Palladium-Catalyzed Carbonylative Transformation of Organic Halides

Aug 18, 2017 - Palladium-Catalyzed Carbonylative Transformation of Organic Halides with Formic Acid as the Coupling Partner and CO Source: Synthesis o...
0 downloads 12 Views 777KB Size
Article pubs.acs.org/joc

Palladium-Catalyzed Carbonylative Transformation of Organic Halides with Formic Acid as the Coupling Partner and CO Source: Synthesis of Carboxylic Acids Fu-Peng Wu,† Jin-Bao Peng,*,† Xinxin Qi,† and Xiao-Feng Wu*,†,‡ †

Department of Chemistry, Zhejiang Sci-Tech University, Hangzhou, Zhejiang 310018, People’s Republic of China Leibniz-Institut für Katalyse e.V. an der, Institution Universität Rostock, Albert-Einstein-Strasse 29a, Rostock 18059, Germany



S Supporting Information *

ABSTRACT: A palladium-catalyzed carbonylative transformation of organic halides with formic acid as the coupling partner to produce carboxylic acids has been developed. With a catalytic amount of DCC as the activator of formic acid, the process can be realized successfully through benzoic formic anhydride as the intermediate. Both vinyl and aryl (pseudo)halides were conveniently transformed into the corresponding acids in good yields.



INTRODUCTION

Among the few available reports, Skrydstrup and co-workers elaborated an elegant palladium-catalyzed hydroxycarbonylation of aryl halides with potassium formate as a mildly basic nucleophile and CO source in 2013 (Scheme 1, a).4c A special

Carboxylic acids represent a large family of important structure units that exist in many natural products, pharmaceuticals, agrochemicals, and materials.1 Several efficient methods have been developed for the synthesis of these structures; oxidation of primary alcohols and aldehydes or hydrolysis of the corresponding nitriles have been considered as the traditional procedures for their preparation. The reaction of organic reagents with CO2, such as the Kolbe−Schmitt reaction and the Grignard reaction, provide good efficiency and regioselectivity, but the functional group compatibility is limited. Thus, palladium-catalyzed hydroxycarbonylation2 of aryl halides provides an efficient and straightforward alternative for preparing such compounds. Despite many advantages, the main drawback for this method is the requirement of highly toxic CO gas and the corresponding high pressure equipment. To circumvent this problem, several CO gas-free carbonylation protocols have been developed based on CO surrogates during the past decades. Metal−carbonyl complexes, formamides, aldehyde, formic acid, and oxalyl chloride were all investigated and successively used as CO sources.3 Among them, formic acid attracted much attention because it can generate one molecule of CO after releasing one molecule of H2O. Many research groups including our group have developed a series of efficient carbonylation reactions employing formic acid as the CO source.4 However, there are two major disadvantages for this method: (1) (over)stoichiometric amounts of activator such as Ac2O and H2SO4 were needed to react with formic acid to release CO; (2) (over)stoichiometric byproduct such as AcOH was generated which might influence the subsequent carbonylation reaction or need an equal amount of base to neutralize. Hydroxycarbonylation with a catalytic amount of activator using formic acid as the CO source was rarely studied. © 2017 American Chemical Society

Scheme 1. Formic Acid-Based Hydroxycarbonylation Reaction with Catalytic Activator

acyl-Pd(II) precatalyst, which delivered substoichiometric CO to initiate the reaction and also acted as the catalyst, was prepared and used in this reaction. Another catalytic hydroxycarbonylation reaction with formic acid was reported by Zhou and co-workers (Scheme 1, b). Using a catalytic amount of acetic anhydride as the initiator, acrylic acids were Received: July 20, 2017 Published: August 18, 2017 9710

