    (k) García-Cuadrado, D.; Braga, A. A. C.; Maseras, F.; Echavarren, A. M. Proton Abstraction Mechanism for the Palladium-Catalyzed Intramolecular Arylation. J. Am. Chem. Soc. 2006, 128, 10661067, DOI: 10.1021/ja056165v .
    (l) Lafrance, M.; Rowley, C. N.; Woo, T. K.; Fagnou, K. Catalytic Intermolecular Direct Arylation of Perfluorobenzenes. J. Am. Chem. Soc. 2006, 128, 87548756, DOI: 10.1021/ja062509l .
    (m) Pascual, S.; Mendoza, P.; Braga, A. A. C.; Maseras, F.; Echavarren, A. M. Bidentate phosphines as ligands in the palladium-catalyzed intramolecular arylation: the intermolecular base-assisted proton abstraction mechanism. Tetrahedron 2008, 64, 60216029, DOI: 10.1016/j.tet.2008.01.056 .
    (n) Kefalidis, C. E.; Baudoin, O.; Clot, E. DFT study of the mechanism of benzocyclobutene formation by palladium-catalysed C(sp3)–H activation: role of the nature of the base and the phosphine. Dalton Trans. 2010, 39, 1052810535, DOI: 10.1039/c0dt00578a .
    (o) Korenaga, T.; Suzuki, N.; Sueda, M.; Shimada, K. Ligand effect on direct arylation by CMD process. J. Organomet. Chem. 2015, 780, 6369, DOI: 10.1016/j.jorganchem.2014.12.017 .
    (p) Holstein, P. M.; Vogler, M.; Larini, P.; Pilet, G.; Clot, E.; Baudoin, O. Efficient Pd0-Catalyzed Asymmetric Activation of Primary and Secondary C–H Bonds Enabled by Modular Binepine Ligands and Carbonate Bases. ACS Catal. 2015, 5, 43004308, DOI: 10.1021/acscatal.5b00898
  6. 6
    Gensch, T.; Hopkinson, M. N.; Glorius, F.; Wencel-Delord, J. Mild metal-catalyzed C–H activation: examples and concepts. Chem. Soc. Rev. 2016, 45, 29002936, DOI: 10.1039/C6CS00075D
  7. 7
    Perutz, R. N.; Sabo-Etienne, S. The σ-CAM Mechanism: σ-Complexes as the Basis of σ-Bond Metathesis at Late-Transition-Metal Centers. Angew. Chem., Int. Ed. 2007, 46, 25782592, DOI: 10.1002/anie.200603224
  8. 8
    (a) Guihaumé, J.; Halbert, S.; Eisenstein, O.; Perutz, R. N. Hydrofluoroarylation of Alkynes with Ni Catalysts. C–H Activation via Ligand-to-Ligand Hydrogen Transfer, an Alternative to Oxidative Addition. Organometallics 2012, 31, 13001314, DOI: 10.1021/om2005673 .
    (b) Tang, S.; Eisenstein, O.; Nakao, Y.; Sakaki, S. Aromatic C–H σ-Bond Activation by Ni0, Pd0, and Pt0 Alkene Complexes: Concerted Oxidative Addition to Metal vs Ligand-to-Ligand H Transfer Mechanism. Organometallics 2017, 36, 27612771, DOI: 10.1021/acs.organomet.7b00256 .
    (c) Yamazaki, K.; Obata, A.; Sasagawa, A.; Ano, Y.; Chatani, N. Computational Mechanistic Study on the Nickel-Catalyzed C–H/N–H Oxidative Annulation of Aromatic Amides with Alkynes: The Role of the Nickel (0) Ate Complex. Organometallics 2019, 38, 248255, DOI: 10.1021/acs.organomet.8b00684
  9. 9
    (a) Shibata, K.; Hasegawa, N.; Fukumoto, Y.; Chatani, N. Ruthenium-Catalyzed Carbonylation of ortho C–H Bonds in Arylacetamides: C–H Bond Activation Utilizing a Bidentate-Chelation System. ChemCatChem 2012, 4, 17331736, DOI: 10.1002/cctc.201200352 .
    (b) Hasegawa, N.; Shibata, K.; Charra, V.; Inoue, S.; Fukumoto, Y.; Chatani, N. Ruthenium-catalyzed cyclocarbonylation of aliphatic amides through the regioselective activation of unactivated C(sp3)–H bonds. Tetrahedron 2013, 69, 44664472, DOI: 10.1016/j.tet.2013.02.006 .
    (c) Misal Castro, L. C.; Obata, A.; Aihara, Y.; Chatani, N. Chelation-Assisted Nickel-Catalyzed Oxidative Annulation via Double C–H Activation/Alkyne Insertion Reaction. Chem. Eur. J. 2016, 22, 13621367, DOI: 10.1002/chem.201504596 .
    (d) He, Z.; Huang, Y. Diverting C–H Annulation Pathways: Nickel-Catalyzed Dehydrogenative Homologation of Aromatic Amides. ACS Catal. 2016, 6, 78147823, DOI: 10.1021/acscatal.6b02477
  10. 10

    For computational studies of Ni-catalyzed C–H functionalization using N,N-bidentate strategy, see:

    (a) Tang, H.; Zhou, B.; Huang, X.-R.; Wang, C.; Yao, J.; Chen, H. Origins of Selective C(sp2)–H Activation Using Transition Metal Complexes with N,N-Bidentate Directing Groups: A Combined Theoretical–Experimental Study. ACS Catal. 2014, 4, 649656, DOI: 10.1021/cs401141k .
    (b) Tang, H.; Huang, X.-R.; Yao, J.; Chen, H. Understanding the Effects of Bidentate Directing Groups: A Unified Rationale for sp2 and sp3 C–H Bond Activations. J. Org. Chem. 2015, 80, 46724682, DOI: 10.1021/acs.joc.5b00580 .
    (c) Xu, Z.-Y.; Jiang, Y.-Y.; Yu, H.-Z.; Fu, Y. Mechanism of Nickel(II)-Catalyzed Oxidative C(sp2)–H/C(sp3)–H Coupling of Benzamides and Toluene Derivatives. Chem. - Asian J. 2015, 10, 24792483, DOI: 10.1002/asia.201500599 .
    (d) Singh, S.; K, S.; Sunoj, R. B. Aliphatic C(sp3)–H Bond Activation Using Nickel Catalysis: Mechanistic Insights on Regioselective Arylation. J. Org. Chem. 2017, 82, 96199626, DOI: 10.1021/acs.joc.7b01672 .
    (e) Omer, H. M.; Liu, P. Computational Study of Ni-Catalyzed C–H Functionalization: Factors That Control the Competition of Oxidative Addition and Radical Pathways. J. Am. Chem. Soc. 2017, 139, 99099920, DOI: 10.1021/jacs.7b03548 .
    (f) Haines, B. E.; Yu, J.-Q.; Musaev, D. G. The mechanism of directed Ni(II)-catalyzed C–H iodination with molecular iodine. Chem. Sci. 2018, 9, 11441154, DOI: 10.1039/C7SC04604A .
    (g) Li, Y.; Zou, L.; Bai, R.; Lan, Y. Ni(I)–Ni(III) vs. Ni(II)–Ni(IV): mechanistic study of Ni-catalyzed alkylation of benzamides with alkyl halides. Org. Chem. Front. 2018, 5, 615622, DOI: 10.1039/C7QO00850C
  11. 11

    For computational studies of other metal-catalyzed C–H functionalization using N,N-bidentate group strategy, see:

    (a) Cross, W. B.; Hope, E. G.; Lin, Y.-H.; Macgregor, S. A.; Singh, K.; Solan, G. A.; Yahyaa, N. N,N-Chelate-control on the regioselectivity in acetate-assisted C–H activation. Chem. Commun. 2013, 49, 19181920, DOI: 10.1039/c3cc38697j .
    (b) Huang, L.; Li, Q.; Wang, C.; Qi, C. Palladium(II)-Catalyzed Regioselective Arylation of Naphthylamides with Aryl Iodides Utilizing a Quinolinamide Bidentate System. J. Org. Chem. 2013, 78, 30303038, DOI: 10.1021/jo400017v .
    (c) Wei, Y.; Tang, H.; Cong, X.; Rao, B.; Wu, C.; Zeng, X. Pd(II)-Catalyzed Intermolecular Arylation of Unactivated C(sp3)–H Bonds with Aryl Bromides Enabled by 8-Aminoquinoline Auxiliary. Org. Lett. 2014, 16, 22482251, DOI: 10.1021/ol500745t .
    (d) Shan, C.; Luo, X.; Qi, X.; Liu, S.; Li, Y.; Lan, Y. Mechanism of Ruthenium-Catalyzed Direct Arylation of C–H Bonds in Aromatic Amides: A Computational Study. Organometallics 2016, 35, 14401445, DOI: 10.1021/acs.organomet.6b00064 .
    (e) Chen, C.; Hao, Y.; Zhang, T.-Y.; Pan, J.-L.; Ding, J.; Xiang, H.-Y.; Wang, M.; Ding, T.-M.; Duan, A.; Zhang, S.-Y. Computational and experimental studies on copper-mediated selective cascade C–H/N–H annulation of electron-deficient acrylamide with arynes. Chem. Commun. 2019, 55, 755758, DOI: 10.1039/C8CC08708C .
    (f) Dewyer, A. L.; Zimmerman, P. M. Simulated Mechanism for Palladium Catalyzed, Directed γ-Arylation of Piperidine. ACS Catal. 2017, 7, 54665477, DOI: 10.1021/acscatal.7b01390
  12. 12
    Shiota, H.; Ano, Y.; Aihara, Y.; Fukumoto, Y.; Chatani, N. Nickel-Catalyzed Chelation-Assisted Transformations Involving Ortho C–H Bond Activation: Regioselective Oxidative Cycloaddition of Aromatic Amides to Alkynes. J. Am. Chem. Soc. 2011, 133, 1495214955, DOI: 10.1021/ja206850s
  13. 13
    For a recent computational study on this reaction, see:Zhang, X.; Zhao, Q.; Fan, J.-Q.; Chen, D.-Z.; Liu, J.-B. A computational mechanistic study of Ni(0)-catalyzed annulation of aromatic amides with alkynes: the effects of directing groups. Org. Chem. Front. 2019, DOI: 10.1039/C8QO01310A
  14. 14

