Decarboxylative Acylation of Carboxylic Acids: Reaction Investigation and Mechanistic Study

酰化 化学 羧酸 有机化学 催化作用
作者
Xiaopeng Wu,Jie Han,Siyu Xia,Weipeng Li,Chengjian Zhu,Jin Xie
出处
期刊:CCS Chemistry [Chinese Chemical Society]
卷期号:4 (7): 2469-2480 被引量:19
标识
DOI:10.31635/ccschem.021.202101197
摘要

Open AccessCCS ChemistryRESEARCH ARTICLE14 Jul 2022Decarboxylative Acylation of Carboxylic Acids: Reaction Investigation and Mechanistic Study Xiaopeng Wu†, Jie Han†, Siyu Xia, Weipeng Li, Chengjian Zhu and Jin Xie Xiaopeng Wu† State Key Laboratory of Coordination Chemistry, Jiangsu Key Laboratory of Advanced Organic Materials, Chemistry and Biomedicine Innovation Center (ChemBIC), School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023 , Jie Han† State Key Laboratory of Coordination Chemistry, Jiangsu Key Laboratory of Advanced Organic Materials, Chemistry and Biomedicine Innovation Center (ChemBIC), School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023 , Siyu Xia State Key Laboratory of Coordination Chemistry, Jiangsu Key Laboratory of Advanced Organic Materials, Chemistry and Biomedicine Innovation Center (ChemBIC), School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023 , Weipeng Li *Corresponding authors: E-mail Address: [email protected] E-mail Address: [email protected] E-mail Address: [email protected] State Key Laboratory of Coordination Chemistry, Jiangsu Key Laboratory of Advanced Organic Materials, Chemistry and Biomedicine Innovation Center (ChemBIC), School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023 , Chengjian Zhu *Corresponding authors: E-mail Address: [email protected] E-mail Address: [email protected] E-mail Address: [email protected] State Key Laboratory of Coordination Chemistry, Jiangsu Key Laboratory of Advanced Organic Materials, Chemistry and Biomedicine Innovation Center (ChemBIC), School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023 State Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, Shanghai 200032 and Jin Xie *Corresponding authors: E-mail Address: [email protected] E-mail Address: [email protected] E-mail Address: [email protected] State Key Laboratory of Coordination Chemistry, Jiangsu Key Laboratory of Advanced Organic Materials, Chemistry and Biomedicine Innovation Center (ChemBIC), School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023 Advanced Catalytic Engineering Research Center of the Ministry of Education, Hunan University, Changsha 410082 https://doi.org/10.31635/ccschem.021.202101197 SectionsSupplemental MaterialAboutAbstractPDF ToolsAdd to favoritesDownload CitationsTrack Citations ShareFacebookTwitterLinked InEmail Ketones serve as one of the most critical building blocks in organic synthesis, involving numerous functional group transformations. Herein, we report an unprecedented photoredox–nickel metallaphotoredox-catalyzed decarboxylative acylation of common aliphatic acids with readily available aromatic and aliphatic thioesters. A wide range of structurally diverse asymmetrical aryl alkyl and dialkyl ketones have been constructed in yields of up to 98% with this strategy. The protocol has excellent reaction selectivity and functional group compatibility, representing a significant step forward in ketone synthesis. The one-pot decarboxylative acylation at the gram scale from two different carboxylic acids and the late-stage application for the synthesis of complex ketones shows its synthetic robustness. Both mechanistic experiments and density functional theory (DFT) calculations suggest that the decarboxylative acylation reaction operates via an underdeveloped Ni(I)–Ni(II)–Ni(I)–Ni(III)–Ni(I) catalytic cycle. Download figure Download PowerPoint Introduction The ketone moiety is involved in many versatile synthetic transformations in organic synthesis and regarded as a core linkage in retrosynthetic analysis.1 In view of the availability and structural diversity of carboxylic acids, the development of new methods of ketone synthesis from carboxylic acids attracts continuing interest. Examples of this type of reaction include Friedel–Crafts acylation,2,3 nucleophilic addition of organometallic compounds to amides,4 and transition-metal-catalyzed cross-coupling reactions.5–10 The combination of photoredox and nickel catalysis recently initiated by MacMillan et al.11 and Molander et al.12 has led to advances in redox-neutral chemical bond formation.13–20 In this context, several research groups21–24 have reported elegant examples of metallaphotoredox-catalyzed C(sp3)-H acylation with amides, anhydrides, or activated 2-pyridylthioesters as acyl surrogates. In these reactions, substrates such as compounds with an α-C(sp3)-H adjacent to a heteroatom or simple alkanes have been employed in a facile hydrogen atom transfer (HAT) process to control the regioselectivity. There