Abstract
Palladium(II)-catalyzed C–H oxidation reactions could streamline the synthesis of pharmaceuticals, agrochemicals, and other complex organic molecules. Existing methods, however, commonly exhibit poor catalyst performance with high Pd loading (e.g., 10 mol%) and a need for (super)stoichiometric quantities of undesirable oxidants, such as benzoquinone and silver(I) salts. The present study probes the mechanism of a representative Pd-catalyzed oxidative C–H arylation reaction and elucidates mechanistic features that undermine catalyst performance, including substrate-consuming side reactions and sequestration of the catalyst as inactive species. Systematic tuning of the quinone co-catalyst overcomes these deleterious features. Use of 2,5-di-tert-butyl-p-benzoquinone enables efficient use of molecular oxygen as the oxidant, high reaction yields, and >1900 turnovers by the palladium catalyst.
One Sentence Summary:
Mechanistic studies lead to a high-performance palladium/quinone co-catalyst system for oxidative C–H arylation using O2 as the oxidant.
Homogeneous palladium-catalyzed oxidation reactions of organic molecules originated in 1959 with the discovery of the Wacker process for oxidative coupling of ethylene and water (1). Pd-catalyzed C–H oxidation reactions for oxidative homocoupling of arenes to biaryls were reported shortly thereafter (2), including examples compatible with O2 as the terminal oxidant (3). The poor catalytic efficiency and low regioselectivity of these methods contributed to development of Pd-catalyzed cross-coupling reactions with pre-oxidized substrates, such as aryl halides, as coupling partners. While Pd-catalyzed cross-coupling reactions have achieved extraordinary success (4), methods for direct oxidative functionalization of C–H bonds could substantially streamline synthetic access to diverse chemical structures (5–8).
Most PdII/0-catalyzed coupling reactions, including both oxidative and non-oxidative examples, fit into two general classes: (i) olefinations, such as Heck, Fujiwara-Moritani, and related reactions that involve coupling with alkenes and generate products via β-hydride elimination from a PdII-alkyl intermediate, and (ii) arylations, such as Suzuki-Miyaura, Negishi, and related reactions that involve coupling with arenes or aryl nucleophiles and generate the products via reductive elimination from a PdII–aryl intermediate. The oxidative reactions (cf. Fig. 1A) are mechanistically similar to non-oxidative cross-coupling reactions, but they often exhibit worse catalytic performance and require the use of one or more stoichiometric oxidants, such as CuII, AgI, and benzoquinone. The representative collection of oxidative arylation reactions in Fig. 1B, emphasizing cases in which the C–H substrate is the limiting reagent, highlights typical catalytic turnover numbers and stoichiometric oxidant(s) (9–16). The initial indole arylation example (9) achieves comparatively high turnovers (32 TOs); however, it requires 3 equiv Cu(OAc)2 as the oxidant and 30 equiv of benzene as a coupling partner. Many other examples feature < 10 TOs (see Fig. S1A in the Supplementary Materials for additional examples). Oxidative olefination reactions with limiting C–H substrate often exhibit similarly poor catalytic performance (see Fig. S1B in the Supplementary Materials for examples). We have noted, however, that Pd-catalyzed oxidation reactions that proceed via β-hydride elimination tend to be more compatible with O2 as an oxidant than reductive elimination reactions (17). This phenomenon is clearly demonstrated in a pair of reactions reported by Yu and coworkers involving PdII/0-catalyzed C–H olefination (18–20) and arylation (21). Both were demonstrated with 1 atm O2 as the oxidant, but the olefination reaction accessed 455 TOs (Fig. 1C, left), whereas the arylation reaction (Fig. 1C, right) achieved only 13 TOs under comparable conditions. Understanding the factors that limit catalyst turnovers in oxidative arylation reactions of this type could play a major role in expanding their utility, and accessing improved catalytic performance could set the stage for large-scale applications that would greatly benefit from low catalyst loading and use of O2 as an oxidant (22). Here, we present a mechanistic study of the Pd-catalyzed C–H arylation reaction in Fig. 1C and use the insights to develop a modified catalyst system capable of accessing >1900 TOs and high yields with O2 as the oxidant.
Fig. 1.
Pd-catalyzed aerobic C–H oxidation reactions, highlighting challenges in catalyst performance. (A) Mechanisms for PdII-catalyzed oxidative olefination and arylation reactions using O2 as the oxidant. (B) Catalytic metrics for representative oxidative arylation reactions (9–16). (C) Comparison of catalyst turnovers in Pd-catalyzed olefination and arylation reactions (19, 21). MPAA = Mono-N-protected amino acid.
The reported arylation reaction in Fig. 1C was adapted to facilitate mechanistic study. Use of the potassium salt of the arene substrate [2-(trifluoromethyl)phenylacetic acid (1)] eliminated the requirement for heterogeneous KHCO3 as a Brønsted base, and arylboronic acid pinacol ester (2) provided a soluble alternative to the ArBF3K coupling partner. An 83% NMR assay yield of the arylation product 3 was obtained with these modifications when the reaction was conducted with 1 atm O2 (Fig. 2A), representing an improvement over the originally reported yield of 65%. The initial experiments confirmed that 1,4-benzoquinone (BQ) is a crucial co-catalyst for the reaction. In the absence of BQ, negligible arylation product was formed, with the majority of the arylboronate undergoing protodeboronation to afford fluorobenzene (ArF–H) (Fig. 2A). Analysis of the reaction time course revealed a kinetic burst at the start of the reaction, with a magnitude corresponding to one turnover of the Pd catalyst, followed by a much slower steady-state rate (Fig. 2B). Increasing the oxygen pressure had no effect on the burst, but it led to an increase in rate following the burst roughly proportional to the change in O2 pressure (0.46 × 10−3 mM/s at 1 atm; 2.5 × 10−3 at 6.4 atm).