    For select examples of chelation-assisted Ni-catalyzed C–H arylation, see:

    (a) Aihara, Y.; Chatani, N. Nickel-Catalyzed Direct Arylation of C(sp3)–H Bonds in Aliphatic Amides via Bidentate-Chelation Assistance. J. Am. Chem. Soc. 2014, 136, 898901, DOI: 10.1021/ja411715v .
    (b) Yokota, A.; Aihara, Y.; Chatani, N. Nickel(II)-Catalyzed Direct Arylation of C–H Bonds in Aromatic Amides Containing an 8-Aminoquinoline Moiety as a Directing Group. J. Org. Chem. 2014, 79, 1192211932, DOI: 10.1021/jo501697n .
    (c) Li, M.; Dong, J.; Huang, X.; Li, K.; Wu, Q.; Song, F.; You, J. Nickel-catalyzed chelation-assisted direct arylation of unactivated C(sp3)–H bonds with aryl halides. Chem. Commun. 2014, 50, 39443946, DOI: 10.1039/C4CC00716F
  15. 15

    For select examples of chelation-assisted Ni-catalyzed C–H alkylation, see:

    (a) Kubo, T.; Chatani, N. Dicumyl Peroxide as a Methylating Reagent in the Ni-Catalyzed Methylation of Ortho C–H Bonds in Aromatic Amides. Org. Lett. 2016, 18, 16981701, DOI: 10.1021/acs.orglett.6b00658 .
    (b) Aihara, Y.; Chatani, N. Nickel-Catalyzed Direct Alkylation of C–H Bonds in Benzamides and Acrylamides with Functionalized Alkyl Halides via Bidentate-Chelation Assistance. J. Am. Chem. Soc. 2013, 135, 53085311, DOI: 10.1021/ja401344e .
    (c) Wu, X.; Zhao, Y.; Ge, H. Nickel-Catalyzed Site-Selective Alkylation of Unactivated C(sp3)–H Bonds. J. Am. Chem. Soc. 2014, 136, 17891792, DOI: 10.1021/ja413131m .
    (d) Song, W.; Lackner, S.; Ackermann, L. Nickel-Catalyzed C–H Alkylations: Direct Secondary Alkylations and Trifluoroethylations of Arenes. Angew. Chem., Int. Ed. 2014, 53, 24772480, DOI: 10.1002/anie.201309584
  16. 16

    For select examples of chelation-assisted Ni-catalyzed C–H benzylation, see:

    (a) Aihara, Y.; Tobisu, M.; Fukumoto, Y.; Chatani, N. Ni(II)-Catalyzed Oxidative Coupling between C(sp2)–H in Benzamides and C(sp3)–H in Toluene Derivatives. J. Am. Chem. Soc. 2014, 136, 1550915512, DOI: 10.1021/ja5095342 .
    (b) Soni, V.; Khake, S. M.; Punji, B. Nickel-Catalyzed C(sp2)–H/C(sp3)–H Oxidative Coupling of Indoles with Toluene Derivatives. ACS Catal. 2017, 7, 42024208, DOI: 10.1021/acscatal.7b01044
  17. 17

    For select examples of chelation-assisted Ni-catalyzed C–H thiolation, see:

    (a) Yan, S.-Y.; Liu, Y.-J.; Liu, B.; Liu, Y.-H.; Shi, B.-F. Nickel-catalyzed thiolation of unactivated aryl C–H bonds: efficient access to diverse aryl sulfides. Chem. Commun. 2015, 51, 40694072, DOI: 10.1039/C4CC10446C .
    (b) Yan, S.-Y.; Liu, Y.-J.; Liu, B.; Liu, Y.-H.; Zhang, Z.-Z.; Shi, B.-F. Nickel-catalyzed direct thiolation of unactivated C(sp3)–H bonds with disulfides. Chem. Commun. 2015, 51, 73417344, DOI: 10.1039/C5CC01436K .
    (c) Lin, C.; Li, D.; Wang, B.; Yao, J.; Zhang, Y. Direct ortho-Thiolation of Arenes and Alkenes by Nickel Catalysis. Org. Lett. 2015, 17, 13281331, DOI: 10.1021/acs.orglett.5b00337 .
    (d) Reddy, V. P.; Qiu, R.; Iwasaki, T.; Kambe, N. Nickel-catalyzed synthesis of diaryl sulfides and sulfones via C–H bond functionalization of arylamides. Org. Biomol. Chem. 2015, 13, 68036813, DOI: 10.1039/C5OB00149H
  18. 18
    Aihara, Y.; Chatani, N. Nickel-Catalyzed Reaction of C–H Bonds in Amides with I2: ortho-Iodination via the Cleavage of C(sp2)–H Bonds and Oxidative Cyclization to β-Lactams via the Cleavage of C(sp3)–H Bonds. ACS Catal. 2016, 6, 43234329, DOI: 10.1021/acscatal.6b00964
  19. 19
    Nakao, Y.; Morita, E.; Idei, H.; Hiyama, T. Dehydrogenative [4 + 2] Cycloaddition of Formamides with Alkynes through Double C–H Activation. J. Am. Chem. Soc. 2011, 133, 32643267, DOI: 10.1021/ja1102037
  20. 20
    (a) Lin, C.; Chen, Z.; Liu, Z.; Zhang, Y. Nickel-Catalyzed Stereoselective Alkenylation of C(sp3)–H Bonds with Terminal Alkynes. Org. Lett. 2017, 19, 850853, DOI: 10.1021/acs.orglett.6b03856 .
    (b) Li, M.; Yang, Y.; Zhou, D.; Wan, D.; You, J. Nickel-Catalyzed Addition-Type Alkenylation of Unactivated, Aliphatic C–H Bonds with Alkynes: A Concise Route to Polysubstituted γ-Butyrolactones. Org. Lett. 2015, 17, 25462549, DOI: 10.1021/acs.orglett.5b01128 .
    (c) Maity, S.; Agasti, S.; Earsad, A. M.; Hazra, A.; Maiti, D. Nickel-Catalyzed Insertion of Alkynes and Electron-Deficient Olefins into Unactivated sp3 C–H Bonds. Chem. - Eur. J. 2015, 21, 1132011324, DOI: 10.1002/chem.201501962 .
    (d) Liu, Y.-H.; Liu, Y.-J.; Yan, S.-Y.; Shi, B.-F. Ni(II)-catalyzed dehydrative alkynylation of unactivated (hetero)aryl C–H bonds using oxygen: a user-friendly approach. Chem. Commun. 2015, 51, 1165011653, DOI: 10.1039/C5CC03729H
  21. 21

    For C–H functionalization reactions where alkyne or alkene act as hydrogen acceptor, see:

    (a) Inoue, S.; Shiota, H.; Fukumoto, Y.; Chatani, N. Ruthenium-Catalyzed Carbonylation at Ortho C–H Bonds in Aromatic Amides Leading to Phthalimides: C–H Bond Activation Utilizing a Bidentate System. J. Am. Chem. Soc. 2009, 131, 68986899, DOI: 10.1021/ja900046z .
    (b) Hasegawa, N.; Charra, V.; Inoue, S.; Fukumoto, Y.; Chatani, N. Highly Regioselective Carbonylation of Unactivated C(sp3)–H Bonds by Ruthenium Carbonyl. J. Am. Chem. Soc. 2011, 133, 80708073, DOI: 10.1021/ja2001709 .
    (c) Song, W.; Ackermann, L. Nickel-catalyzed alkyne annulation by anilines: versatile indole synthesis by C–H/N–H functionalization. Chem. Commun. 2013, 49, 66386640, DOI: 10.1039/c3cc43915a .
    (d) Obata, A.; Ano, Y.; Chatani, N. Nickel-catalyzed C–H/N–H annulation of aromatic amides with alkynes in the absence of a specific chelation system. Chem. Sci. 2017, 8, 66506655, DOI: 10.1039/C7SC01750B
  22. 22

    For Rh-catalyzed C–H functionalization with internal alkyne, see:

    (a) Mochida, S.; Nobuyoshi, U.; Koji, H.; Tetsuya, S.; Masahiro, M. Rhodium-catalyzed Oxidative Coupling/Cyclization of Benzamides with Alkynes via C–H Bond Cleavage. Chem. Lett. 2010, 39, 744746, DOI: 10.1246/cl.2010.744 .
    (b) Hyster, T. K.; Rovis, T. Rhodium-Catalyzed Oxidative Cycloaddition of Benzamides and Alkynes via C–H/N–H Activation. J. Am. Chem. Soc. 2010, 132, 1056510569, DOI: 10.1021/ja103776u .
    (c) Guimond, N.; Gorelsky, S. I.; Fagnou, K. Rhodium(III)-Catalyzed Heterocycle Synthesis Using an Internal Oxidant: Improved Reactivity and Mechanistic Studies. J. Am. Chem. Soc. 2011, 133, 64496457, DOI: 10.1021/ja201143v .
    (d) Hyster, T. K.; Rovis, T. An improved catalyst architecture for rhodium(III) catalyzed C–H activation and its application to pyridone synthesis. Chem. Sci. 2011, 2, 16061610, DOI: 10.1039/C1SC00235J .
    (e) Shan, G.; Flegel, J.; Li, H.; Merten, C.; Ziegler, S.; Antonchick, A. P.; Waldmann, H. C–H Bond Activation for the Synthesis of Heterocyclic Atropisomers Yields Hedgehog Pathway Inhibitors. Angew. Chem., Int. Ed. 2018, 57, 1425014254, DOI: 10.1002/anie.201809680 .
    (f) Shibata, K.; Natsui, S.; Chatani, N. Rhodium-Catalyzed Alkenylation of C–H Bonds in Aromatic Amides with Alkynes. Org. Lett. 2017, 19, 22342237, DOI: 10.1021/acs.orglett.7b00709
  23. 23