is still, however, a great demand for the development of a practical protocol using a readily available chemical as an alkyl radical source,25–27 with the aim to construct complex ketones without redundant protection and deprotection processes. Carboxylic acids can realize a synthetase-catalyzed formal decarboxylative acylation by transfer of an acyl group from coenzyme A (CoA)-activated thioesters to deliver a chain-extended ketone (Scheme 1a).28 We questioned if it is possible to use metallaphotoredox catalysis with readily available thioesters to develop a decarboxylative acylation strategy of aliphatic carboxylic acids. In such a reaction, a transition-metal catalytic unit is responsible for the activation of thioesters (Scheme 1b). Additionally, in recent years, visible-light-mediated radical decarboxylation is a powerful strategy for forming C–C or C–X bonds,29–36 such as decarboxylative alkynylation, alkenylation, alkylation, arylation, fluorination, and borylation. However, the decarboxylative acylation of aliphatic carboxylic acids has not been reported to date.37 Given the structural diversity of carboxylic acids, the success of a metallaphotoredox-catalyzed decarboxylative acylation protocol could potentially enhance the synthesis of complex ketones. More significantly, decarboxylative acylation can be used to construct ketones from two different carboxylic acids in a one-pot operation, thus expanding the existing ketone synthetic routes. Scheme 1 | Synthesis of ketones from two carboxylic acids by decarboxylative acylation. Download figure Download PowerPoint Based on these assumptions, a plausible mechanism is proposed in Scheme 2. Initially, irradiation of the iridium(III) photocatalyst (PC) Ir[dF(CF3)ppy]2(dtbbpy)PF6 produces a photoexcited *IrIII species. A carboxylate anion could participate in a single electron transfer (SET) with the photoexcited PC [E1/2red [*IrIII/IrII] = + 1.21 V vs saturated calomel electrode (SCE)]38,39 to generate an alkyl radical (I) and CO2. In the presence of N-containing ligands, the nickel-based catalyst precursor can be reduced to an LNiX species (II),40 which undergoes radical recombination with generated alkyl radical (I) to produce alkyl-Ni(II) species (III). This can accept an additional electron from the Ir(II) species (E [IrII/IrIII] = −1.37 V vs SCE)38,39 to furnish a highly active R–Ni(I) species (IV).41 Subsequent oxidative addition of the thioester C–S bond to the R–Ni(I) species forms the Ni(III) species (V).42 Finally, reductive elimination would form the requisite C–C bond delivering the desired ketone product, completing the catalytic cycle. Scheme 2 | Proposed mechanism of decarboxylative acylation. Download figure Download PowerPoint Experimental Methods General procedure for ketone synthesis Ir[dF(CF3)ppy]2(dtbbpy)PF6 (2.2 mg, 2.0 mol %), NiCl2·1,2-dimethoxyethane (NiCl2·DME) (2.2 mg, 10 mol %), 4,7-dimethyl-1,10-phenanthroline (4.2 mg, 20 mol %), carboxylic acid (0.3 mmol, 3.0 equiv), and anhydrous powder Cs2CO3 (97.5 mg, 0.3 mmol, 3.0 equiv) were added to an oven-dried 10 mL Schlenk tube equipped with a poly(tetrafluoroethylene) (PTFE) stir bar. The tube was evacuated and backfilled with argon three times. Subsequently, the thioester (0.1 mmol, 1.0 equiv) and CH3CN (3.0 mL) were added into this Schlenk tube under argon. The tube was then sealed and placed approximately 5 cm from two 45-W blue light-emitting diodes (LEDs) (λmax = 455 nm). The reaction mixture was stirred for 21–48 h at room temperature (rt) (a fan and air-conditioners were used to keep the reaction temperature close to 25 °C). After completion, the reaction mixture was removed from the light, diluted with water, and the aqueous layer was extracted with EtOAc (3 × 2.0 mL). The combined organic layers were washed with brine, dried over anhydrous Na2SO4, filtered, and concentrated. The residue was purified by flash chromatography on silica gel to afford the corresponding ketone products. 1a, R = nBu; the other substrates, R = tBu, unless otherwise noted. Density functional theory calculation All the calculations were carried out using density functional theory (DFT) and an ultrafine grid as implemented in Gaussian16 program package. Gibbs free energies were calculated at PBE0-D3/def2-SVP/def2-TZVP/SMD(CH3CN) level of theory (see the detailed computational methods and reference in the Supporting Information). Results and Discussion Initially, the reaction of 3-(3,4-dimethoxyphenyl)propanoic acid ( 1a) and S-butyl benzothioate ( 2a) was selected for a model reaction and to optimize the standard reaction conditions. As shown in Table 1, the optimized reaction conditions include the use of 2 mol % of Ir[dF(CF3)ppy]2(dtbbpy)PF6 as the PC, 10 mol % of the nickel catalyst together with 20 mol % of 4,7-dimethyl-1,10-phenanthroline ( L4) as the ligand in CH3CN. These conditions deliver the desired ketone ( 3a) in 98% yield under the irradiation of blue LEDs at rt. The other PCs screened failed to promote the decarboxylative acylation possibly because the strong oxidation potential of the photoexcited *Ir-species can facilitate the single electron oxidation of aliphatic acids (entries 2 and 3, Table 1). Phenanthrolines are better as ligands than substituted bipyridines, and among the ligands examined 4,7-dimethyl-1,10-phenanthroline ( L4) gave good yields (entries 4–7, Table 1). Control experiments (entries 8–10, Table 1) showed that PC, base, nickel catalyst, and ligand are indispensable for the decarboxylative acylation reaction. Table 1 | Optimization of Reaction Conditionsa Entry Variation of Standard Conditions Yield (%)b 1 None 98 (81) 2 PC2 instead of PC1 70 (53) 3 PC3, PC4 or PC5 instead of PC1 ND 4 L1 instead of L4 36 5 L2 instead of L4 Trace 6 L3 instead of L4 76 (55) 7 L5 instead of L4 ND 8 Without PC ND 9 Without nickel/ligand ND 10 Without base ND aStandard reaction conditions: Ir[dF(CF3)ppy]2(dtbbpy)PF6 (2 mol %), NiCl2·DME (10 mol %), 4,7-dimethyl-1,10-phenanthroline ( L4) (20 mol %), 1a (0.1 mmol), 2a (0.3 mmol), Cs2CO3 (3.0 equiv), CH3CN (3 mL), blue LEDs, 21 h, rt. ND, not detected. bGas chromatography (GC) yield using biphenyl as an internal standard and isolated yield is shown in parentheses. With the optimal reaction conditions in hand, we examined the scope of commercial carboxylic acids that can construct aryl alkyl ketones with this acylation reaction (Scheme 3). A variety of carboxylic acids bearing electron-rich and -poor functional groups on the phenyl ring undergo decarboxylative acylation smoothly, giving rise to the corresponding ketones in moderate to excellent yields of up to 98% ( 3a– 3y). The mild reaction conditions accommodate an array of useful functional groups including ethers ( 3c– 3f and 3h), an ester ( 3w), amides ( 3u and 3v), and an acetal ( 3g). Strongly electron-withdrawing groups like methylsulfonyl ( 3p) and ester ( 3w) are well tolerated. Halogen- (F, Cl, and Br) and trifluoromethyl (CF3)-substituted phenylacetic acids ( 3j– 3m, 47–61%) are compatible with the reaction process, which opens the possibility of further downstream conversion. Notably, carboxylic acid substrates containing aryl bromide ( 3k) and free N–H groups ( 3q and 3r), which are known to be sensitive in traditional metal-catalyzed cross-coupling reactions, appear to be tolerated, and heterocyclic carboxylic acids are also competent coupling starting materials ( 3r– 3t). Biologically important molecules such as actarit ( 3v, 73%), indomethacin ( 3x, 77%), repaglinide precursor ( 3w, 49%), and zolpidem precursor ( 3y, 82%) can also be used in this reaction. Additionally, α-ketocarboxylic acids ( 3z), piperidine formic acid ( 3aa), pyrrolidine formic acid ( 3bb), proline ( 3cc), and azacyclobutanoic acid ( 3dd) give the corresponding aryl alkyl ketones in 40–60% yields. Unfortunately, the employment of a tertiary carboxylic acid failed to give rise to the desired product ( 3mm). In general, direct decarboxylation generating an alkyl radical can ensure good regioselectivity and complement the reported C–H acylation protocols.21–24 Scheme 3 | Scope of aryl alkyl ketones. Standard reaction conditions: Ir[dF(CF3)ppy]2(dtbbpy)PF6 (2 mol %), NiCl2·DME (10 mol %), 4,7-dimethyl-1,10-phenanthroline (L4) (20 mol %), carboxylic acid (1) (0.3 mmol), thioesters (0.1 mmol), Cs2CO3 (0.3 mmol), CH3CN (3 mL), blue LEDs, 21 h, rt. Isolated yield is shown. aYield determined by GC-MS using biphenyl as an internal standard. b–StBu instead of –SnBu. Download figure Download PowerPoint We also evaluated the nature of thioesters in this photoredox–nickel mediated cross-coupling reaction. As shown in Scheme 3, both electron-rich and -deficient groups on the phenyl ring have little effect on the insertion of the C–S bond and produce the desired ketones ( 3ee– 3ll) in moderate to good yields. Thioesters containing heterocycles such as furan and thiophene can participate readily in this acylation, furnishing the desired products ( 3jj and 3kk). In view of our previous aryl alkyl ketone synthesis routes,43,44 the flexibility of the alkyl source is a useful addition. Construction of more complex asymmetrical dialkyl ketones was attempted. As shown in Scheme 4, the commercially available short-chain thioesters could be converted into the desired alkyl ketones ( 4a– 4c) in yields as high as 91%. The α-substituted thioesters ( 4d) also gave acceptable yields. Aliphatic thioesters containing a carbonyl group are compatible, providing the 1,5-dione ( 4g) or the 1,7-dione ( 4h) in 95% and 73% yield, respectively. This can enhance the classical Weinreb ketone synthesis, which entails reactive