Fig. 2.
Kinetic and spectroscopic analysis of the Pd-catalyzed C–H arylation reaction. (A) Oxidative arylation conditions (R = 2-(trifluoromethyl)C6H4CH2-). (B) Kinetic data showing a burst and the influence of O2 pressure on the steady state arylation rate. (C) Simplified mechanism rationalizing the kinetic burst and BQ inhibition of steady state turnover. (D) Enhanced steady state rate observed with 2,5-tBu2BQ. (E) Reaction progress obtained from 19F NMR spectroscopic analysis of the reaction mixture to which different reaction components were added sequentially. Interpolative curve fits are included as a guide. (F) Kinetic data [2,5-tBu2BQ] for formation of 3 and ArF–H, with correlations reflecting a hyperbolic curve fit [y = ax/(1 + bx)]. (G) Spectroscopic observation of palladacycle resting state under 6.9 atm O2, with linear and interpolative curve fits as a guide. Conditions: (A) [1] = 200 mM, 24 h (B, D, F) [1] = 100 mM, [Pd]t = 5 mM, [Boc-Val-OH] = 10 mM, [quinone] = 20 mM, 85 °C, 1 mL, 1 atm O2. (E, G) [1] = 100 mM, [Pd]t = 10 mM, [Boc-Val-OH] = 20 mM, [2,5-tBu2BQ] = 40 mM, 64 °C, 0.55 mL.
The two-stage catalytic mechanism in Fig. 2C provided a framework for preliminary interpretation of these results. The kinetic burst could arise from fast stoichiometric oxidative coupling of the substrates by PdII/BQ, while the post-burst phase could arise from turnover-limiting aerobic oxidation of the reduced Pd catalyst. BQ is commonly used to oxidize Pd0 in Pd-catalyzed oxidation reactions (23); however, Brønsted acid is needed to promote this reaction (24). We speculated that the basic conditions associated with the present arylation reaction prevent BQ from serving as an effective oxidant, and it instead coordinates to Pd0 and inhibits its reoxidation by O2. This mechanistic hypothesis prompted us to test sterically modified benzoquinone derivatives that might undergo more facile displacement by O2 (Fig. 2D) (25), thereby facilitating catalytic turnover. Use of tert-butyl-p-benzoquinone (tBuBQ) had virtually no impact on the reaction time course; the reaction still exhibited a kinetic burst, followed by slow steady-state turnover. In contrast, use of 2,5-di-tert-butyl-p-benzoquinone (2,5-tBu2BQ) led to a sustained high rate, without a slowdown after the initial catalyst turnover (Fig. 2D, blue).
The reaction was then interrogated by 19F NMR spectroscopy in order to gain insights into the role of the different catalyst components. The C–H substrate 1 and PdII(OAc)2 were dissolved in tAmylOH and added to an NMR tube under ambient air. No reaction was observed upon warming this mixture to 64 °C, consistent with previous results showing that C–H activation is very slow in the absence of a mono-N-protected amino acid (MPAA) ligand (8,19,26,27). Addition of the MPAA ligand initiated immediate formation of the previously characterized palladacycle derived from activation of the substrate C–H bond (Fig. 2E, black, see Supplementary Materials Section 5 for details) (27). Subsequent addition of the arylboronic ester coupling partner led to formation of protodeboronation side product, ArF–H (Fig. 2E, blue). No C–H arylation product 3 was observed at this stage, but arose only after addition of 2,5-tBu2BQ (Fig. 2E, red). These observations align with the catalytic data in Fig. 2A showing that ArF–H is the product obtained when the reaction is conducted in the absence of BQ. Systematic variation of the 2,5-tBu2BQ concentration provided evidence for direct competition between the formation of ArF–H and 3, with the latter promoted by higher quinone concentrations (Fig. 2F).
These experiments were further complemented by spectroscopic and kinetic analysis of the fully constituted catalytic reaction mixture. Use of a sealed NMR tube with higher initial O2 pressure (6.9 atm) minimized complications arising from depletion of dissolved O2 in the NMR solution (e.g., which accounts for decay of the palladacycle after 2.5 h in Fig. 2E). Under these conditions, the palladacycle forms rapidly and accounts for all of the Pd in solution (Fig. 2G, see Supplementary Materials Section 6 for details). This species persists as the catalyst resting state, while the arylation product 3 steadily appears in parallel with smaller quantities of the ArF–H byproduct. Initial-rate kinetic studies revealed a negligible deuterium kinetic isotope effect (kH/kD = 1.2 ± 0.1), based on independent rates measured for arylation of 1 and 1-d1, the latter deuterated at the site targeted by C–H activation, ortho to the directing group. Under conditions of saturating [2,5-tBu2BQ], the catalytic rate law exhibits a first-order dependence on [Pd], a zero-order dependence on O2 pressure, and saturation dependences on the coupling partners [1] and [2] (Figs. S2, S8-10 in the Supplementary Materials).