    For Ru-catalyzed C–H functionalization with internal alkynes, see:

    (a) Allu, S.; Swamy, K. C. K. Ruthenium-Catalyzed Synthesis of Isoquinolones with 8-Aminoquinoline as a Bidentate Directing Group in C–H Functionalization. J. Org. Chem. 2014, 79, 39633972, DOI: 10.1021/jo500424p .
    (b) Kaishap, P. P.; Duarah, G.; Chetiab, D.; Gogoi, S. Ru(II)-Catalyzed annulation of benzamidines and alkynes by C–H/N–H activation: a facile synthesis of 1-aminoisoquinolines. Org. Biomol. Chem. 2017, 15, 34913498, DOI: 10.1039/C7OB00389G
  24. 24

    For Pd-catalyzed C–H functionalization with internal alkynes, see:

    (a) Shu, Z.; Guo, Y.; Li, W.; Wang, B. Pd/C-catalyzed synthesis of N-aryl and N-alkyl isoquinolones via CH/NH activation. Catal. Today 2017, 297, 292297, DOI: 10.1016/j.cattod.2017.02.005 .
    (b) Sharma, N.; Saha, R.; Parveen, N.; Sekar, G. Palladium-Nanoparticles-Catalyzed Oxidative Annulation of Benzamides with Alkynes for the Synthesis of Isoquinolones. Adv. Synth. Catal. 2017, 359, 19471958, DOI: 10.1002/adsc.201601137
  25. 25

    For Co-catalyzed C–H functionalization with internal alkyne, see:

    (a) Grigorjeva, L.; Daugulis, O. Cobalt-catalyzed, aminoquinoline-directed C(sp2)-H bond alkenylation by alkynes. Angew. Chem., Int. Ed. 2014, 53, 1020910212, DOI: 10.1002/anie.201404579 .
    (b) Nguyen, T. T.; Grigorjeva, L.; Daugulis, O. Cobalt-Catalyzed, Aminoquinoline-Directed Functionalization of Phosphinic Amide sp2 C–H Bonds. ACS Catal. 2016, 6, 551554, DOI: 10.1021/acscatal.5b02391 .
    (c) Manoharan, R.; Jeganmohan, M. Cobalt-catalyzed cyclization of benzamides with alkynes: a facile route to isoquinolones with hydrogen evolution. Org. Biomol. Chem. 2018, 16, 83848389, DOI: 10.1039/C8OB01924J .
    (d) Zhai, S.; Qiu, S.; Chen, X.; Wu, J.; Zhao, H.; Tao, C.; Li, Y.; Cheng, B.; Wang, H.; Zhai, H. 2-(1-Methylhydrazinyl)pyridine as a reductively removable directing group in a cobalt-catalyzed C(sp2)–H bond alkenylation/annulation cascade. Chem. Commun. 2018, 54, 98101, DOI: 10.1039/C7CC08533H
  26. 26

    In the presence of a strong base, such as t-BuOK, N–H deprotonation would form a nickel(0) ate complex. A recent computational study indicates the subsequent C–H metalation from the nickel(0) ate complex occurs via a ligand-to-ligand hydrogen transfer mechanism. See ref (8c).

  27. 27

    For select reviews and crystal structures of Ni-hydride complexes, see:

    (a) Eberhardt, N. A.; Guan, H. Nickel Hydride Complexes. Chem. Rev. 2016, 116, 83738426, DOI: 10.1021/acs.chemrev.6b00259 .
    (b) Matson, E. M.; Martinez, G. E.; Ibrahim, A. D.; Jackson, B. J.; Bertke, J. A.; Fout, A. R. Nickel(II) Pincer Carbene Complexes: Oxidative Addition of an Aryl C–H Bond to Form a Ni(II) Hydride. Organometallics 2015, 34, 399407, DOI: 10.1021/om5007177 .
    (c) Clement, N. D.; Cavell, K. J.; Jones, C.; Elsevier, C. J. Oxidative Addition of Imidazolium Salts to Ni0 and Pd0: Synthesis and Structural Characterization of Unusually Stable Metal–Hydride Complexes. Angew. Chem., Int. Ed. 2004, 43, 1277
  28. 28
    Crabtree, R. H. Dihydrogen Complexation. Chem. Rev. 2016, 116, 87508769, DOI: 10.1021/acs.chemrev.6b00037
  29. 29
    (a) She, L.; Li, X.; Sun, H.; Ding, J.; Frey, M.; Klein, H.-F. Insertion of Alkynes into Ni–H Bonds: Synthesis of Novel Vinyl Nickel(II) and Dinuclear Vinyl Nickel(II) Complexes Containing a [P, S]-Ligand. Organometallics 2007, 26, 566570, DOI: 10.1021/om0606340 .
    (b) Zhou, H.; Sun, H.; Zheng, T.; Zhang, S.; Li, X. Synthesis of Vinylnickel and Nickelacyclopropane Complexes Containing a Chelate [P,Se]-Ligand. Eur. J. Inorg. Chem. 2015, 31393145, DOI: 10.1002/ejic.201500293 .
    (c) Xue, B.; Sun, H.; Ren, S.; Li, X.; Fuhr, O. Vinyl/Phenyl Exchange Reaction within Vinyl Nickel Complexes Bearing Chelate [P, S]-Ligands. Organometallics 2017, 36, 42464255, DOI: 10.1021/acs.organomet.7b00671
  30. 30

    An alternative pathway involving oxidative addition of ortho C(sp2)–H bond with the Ni(0) catalyst was considered computationally. The barrier of the C–H oxidative addition is 19.5 kcal/mol with respect to amide 1 and Ni(cod)2. However, the subsequent steps in this pathway were highly disfavored. Both alkyne insertion into Ni–H bond and N–H deprotonation to form the five-membered Ni(II) metallacycle 9 require very high activation barriers of 44.5 and 67.2 kcal/mol with respect to 1 and Ni(cod)2, respectively. Based on these results, this oxidative addition pathway was ruled out (see Supporting Information for more details).

  31. 31
    (a) Brookhart, M.; Green, M. L. H.; Parkin, G. Agostic interactions in transition metal compounds. Proc. Natl. Acad. Sci. USA 2007, 104, 69086914, DOI: 10.1073/pnas.0610747104 .
    (b) Jongbloed, L. S.; García-Lopez, D.; Heck, R. V.; Siegler, M. A.; Carbo, J. J.; Vlugt, J. I. V. D. Arene C(sp2)-H Metalation at NiII Modeled with a Reactive PONCPh Ligand. Inorg. Chem. 2016, 55, 80418047, DOI: 10.1021/acs.inorgchem.6b01162 .
    (c) Lein, M. Characterization of agostic interactions in theory and computation. Coord. Chem. Rev. 2009, 253, 625634, DOI: 10.1016/j.ccr.2008.07.007 .
    (d) Etienne, M.; Weller, A. S. Intramolecular C–C agostic complexes: C–C sigma interactions by another name. Chem. Soc. Rev. 2014, 43, 242259, DOI: 10.1039/C3CS60295H .
    (e) Beattie, D. D.; Bowes, E. G.; Drover, M. W.; Love, J. A.; Schafer, L. L. Oxidation State Dependent Coordination Modes: Accessing an Amidate-Supported Nickel(I) δ-bis(C–H) Agostic Complex. Angew. Chem. 2016, 128, 1348413489, DOI: 10.1002/ange.201607243 .
    (f) Crabtree, R. H. Sigma Bonds as Ligand Donor Groups in Transition Metal Complexes. The Chemical Bond III; Mingos, D.M.P., Ed.; Springer International Publishing: Switzerland, 2015; Vol. 171, pp 6378.
  32. 32

    An alternatively pathway to form intermediate 11 from phosphine-bound Ni(II)-hydride complex 5 via σ-bond metathesis followed by PPh3 decomplexation requires a much higher barrier (66.0 kcal/mol with respect to aromatic amide 1 and Ni(cod)2 catalyst). This very high barrier is due to the absence of an agostic interaction and unfavorable steric effects of the additional PPh3 ligand.

  33. 33

    PPh3 coordination to intermediate 8 forms an off-cycle phosphine-bound alkenyl-Ni(II) complex 7. Complex 7 is 11.2 kcal/mol more stable than 8.

  34. 34
    (a) Another mechanism for C–H metalation involves σ-bond metathesis of the Ni–N bond in intermediate 8 with the ortho C–H bond to form a five-membered alkenyl-nickelacycle (see Supporting Information for details). This process requires an activation barrier of 39 kcal/mol with respect to the separate reactants and Ni(cod)2, and thus can be ruled out.
    (b) In our calculations, we could not locate neither the transition state structure for the oxidative addition of the ortho C(sp2)–H bond from alkenyl-Ni(II) complex 8 nor the resulting Ni(IV)-hydride complex. All attempts to locate these structures resulted in TS3, 8, or 9. Intrinsic reaction coordinate (IRC) calculations were carried out for TS3 to confirm that it connects to complexes 8 and 9.
  35. 35

    We also computationally considered the use of cis-2-butene rather than 2-butyne as the H2 acceptor to promote the C–H metalation. In this alternatively pathway, the barrier of σ-bond metathesis is 34.8 kcal/mol with respect to the separate reactants and Ni(cod)2, and thus this pathway is less favorable than that using alkyne as H2 acceptor (see Supporting Information for details).