ketones being attacked by organometallics. Thioesters prepared from cyclic carboxylic acids such as indene carboxylic acid ( 4l), cyclohexane formic acid ( 4j) and small-ring carboxylic acids ( 4i, 85%; 4k, 84%) also work well. The use of other substituted thioesters in place of S-(tert-butyl)ethane thioate afford the desired product ( 4a) in moderate yields, indicating structural changes in the thioester are possible for metallaphotoredox decarboxylative acylation. Scheme 4 | Scope of asymmetrical alkyl ketones. Standard reaction conditions: Ir[dF(CF3)ppy]2(dtbbpy)PF6 (2 mol %), NiCl2·DME (10 mol %), 4,7-dimethyl-1,10-phenanthroline (L4) (20 mol %), carboxylic acid (0.3 mmol), thioesters (0.1 mmol), Cs2CO3 (0.3 mmol), CH3CN (3 mL), blue LEDs, 21 h, rt. Isolated yield is shown. a–SEt instead of –StBu in thioester. b–SMe instead of –StBu in thioester. Download figure Download PowerPoint The high efficiency and good functional group tolerance of this photoredox–nickel-catalyzed reaction encouraged us to explore its late-stage application in the synthesis of complex ketones (Scheme 5). Some drug molecules such as ketoprofen ( 5a), loxoprofen ( 5b), oxaprozin ( 5c), and indomethacin ( 5f and 5h) selectively produce the desired ketones in moderate yield under standard conditions. When complex aliphatic acids and the corresponding thioesters were subjected to this protocol, they successfully constructed the related dialkyl ketones ( 5f– 5i). Scheme 5 | Late-stage application of the reaction for the synthesis of complex ketones. Reaction conditions: Ir[dF(CF3)ppy]2(dtbbpy)PF6 (2 mol %), NiCl2·DME (10 mol %), 4,7-dimethyl-1,10-phenanthroline (L4) (20 mol %), carboxylic acid (0.3 mmol), thioesters (0.1 mmol), Cs2CO3 (0.3 mmol), CH3CN (3 mL), blue LEDs, 21 h, rt. Isolated yield is shown. The structure in black represents the thioesters, blue represents the carboxylic acids, and the bond in red represents the newly formed bond. Download figure Download PowerPoint To further demonstrate the practicality of this protocol, experiments on gram scale were carried out (Scheme 6). The corresponding thioesters were prepared in situ by treatment of carboxylic acids with oxalyl chloride and thiols, and were used directly without purification in the subsequent decarboxylative acylation procedure. The synthetic route provides satisfactory access to linkages in the form of ketones connecting two functional fragments from different carboxylic acids. Scheme 6 | Gram-scale ketone synthesis in one-pot operation. Reaction conditions: (1) RCOOH (1.0 equiv), oxalyl chloride (1.5 equiv), CH2Cl2, 0 °C to rt. (2) thiol (1.5 equiv), CH2Cl2, rt. (3) Ir[dF(CF3)ppy]2(dtbbpy)PF6 (0.5 mol %), NiCl2·DME (10 mol %), L4 (20 mol %), RCOOH (1.5 equiv), Cs2CO3 (3.0 equiv), CH3CN (0.133 M), blue LEDs, 48 h, rt. Download figure Download PowerPoint Cyclic voltammetry (CV) measurements of 2a showed that the high oxidation potential of carboxylic acids ( 2a) is greatly diminished after the formation in situ of cesium salts (E = +0.72 V vs SCE). Consequently, they can be oxidized by an excited-state *IrIII-complex as shown in Scheme 2 (E1/2red [*IrIII/IrII] = +1.21 V vs SCE). Our luminescence quenching experimental results further indicate that the photoexcited *IrIII-complex is quenched by carboxylate cesium salts (see Supporting Information for details). A free radical trapping experiment with 2,2,6,6-tetramethylpiperidin-1-oxyl (TEMPO) showed that a radical process was highly likely (Scheme 7a), and this was further verified by electron paramagnetic resonance (EPR) experiments with N-tert-butyl-α-phenylnitrone (Scheme 7b). Interestingly, the control experiments with a prepared nickel(II) acyl intermediate ( 6) and a carboxylic acid ( 2c) under standard conditions could not successfully produce the target product. Only the decarboxylation byproduct of the carboxylic acid ( 2c) could be detected (Scheme 7c), suggesting that the radical addition to a Ni(II) species42,45 in the decarboxylative acylation is less likely. Scheme 7 | (a–f) Control experiments. HRMS, high-resolution mass spectrometry. Download figure Download PowerPoint We wished to experimentally control the sequence of the two processes, oxidative addition and radical addition, by varying the order of how the reaction components are added. Combining the thioester ( 8) and stoichiometric amounts of Ni(cod)2 (Scheme 7d) forces thioester ( 8) to undergo the oxidative addition first, leading to a low yield of expected product (16%). Delaying addition of the thioester ( 8) after prestirring of the Ni(II) catalyst and carboxylic acid ( 2c) under irradiation pre-generates the alkyl nickel(I) species, driving the reaction to the radical addition pathway first (Scheme 7e). As a result, continuous irradiation (2–10 min) promotes the radical addition into