DOI: 10.1021/acs.joc.7b01808 J. Org. Chem. 2017, 82, 9710−9714

Article

The Journal of Organic Chemistry

sealed tube, which gives a theoretical CO partial pressure of 0.1 bar when 0.2 equiv of DCC was used. This also agrees with the previous report from Skrydstrup4c that a minimum of 0.1 bar carbon monoxide partial pressure was essential for effective convention. This is explainable because the coordination and release of CO to the vinyl palladium complex is a reversible process. Lower CO partial pressure would facilitate the release of CO from the palladium and thus decrease the hydroxycarboxylation yield. Additionally, no desired acid product can be obtained in the absence of DCC, but some 1,4-diphenylbuta-1,3-diene can be detected as the homocoupling product (Table 1, entry 6). Some other bidentate phosphine ligands were tested as well; decreased yield was obtained (Table 1, entryies 7−11). Only a cinnamic acid yield of 6% could be detected with PPh3 as the ligand (Table 1, entry 12). With the optimized reaction conditions in hand (Table 1, entry 2), we began to investigate the generality of the carboxylation reaction with a series of substituted vinyl iodides. As shown in Table 2, various substituted 2-phenylvinyl iodides

synthesized by a palladium-catalyzed hydrocarboxylation reaction of alkynes with formic acid.4d Recently, we have developed a novel palladium-catalyzed hydroxycarbonylation of aryl halides with the in situ generation of CO and H2O. Benzoic acid was synthesized without the addition of CO gas.5 However, stoichiometric acetic anhydride and aliphatic alcohol were needed for this transformation. Lately, we found that benzoic acid could be generated via decarbonylation of the benzoic formic anhydride, where the property of the ligand played a determinative role.4h In view of the reaction mechanism, we assume that a catalytic amount of DCC would be possible as well. The CO from the decomposed benzoic formic anhydride would be applicable for the next catalytic cycle. Furthermore, the carboxylation of vinyl (pseudo)halides provides a direct method for the preparation of acrylic acid, which is a versatile material in industry. With these considerations above, herein we report a newly developed palladium-catalyzed hydroxycarbonylation procedure with formic acid using a catalytic amount of DCC as the activator (Scheme 1, c).



Table 2. Substrate Scope of Vinyl Iodidesa

RESULTS AND DISCUSSION The initial investigation was performed with 2-phenylvinyl iodide as the starting material. We chose Xantphos as the ligand because the large bite angle bidentate phosphine ligand facilitates the formation of anhydride and subsequently provides the acid product. To our delight, by heating a solution of vinyl iodide and formic acid in the presence of Pd(OAc)2 with only 0.2 equiv of DCC as the activator in DMF for 20 h, the carboxylation product cinnamic acid was produced in 92% yield (Table 1, entry 2). Reducing the loading of DCC to 0.1 equiv gave a decreased yield of 75% (Table 1, entry 1). However, further increasing of the amount of DCC did not improve the reaction yield (Table 1, entries 3−5). It should be mentioned that all the reactions were conducted in a 20 mL Table 1. Optimizing the Reaction Conditionsa

entry

ligand

DCC (equiv)

yield (%)

1 2 3 4 5 6 7 8 9 10 11 12

Xantphos Xantphos Xantphos Xantphos Xantphos Xantphos DPPE DPPP DPPPe DPPF DPEphos PPh3

0.1 0.2 0.4 0.6 1.0 0 0.2 0.2 0.2 0.2 0.2 0.2

75 92 84 81 84 0 30 27 35 48 81 6

a Reaction conditions: vinyl iodides (0.5 mmol), Pd(OAc)2 (3 mol %), Xantphos (3 mol %), HCOOH (3.5 mmol), DCC (0.1 mmol), Et3N (1.0 mmol), DMF (1 mL), 100 °C, 20 h, isolated yields.