The experimental data presented above indicate that the palladacycle is the catalyst resting state and transmetalation of ArF from the arylboronate to palladium is the turnover-limiting step when 2,5-tBu2BQ is used as the quinone co-catalyst. The dependence of the rate on substrate 1 is rationalized by the ability of basic additives to promote the transmetalation step (28). The proposed mechanism in Fig. 3A incorporates these features, together with other steps in the catalytic reaction and steps leading to byproduct formation. The reaction is initiated by MPAA-promoted C–H activation of the substrate in I to form palladacycle II (26), which is observed by 19F NMR spectroscopy. A recent study indicates that the MPAA ligand plays a catalytic role in the C–H activation step and does not influence other steps in the mechanism (27). Moreover, MPAA undergoes facile exchange with substrate-derived carboxylates and likely does not remain coordinated to Pd throughout the cycle. Transmetalation from the arylboronate to the palladacycle generates the diaryl-PdII species III, which can undergo protodemetalation to form byproduct ArF–H and regenerate the palladacycle II or undergo quinone-promoted reductive elimination to generate the arylation product 3 and the Pd0-quinone adduct IV. The PdII catalyst is regenerated via reaction of IV with O2.
Fig. 3.
Catalytic cycle consistent with the experimental data and computational analysis. (A) The proposed catalytic mechanism for C–H arylation of 1 (R = 2-(trifluoromethyl)C6H4CH2-), and (B) calculated free energy diagram for protodemetalation and reductive elimination pathways in the absence and presence of quinones. See text and Supplementary Materials for details of the computational methods. TOF = Turnover Frequency, TON = Turnover Numbers
The proposed mechanism in Fig. 3A highlights the delicate balance between the beneficial and deleterious effects of the quinone co-catalyst. It is needed to promote reductive elimination (29–31) and avoid catalytic degradation of the arylboronate; however, it can also inhibit catalytic turnover by coordinating too strongly to Pd0. Further insights into these roles of quinone were obtained from density functional theory calculations (Fig. 3b). The relative energies presented in Fig. 3B were calculated at the M06-(IEF-PCM)/BS2 [BS2 = 6–311+G(d,p) (for all atoms except Pd) and SDD (for Pd)] levels of theory by using geometries and enthalpy and entropy corrections calculated at the B3LYP-D3BJ/BS1 level of theory [BS1 = 6–31G(d,p) (for all atoms except Pd) and LanL2dz (for Pd)] (see Section 8 in the Supplementary Materials for details).
Specifically, the PdII-diaryl species, Pd(Ar)(Ar’) (cf. III in Fig. 3A) was interrogated to compare the energetics of the protodemetalation step with C–C reductive elimination in the absence and presence of quinone (BQ and 2,5-tBu2BQ) (Fig. 3b). Neither BQ nor 2,5-tBu2BQ form favorable adducts with the Pd(Ar)(Ar’) species (ΔG° = +2.9 and =3.9 kcal/mol, respectively); however, both lower the kinetic barrier for reductive elimination relative to the quinone-free barrier (ΔΔG‡ = –4.4 and –3.2 kcal/mol for BQ and 2,5-tBu2BQ). The transition-state energies for quinone-promoted reductive elimination is very similar to the transition-state energy for the protodemetalation pathway, consistent with the competitive formation of ArF–H and product 3 (cf. Fig. 2f).
The mechanistic insights provided above set the stage for efforts to optimize the catalytic performance. Key screening data are summarized in Fig. 4A, with full details provided in Section 9 of the Supplementary Materials. Use of 2,5-tBu2BQ with 5% mol Pd(OAc)2 and 3 atm O2 led to >99% yield of 3, corresponding 20 TOs (Fig. 4a, entry 1). Reducing the catalyst loading to 0.15 mol% Pd with the same ratio of Pd:MPAA:2,5-tBu2BQ (1:2:4) resulted in 147 TOs. The reaction instead generated large quantities of ArF– H and 4,4’-difluorobiphenyl (ArF–ArF) (entry 2). The latter byproduct arises from oxidative homocoupling of arylboronate (32) and is rationalized by inefficient C–H activation of 1 at low Pd/MPAA loading, due to relatively weak binding of MPAA to PdII (27). This complication was addressed by increasing the MPAA loading to 10 mol%, which enhances the rate of C–H activation without increasing the Pd loading. These conditions supported effective C–H activation and increased the Pd TOs to 373 (Fig. 4a, entry 3). The quinone loading and identity of the MPAA ligand were then varied in order to access higher reaction yields. The most dramatic effect was observed upon replacing Boc-Val-OH with Ac-Ile-OH as the MPAA ligand, which led to 667 TOs and a quantitative yield of 3 (entry 5).
Fig. 4.
Catalytic performance with low catalyst loading. (A) Catalyst optimization data showing product ratios and catalyst turnover numbers (TON) for the oxidative coupling of 1 and 2. (B) Observed turnovers of 3 with different quinones/alkenes. (C) Application of mechanistic insights to other substrates and coupling partners. Conditions: (A,B) see conditions shown in above and below the arrow in A and B, 0.5 mL scale, Data based on NMR assay yield. (C) Identical to the conditions used in B except 0.5 mol% Pd was used with 30 mol% 2,5-tBu2BQ. [**] Reported catalyst turnovers with 5 mol % Pd(OAc)2 and 20 mol % BQ (21). [^] A 93% isolated yield was obtained. [†] with 0.05 mol% Pd(O2CR)2 and 72 hours reaction time. [‡] A 96% isolated yield was obtained. [∫] Yields shown are isolated product yields. [#] 30 mol% 2,6-tBu2BQ was used. [§] 5 equivalents of methylboronic acid was used. [¶] K2CO3 was used as the base and no Ac-Ile-OH added.