  36. 36
    McCarren, P. R.; Liu, P.; Cheong, P. H.-Y.; Jamison, T. F.; Houk, K. N. Mechanism and Transition-State Structures for Nickel-Catalyzed Reductive Alkyne–Aldehyde Coupling Reactions. J. Am. Chem. Soc. 2009, 131, 66546655, DOI: 10.1021/ja900701g
  37. 37

    In addition, the N in 2-pyridinylmethylamine is a better donor that electronically promotes the reductive elimination via TS6.

  38. 38
    Liang, L.-C.; Pin-Shu Chien, P.-S.; Lee, P.-Y. Phosphorus and Olefin Substituent Effects on the Insertion Chemistry of Nickel(II) Hydride Complexes Containing Amido Diphosphine Ligands. Organometallics 2008, 27, 30823093, DOI: 10.1021/om701294a
  39. 39
    (a) Kogut, E.; Zeller, A.; Warren, T. H.; Strassner, T. Structure and Dynamics of Neutral β-H Agostic Nickel Alkyls: A Combined Experimental and Theoretical Study. J. Am. Chem. Soc. 2004, 126, 1198411994, DOI: 10.1021/ja0477221 .
    (b) Scherer, W.; Herz, V.; Brück, A.; Hauf, C.; Reiner, F.; Altmannshofer, S.; Leusser, D.; Stalke, D. The Nature of β-Agostic Bonding in Late-Transition-Metal Alkyl Complexes. Angew. Chem., Int. Ed. 2011, 50, 28452849, DOI: 10.1002/anie.201006065
  40. 40
    Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision D.01; Gaussian, Inc.: Wallingford, CT, 2010.
  41. 41
    Legault, C. Y. CYLview, version 1.0b; Universitede Sherbrooke: Quebec, Canada, 2009; http://www.cylview.org.
  42. 42
    (a) Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 56485652, DOI: 10.1063/1.464913 .
    (b) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785789, DOI: 10.1103/PhysRevB.37.785
  43. 43
    (a) Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theoretical Chemistry Accounts; Springer, 2008; Vol. 120, pp 215241.
    (b) Zhao, Y.; Truhlar, D. G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem. Res. 2008, 41, 157167, DOI: 10.1021/ar700111a
  44. 44
    Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 63786396, DOI: 10.1021/jp810292n

Cited By


This article is cited by 4 publications.

  1. Chaoshen Zhang, Ruihua Zhao, Wasihun Menberu Dagnaw, Zheyuan Liu, Yu Lu, Zhi-Xiang Wang. Density Functional Theory Mechanistic Insight into the Base-Free Nickel-Catalyzed Suzuki–Miyaura Cross-Coupling of Acid Fluoride: Concerted versus Stepwise Transmetalation. The Journal of Organic Chemistry 2019, 84 (21) , 13983-13991. DOI: 10.1021/acs.joc.9b02154.
  2. Yan‐Hua Liu, Yu‐Nong Xia, Bing‐Feng Shi. Ni‐Catalyzed Chelation‐Assisted Direct Functionalization of Inert C—H Bonds. Chinese Journal of Chemistry 2020, 37 DOI: 10.1002/cjoc.201900468.
  3. Andreas Berkefeld, Markus Fröhlich, Mike Kordan, Gerald Hörner, Hartmut Schubert. Selective metalation of phenol-type proligands for preparative organometallic chemistry. Chemical Communications 2020, 56 (28) , 3987-3990. DOI: 10.1039/D0CC01254H.
  4. Shrikant M. Khake, Naoto Chatani. Chelation-Assisted Nickel-Catalyzed C−H Functionalizations. Trends in Chemistry 2019, 1 (5) , 524-539. DOI: 10.1016/j.trechm.2019.06.002.
  • Abstract

    Scheme 1

    Scheme 1. Mechanisms of Transition-Metal-Catalyzed C–H Functionalization of Amides Using N,N-Bidentate Directing Groups

    Scheme 2

    Scheme 2. Ni-Catalyzed C–H Functionalization of Amides with Alkynes Using N,N-Bidentate Directing Groups

    Scheme 3

    Scheme 3. Proposed Mechanisms of the Ni-Catalyzed Ortho C(sp2)–H Oxidative Cycloaddition Reaction

    Figure 1

    Figure 1. Mechanisms of the C–H metalation steps in the reaction of amide 1 with 2-butyne. (a) Computed reaction energy profiles of Ni-catalyzed ortho C(sp2)–H metalation. (b) Optimized structures of σ-C–H complexes and σ-CAM transition states with select bond distances shown in Å. All energies are with respect to the separate reactants and Ni(cod)2.

    Figure 2

    Figure 2. Mechanisms of the C–C and C–N bond formation steps from the nickelacycle intermediate 9. (a) Computed reaction energy profiles of C–C and C–N bond formation mechanisms. (b) Optimized structures of transition states with select bond distances shown in Å. All energies are with respect to the separate reactants and Ni(cod)2.

    Figure 3

    Figure 3. Effects of directing group on the C–N bond reductive elimination. (a) C–N reductive elimination with 2-pyridinylmethylamine directing group. (b) C–N reductive elimination with 8-aminoquinoline directing group. (c) Optimized structures of C–N reductive elimination transition states with select bond distances in angstrom (Å) and bond angles shown in degree. All energies are with respect to the phosphine-bound seven-membered nickelacycles 15 and 20.

    Scheme 4

    Scheme 4. Experimentally Observed Regioselectivity with Internal Aryl Alkynes(12)

    Figure 4

    Figure 4. Regioselectivity-determining insertion transition states with (a) alkyne 23 and (b) alkyne 24.

    Figure 5

    Figure 5. Reaction energy profile of the Ni(II)-hydride-catalyzed isomerization of cis-2-butene to trans-2-butene. All energies are with respect to the Ni(II)-hydride complex 5 and cis-2-butene.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 44 other publications.

    1. 1
      (a) Godula, K.; Sames, D. C-H Bond Functionalization in Complex Organic Synthesis. Science 2006, 312, 6772, DOI: 10.1126/science.1114731 .
      (b) Albrecht, M. Cyclometalation Using d-Block Transition Metals: Fundamental Aspects and Recent Trends. Chem. Rev. 2010, 110, 576623, DOI: 10.1021/cr900279a .
      (c) Balcells, D.; Clot, E.; Eisenstein, O. C–H Bond Activation in Transition Metal Species from a Computational Perspective. Chem. Rev. 2010, 110, 749823, DOI: 10.1021/cr900315k .
      (d) Yeung, C. S.; Dong, V. M. Catalytic Dehydrogenative Cross-Coupling: Forming Carbon–Carbon Bonds by Oxidizing Two Carbon–Hydrogen Bonds. Chem. Rev. 2011, 111, 12151292, DOI: 10.1021/cr100280d .
      (e) Brückl, T.; Baxter, R. D.; Ishihara, Y.; Baran, P. S. Innate and Guided C–H Functionalization Logic. Acc. Chem. Res. 2012, 45, 826839, DOI: 10.1021/ar200194b .
      (f) Pototschnig, G.; Maulide, N.; Schnürch, M. Direct Functionalization of C–H Bonds by Iron, Nickel, and Cobalt Catalysis. Chem. Eur. J. 2017, 23, 92069232, DOI: 10.1002/chem.201605657 .
      (g) Chen, Z.; Wang, B.; Zhang, J.; Yu, W.; Liu, Z.; Zhang, Y. Transition metal-catalyzed C–H bond functionalizations by the use of diverse directing groups. Org. Chem. Front. 2015, 2, 11071295, DOI: 10.1039/C5QO00004A .
      (h) Wencel-Delord, J.; Dröge, T.; Liu, F.; Glorius, F. Towards mild metal-catalyzed C–H bond activation. Chem. Soc. Rev. 2011, 40, 47404761, DOI: 10.1039/c1cs15083a .
      (i) Davies, H. M. L.; Morton, D. Recent Advances in C–H Functionalization. J. Org. Chem. 2016, 81, 343350, DOI: 10.1021/acs.joc.5b02818
    2. 2
      (a) Zaitsev, V. G.; Shabashov, D.; Daugulis, O. Highly Regioselective Arylation of sp3 C–H Bonds Catalyzed by Palladium Acetate. J. Am. Chem. Soc. 2005, 127, 1315413155, DOI: 10.1021/ja054549f .
      (b) Nadres, E. T.; Santos, G. I. F.; Shabashov, D.; Daugulis, O. Scope and Limitations of Auxiliary-Assisted, Palladium-Catalyzed Arylation and Alkylation of sp2 and sp3 C–H Bonds. J. Org. Chem. 2013, 78, 96899714, DOI: 10.1021/jo4013628 .
      (c) Daugulis, O.; Roane, J.; Tran, L. D. Bidentate, Monoanionic Auxiliary-Directed Functionalization of Carbon–Hydrogen Bonds. Acc. Chem. Res. 2015, 48, 10531064, DOI: 10.1021/ar5004626
    3. 3

      For reviews on chelation-assisted C–H functionalization, see:

      (a) Rouquet, G.; Chatani, N. Catalytic Functionalization of C(sp2)–H and C(sp3)–H Bonds by Using Bidentate Directing Groups. Angew. Chem., Int. Ed. 2013, 52, 1172611743, DOI: 10.1002/anie.201301451 .
      (b) Rit, R. K.; Yadav, M. R.; Ghosh, K.; Sahoo, A. K. Reusable directing groups [8-aminoquinoline, picolinamide, sulfoximine] in C(sp3)–H bond activation: present and future. Tetrahedron 2015, 71, 44504459, DOI: 10.1016/j.tet.2015.03.085 .
      (c) Yang, X.; Shan, G.; Wang, L.; Rao, Y. Recent advances in transition metal (Pd, Ni)-catalyzed C(sp3)–H bond activation with bidentate directing groups. Tetrahedron Lett. 2016, 57, 819836, DOI: 10.1016/j.tetlet.2016.01.009 .
      (d) He, R.; Huang, Z.-T.; Zheng, Q.-Y.; Wang, C. Isoquinoline skeleton synthesis via chelation-assisted C–H activation. Tetrahedron Lett. 2014, 55, 57055713, DOI: 10.1016/j.tetlet.2014.08.077
    4. 4