were subjected to the optimized reaction conditions, the carboxylation reaction proceeded smoothly, and the corresponding cinnamic acids were generated in high to excellent yields (Scheme 2, a−h). The electronic property and position of the substituent group on the benzene ring showed little influence on the yields. Then various aryl iodides were studied under our standard conditions. As summarized in Table 3, a series of aryl iodides were tested in this reaction and the corresponding carboxylic acids were isolated in good yields in general. Both electrondonating groups (Table 3, 4b−f) and electron-withdrawing groups (Table 3, 4g−j) substitution at different positions were well tolerated and produced the corresponding acids in good yields. Generally, substrates with electron-withdrawing groups gave slightly higher yield than that of electron-donating groups. Notably, aryl iodides with fluoro- and chloro-substitutions were also tolerated in this reaction and delivered the corresponding product in 82% yields (Table 3, 4g and 4h). However, p-

a

Reaction conditions: vinyl iodide (0.5 mmol), formic acid (3.5 mmol), Pd(OAc)2 (3 mol %), ligand (3 mol %), Et3N (1.0 mmol), 100 °C, 20 h. DPPE = 1,2-bis(diphenylphosphino)ethane. DPPP = 1,3bis(diphenylphosphino)propane. DPPPe = 1,5-bis(diphenylphosphino)pentane. DPPF = 1,1′-bis(diphenylphosphino)ferrocene. DPEphos = bis(2-diphenylphosphino)diphenyl ether. Xantphos = 4,5bis(diphenylphosphino)-9,9-dimethylxanthene. 9711

DOI: 10.1021/acs.joc.7b01808 J. Org. Chem. 2017, 82, 9710−9714

Article

The Journal of Organic Chemistry Table 4. Substrate Scope of Aryl Bromidesa

Scheme 2. Carbonylative Transformation of pBromoiodobenzene

Table 3. Substrate Scope of Aryl Iodidesa a

Reaction conditions: aryl bromides (0.5 mmol), Pd(OAc)2 (3 mol %), Xantphos (3 mol %), HCOOH (3.5 mmol), DCC (0.1 mmol), Et3N (1.0 mmol), DMF (1 mL), 100 °C, 20 h, isolated yields.

Scheme 3. Reaction of Phenyl Triflate and Gram-Scale Synthesis of Benzoic Acid

a Reaction conditions: aryl iodides (0.5 mmol), Pd(OAc)2 (3 mol %), Xantphos (3 mol %), HCOOH (3.5 mmol), DCC (0.1 mmol), Et3N (1.0 mmol), DMF (1 mL), 100 °C, 20 h, isolated yields.

Scheme 4. Plausible Reaction Mechanism

bromoiodobenzene gave an nonseparatable mixture of benzoic acid, 4-bromobenzoic acid, and terephthalic acid (1:0.06:0.1; Scheme 2). This phenomena implies that aryl bromides are applicable substrates as well. Furthermore, 2-iodonaphthalene and 3-iodothiophene were also amenable to this reaction and provided the desired products in 97% and 46% yields, respectively (Table 3, 4k and 4l). Aryl bromides, which are more abundant and less expensive than aryl iodides, usually represent more challenging materials for palladium-catalyzed carbonylation owing to their higher activation barrier for the C−Br bond activation. To our delight, various substituted aryl bromides were conveniently applied in this reaction and the corresponding carboxylic acids were generated in moderate to good yields (Table 4). In addition to aryl iodides and bromides, when phenyl triflite was subjected to the optimized conditions, the carboxylation reaction proceeded successfully and the desired benzoic acid was obtained in 89% yield (Scheme 3, eq 1). We also conducted a gram-scale reaction to demonstrate the synthetic utility of this method. When the reaction was performed on 20 mmol scale, benzoic acid was obtained in 68% yield (Scheme 3, eq 2). However, no conversion of substrate could be observed when chlorobenzene or 4-chlorobenzotrifluoride was applied. Based on these results and precedent reports, a possible reaction mechanism is proposed in Scheme 4. Initially, oxidative addition of the in situ-generated Pd0 to the C−X

bond gave an arylpalladium complex 6. Then coordination and insertion of one molecule of carbon monoxide, which was generated in situ from the reaction of formic acid with DCC as the activator, into 6 forms acylpalladium intermediate 7. Subsequently, the ligand exchange of the iodide with formic acid would afford the acylpalladium formic acid complex 8. The steric effect of large bite angle bisphosphine-ligated palladium would enforce an η1 binding mode through one oxygen atom. 9712