With the identification of conditions compatible with low Pd loading, we revisited the influence of the quinone identity. The results reinforced previous observations (Fig. 4b). Little product was observed with the parent BQ and with tBuBQ (A and B), while higher catalytic turnover was evident with 2,5-Me2BQ (C, 230 TOs). Further increase in the number or size of the substituents conferred additional benefits, as evident with Me4BQ (F, 574 TOs) and 2,6- tBu2BQ, and 2,5- tBu2BQ (G and H, 667 TOs). The latter two examples led to quantitative product yield at 0.15 mol% Pd loading, prompting reassessment with 0.05 mol% Pd. The beneficial effect of these sterically encumbered quinones is evident from the 1520 turnovers with the 2,6-tBu2BQ and 1960 turnovers with the 2,5-tBu2BQ, the latter corresponding to a near-quantitative product yield.
This survey also provided an opportunity to test an unexpected conclusion of the mechanistic studies above. The proposed catalytic mechanism in Fig. 3A features a non-redox role for the quinone, suggesting that redox-inactive electron-deficient alkenes should also support catalytic turnover. This hypothesis was validated by the use of dimethyl fumarate (D) and diisopropyl fumarate (E), which supported 390 and 440 turnovers respectively. Although this performance is not better than that observed with the sterically encumbered quinones (F–H), it is much better than BQ, the most commonly used co-catalyst in Pd-catalyzed aerobic oxidation reactions.
The optimized catalyst conditions were tested with a representative collection of other substrate partners with 0.5 mol% Pd (Fig. 4C). Several arylboronic esters, including methyl benzoate and protected aniline derivatives, and methylboronic acid react with the parent C–H substrate in good-to-excellent isolated yields (4–7). The 1-naphthylacetate and benzoate C–H substrates (8–10) also proceed in good yield. Substrates with neutral nitrogen direction groups, including the rather heterocyclic sildenafil precursor, were also effective (11, 12), albeit with more modest yields.
Overall, the results above highlight the delicate balancing act of quinone co-catalysts in these reactions. The quinone is needed to promote the kinetically difficult reductive elimination step and prevent deleterious consumption of the arylboronate via protodemetalation, but it also must avoid catalytic inhibition, which can arise if it coordinates too strongly to Pd0. Tuning of the quinone structure allows an effective balance to be found among these roles. These considerations provide relevant context for comparison with the closely related olefination reactions (Fig. 1A and 1C). The β-hydride elimination and H–X reductive elimination steps in olefination reactions are kinetically more facile than the C–C reductive elimination step in the arylation reactions, and olefination reactions lack the deleterious protodemetalation side-reaction that hinders the arylation reaction. Thus, olefinations often do not employ a quinone co-catalyst and tend to be more amenable to aerobic catalytic turnover. The insights from this study show how strategic integration of a redox inactive quinone or related co-catalyst in Pd-catalyzed arylation reactions could provide the basis for highly effective catalytic performance, enabling low Pd loading and effective use of O2 as the oxidant. The promising extension of these results to other substrates (Fig. 4C) set the stage for further application of the concepts described herein.
Supplementary Material
Acknowledgments:
We thank Jin-Quan Yu (Scripps), Donna Blackmond (Scripps), and other members in the NSF Center for C–H Functionalization for stimulating discussions during the course of this project. We also appreciate valuable discussions with Joshua Buss and David Bruns (UW-Madison).
Funding: C.A.S. was supported by an NSF predoctoral fellowship (DGE-1747503), and other financial support was provided by NSF under the CCI Center for C–H Functionalization (CHE-1700982). Spectroscopic instrumentation was supported by a gift from Paul J. Bender, NSF (CHE-1048642), and NIH (1S10 OD020022-1), and the authors gratefully acknowledge the use of resources of the Cherry L. Emerson Center for Scientific Computation (Emory University).
Footnotes
Competing interests: The authors declare no competing interests.
Data and materials availability: All data are available in the main text or the supplementary materials.