      For reviews on Ni-catalyzed C–H functionalization, see:

      (a) Tasker, S. Z.; Standley, E. A.; Jamison, T. F. Recent advances in homogeneous nickel catalysis. Nature 2014, 509, 299309, DOI: 10.1038/nature13274 .
      (b) Chatani, N. Nickel-Catalyzed C–H Bond Functionalization Utilizing an N,N′-Bidentate Directing Group. Top. Organomet. Chem. 2016, 56, 1946.
      (c) Aihara, Y.; Wuelbern, J.; Chatani, N. The Nickel(II)-Catalyzed Direct Benzylation, Allylation, Alkylation, and Methylation of CH Bonds in Aromatic Amides Containing an 8-Aminoquinoline Moiety as the Directing Group. Bull. Chem. Soc. Jpn. 2015, 88, 438446, DOI: 10.1246/bcsj.20140387 .
      (d) Castro, L. C. M.; Chatani, N. Nickel Catalysts/N,N′-Bidentate Directing Groups: An Excellent Partnership in Directed C–H Activation Reactions. Chem. Lett. 2015, 44, 410421, DOI: 10.1246/cl.150024 .
      (e) Cai, X.; Xie, B. Recent advances in nickel-catalyzed C–H bond functionalized reactions. ARKIVOC 2015, i, 184, DOI: 10.3998/ark.5550190.p008.915 .
      (f) Yamaguchi, J.; Muto, K.; Itami, K. Nickel-Catalyzed Aromatic C-H Functionalization. Top. Curr. Chem. 2016, 374, 55, DOI: 10.1007/s41061-016-0053-z .
      (g) Yamaguchi, J.; Muto, K.; Itami, K. Recent Progress in Nickel-Catalyzed Biaryl Coupling. Eur. J. Org. Chem. 2013, 1930, DOI: 10.1002/ejoc.201200914 .
      (h) Khan, M. S.; Haque, A.; Al-Suti, M. K.; Raithby, P. R. Recent advances in the application of group-10 transition metal based catalysts in C-H activation and functionalization. J. Organomet. Chem. 2015, 793, 114133, DOI: 10.1016/j.jorganchem.2015.03.023 .
      (i) Johnson, S. A. Nickel complexes for catalytic C–H bond functionalization. Dalton Trans. 2015, 44, 1090510913, DOI: 10.1039/C5DT00032G
    5. 5
      (a) Davies, D. L.; Donald, S. M. A.; Macgregor, S. A. Computational Study of the Mechanism of Cyclometalation by Palladium Acetate. J. Am. Chem. Soc. 2005, 127, 1375413755, DOI: 10.1021/ja052047w .
      (b) Lapointe, D.; Fagnou, K. Computational Study of the Mechanism of Cyclometalation by Palladium Acetate. Chem. Lett. 2010, 39, 11181126, DOI: 10.1246/cl.2010.1118 .
      (c) Lafrance, M.; Gorelsky, S. I.; Fagnou, K. High-Yielding Palladium-Catalyzed Intramolecular Alkane Arylation: Reaction Development and Mechanistic Studies. J. Am. Chem. Soc. 2007, 129, 1457014571, DOI: 10.1021/ja076588s .
      (d) Gorelsky, S. I.; Lapointe, D.; Fagnou, K. Analysis of the Concerted Metalation-Deprotonation Mechanism in Palladium-Catalyzed Direct Arylation Across a Broad Range of Aromatic Substrates. J. Am. Chem. Soc. 2008, 130, 1084810849, DOI: 10.1021/ja802533u .
      (e) Guihaumé, J.; Clot, E.; Eisenstein, O.; Perutz, R. N. Importance of palladium–carbon bond energies in direct arylation of polyfluorinated benzenes. Dalton Trans. 2010, 39, 1051010519, DOI: 10.1039/c0dt00296h .
      (f) Gorelsky, S. I.; Lapointe, D.; Fagnou, K. Analysis of the Palladium-Catalyzed (Aromatic)C–H Bond Metalation–Deprotonation Mechanism Spanning the Entire Spectrum of Arenes. J. Org. Chem. 2012, 77, 658668, DOI: 10.1021/jo202342q .
      (g) Petit, A.; Flygare, J.; Miller, A. T.; Winkel, G.; Ess, D. H. Transition-State Metal Aryl Bond Stability Determines Regioselectivity in Palladium Acetate Mediated C–H Bond Activation of Heteroarenes. Org. Lett. 2012, 14, 36803683, DOI: 10.1021/ol301521n .
      (h) Dang, Y.; Qu, S.; Nelson, J. W.; Pham, H. D.; Wang, Z.-X.; Wang, X. The Mechanism of a Ligand-Promoted C(sp3)–H Activation and Arylation Reaction via Palladium Catalysis: Theoretical Demonstration of a Pd(II)/Pd(IV) Redox Manifold. J. Am. Chem. Soc. 2015, 137, 20062014, DOI: 10.1021/ja512374g .
      (i) Dang, Y.; Deng, X.; Guo, J.; Song, C.; Hu, W.; Wang, Z.-X. Unveiling Secrets of Overcoming the “Heteroatom Problem” in Palladium-Catalyzed Aerobic C–H Functionalization of Heterocycles: A DFT Mechanistic Study. J. Am. Chem. Soc. 2016, 138, 27122723, DOI: 10.1021/jacs.5b12112 .
      (j) Davies, D. L.; Macgregor, S. A.; McMullin, C. L. Computational Studies of Carboxylate-Assisted C–H Activation and Functionalization at Group 8–10 Transition Metal Centers. Chem. Rev. 2017, 117, 86498709, DOI: 10.1021/acs.chemrev.6b00839 .
      (k) García-Cuadrado, D.; Braga, A. A. C.; Maseras, F.; Echavarren, A. M. Proton Abstraction Mechanism for the Palladium-Catalyzed Intramolecular Arylation. J. Am. Chem. Soc. 2006, 128, 10661067, DOI: 10.1021/ja056165v .
      (l) Lafrance, M.; Rowley, C. N.; Woo, T. K.; Fagnou, K. Catalytic Intermolecular Direct Arylation of Perfluorobenzenes. J. Am. Chem. Soc. 2006, 128, 87548756, DOI: 10.1021/ja062509l .
      (m) Pascual, S.; Mendoza, P.; Braga, A. A. C.; Maseras, F.; Echavarren, A. M. Bidentate phosphines as ligands in the palladium-catalyzed intramolecular arylation: the intermolecular base-assisted proton abstraction mechanism. Tetrahedron 2008, 64, 60216029, DOI: 10.1016/j.tet.2008.01.056 .
      (n) Kefalidis, C. E.; Baudoin, O.; Clot, E. DFT study of the mechanism of benzocyclobutene formation by palladium-catalysed C(sp3)–H activation: role of the nature of the base and the phosphine. Dalton Trans. 2010, 39, 1052810535, DOI: 10.1039/c0dt00578a .
      (o) Korenaga, T.; Suzuki, N.; Sueda, M.; Shimada, K. Ligand effect on direct arylation by CMD process. J. Organomet. Chem. 2015, 780, 6369, DOI: 10.1016/j.jorganchem.2014.12.017 .
      (p) Holstein, P. M.; Vogler, M.; Larini, P.; Pilet, G.; Clot, E.; Baudoin, O. Efficient Pd0-Catalyzed Asymmetric Activation of Primary and Secondary C–H Bonds Enabled by Modular Binepine Ligands and Carbonate Bases. ACS Catal. 2015, 5, 43004308, DOI: 10.1021/acscatal.5b00898
    6. 6
      Gensch, T.; Hopkinson, M. N.; Glorius, F.; Wencel-Delord, J. Mild metal-catalyzed C–H activation: examples and concepts. Chem. Soc. Rev. 2016, 45, 29002936, DOI: 10.1039/C6CS00075D
    7. 7
      Perutz, R. N.; Sabo-Etienne, S. The σ-CAM Mechanism: σ-Complexes as the Basis of σ-Bond Metathesis at Late-Transition-Metal Centers. Angew. Chem., Int. Ed. 2007, 46, 25782592, DOI: 10.1002/anie.200603224
    8. 8
      (a) Guihaumé, J.; Halbert, S.; Eisenstein, O.; Perutz, R. N. Hydrofluoroarylation of Alkynes with Ni Catalysts. C–H Activation via Ligand-to-Ligand Hydrogen Transfer, an Alternative to Oxidative Addition. Organometallics 2012, 31, 13001314, DOI: 10.1021/om2005673 .
      (b) Tang, S.; Eisenstein, O.; Nakao, Y.; Sakaki, S. Aromatic C–H σ-Bond Activation by Ni0, Pd0, and Pt0 Alkene Complexes: Concerted Oxidative Addition to Metal vs Ligand-to-Ligand H Transfer Mechanism. Organometallics 2017, 36, 27612771, DOI: 10.1021/acs.organomet.7b00256 .
      (c) Yamazaki, K.; Obata, A.; Sasagawa, A.; Ano, Y.; Chatani, N. Computational Mechanistic Study on the Nickel-Catalyzed C–H/N–H Oxidative Annulation of Aromatic Amides with Alkynes: The Role of the Nickel (0) Ate Complex. Organometallics 2019, 38, 248255, DOI: 10.1021/acs.organomet.8b00684
    9. 9
      (a) Shibata, K.; Hasegawa, N.; Fukumoto, Y.; Chatani, N. Ruthenium-Catalyzed Carbonylation of ortho C–H Bonds in Arylacetamides: C–H Bond Activation Utilizing a Bidentate-Chelation System. ChemCatChem 2012, 4, 17331736, DOI: 10.1002/cctc.201200352 .
      (b) Hasegawa, N.; Shibata, K.; Charra, V.; Inoue, S.; Fukumoto, Y.; Chatani, N. Ruthenium-catalyzed cyclocarbonylation of aliphatic amides through the regioselective activation of unactivated C(sp3)–H bonds. Tetrahedron 2013, 69, 44664472, DOI: 10.1016/j.tet.2013.02.006 .
      (c) Misal Castro, L. C.; Obata, A.; Aihara, Y.; Chatani, N. Chelation-Assisted Nickel-Catalyzed Oxidative Annulation via Double C–H Activation/Alkyne Insertion Reaction. Chem. Eur. J. 2016, 22, 13621367, DOI: 10.1002/chem.201504596 .
      (d) He, Z.; Huang, Y. Diverting C–H Annulation Pathways: Nickel-Catalyzed Dehydrogenative Homologation of Aromatic Amides. ACS Catal. 2016, 6, 78147823, DOI: 10.1021/acscatal.6b02477
    10. 10