DOI: 10.1021/acs.joc.7b01808 J. Org. Chem. 2017, 82, 9710−9714

Article

The Journal of Organic Chemistry

= 8.4 Hz, 2H), 6.55 (d, J = 16.0 Hz, 1H). 13C NMR (101 MHz, DMSO) δ 167.90, 142.94, 135.17, 133.64, 130.33, 129.35, 120.53. (E)-3-(4-Methoxyphenyl)acrylic acid (2g): 7 87% yield, 77 mg, eluent (petroleum ether/EtOAc: 3:2), 1H NMR (400 MHz, DMSO) δ 12.24 (s, 1H), 7.63 (d, J = 8.6 Hz, 2H), 7.56 (d, J = 16.0 Hz, 1H), 6.96 (d, J = 8.6 Hz, 2H), 6.38 (d, J = 16.0 Hz, 1H), 3.78 (s, 3H). 13C NMR (101 MHz, DMSO) δ 168.32, 161.39, 144.19, 130.37, 127.28, 116.96, 114.79, 55.71. (E)-3-(4-(tert-Butyl)phenyl)acrylic acid (2h): 7 78% yield, 80 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 12.38 (s, 1H), 7.62−7.54 (m, 3H), 7.41 (d, J = 8.1 Hz, 2H), 6.48 (d, J = 16.0 Hz, 1H), 1.26 (s, 9H). 13C NMR (101 MHz, DMSO) δ 168.16, 153.51, 144.23, 131.97, 128.46, 126.14, 118.78, 34.98, 31.33. Benzoic acid (4a): 4h 84% yield, 51 mg, eluent (petroleum ether/ EtOAc: 2:1),1H NMR (400 MHz, DMSO) δ 12.96 (s, 1H), 8.00−7.92 (m, 2H), 7.61 (d, J = 7.5 Hz, 1H), 7.50 (t, J = 7.6 Hz, 2H). 13C NMR (101 MHz, DMSO) δ 167.79, 133.31, 131.21, 129.73, 129.02. 2-Methylbenzoic acid (4b): 4h 83% yield, 57 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, CDCl3) δ 12.23 (s, 1H), 8.14 (t, J = 9.8 Hz, 1H), 7.50 (t, J = 7.5 Hz, 1H), 7.33 (t, J = 7.6 Hz, 2H), 2.72 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 173.74, 141.43, 133.02, 131.98, 131.66, 128.37, 125.91, 22.20. 3-Methylbenzoic acid (4c): 4h 82% yield, 56 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, CDCl3) δ 12.49 (s, 1H), 7.98 (d, J = 7.2 Hz, 2H), 7.59−7.36 (m, 2H), 2.46 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 172.87, 138.33, 134.65, 130.75, 129.30, 128.42, 127.43, 21.28. 4-Methylbenzoic acid (4d): 4h 82% yield, 56 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, CDCl3) δ 11.11 (s, 1H), 8.04 (d, J = 8.0 Hz, 2H), 7.30 (d, J = 7.8 Hz, 2H), 2.46 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 172.53, 144.68, 130.28, 129.24, 126.63, 21.79. 