References
- (1).Smidt J, Hafner W, Jira R, Sedlmeier J, Sieber R, Rüttinger R, Kojer H, Katalytische umsetzungen von olefinen an platinmetall-verbindungen das consortium-verfahren zur herstellung von acetaldehyd. Angew. Chem 71, 176–182 (1959). [Google Scholar]
- (2).van Helden R, Verberg G, The oxidative coupling of aromatic compounds with palladium salts. Rec. Trav. Chim. Pays-Bas 84, 1263–1273 (1965). [Google Scholar]
- (3).Stahl SS, Palladium oxidase catalysis: Selective oxidation of organic chemicals by direct dioxygen-coupled turnover. Angew. Chem. Int. Ed 43, 3400–3420 (2004). [DOI] [PubMed] [Google Scholar]
- (4).Johansson-Seechurn CCC, Kitching MO, Colacot TJ, Snieckus V, Palladium-catalyzed cross-coupling: A historical contextual perspective to the 2010 Nobel Prize. Angew. Chem. Int. Ed 51, 5062–5085 (2012). [DOI] [PubMed] [Google Scholar]
- (5).Chen X, Engle KM, Wang D-H, Yu J-Q, Palladium(II)-catalyzed C-H activation/C-C cross-coupling reactions: versatility and practicality. Angew. Chem. Int. Ed 48, 5094–5115 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (6).Lyons TW, Sanford MS, Palladium-catalyzed ligand-directed C−H functionalization reactions. Chem. Rev 110, 1147–1169 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Wang D, Weinstein AB, White PB, Stahl SS, Ligand-promoted palladium-catalyzed aerobic oxidation reactions. Chem. Rev 118, 2636–2679 (2017). [DOI] [PubMed] [Google Scholar]
- (8).He J, Wasa M, Chan KSL, Shao Q, Yu J-Q, Palladium-catalyzed transformations of alkyl C–H bonds. Chem. Rev 117, 8754–8786 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Stuart DR, Fagnou K, The catalytic cross-coupling of unactivated arenes. Science. 316, 1172–1175 (2007). [DOI] [PubMed] [Google Scholar]
- (10).Potavathri S, Pereira KC, Gorelsky SI, Pike A, LeBris AP, DeBoef B, Regioselective oxidative arylation of indoles bearing N-alkyl protecting froups: Dual C−H functionalization via a concerted metalation−deprotonation mechanism. J. Am. Chem. Soc 132, 14676–14681 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (11).Dai H-X, Stepan AF, Plummer MS, Zhang Y-H, Yu J-Q, Divergent C–H functionalizations directed by sulfonamide pharmacophores: Late-stage diversification as a tool for drug discovery. J. Am. Chem. Soc 133, 7222–7228 (2011). [DOI] [PubMed] [Google Scholar]
- (12).Yang Y, Qiu X, Zhao Y, Mu Y, Shi Z, Palladium-catalyzed C–H arylation of indoles at the C7 position. J. Am. Chem. Soc 138, 495–498 (2016). [DOI] [PubMed] [Google Scholar]
- (13).Rosen BR, Simke LR, Thuy-Boun PS, Dixon DD, Yu J-Q, Baran PS, C–H functionalization logic enables synthesis of (+)-Hongoquercin A and related compounds. Angew. Chem. Int. Ed 52, 7317–7320 (2013). [DOI] [PubMed] [Google Scholar]
- (14).Chan KSL, Wasa M, Chu L, Laforteza BN, Miura M, Yu J-Q, Ligand-enabled cross-coupling of C(sp3)–H bonds with arylboron reagents via Pd(II)/Pd(0) catalysis. Nat. Chem 6, 146–150 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (15).Jain P, Verma P, Xia G, Yu J-Q, Enantioselective amine α-functionalization via palladium-catalysed C–H arylation of thioamides. Nat. Chem 9, 140–144 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (16).Rodrigalvarez J, Nappi M, Azuma H, Flodén NJ, Burns ME, Gaunt MJ, Catalytic C(sp3)–H bond activation in tertiary alkylamines. Nat. Chem 12, 76–81 (2020). [DOI] [PubMed] [Google Scholar]
- (17).Campbell AN, Stahl SS, Overcoming the “oxidant problem”: Strategies to use O2 as the oxidant in organometallic C–H oxidation reactions catalyzed by Pd (and Cu). Acc. Chem. Res 45, 851–863 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).Wang D-H, Engle KM, Shi B-F, Yu J-Q, Ligand-enabled reactivity and selectivity in a synthetically versatile aryl C–H olefination. Science. 327, 315–319 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (19).Engle KM, Wang D-H, Yu J-Q, Ligand-accelerated C−H activation reactions: Evidence for a switch of mechanism. J. Am. Chem. Soc 132, 14137–14151 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (20).Baxter RD, Sale D, Engle KM, Yu J-Q, Blackmond DG, Mechanistic rationalization of unusual kinetics in Pd-catalyzed C–H olefination. J. Am. Chem. Soc 134, 4600–4606 (2012). [DOI] [PubMed] [Google Scholar]
- (21).Engle KM, Thuy-Boun PS, Dang M, Yu J-Q, Ligand-accelerated cross-coupling of C(sp2)–H bonds with arylboron reagents. J. Am. Chem. Soc 133, 18183–18193 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).Stahl SS, Alsters PL, Liquid phase aerobic oxidation catalysis: Industrial applications and academic perspectives (Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, 2016) [Google Scholar]