the nickel center to ensure the equivalent conversion to product (85–92%), while prolonged irradiation (30 min) may lead to a large amount of byproduct 7. Interestingly, without light irradiation, the in situ generated alkyl nickel species also underwent oxidation addition with the thioester, but only generated the desired product in trace yield (Scheme 7f). Thus, the pathway of initial radical addition is more feasible than initial oxidation addition. With the strong experimental evidences of our proposed mechanism, we further explored various mechanistic possibilities through quantum chemical calculations performed on selected model reactions as shown in Figures 1 and 2. Formation of 3a and 3c have been considered for comparison. All the calculations were carried out using DFT and an ultrafine grid as implemented in Gaussian16 program package (see Supporting Information for the full reference and computational details). Figure 1 | Free energy profiles of paths 1 and 2 for the nickel catalytic cycle of Scheme 2. Gibbs free energies were calculated at PBE0-D3/def2-SVP/def2-TZVP/SMD(CH3CN) level of theory. For each pathway, only MEP is shown here. LE, ligand exchange; MEP, minimum energy path; OA, oxidative addition; RA, radical addition; RE, reductive elimination. Download figure Download PowerPoint Based on the previous mechanism, we considered a reductive quenching cycle for the photoredox catalytic cycle Ir(III)−*Ir(III)–Ir(II) sequence of iridium oxidation states (Scheme 2), as well as two possible reaction pathways for the nickel catalytic cycle with a Ni(I) species as the active catalyst (Figures 1 and 2), that is, Ni(I)−Ni(0)−Ni(II)−Ni(III) (paths 1 and 1′) and Ni(I)−Ni(II)−Ni(I)−Ni(III) (path 2 and 2′). Both Ni(I)–Cl and Ni(I)–SnBu might serve as the active catalysts in the nickel catalytic cycle. Ni(I)–SnBu is excluded since the activation energies of all its relevant SET steps in the photoredox cycle are ∼10 kcal/mol higher than that of Ni(I)–Cl (see Supporting Information Figure S18). Regarding the formations of radicals 10 (unreactive alkyl radical) and 9 (benzylic radical), the calculated free energy barriers for the corresponding SET steps are negligible (0.0 and 1.1 kcal/mol). However, the relative values of their free energy changes (−18.5 vs −29.0 kcal/mol) indicate that the generation of 9 is thermodynamically more favorable. Figure 2 | Free energy profiles of pathways paths 1′ and 2′ for the nickel catalytic cycle of Scheme 2. Gibbs free energies were calculated at PBE0-D3/def2-SVP/def2-TZVP/SMD(CH3CN) level of theory. For each pathway, only MEP is shown here. LE, ligand exchange; OA, oxidative addition; RA, radical addition; RE, reductive elimination. Download figure Download PowerPoint For the acylation involved with 9, in path 1, Ni(I)–Cl initiates with SET and complexation with 1a to form Ni(0) species 1A, the free energy change of the process is 12.1 kcal/mol within the photoredox cycle (SET1, Figure 1). 1A continues to undergo the oxidative addition with 1a to generate Ni(II) species 1B, with an energy barrier as low as 3.5 kcal/mol. 1B subsequently binds to radical 9 via a transition-state 1B-TS, and forms an intermediate 1C. The reaction proceeds with reductive elimination of 1C, resulting in the target product ( 3a) and Ni(I)-SnBu. Finally, the active catalysis is regenerated by ligand exchange between –SnBu and Cl–, with a free energy of 1.4 kcal/mol. The rate-determining step of path 1 is from Ni(I)–Cl to 1A-TS, with an overall 15.6 kcal/mol barrier. As shown in Figure 1, path 2 starts from a barrierless radical addition of Ni(I)-Cl with 9, leading to an intermediate 2A. The Cl− anion is detached from 2A by SET with Ir(II), resulting in Ni(I) species 2B. 2B binds with 1a and then undergoes oxidative addition via 2C and 2C-TS, leading to a Ni(III) intermediate species 2D, which is 7.9 kcal/mol lower than 1C in path 1. The Ni(III) intermediate 2D is lower in energy than 2A, thus the reaction rate of path 2 is determined by the SET step 2A to 2B, with an energy barrier of 10.3 kcal/mol. Since the overall energy barrier of path 1 is higher than that of path 2, path 2 seems to be theoretically more feasible. For the acylation involved with 10, both paths 1′ and 2′ of the reaction are still operative under the reaction conditions. Similarly, path 2′ (12.6 kcal/mol) is energetically more favorable than path 1′ (15.6 kcal/mol). It is noted that radical addition of Ni(I)-Cl and 10 in path 2′ generates a more stable intermediate 2A′ (−14.7 kcal/mol) with a π–π stacked-like motif (Figure 2), whereas the binding of Ni(I)-Cl and 9 in path 2 forms a mostly square planar geometry 2A (−5.8 kcal/mol). Therefore, the radical addition step ( RA2′) of path 2′ is much more exothermic than that ( RA2) of path 2. 