4-Methoxybenzoic acid (4e): 4h 67% yield, 51 mg, eluent (petroleum ether/EtOAc: 1:1), 1H NMR (400 MHz, DMSO) δ 12.64 (s, 1H), 7.90 (d, J = 8.8 Hz, 2H), 7.01 (d, J = 8.8 Hz, 2H), 3.82 (s, 3H). 13C NMR (101 MHz, DMSO) δ 167.48, 163.30, 131.80, 123.43, 114.24, 55.85. [1,1′-Biphenyl]-4-carboxylic acid (4f): 4h 63% yield, 62 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 12.99 (s, 1H), 8.04 (d, J = 8.1 Hz, 2H), 7.79 (d, J = 8.2 Hz, 2H), 7.72 (d, J = 7.6 Hz, 2H), 7.50 (t, J = 7.5 Hz, 2H), 7.42 (t, J = 7.3 Hz, 1H). 13C NMR (101 MHz, DMSO) δ 167.63, 144.78, 139.50, 130.44, 130.10, 129.53, 128.74, 127.42, 127.26. 4-Fluorobenzoic acid (4g): 4h 82% yield, 57 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 13.04 (s, 1H), 8.08−7.97 (m, 2H), 7.31 (dd, J = 12.3, 5.4 Hz, 2H). 13C NMR (101 MHz, DMSO) δ 166.83, 166.62, 164.13, 162.17, 132.61, 132.51, 127.81, 116.17, 115.95. 4-Chlorobenzoic acid (4h): 4h 82% yield, 64 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 13.19 (s, 1H), 7.93 (d, J = 8.4 Hz, 2H), 7.53 (d, J = 8.4 Hz, 2H). 13C NMR (101 MHz, DMSO) δ 166.93, 138.26, 131.57, 130.08, 129.14. 4-Acetylbenzoic acid (4i): 4h 71% yield, 58 mg, eluent (petroleum ether/EtOAc: 3:2), 1H NMR (400 MHz, DMSO) δ 13.29 (s, 1H), 8.04 (s, 4H), 2.62 (s, 3H). 13C NMR (101 MHz, DMSO) δ 197.65, 166.61, 139.76, 134.52, 129.50, 128.25, 26.93. 4-(Methoxycarbonyl)benzoic acid (4j): 4h 81% yield, 73 mg, eluent (petroleum ether/EtOAc: 3:2), 1H NMR (400 MHz, DMSO) δ 13.29 (s, 1H), 8.03 (d, J = 2.2 Hz, 4H), 3.87 (s, 3H). 13C NMR (101 MHz, DMSO) δ 167.01, 166.03, 135.27, 133.56, 129.99, 129.74, 52.83. 1-Naphthoic acid (4k): 4h 97% yield, 84 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 13.18 (s, 1H), 8.92 (d, J = 8.6 Hz, 1H), 8.24−8.16 (m, 1H), 8.14 (d, J = 8.1 Hz, 1H), 8.00 (d, J = 8.1 Hz, 1H), 7.65 (dd, J = 8.3, 7.0 Hz, 1H), 7.58 (dt, J = 7.7, 4.6 Hz, 2H). 13C NMR (101 MHz, DMSO) δ 169.18, 133.95, 133.39, 131.20, 130.36, 129.06, 128.21, 128.02, 126.64, 126.06, 125.33. Thiophene-3-carboxylic acid (4l): 4h 46% yield, 30 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 12.70 (s, 1H), 8.26 (d, J = 2.9 Hz, 1H), 7.61 (dd, J = 4.9, 3.0 Hz, 1H), 7.45−