- (23).Bäckvall J-E, Hopkins RB, Grennberg H, Mader M, Awasthi AK, Multistep electron transfer in palladium-catalyzed aerobic oxidations via a metal macrocycle quinone system. J. Am. Chem. Soc 112, 5160–5166 (1990). [Google Scholar]
- (24).Grennberg H, Gogoll A, Bäckvall JE, Acid-induced transformation of palladium(0)-benzoquinone complexes to palladium(II) and hydroquinone. Organometallics. 12, 1790–1793 (1993). [Google Scholar]
- (25).Popp BV, Stahl SS, “Oxidatively induced” reductive elimination of dioxygen from an η2-peroxopalladium(II) complex promoted by electron-deficient alkenes. J. Am. Chem. Soc 128, 2804–2805 (2006). [DOI] [PubMed] [Google Scholar]
- (26).Engle KM, The mechanism of palladium(II)-mediated C–H cleavage with mono-N-protected amino acid (MPAA) ligands: Origins of rate acceleration. Pure Appl. Chem 88, 4520–20 (2016). [Google Scholar]
- (27).Salazar CA, Gair JJ, Flesch KN, Guzei IA, Lewis JC, Stahl SS, Catalytic behavior of mono-N-protected amino-acid ligands in ligand-accelerated C−H activation by palladium(II). Angew. Chem. Int. Ed 59, 10873–10877 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (28).Lennox AJJ, Lloyd-Jones GC, Selection of boron reagents for Suzuki–Miyaura coupling. Chem. Soc. Rev 43, 412–443 (2014). [DOI] [PubMed] [Google Scholar]
- (29).Hull KL, Sanford MS, Catalytic and highly regioselective cross-coupling of aromatic C−H substrates. J. Am. Chem. Soc 129, 11904–11905 (2007). [DOI] [PubMed] [Google Scholar]
- (30).Sanhueza IA, Wagner AM, Sanford MS, Schoenebeck F, On the role of anionic ligands in the site-selectivity of oxidative C-H functionalization reactions of arenes. Chem. Sci 4, 2767–2775 (2013). [Google Scholar]
- (31).Vasseur A, Muzart J, Le Bras J, Ubiquitous Benzoquinones, Multitalented Compounds for Palladium-Catalyzed Oxidative Reactions. Eur. J. Org. Chem 2015, 4053–4069 (2015). [Google Scholar]
- (32).Adamo C, Amatore C, Ciofini I, Jutand A, Lakmini H, Mechanism of the Palladium-Catalyzed Homocoupling of Arylboronic Acids: Key Involvement of a Palladium Peroxo Complex. J. Am. Chem. Soc 128, 6829–6836 (2006). [DOI] [PubMed] [Google Scholar]
- (33).Takahashi M, Masui K, Sekiguchi H, Kobayashi N, Mori A, Funahashi M, Tamaoki N, Palladium-catalyzed C−H homocoupling of bromothiophene derivatives and synthetic application to well-defined oligothiophenes. J. Am. Chem. Soc 128, 10930–10933 (2006). [DOI] [PubMed] [Google Scholar]
- (34).Hull KL, Sanford MS, Catalytic and highly regioselective cross-coupling of aromatic C−H substrates. J. Am. Chem. Soc 129, 11904–11905 (2007). [DOI] [PubMed] [Google Scholar]
- (35).Stuart DR, Villemure E, Fagnou K, Elements of regiocontrol in palladium-catalyzed oxidative arene cross-coupling. J. Am. Chem. Soc 129, 12072–12073 (2007). [DOI] [PubMed] [Google Scholar]
- (36).Cho SH, Hwang SJ, Chang S, Palladium-catalyzed C−H functionalization of pyridine N-oxides: Highly selective alkenylation and direct arylation with unactivated arenes. J. Am. Chem. Soc 130, 9254–9256 (2008). [DOI] [PubMed] [Google Scholar]
- (37).He C-Y, Fan S, Zhang X, Pd-catalyzed oxidative cross-coupling of perfluoroarenes with aromatic heterocycles. J. Am. Chem. Soc 132, 12850–12852 (2010). [DOI] [PubMed] [Google Scholar]
- (38).Zhao X, Yeung CS, Dong VM, Palladium-catalyzed ortho-arylation of o-phenylcarbamates with simple arenes and sodium persulfate. J. Am. Chem. Soc 132, 5837–5844 (2010). [DOI] [PubMed] [Google Scholar]
- (39).Kirchberg S, Tani S, Ueda K, Yamaguchi J, Studer A, Itami K, Oxidative biaryl coupling of thiophenes and thiazoles with arylboronic acids through palladium catalysis: otherwise difficult C4selective C-H arylation enabled by boronic acids. Angew. Chem. Int. Ed 50, 2387–2391 (2011). [DOI] [PubMed] [Google Scholar]
- (40).Mandal D, Yamaguchi AD, Yamaguchi J, Itami K, Synthesis of Dragmacidin D via direct C–H couplings. J. Am. Chem. Soc 133, 19660–19663 (2011). [DOI] [PubMed] [Google Scholar]
- (41).Pintori DG, Greaney MF, Intramolecular oxidative C−H coupling for medium-ring synthesis. J. Am. Chem. Soc 133, 1209–1211 (2011). [DOI] [PubMed] [Google Scholar]
- (42).Wang Z, Li K, Zhao D, Lan J, You J, Palladium-catalyzed oxidative C–H/C–H cross-coupling of indoles and pyrroles with heteroarenes. Angew. Chem. Int. Ed 50, 5365–5369 (2011). [DOI] [PubMed] [Google Scholar]
- (43).Gao D-W, Shi Y-C, Gu Q, Zhao Z-L, You S-L, Enantioselective synthesis of planar chiral ferrocenes via palladium-catalyzed direct coupling with arylboronic acids. J. Am. Chem. Soc 135, 86–89 (2013). [DOI] [PubMed] [Google Scholar]