2B′ binds with 1a and then undergoes oxidative addition via 2C′ and 2C′-TS, leading to a Ni(III) intermediate 1C′ and merging reaction paths 1′ and 2′. Moreover, the Ni(III) intermediate 1C′ is higher in energy than 2A′, which is contrary to the scenario of path 2 (G( 2D) < G( 2A)). Consequently, the reaction rate of path 2′ is determined from 2A′ to 1C′-TS, with an overall energy barrier of 12.6 kcal/mol. As shown in Figures 1 and 2, path 2 and 2′ are dynamically more favorable than path 1 and 1′, and the more stable benzylic radical can undergo this reaction more easily than unreactive alkyl radical (10.3 kcal/mol in path 2 vs 12.6 kcal/mol in path 2′). Conclusion We have developed the first photoredox–nickel metallaphotoredox-catalyzed decarboxylative acylation of aliphatic carboxylic acids with readily available aromatic and aliphatic thioesters. This protocol allows the processing of a wide range of structurally diverse aryl alkyl ketones and asymmetrical dialkyl ketones, delivering moderate to good yields. This general and practical protocol has excellent chemoselectivity and functional group compatibility, and represents significant progress in decarboxylative functionalization of carboxylic acids. The one-pot decarboxylative acylation on a gram scale from two different carboxylic acids and late-stage application for the synthesis of complex ketones further enhances its synthetic utility. Both experimental and theoretical investigation support that decarboxylative acylation prefers to initiate with radical addition to Ni(I)-species, which suggests an underdeveloped Ni(I)−Ni(II)−Ni(I)−Ni(III)–Ni(I) catalytic cycle in metallaphotoredox catalysis. Supporting Information Supporting Information is available and includes computational details, Figures S1–S19, Tables S1–S7, general procedures, analytic data, gas chromatography mass spectrometry (GC-MS), high-performance liquid chromatography (HPLC), 1H NMR, 13C NMR, and 19F NMR spectra. Conflict of Interest There is no conflict of interest to report. Funding Information The authors are grateful for the financial support from the National Natural Science Foundation of China (grant nos. 22001117, 21971108, 21971111, and 21732003), the Natural Science Foundation of Jiangsu Province (grant no. BK20190006), Fundamental Research Funds for the Central Universities (0205/14380252), "Innovation & Entrepreneurship Talents Plan" of Jiangsu Province, and Foundation of Advanced Catalytic Engineering Research Center of the Ministry of Education of Hunan University. Acknowledgments The authors wish to acknowledge Dongping Wang, Yijie He, Kai Liu, Yantao Li for reproducing the experimental procedures for products ( 3a, 3r, 4g, and 5d). All theoretical calculations were performed at the High-Performance Computing Center (HPCC) of Nanjing University and Prof. Guoqiang Wang at Institute of Theoretical and Computational Chemistry is warmly acknowledged for his help on DFT calculations. References 1. Ma K.; Martin B. S.; Yin X.; Dai M.Natural Product Syntheses via Carbonylative Cyclizations.Nat. Prod. Rep.2019, 36, 174–219. Google Scholar 2. Vekariya R. H.; Aube J.Hexafluoro-2-Propanol-Promoted Intermolecular Friedel-Crafts Acylation Reaction.Org. Lett.2016, 18, 3534–3537. Google Scholar 3. Schmidt N. G.; Pavkov-Keller T.; Richter N.; Wiltschi B.; Gruber K.; Kroutil W.Biocatalytic Friedel–Crafts Acylation and Fries Reaction.Angew. Chem. Int. Ed.2017, 56, 7615–7619. Google Scholar 4. Nahm S.; Weinreb S. M.N-Methoxy-n-Methylamides as Effective Acylating Agents.Tetrahedron Lett.1981, 22, 3815–3818. Google Scholar 5. Buchspies J.; Szostak M.Recent Advances in Acyl Suzuki Cross-Coupling.Catalysts2019, 9, 53–75. Google Scholar 6. Wang J.; Hoerrner M. E.; Watson M. P.; Weix D. J.Nickel-Catalyzed Synthesis of Dialkyl Ketones from the Coupling of N-Alkyl Pyridinium Salts with Activated Carboxylic Acids.Angew. Chem. Inte. Ed.2020, 59, 13484–13489. Google Scholar 7. Ishii T.; Kakeno Y.; Nagao K.; Ohmiya H.N-Heterocyclic Carbene-Catalyzed Decarboxylative Alkylation of Aldehydes.J. Am. Chem. Soc.2019, 141, 3854–3858. Google Scholar 8. Cherney A. H.; Kadunce N. T.; Reisman S. E.Catalytic Asymmetric Reductive Acyl Cross-Coupling: Synthesis of Enantioenriched Acyclic α,α-Disubstituted Ketones.J. Am. Chem. Soc.2013, 135, 7442–7445. Google Scholar 9. Ni S.; Padial M. M.; Kingston C.; Vantourout J. C.; Schmitt D. C.; Edwards J. T.; Kruszyk M. M.; Merchant R. R.; Mykhailiuk P. K.; Sanchez B. B.; Yang S.; Perry M. A.; Gallego G. M.; Mousseau J. J.; Collins M. R.; Cherney R. J.; Lebed P. S.; Chen J. S.; Qin T.; Baran P. S.A Radical Approach to Anionic Chemistry: Synthesis of Ketones, Alcohols, and Amines.J. Am. Chem. Soc.2019, 141, 6726–6739. Google Scholar 10. Yin H.; Zhao C.; You H.; Lin K.; Gong H.Mild Ketone Formation via Ni-Catalyzed Reductive Coupling of Unactivated Alkyl Halides with Acid Anhydrides.Chem. Commun.2012, 48, 7034–7036. Google Scholar 11. Zuo Z.; Ahneman D. T.; Chu L.; Terrett J. A.; Doyle A. G.; MacMillan D. W. C.Merging Photoredox with Nickel Catalysis: Coupling of Alpha-Carboxyl sp3-Carbons with Aryl Halides.Science2014, 345, 437–440. Google Scholar 12. Tellis J. C.; Primer D. N.; Molander G. A.Single-Electron Transmetalation in Organoboron Cross-Coupling by Photoredox/Nickel Dual Catalysis.Science2014, 345, 433–436. Google Scholar 13. Milligan J. A.; Phelan J. P.; Badir S. O.; Molander G. A.Alkyl Carbon–Carbon Bond Formation by Nickel/Photoredox Cross-Coupling.Angew. Chem. Int. Ed.2019, 58, 6152–6163. Google Scholar 14. Yu X.