      For computational studies of Ni-catalyzed C–H functionalization using N,N-bidentate strategy, see:

      (a) Tang, H.; Zhou, B.; Huang, X.-R.; Wang, C.; Yao, J.; Chen, H. Origins of Selective C(sp2)–H Activation Using Transition Metal Complexes with N,N-Bidentate Directing Groups: A Combined Theoretical–Experimental Study. ACS Catal. 2014, 4, 649656, DOI: 10.1021/cs401141k .
      (b) Tang, H.; Huang, X.-R.; Yao, J.; Chen, H. Understanding the Effects of Bidentate Directing Groups: A Unified Rationale for sp2 and sp3 C–H Bond Activations. J. Org. Chem. 2015, 80, 46724682, DOI: 10.1021/acs.joc.5b00580 .
      (c) Xu, Z.-Y.; Jiang, Y.-Y.; Yu, H.-Z.; Fu, Y. Mechanism of Nickel(II)-Catalyzed Oxidative C(sp2)–H/C(sp3)–H Coupling of Benzamides and Toluene Derivatives. Chem. - Asian J. 2015, 10, 24792483, DOI: 10.1002/asia.201500599 .
      (d) Singh, S.; K, S.; Sunoj, R. B. Aliphatic C(sp3)–H Bond Activation Using Nickel Catalysis: Mechanistic Insights on Regioselective Arylation. J. Org. Chem. 2017, 82, 96199626, DOI: 10.1021/acs.joc.7b01672 .
      (e) Omer, H. M.; Liu, P. Computational Study of Ni-Catalyzed C–H Functionalization: Factors That Control the Competition of Oxidative Addition and Radical Pathways. J. Am. Chem. Soc. 2017, 139, 99099920, DOI: 10.1021/jacs.7b03548 .
      (f) Haines, B. E.; Yu, J.-Q.; Musaev, D. G. The mechanism of directed Ni(II)-catalyzed C–H iodination with molecular iodine. Chem. Sci. 2018, 9, 11441154, DOI: 10.1039/C7SC04604A .
      (g) Li, Y.; Zou, L.; Bai, R.; Lan, Y. Ni(I)–Ni(III) vs. Ni(II)–Ni(IV): mechanistic study of Ni-catalyzed alkylation of benzamides with alkyl halides. Org. Chem. Front. 2018, 5, 615622, DOI: 10.1039/C7QO00850C
    11. 11

      For computational studies of other metal-catalyzed C–H functionalization using N,N-bidentate group strategy, see:

      (a) Cross, W. B.; Hope, E. G.; Lin, Y.-H.; Macgregor, S. A.; Singh, K.; Solan, G. A.; Yahyaa, N. N,N-Chelate-control on the regioselectivity in acetate-assisted C–H activation. Chem. Commun. 2013, 49, 19181920, DOI: 10.1039/c3cc38697j .
      (b) Huang, L.; Li, Q.; Wang, C.; Qi, C. Palladium(II)-Catalyzed Regioselective Arylation of Naphthylamides with Aryl Iodides Utilizing a Quinolinamide Bidentate System. J. Org. Chem. 2013, 78, 30303038, DOI: 10.1021/jo400017v .
      (c) Wei, Y.; Tang, H.; Cong, X.; Rao, B.; Wu, C.; Zeng, X. Pd(II)-Catalyzed Intermolecular Arylation of Unactivated C(sp3)–H Bonds with Aryl Bromides Enabled by 8-Aminoquinoline Auxiliary. Org. Lett. 2014, 16, 22482251, DOI: 10.1021/ol500745t .
      (d) Shan, C.; Luo, X.; Qi, X.; Liu, S.; Li, Y.; Lan, Y. Mechanism of Ruthenium-Catalyzed Direct Arylation of C–H Bonds in Aromatic Amides: A Computational Study. Organometallics 2016, 35, 14401445, DOI: 10.1021/acs.organomet.6b00064 .
      (e) Chen, C.; Hao, Y.; Zhang, T.-Y.; Pan, J.-L.; Ding, J.; Xiang, H.-Y.; Wang, M.; Ding, T.-M.; Duan, A.; Zhang, S.-Y. Computational and experimental studies on copper-mediated selective cascade C–H/N–H annulation of electron-deficient acrylamide with arynes. Chem. Commun. 2019, 55, 755758, DOI: 10.1039/C8CC08708C .
      (f) Dewyer, A. L.; Zimmerman, P. M. Simulated Mechanism for Palladium Catalyzed, Directed γ-Arylation of Piperidine. ACS Catal. 2017, 7, 54665477, DOI: 10.1021/acscatal.7b01390
    12. 12
      Shiota, H.; Ano, Y.; Aihara, Y.; Fukumoto, Y.; Chatani, N. Nickel-Catalyzed Chelation-Assisted Transformations Involving Ortho C–H Bond Activation: Regioselective Oxidative Cycloaddition of Aromatic Amides to Alkynes. J. Am. Chem. Soc. 2011, 133, 1495214955, DOI: 10.1021/ja206850s
    13. 13
      For a recent computational study on this reaction, see:Zhang, X.; Zhao, Q.; Fan, J.-Q.; Chen, D.-Z.; Liu, J.-B. A computational mechanistic study of Ni(0)-catalyzed annulation of aromatic amides with alkynes: the effects of directing groups. Org. Chem. Front. 2019, DOI: 10.1039/C8QO01310A
    14. 14

      For select examples of chelation-assisted Ni-catalyzed C–H arylation, see:

      (a) Aihara, Y.; Chatani, N. Nickel-Catalyzed Direct Arylation of C(sp3)–H Bonds in Aliphatic Amides via Bidentate-Chelation Assistance. J. Am. Chem. Soc. 2014, 136, 898901, DOI: 10.1021/ja411715v .
      (b) Yokota, A.; Aihara, Y.; Chatani, N. Nickel(II)-Catalyzed Direct Arylation of C–H Bonds in Aromatic Amides Containing an 8-Aminoquinoline Moiety as a Directing Group. J. Org. Chem. 2014, 79, 1192211932, DOI: 10.1021/jo501697n .
      (c) Li, M.; Dong, J.; Huang, X.; Li, K.; Wu, Q.; Song, F.; You, J. Nickel-catalyzed chelation-assisted direct arylation of unactivated C(sp3)–H bonds with aryl halides. Chem. Commun. 2014, 50, 39443946, DOI: 10.1039/C4CC00716F
    15. 15

      For select examples of chelation-assisted Ni-catalyzed C–H alkylation, see:

      (a) Kubo, T.; Chatani, N. Dicumyl Peroxide as a Methylating Reagent in the Ni-Catalyzed Methylation of Ortho C–H Bonds in Aromatic Amides. Org. Lett. 2016, 18, 16981701, DOI: 10.1021/acs.orglett.6b00658 .
      (b) Aihara, Y.; Chatani, N. Nickel-Catalyzed Direct Alkylation of C–H Bonds in Benzamides and Acrylamides with Functionalized Alkyl Halides via Bidentate-Chelation Assistance. J. Am. Chem. Soc. 2013, 135, 53085311, DOI: 10.1021/ja401344e .
      (c) Wu, X.; Zhao, Y.; Ge, H. Nickel-Catalyzed Site-Selective Alkylation of Unactivated C(sp3)–H Bonds. J. Am. Chem. Soc. 2014, 136, 17891792, DOI: 10.1021/ja413131m .
      (d) Song, W.; Lackner, S.; Ackermann, L. Nickel-Catalyzed C–H Alkylations: Direct Secondary Alkylations and Trifluoroethylations of Arenes. Angew. Chem., Int. Ed. 2014, 53, 24772480, DOI: 10.1002/anie.201309584
    16. 16

      For select examples of chelation-assisted Ni-catalyzed C–H benzylation, see:

      (a) Aihara, Y.; Tobisu, M.; Fukumoto, Y.; Chatani, N. Ni(II)-Catalyzed Oxidative Coupling between C(sp2)–H in Benzamides and C(sp3)–H in Toluene Derivatives. J. Am. Chem. Soc. 2014, 136, 1550915512, DOI: 10.1021/ja5095342 .
      (b) Soni, V.; Khake, S. M.; Punji, B. Nickel-Catalyzed C(sp2)–H/C(sp3)–H Oxidative Coupling of Indoles with Toluene Derivatives. ACS Catal. 2017, 7, 42024208, DOI: 10.1021/acscatal.7b01044
    17. 17