- (44).Thuy-Boun PS, Villa G, Dang D, Richardson P, Su S, Yu J-Q, Ligand-accelerated ortho-C–H alkylation of arylcarboxylic acids using alkyl boron reagents. J. Am. Chem. Soc 135, 17508–17513 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (45).Xiao K-J, Lin DW, Miura M, Zhu R-Y, Gong W, Wasa M, Yu J-Q, Palladium(II)-catalyzed enantioselective C(sp3)–H activation using a chiral hydroxamic acid ligand. J. Am. Chem. Soc 136, 8138–8142 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (46).He C, Gaunt MJ, Ligand-enabled catalytic C–H arylation of aliphatic amines by a four-membered-ring cyclopalladation pathway. Angew. Chem. Int. Ed 54, 15840–15844 (2015). [DOI] [PubMed] [Google Scholar]
- (47).Cheng Y, Li G, Liu Y, Shi Y, Gao G, Wu D, Lan J, You J, Unparalleled ease of access to a library of biheteroaryl fluorophores via oxidative cross-coupling reactions: Discovery of photostable NIR probe for mitochondria. J. Am. Chem. Soc 138, 4730–4738 (2016). [DOI] [PubMed] [Google Scholar]
- (48).Gao D-W, Gu Q, You S-L, An enantioselective oxidative C–H/C–H cross-coupling reaction: Highly efficient method to prepare planar chiral ferrocenes. J. Am. Chem. Soc 138, 2544–2547 (2016). [DOI] [PubMed] [Google Scholar]
- (49).Xiao K-J, Chu L, Chen G, Yu J-Q, Kinetic resolution of benzylamines via palladium(II)-catalyzed C–H cross-coupling. J. Am. Chem. Soc 138, 7796–7800 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Qi C, Wang W, Reichl KD, McNeely J, Porco JA, Total synthesis of Aurofusarin: Studies on the atropisomeric stability of bis-naphthoquinones. Angew. Chem. Int. Ed 57, 2101–2104 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (51).Shao Q, Wu Q-F, He J, Yu J-Q, Enantioselective γ-C(sp3)–H activation of alkyl amines via Pd(II)/Pd(0) catalysis. J. Am. Chem. Soc 140, 5322–5325 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (52).Hu L, Shen P-X, Shao Q, Hong K, Qiao JX, Yu J-Q, PdII-catalyzed enantioselective C(sp3)−H activation/cross-coupling reactions of free carboxylic acids. Angew. Chem. Int. Ed 58, 2134–2138 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (53).Ye M, Gao G-L, Yu J-Q, Ligand-promoted C-3 selective C–H olefination of pyridines with Pd catalysts. J. Am. Chem. Soc 133, 6964–6967 (2011). [DOI] [PubMed] [Google Scholar]
- (54).Yang G, Lindovska P, Zhu D, Kim J, Wang P, Tang R-Y, Movassaghi M, Yu J-Q, Pd(II)-catalyzed meta-C–H olefination, arylation, and acetoxylation of indolines using a U-shaped template. J. Am. Chem. Soc 136, 10807–10813 (2014). [DOI] [PubMed] [Google Scholar]
- (55).He J, Li S, Deng Y, Fu H, Laforteza BN, Spangler JE, Homs A, Yu J-Q, Ligand-controlled C(sp3)–H arylation and olefination in synthesis of unnatural chiral α–amino acids. Science. 343, 1216 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (56).Deb A, Bag S, Kancherla R, Maiti D, Palladium-catalyzed aryl C–H olefination with unactivated, aliphatic alkenes. J. Am. Chem. Soc 136, 13602–13605 (2014). [DOI] [PubMed] [Google Scholar]
- (57).Calleja J, Pla D, Gorman TW, Domingo V, Haffemayer B, Gaunt MJ, A steric tethering approach enables palladium-catalysed C–H activation of primary amino alcohols. Nat. Chem 7, 1009–1016 (2015). [DOI] [PubMed] [Google Scholar]
- (58).Jiang H, He J, Liu T, Yu J-Q, Ligand-enabled γ-C(sp3)–H olefination of amines: en route to pyrrolidines. J. Am. Chem. Soc 138, 2055–2059 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (59).Liang Q-J, Yang C, Meng F-F, Jiang B, Xu Y-H, Loh T-P, Chelation versus non-chelation control in the stereoselective alkenyl sp2 C−H bond functionalization reaction. Angew. Chem. Int. Ed 56, 5091–5095 (2017). [DOI] [PubMed] [Google Scholar]
- (60).Gorsline BJ, Wang L, Ren P, Carrow BP, C–H alkenylation of heteroarenes: Mechanism, rate, and selectivity changes enabled by thioether ligands. J. Am. Chem. Soc 139, 9605–9614 (2017). [DOI] [PubMed] [Google Scholar]
- (61).Zhang Z, Tanaka K, Yu J-Q, Remote site-selective C–H activation directed by a catalytic bifunctional template. Nature. 543, 538–542 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (62).Wang P, Verma P, Xia G, Shi J, Qiao JX, Tao S, Cheng PTW, Poss MA, Farmer ME, Yeung K-S, Yu J-Q, Ligand-accelerated non-directed C–H functionalization of arenes. Nature. 551, 489–493 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (63).Zhuang Z, Yu C-B, Chen G, Wu Q-F, Hsiao Y, Joe CL, Qiao JX, Poss MA, Yu J-Q, Ligand-enabled β-C(sp3)–H olefination of free carboxylic acids. J. Am. Chem. Soc 140, 10363–10367 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (64).Chen H, Wedi P, Meyer T, Tavakoli G, van Gemmeren M, Dual ligand-enabled nondirected C−H olefination of arenes. Angew. Chem. Int. Ed 57, 2497–2501 (2018). [DOI] [PubMed] [Google Scholar]