-Y.; Chen J.-R.; Xiao W.-J.Visible Light-Driven Radical-Mediated C-C Bond Cleavage/Functionalization in Organic Synthesis.Chem. Rev.2021, 121, 506–561. Google Scholar 15. Zhu C.; Yue H.; Jia J.; Rueping M.Recent Advances in Nickel-Catalyzed C-Heteroatom Cross-Coupling Reactions under Mild Conditions via Facilitated Reductive Elimination.Angew. Chem. Int. Ed.2021, 60, 17810–17831. Google Scholar 16. Lipp A., Badir S. O.; Molander G. A.Stereoinduction in Metallaphotoredox Catalysis.Angew. Chem. Int. Ed.2021, 60, 1714–1726. Google Scholar 17. Zhu C., Yue H., Maity B., Atodiresei I., Cavallo L.; Rueping M.A Multicomponent Synthesis of Stereodefined Olefins via Nickel Catalysis and Single Electron/Triplet Energy Transfer.Nat. Catal.2019, 2, 678–687. Google Scholar 18. Zhang L.; Si X.; Yang Y.; Zimmer M.; Witzel S.; Sekine K.; Rudolph M.; Hashmi A. S. K.The Combination of Benzaldehyde and Nickel-Catalyzed Photoredox C(sp3)-H Alkylation/Arylation.Angew. Chem. Int. Ed.2019, 58, 1823–1827. Google Scholar 19. Thullen S. M.; Treacy S. M.; Rovis T.Regioselective Alkylative Cross-Coupling of Remote Unactivated C(sp3)-H Bonds.J. Am. Chem. Soc.2019, 141, 14062–14067. Google Scholar 20. Zhang Z.; Hu X.Arylsilylation of Electron-Deficient Alkenes via Cooperative Photoredox and Nickel Catalysis.ACS Catal.2019, 10, 777–782. Google Scholar 21. Joe C. L.; Doyle A. G.Direct Acylation of C(sp3)-H Bonds Enabled by Nickel and Photoredox Catalysis.Angew. Chem. Int. Ed.2016, 55, 4040–4043. Google Scholar 22. Lee G. S.; Won J.; Choi S.; Baik M.-H.; Hong S. H.Synergistic Activation of Amides and Hydrocarbons for Direct C(sp3)–H Acylation Enabled by Metallaphotoredox Catalysis.Angew. Chem. Int. Ed.2020, 59, 16933–16942. Google Scholar 23. Kawasaki T.; Ishida N.; Murakami M.Dehydrogenative Coupling of Benzylic and Aldehydic C-H Bonds.J. Am. Chem. Soc.2020, 142, 3366–3370. Google Scholar 24. Shu X.; Huan L.; Huang Q.; Huo H.Direct Enantioselective C(sp3)–H Acylation for the Synthesis of α-Amino Ketones.J. Am. Chem. Soc.2020, 142, 19058–19064. Google Scholar 25. Crespi S.; Fagnoni M.Generation of Alkyl Radicals: From the Tyranny of Tin to the Photon Democracy.Chem. Rev.2020, 120, 9790–9833. Google Scholar 26. Xu W.; Ma J.; Yuan X. A.; Dai J.; Xie J.; Zhu C.; Synergistic Catalysis for the Umpolung Trifluoromethylthiolation of Tertiary Ethers.Angew. Chem. Int. Ed.2018, 57, 10357–10361. Google Scholar 27. Guo L.; Yuan M.; Zhang Y.; Wang F.; Zhu S.; Gutierrez O.; Chu L.General Method for Enantioselective Three-Component Carboarylation of Alkenes Enabled by Visible-Light Dual Photoredox/Nickel Catalysis.J. Am. Chem. Soc.2020, 142, 20390–20399. Google Scholar 28. Röhrig F.; Schulze A.The Multifaceted Roles of Fatty Acid Synthesis in Cancer.Nat Rev. Cancer2016, 16, 732–749. Google Scholar 29. Yue H.; Zhu C.; Kancherla R.; Liu F.; Rueping M.Regioselective Hydroalkylation and Arylalkylation of Alkynes by Photoredox/Nickel Dual Catalysis: Application and Mechanism.Angew. Chem. Int. Ed.2020, 59, 5738–5746. Google Scholar 30. Cao H.; Jiang H.; Feng H.; Kwan J. M. C.; Liu X.; Wu J.Photo-Induced Decarboxylative Heck-Type Coupling of Unactivated Aliphatic Acids and Terminal Alkenes in the Absence of Sacrificial Hydrogen Acceptors.J. Am. Chem. Soc.2018, 140, 16360–16367. Google Scholar 31. Sun X.; Chen J.; Ritter T.Catalytic Dehydrogenative Decarboxyolefination of Carboxylic Acids.Nat. Chem.2018, 10, 1229–1233. Google Scholar 32. Zhou Q. Q.; Guo W.; Ding W.; Wu X.; Chen X.; Lu L. Q.; Xiao W. J.Decarboxylative Alkynylation and Carbonylative Alkynylation of Carboxylic Acids Enabled by Visible-Light Photoredox Catalysis.Angew. Chem. Int. Ed.2015, 54, 11196–11199. Google Scholar 33. Ernouf G.; Chirkin E.; Rhyman L.; Ramasami P.; Cintrat J.-C.Photochemical Strain-Release-Driven Cyclobutylation of C(sp3)-Centered Radicals.Angew. Chem. Int. Ed.2020, 59, 2618–2622. Google Scholar 34. Zuo Z.; Gong H.; Li W.; Choi J.; Fu G. C.; MacMillan D. W. C.Enantioselective Decarboxylative Arylation of Alpha-Amino Acids via the Merger of Photoredox and Nickel Catalysis.Angew. Chem. Int. Ed.2016, 138, 1832–1835. Google Scholar 35. Fu M.