      For select examples of chelation-assisted Ni-catalyzed C–H thiolation, see:

      (a) Yan, S.-Y.; Liu, Y.-J.; Liu, B.; Liu, Y.-H.; Shi, B.-F. Nickel-catalyzed thiolation of unactivated aryl C–H bonds: efficient access to diverse aryl sulfides. Chem. Commun. 2015, 51, 40694072, DOI: 10.1039/C4CC10446C .
      (b) Yan, S.-Y.; Liu, Y.-J.; Liu, B.; Liu, Y.-H.; Zhang, Z.-Z.; Shi, B.-F. Nickel-catalyzed direct thiolation of unactivated C(sp3)–H bonds with disulfides. Chem. Commun. 2015, 51, 73417344, DOI: 10.1039/C5CC01436K .
      (c) Lin, C.; Li, D.; Wang, B.; Yao, J.; Zhang, Y. Direct ortho-Thiolation of Arenes and Alkenes by Nickel Catalysis. Org. Lett. 2015, 17, 13281331, DOI: 10.1021/acs.orglett.5b00337 .
      (d) Reddy, V. P.; Qiu, R.; Iwasaki, T.; Kambe, N. Nickel-catalyzed synthesis of diaryl sulfides and sulfones via C–H bond functionalization of arylamides. Org. Biomol. Chem. 2015, 13, 68036813, DOI: 10.1039/C5OB00149H
    18. 18
      Aihara, Y.; Chatani, N. Nickel-Catalyzed Reaction of C–H Bonds in Amides with I2: ortho-Iodination via the Cleavage of C(sp2)–H Bonds and Oxidative Cyclization to β-Lactams via the Cleavage of C(sp3)–H Bonds. ACS Catal. 2016, 6, 43234329, DOI: 10.1021/acscatal.6b00964
    19. 19
      Nakao, Y.; Morita, E.; Idei, H.; Hiyama, T. Dehydrogenative [4 + 2] Cycloaddition of Formamides with Alkynes through Double C–H Activation. J. Am. Chem. Soc. 2011, 133, 32643267, DOI: 10.1021/ja1102037
    20. 20
      (a) Lin, C.; Chen, Z.; Liu, Z.; Zhang, Y. Nickel-Catalyzed Stereoselective Alkenylation of C(sp3)–H Bonds with Terminal Alkynes. Org. Lett. 2017, 19, 850853, DOI: 10.1021/acs.orglett.6b03856 .
      (b) Li, M.; Yang, Y.; Zhou, D.; Wan, D.; You, J. Nickel-Catalyzed Addition-Type Alkenylation of Unactivated, Aliphatic C–H Bonds with Alkynes: A Concise Route to Polysubstituted γ-Butyrolactones. Org. Lett. 2015, 17, 25462549, DOI: 10.1021/acs.orglett.5b01128 .
      (c) Maity, S.; Agasti, S.; Earsad, A. M.; Hazra, A.; Maiti, D. Nickel-Catalyzed Insertion of Alkynes and Electron-Deficient Olefins into Unactivated sp3 C–H Bonds. Chem. - Eur. J. 2015, 21, 1132011324, DOI: 10.1002/chem.201501962 .
      (d) Liu, Y.-H.; Liu, Y.-J.; Yan, S.-Y.; Shi, B.-F. Ni(II)-catalyzed dehydrative alkynylation of unactivated (hetero)aryl C–H bonds using oxygen: a user-friendly approach. Chem. Commun. 2015, 51, 1165011653, DOI: 10.1039/C5CC03729H
    21. 21

      For C–H functionalization reactions where alkyne or alkene act as hydrogen acceptor, see:

      (a) Inoue, S.; Shiota, H.; Fukumoto, Y.; Chatani, N. Ruthenium-Catalyzed Carbonylation at Ortho C–H Bonds in Aromatic Amides Leading to Phthalimides: C–H Bond Activation Utilizing a Bidentate System. J. Am. Chem. Soc. 2009, 131, 68986899, DOI: 10.1021/ja900046z .
      (b) Hasegawa, N.; Charra, V.; Inoue, S.; Fukumoto, Y.; Chatani, N. Highly Regioselective Carbonylation of Unactivated C(sp3)–H Bonds by Ruthenium Carbonyl. J. Am. Chem. Soc. 2011, 133, 80708073, DOI: 10.1021/ja2001709 .
      (c) Song, W.; Ackermann, L. Nickel-catalyzed alkyne annulation by anilines: versatile indole synthesis by C–H/N–H functionalization. Chem. Commun. 2013, 49, 66386640, DOI: 10.1039/c3cc43915a .
      (d) Obata, A.; Ano, Y.; Chatani, N. Nickel-catalyzed C–H/N–H annulation of aromatic amides with alkynes in the absence of a specific chelation system. Chem. Sci. 2017, 8, 66506655, DOI: 10.1039/C7SC01750B
    22. 22

      For Rh-catalyzed C–H functionalization with internal alkyne, see:

      (a) Mochida, S.; Nobuyoshi, U.; Koji, H.; Tetsuya, S.; Masahiro, M. Rhodium-catalyzed Oxidative Coupling/Cyclization of Benzamides with Alkynes via C–H Bond Cleavage. Chem. Lett. 2010, 39, 744746, DOI: 10.1246/cl.2010.744 .
      (b) Hyster, T. K.; Rovis, T. Rhodium-Catalyzed Oxidative Cycloaddition of Benzamides and Alkynes via C–H/N–H Activation. J. Am. Chem. Soc. 2010, 132, 1056510569, DOI: 10.1021/ja103776u .
      (c) Guimond, N.; Gorelsky, S. I.; Fagnou, K. Rhodium(III)-Catalyzed Heterocycle Synthesis Using an Internal Oxidant: Improved Reactivity and Mechanistic Studies. J. Am. Chem. Soc. 2011, 133, 64496457, DOI: 10.1021/ja201143v .
      (d) Hyster, T. K.; Rovis, T. An improved catalyst architecture for rhodium(III) catalyzed C–H activation and its application to pyridone synthesis. Chem. Sci. 2011, 2, 16061610, DOI: 10.1039/C1SC00235J .
      (e) Shan, G.; Flegel, J.; Li, H.; Merten, C.; Ziegler, S.; Antonchick, A. P.; Waldmann, H. C–H Bond Activation for the Synthesis of Heterocyclic Atropisomers Yields Hedgehog Pathway Inhibitors. Angew. Chem., Int. Ed. 2018, 57, 1425014254, DOI: 10.1002/anie.201809680 .
      (f) Shibata, K.; Natsui, S.; Chatani, N. Rhodium-Catalyzed Alkenylation of C–H Bonds in Aromatic Amides with Alkynes. Org. Lett. 2017, 19, 22342237, DOI: 10.1021/acs.orglett.7b00709
    23. 23

      For Ru-catalyzed C–H functionalization with internal alkynes, see:

      (a) Allu, S.; Swamy, K. C. K. Ruthenium-Catalyzed Synthesis of Isoquinolones with 8-Aminoquinoline as a Bidentate Directing Group in C–H Functionalization. J. Org. Chem. 2014, 79, 39633972, DOI: 10.1021/jo500424p .
      (b) Kaishap, P. P.; Duarah, G.; Chetiab, D.; Gogoi, S. Ru(II)-Catalyzed annulation of benzamidines and alkynes by C–H/N–H activation: a facile synthesis of 1-aminoisoquinolines. Org. Biomol. Chem. 2017, 15, 34913498, DOI: 10.1039/C7OB00389G
    24. 24

      For Pd-catalyzed C–H functionalization with internal alkynes, see:

      (a) Shu, Z.; Guo, Y.; Li, W.; Wang, B. Pd/C-catalyzed synthesis of N-aryl and N-alkyl isoquinolones via CH/NH activation. Catal. Today 2017, 297, 292297, DOI: 10.1016/j.cattod.2017.02.005 .
      (b) Sharma, N.; Saha, R.; Parveen, N.; Sekar, G. Palladium-Nanoparticles-Catalyzed Oxidative Annulation of Benzamides with Alkynes for the Synthesis of Isoquinolones. Adv. Synth. Catal. 2017, 359, 19471958, DOI: 10.1002/adsc.201601137
    25. 25

      For Co-catalyzed C–H functionalization with internal alkyne, see:

      (a) Grigorjeva, L.; Daugulis, O. Cobalt-catalyzed, aminoquinoline-directed C(sp2)-H bond alkenylation by alkynes. Angew. Chem., Int. Ed. 2014, 53, 1020910212, DOI: 10.1002/anie.201404579 .
      (b) Nguyen, T. T.; Grigorjeva, L.; Daugulis, O. Cobalt-Catalyzed, Aminoquinoline-Directed Functionalization of Phosphinic Amide sp2 C–H Bonds. ACS Catal. 2016, 6, 551554, DOI: 10.1021/acscatal.5b02391 .
      (c) Manoharan, R.; Jeganmohan, M. Cobalt-catalyzed cyclization of benzamides with alkynes: a facile route to isoquinolones with hydrogen evolution. Org. Biomol. Chem. 2018, 16, 83848389, DOI: 10.1039/C8OB01924J .
      (d) Zhai, S.; Qiu, S.; Chen, X.; Wu, J.; Zhao, H.; Tao, C.; Li, Y.; Cheng, B.; Wang, H.; Zhai, H. 2-(1-Methylhydrazinyl)pyridine as a reductively removable directing group in a cobalt-catalyzed C(sp2)–H bond alkenylation/annulation cascade. Chem. Commun. 2018, 54, 98101, DOI: 10.1039/C7CC08533H
    26. 26

      In the presence of a strong base, such as t-BuOK, N–H deprotonation would form a nickel(0) ate complex. A recent computational study indicates the subsequent C–H metalation from the nickel(0) ate complex occurs via a ligand-to-ligand hydrogen transfer mechanism. See ref (8c).