- (65).Achar TK, Zhang X, Mondal R, Shanavas MS, Maiti S, Maity S, Pal N, Paton RS, Maiti D, Palladium-catalyzed directed meta-selective C−H allylation of arenes: Unactivated internal olefins as allyl surrogates. Angew. Chem. Int. Ed 58, 10353–10360 (2019). [DOI] [PubMed] [Google Scholar]
- (66).Ghosh KK, Uttry A, Mondal A, Ghiringhelli F, Wedi P, van Gemmeren M, Ligand-enabled γ-C(sp3)−H olefination of free carboxylic acids. Angew. Chem. Int. Ed (2020), doi: 10.1002/anie.202002362. [DOI] [PMC free article] [PubMed]
- (67).Fan Z, Bay KL, Chen X, Zhuang Z, Park HS, Yeung K-S, Houk KN, Yu J-Q, Rational development of remote C−H functionalization of biphenyl: Experimental and computational studies. Angew. Chem. Int. Ed 59, 4770–4777 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (68).Osterberg PM, Niemeier JK, Welch CJ, Hawkins JM, Martinelli JR, Johnson TE, Root TW, Stahl SS, Experimental limiting oxygen concentrations for nine organic solvents at temperatures and pressures relevant to aerobic oxidations in the pharmaceutical industry. Org. Process Res. Dev 19, 1537–1543 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (69).Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Scalmani G, Barone V, Mennucci B, Petersson GA, Nakatsuji H, Caricato M, Li X, Hratchian HP, Izmaylov AF, Bloino J, Zheng G, Sonnenberg JL, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Vreven T, Montgomery JA Jr., Peralta JE, Ogliaro F, Bearpark M, Heyd JJ, Brothers E, Kudin KN, Staroverov VN, Keith T, Kobayashi R, Normand J, Raghavachari K, Rendell A, Burant JC, Iyengar SS, Tomasi J, Cossi M, Rega N, Millam JM, Klene M, Knox JE, Cross JB, Bakken V, Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Martin RL, Morokuma K, Zakrzewski VG, Voth GA, Salvador P, Dannenberg JJ, Dapprich S, Daniels AD, Farkas O, Foresman JB, Ortiz JV, Cioslowski J, and Fox DJ, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT, (2009). [Google Scholar]
- (70).Hay PJ, Wadt WR, Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J. Chem. Phys 82, 270–283 (1985). [Google Scholar]
- (71).Wadt WR, Hay PJ, Ab initio effective core potentials for molecular calculations. Potentials for main group elements Na to Bi. J. Chem. Phys 82, 284–298 (1985). [Google Scholar]
- (72).Becke AD, Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 38, 3098–3100 (1988). [DOI] [PubMed] [Google Scholar]
- (73).Lee C, Yang W, Parr RG, Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 37, 785–789 (1988). [DOI] [PubMed] [Google Scholar]
- (74).Becke AD, A new mixing of Hartree–Fock and local density-functional theories. J. Chem. Phys 98, 1372–1377 (1993). [Google Scholar]
- (75).Grimme S, Antony J, Ehrlich S, Krieg H, A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys 132, 154104 (2010). [DOI] [PubMed] [Google Scholar]
- (76).Cancès E, Mennucci B, Tomasi J, A new integral equation formalism for the polarizable continuum model: Theoretical background and applications to isotropic and anisotropic dielectrics. J. Chem. Phys 107, 3032–3041 (1997). [Google Scholar]
- (77).Mennucci B, Tomasi J, Continuum solvation models: A new approach to the problem of solute’s charge distribution and cavity boundaries. J. Chem. Phys 106, 5151–5158 (1997). [Google Scholar]
- (78).Scalmani G, Frisch MJ, Continuous surface charge polarizable continuum models of solvation. I. General formalism. J. Chem. Phys 132, 114110 (2010). [DOI] [PubMed] [Google Scholar]
- (79).Dolg M, Wedig U, Stoll H, Preuss H, Energy-adjusted ab initio pseudopotentials for the first row transition elements. J. Chem. Phys 86, 866–872 (1987). [Google Scholar]
- (80).Simonetti M, Cannas DM, Panigrahi A, Kujawa S, Kryjewski M, Xie P, Larrosa I, Ruthenium-Catalyzed C−H Arylation of Benzoic Acids and Indole Carboxylic Acids with Aryl Halides. Chem. Eur. J 23, 549–553 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- (81).Cai G, Fu Y, Li Y, Wan X, Shi Z, Indirect ortho Functionalization of Substituted Toluenes through ortho Olefination of N, N-Dimethylbenzylamines Tuned by the Acidity of Reaction Conditions. J. Am. Chem. Soc 129, 7666–7673 (2007). [DOI] [PubMed] [Google Scholar]
- (82).Zhang J-C, Shi J-L, Wang B-Q, Hu P, Zhao K-Q, Shi Z-J, Direct Oxidative Arylation of Aryl C–H Bonds with Aryl Boronic Acids via Pd Catalysis Directed by the N,N-Dimethylaminomethyl Group. Chem. Asian J 10, 840–843 (2015). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