-C.; Shang R.; Zhao B.; Wang B.; Fu Y.Photocatalytic Decarboxylative Alkylations Mediated by Triphenylphosphine and Sodium Iodide.Science2019, 363, 1429–1434. Google Scholar 36. Zeng X.; Yan W.; Zacate S. B.; Chao T. H.; Sun X.; Cao Z.; Bradford K. G. E.; Paeth M.; Tyndall S. B.; Yang K.; Kuo T. C.; Cheng M. J.; Liu W.Copper-Catalyzed Decarboxylative Difluoromethylation.J. Am. Chem. Soc.2019, 141, 11398–11403. Google Scholar 37. Guo L.-N.; Wang H.; Duan X.-H.Recent Advances in Catalytic Decarboxylative Acylation Reactions via a Radical Process.Org. Biomol. Chem.2016, 14, 7380–7391. Google Scholar 38. Prier C. K.; Rankic D. A.; MacMillan D. W.Visible Light Photoredox Catalysis with Transition Metal Complexes: Applications in Organic Syntheisis.Chem. Rev.2013, 113, 5322–5363. Google Scholar 39. Wang C.-S.; Dixneuf P. H.; Soulé J.-F.Photoredox Catalysis for Building C–C Bonds from C(sp2)–H Bonds.Chem. Rev.2018, 118, 7532–7585. Google Scholar 40. Maity B.; Zhu C.; Yue H.; Huang L.; Harb M.; Minenkov Y.; Rueping M.; Cavallo L.Mechanistic Insight into the Photoredox-Nickel-HAT Triple Catalyzed Arylation and Alkylation of Alpha-Amino Csp3-H Bonds.J. Am. Chem. Soc.2020, 142, 16942–16952. Google Scholar 41. Yuan M.; Song Z.; Badir S.O.; Molander G. A.; Gutierrez O.On the Nature of C(sp3)-C(sp2) Bond Formation in Nickel-Catalyzed Tertiary Radical Cross-Couplings: A Case Study of Ni/Photoredox Catalytic Cross-Coupling of Alkyl Radicals and Aryl Halides.J. Am. Chem. Soc.2020, 142, 7225–7234. Google Scholar 42. Gutierrez O.; Tellis J. C.; Primer D. N.; Molander G. A.; Kozlowski M. C.Nickel-Catalyzed Cross-Coupling of Photoredox-Generated Radicals: Uncovering a General Manifold for Stereoconvergence in Nickel-Catalyzed Cross-Couplings.J. Am. Chem. Soc.2015, 137, 4896–4899. Google Scholar 43. Ruzi R.; Liu K.; Zhu C.; Xie J.Upgrading Ketone Synthesis Direct from Carboxylic Acids and Organohalides.Nat. Commun.2020, 11, 3312. Google Scholar 44. Zhang M.; Xie J.; Zhu C.A General Deoxygenation Approach for Synthesis of Ketones from Aromatic Carboxylic Acids and Alkenes.Nat. Commun.2018, 9, 3517. Google Scholar 45. Stache E. E.; Rovis T.; Doyle A. G.Dual Nickel- and Photoredox-Catalyzed Enantioselective Desymmetrization of Cyclic Meso-Anhydrides.Angew. Chem. Int. Ed.2017, 56, 3679–3683. Google Scholar Previous articleNext article FiguresReferencesRelatedDetails Issue AssignmentVolume 4Issue 7Page: 2469-2480Supporting Information Copyright & Permissions© 2021 Chinese Chemical SocietyKeywordsdual-catalysisdecarboxylative acylationcarboxylic acidmetallaphotoredoxketones synthesisAcknowledgmentsThe authors wish to acknowledge Dongping Wang, Yijie He, Kai Liu, Yantao Li for reproducing the experimental procedures for products (3a, 3r, 4g, and 5d). All theoretical calculations were performed at the High-Performance Computing Center (HPCC) of Nanjing University and Prof. Guoqiang Wang at Institute of Theoretical and Computational Chemistry is warmly acknowledged for his help on DFT calculations. Downloaded 3,836 times PDF DownloadLoading ...
最长约 10秒,即可获得该文献文件

科研通智能强力驱动
Strongly Powered by AbleSci AI
更新
大幅提高文件上传限制,最高150M (2024-4-1)

科研通是完全免费的文献互助平台,具备全网最快的应助速度,最高的求助完成率。 对每一个文献求助,科研通都将尽心尽力,给求助人一个满意的交代。
实时播报
蓝意完成签到,获得积分0
16秒前
光亮若翠完成签到,获得积分10
16秒前
小美完成签到 ,获得积分10
18秒前
安安滴滴完成签到 ,获得积分10
20秒前
红茸茸羊完成签到 ,获得积分10
23秒前
满意人英完成签到,获得积分10
24秒前
xiaodong发布了新的文献求助10
32秒前
轩辕远航完成签到 ,获得积分10
35秒前
zz完成签到 ,获得积分10
37秒前
xiaodong完成签到,获得积分10
46秒前
杪夏二八完成签到 ,获得积分10
56秒前
lesen完成签到,获得积分10
1分钟前
lemon完成签到 ,获得积分10
1分钟前
没用的三轮完成签到,获得积分10
1分钟前
研友_Z1eDgZ完成签到,获得积分10
1分钟前
终究是残念完成签到,获得积分10
1分钟前
Amy完成签到 ,获得积分10
1分钟前
面包完成签到 ,获得积分10
1分钟前
tianliyan完成签到 ,获得积分10
1分钟前
a46539749完成签到 ,获得积分10
1分钟前
执着晓亦完成签到 ,获得积分10
2分钟前
孙哈哈完成签到 ,获得积分10
2分钟前
善学以致用应助zjh采纳,获得10
2分钟前
缥缈映安完成签到 ,获得积分10
3分钟前
wlei完成签到,获得积分10
3分钟前
isedu完成签到,获得积分10
3分钟前
2024kyt完成签到 ,获得积分10
3分钟前
zjh发布了新的文献求助10
3分钟前
天天赚积分完成签到,获得积分0
3分钟前
zcydbttj2011完成签到 ,获得积分10
3分钟前
zjh完成签到,获得积分10
3分钟前
rafa完成签到 ,获得积分10
4分钟前
4分钟前
蛋妮完成签到 ,获得积分10
4分钟前
唯有一个心完成签到 ,获得积分10
4分钟前
loren313完成签到,获得积分0
4分钟前
铜豌豆完成签到 ,获得积分10
5分钟前
5分钟前
CUN完成签到,获得积分10
6分钟前
合适醉蝶完成签到 ,获得积分10
6分钟前
高分求助中
Licensing Deals in Pharmaceuticals 2019-2024 3000
Effect of reactor temperature on FCC yield 2000
Very-high-order BVD Schemes Using β-variable THINC Method 1000
PraxisRatgeber: Mantiden: Faszinierende Lauerjäger 800
A new species of Coccus (Homoptera: Coccoidea) from Malawi 500
Impiego dell’associazione acetazolamide/pentossifillina nel trattamento dell’ipoacusia improvvisa idiopatica in pazienti affetti da glaucoma cronico 480
Geochemistry, 2nd Edition 地球化学经典教科书第二版,不要epub版本 431
热门求助领域 (近24小时)
化学 医学 生物 材料科学 工程类 有机化学 生物化学 物理 内科学 纳米技术 计算机科学 化学工程 复合材料 基因 遗传学 催化作用 物理化学 免疫学 量子力学 细胞生物学
热门帖子
关注 科研通微信公众号,转发送积分 3291547
求助须知:如何正确求助?哪些是违规求助? 2928027
关于积分的说明 8435109
捐赠科研通 2599859
什么是DOI,文献DOI怎么找? 1418806
科研通“疑难数据库(出版商)”最低求助积分说明 660150
邀请新用户注册赠送积分活动 642771