    27. 27

      For select reviews and crystal structures of Ni-hydride complexes, see:

      (a) Eberhardt, N. A.; Guan, H. Nickel Hydride Complexes. Chem. Rev. 2016, 116, 83738426, DOI: 10.1021/acs.chemrev.6b00259 .
      (b) Matson, E. M.; Martinez, G. E.; Ibrahim, A. D.; Jackson, B. J.; Bertke, J. A.; Fout, A. R. Nickel(II) Pincer Carbene Complexes: Oxidative Addition of an Aryl C–H Bond to Form a Ni(II) Hydride. Organometallics 2015, 34, 399407, DOI: 10.1021/om5007177 .
      (c) Clement, N. D.; Cavell, K. J.; Jones, C.; Elsevier, C. J. Oxidative Addition of Imidazolium Salts to Ni0 and Pd0: Synthesis and Structural Characterization of Unusually Stable Metal–Hydride Complexes. Angew. Chem., Int. Ed. 2004, 43, 1277
    28. 28
      Crabtree, R. H. Dihydrogen Complexation. Chem. Rev. 2016, 116, 87508769, DOI: 10.1021/acs.chemrev.6b00037
    29. 29
      (a) She, L.; Li, X.; Sun, H.; Ding, J.; Frey, M.; Klein, H.-F. Insertion of Alkynes into Ni–H Bonds: Synthesis of Novel Vinyl Nickel(II) and Dinuclear Vinyl Nickel(II) Complexes Containing a [P, S]-Ligand. Organometallics 2007, 26, 566570, DOI: 10.1021/om0606340 .
      (b) Zhou, H.; Sun, H.; Zheng, T.; Zhang, S.; Li, X. Synthesis of Vinylnickel and Nickelacyclopropane Complexes Containing a Chelate [P,Se]-Ligand. Eur. J. Inorg. Chem. 2015, 31393145, DOI: 10.1002/ejic.201500293 .
      (c) Xue, B.; Sun, H.; Ren, S.; Li, X.; Fuhr, O. Vinyl/Phenyl Exchange Reaction within Vinyl Nickel Complexes Bearing Chelate [P, S]-Ligands. Organometallics 2017, 36, 42464255, DOI: 10.1021/acs.organomet.7b00671
    30. 30

      An alternative pathway involving oxidative addition of ortho C(sp2)–H bond with the Ni(0) catalyst was considered computationally. The barrier of the C–H oxidative addition is 19.5 kcal/mol with respect to amide 1 and Ni(cod)2. However, the subsequent steps in this pathway were highly disfavored. Both alkyne insertion into Ni–H bond and N–H deprotonation to form the five-membered Ni(II) metallacycle 9 require very high activation barriers of 44.5 and 67.2 kcal/mol with respect to 1 and Ni(cod)2, respectively. Based on these results, this oxidative addition pathway was ruled out (see Supporting Information for more details).

    31. 31
      (a) Brookhart, M.; Green, M. L. H.; Parkin, G. Agostic interactions in transition metal compounds. Proc. Natl. Acad. Sci. USA 2007, 104, 69086914, DOI: 10.1073/pnas.0610747104 .
      (b) Jongbloed, L. S.; García-Lopez, D.; Heck, R. V.; Siegler, M. A.; Carbo, J. J.; Vlugt, J. I. V. D. Arene C(sp2)-H Metalation at NiII Modeled with a Reactive PONCPh Ligand. Inorg. Chem. 2016, 55, 80418047, DOI: 10.1021/acs.inorgchem.6b01162 .
      (c) Lein, M. Characterization of agostic interactions in theory and computation. Coord. Chem. Rev. 2009, 253, 625634, DOI: 10.1016/j.ccr.2008.07.007 .
      (d) Etienne, M.; Weller, A. S. Intramolecular C–C agostic complexes: C–C sigma interactions by another name. Chem. Soc. Rev. 2014, 43, 242259, DOI: 10.1039/C3CS60295H .
      (e) Beattie, D. D.; Bowes, E. G.; Drover, M. W.; Love, J. A.; Schafer, L. L. Oxidation State Dependent Coordination Modes: Accessing an Amidate-Supported Nickel(I) δ-bis(C–H) Agostic Complex. Angew. Chem. 2016, 128, 1348413489, DOI: 10.1002/ange.201607243 .
      (f) Crabtree, R. H. Sigma Bonds as Ligand Donor Groups in Transition Metal Complexes. The Chemical Bond III; Mingos, D.M.P., Ed.; Springer International Publishing: Switzerland, 2015; Vol. 171, pp 6378.
    32. 32

      An alternatively pathway to form intermediate 11 from phosphine-bound Ni(II)-hydride complex 5 via σ-bond metathesis followed by PPh3 decomplexation requires a much higher barrier (66.0 kcal/mol with respect to aromatic amide 1 and Ni(cod)2 catalyst). This very high barrier is due to the absence of an agostic interaction and unfavorable steric effects of the additional PPh3 ligand.

    33. 33

      PPh3 coordination to intermediate 8 forms an off-cycle phosphine-bound alkenyl-Ni(II) complex 7. Complex 7 is 11.2 kcal/mol more stable than 8.

    34. 34
      (a) Another mechanism for C–H metalation involves σ-bond metathesis of the Ni–N bond in intermediate 8 with the ortho C–H bond to form a five-membered alkenyl-nickelacycle (see Supporting Information for details). This process requires an activation barrier of 39 kcal/mol with respect to the separate reactants and Ni(cod)2, and thus can be ruled out.
      (b) In our calculations, we could not locate neither the transition state structure for the oxidative addition of the ortho C(sp2)–H bond from alkenyl-Ni(II) complex 8 nor the resulting Ni(IV)-hydride complex. All attempts to locate these structures resulted in TS3, 8, or 9. Intrinsic reaction coordinate (IRC) calculations were carried out for TS3 to confirm that it connects to complexes 8 and 9.
    35. 35

      We also computationally considered the use of cis-2-butene rather than 2-butyne as the H2 acceptor to promote the C–H metalation. In this alternatively pathway, the barrier of σ-bond metathesis is 34.8 kcal/mol with respect to the separate reactants and Ni(cod)2, and thus this pathway is less favorable than that using alkyne as H2 acceptor (see Supporting Information for details).

    36. 36
      McCarren, P. R.; Liu, P.; Cheong, P. H.-Y.; Jamison, T. F.; Houk, K. N. Mechanism and Transition-State Structures for Nickel-Catalyzed Reductive Alkyne–Aldehyde Coupling Reactions. J. Am. Chem. Soc. 2009, 131, 66546655, DOI: 10.1021/ja900701g
    37. 37

      In addition, the N in 2-pyridinylmethylamine is a better donor that electronically promotes the reductive elimination via TS6.

    38. 38
      Liang, L.-C.; Pin-Shu Chien, P.-S.; Lee, P.-Y. Phosphorus and Olefin Substituent Effects on the Insertion Chemistry of Nickel(II) Hydride Complexes Containing Amido Diphosphine Ligands. Organometallics 2008, 27, 30823093, DOI: 10.1021/om701294a
    39. 39
      (a) Kogut, E.; Zeller, A.; Warren, T. H.; Strassner, T. Structure and Dynamics of Neutral β-H Agostic Nickel Alkyls: A Combined Experimental and Theoretical Study. J. Am. Chem. Soc. 2004, 126, 1198411994, DOI: 10.1021/ja0477221 .
      (b) Scherer, W.; Herz, V.; Brück, A.; Hauf, C.; Reiner, F.; Altmannshofer, S.; Leusser, D.; Stalke, D. The Nature of β-Agostic Bonding in Late-Transition-Metal Alkyl Complexes. Angew. Chem., Int. Ed. 2011, 50, 28452849, DOI: 10.1002/anie.201006065
    40. 40
      Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision D.01; Gaussian, Inc.: Wallingford, CT, 2010.
    41. 41
      Legault, C. Y. CYLview, version 1.0b; Universitede Sherbrooke: Quebec, Canada, 2009; http://www.cylview.org.
    42. 42
      (a) Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 56485652, DOI: 10.1063/1.464913 .
      (b) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785789, DOI: 10.1103/PhysRevB.37.785
    43. 43
      (a) Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theoretical Chemistry Accounts; Springer, 2008; Vol. 120, pp 215241.
      (b) Zhao, Y.; Truhlar, D. G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem. Res. 2008, 41, 157167, DOI: 10.1021/ar700111a
    44. 44
      Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 63786396, DOI: 10.1021/jp810292n
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b00030.

    • Reaction energy profiles of amide N–H oxidative addition with different ligands (Figure S1), alternative C−H metalation mechanisms with cis-2-butene (Figure S2), via σ-bond metathesis of phosphine-bound Ni(II)-hydride (Figure S3), via oxidative addition to Ni(0) (Figure S4), via deprotonation by the amide N (Figure S5), and C−N reductive elimination with model substrates (Figure S6); Cartesian coordinates of optimized geometries (PDF)

    Terms & Conditions

    Electronic Supporting Information files are available without a subscription to ACS Web Editions. The American Chemical Society holds a copyright ownership interest in any copyrightable Supporting Information. Files available from the ACS website may be downloaded for personal use only. Users are not otherwise permitted to reproduce, republish, redistribute, or sell any Supporting Information from the ACS website, either in whole or in part, in either machine-readable form or any other form without permission from the American Chemical Society. For permission to reproduce, republish and redistribute this material, requesters must process their own requests via the RightsLink permission system. Information about how to use the RightsLink permission system can be found at http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

OOPS

You have to login with your ACS ID befor you can login with your Mendeley account.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect

This website uses cookies to improve your user experience. By continuing to use the site, you are accepting our use of cookies. Read the ACS privacy policy.

CONTINUE