Subsequently, reductive elimination delivered the benzoic formic anhydride 9. Meanwhile, Pd0 was regenerated and used for the next catalyst cycle. The benzoic formic anhydride 9 then decomposed to generate the terminal acid product 4 and release CO at the same time for follow-up reactions.4c,6 In summary, we have developed a palladium-catalyzed carbonylative transformation reaction of organic halides with formic acid as the CO source and coupling partner. With a catalytic amount of DCC as the activator, the reaction was initiated and CO was generated from the decarbonylation of the formed benzoic formic anhydride and used in the next catalytic cycle. Vinyl and aryl (pseudo)halides were all suitable substrates and transformed into the corresponding acids in good yields.



EXPERIMENTAL SECTION

General Information. Unless otherwise noted, all reactions were carried out under a nitrogen atmosphere. All chemicals were from commercial sources and used as received without further purification. All solvents were purified and dried by standard techniques and distilled prior to use. Column chromatography was performed on silica gel (200−300 mesh) using petroleum ether (bp 60−90 °C) and ethyl acetate as eluent. 1H and 13C NMR spectra were taken on 400 MHz instruments, and spectral data were reported in ppm relative to tetramethylsilane (TMS) as internal standard and CDCl3 as solvent. Mass spectra (MS) were measured on spectrometer by direct inlet at 70 eV. General Procedure. Pd(OAc)2 (3 mol %) and Xantphos (3 mol %) were added to an oven-dried tube which was then placed under vacuum and refilled with nitrogen three times. A solution of formic acid (3.5 mmol) and organic halide (0.5 mmol) in DMF (1.0 mL) was added to the reaction tube. After DCC (0.1 mmol) and Et3N (1.0 mmol) were added, the tube was sealed and the mixture was stirred at 100 °C for 20 h. After the reaction was completed, the reaction mixture was filtered and concentrated under vacuum. The crude product was purified by column chromatography on silica gel to afford the corresponding product. Cinnamic acid (2a): 7 92% yield, 68 mg, eluent (petroleum ether/ EtOAc: 2:1),1H NMR (400 MHz, DMSO) δ 12.41 (s, 1H), 7.69 (dd, J = 6.3, 2.6 Hz, 2H), 7.60 (d, J = 16.0 Hz, 1H), 7.46−7.38 (m, 3H), 6.54 (d, J = 16.0 Hz, 1H). 13C NMR (101 MHz, DMSO) δ 168.07, 144.37, 134.70, 130.68, 129.36, 128.66, 119.72. (E)-3-(o-Tolyl)acrylic acid (2b): 7 90% yield, 73 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 12.46 (s, 1H), 7.83 (d, J = 15.9 Hz, 1H), 7.68 (d, J = 7.6 Hz, 1H), 7.33−7.26 (m, 1H), 7.26−7.19 (m, 2H), 6.42 (d, J = 15.9 Hz, 1H), 2.37 (s, 3H). 13C NMR (101 MHz, DMSO) δ 168.03, 141.59, 137.60, 133.36, 131.14, 130.39, 126.90, 126.86, 120.63, 19.71. (E)-3-(m-Tolyl)acrylic acid (2c): 7 88% yield, 71 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 12.39 (s, 1H), 7.55 (dd, J = 16.0, 4.8 Hz, 1H), 7.47 (d, J = 8.3 Hz, 2H), 7.34−7.20 (m, 2H), 6.50 (ddd, J = 16.0, 3.8, 2.3 Hz, 1H), 2.37−2.29 (m, 3H). 13C NMR (101 MHz, DMSO) δ 168.07, 144.51, 138.60, 134.60, 131.38, 129.24, 129.10, 125.88, 119.45, 21.28. (E)-3-(p-Tolyl)acrylic acid (2d): 7 89% yield, 72 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 12.34 (s, 1H), 7.56 (dd, J = 12.1, 5.6 Hz, 3H), 7.21 (d, J = 7.9 Hz, 2H), 6.47 (d, J = 16.0 Hz, 1H), 2.31 (s, 3H). 13C NMR (101 MHz, DMSO) δ 168.17, 144.38, 140.58, 131.96, 129.96, 128.62, 118.55, 21.43. (E)-3-(4-Fluorophenyl)acrylic acid (2e): 7 79% yield, 66 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 12.41 (s, 1H), 7.75 (dd, J = 8.4, 5.7 Hz, 2H), 7.60 (d, J = 16.0 Hz, 1H), 7.23 (t, J = 8.8 Hz, 2H), 6.49 (d, J = 16.0 Hz, 1H). 13C NMR (101 MHz, DMSO) δ 168.00, 164.84, 162.37, 143.14, 131.36, 131.33, 130.96, 130.88, 119.54, 116.41, 116.20. (E)-3-(4-Chlorophenyl)acrylic acid (2f): 7 89% yield, 81 mg, eluent (petroleum ether/EtOAc: 2:1), 1H NMR (400 MHz, DMSO) δ 12.49 (s, 1H), 7.70 (d, J = 8.4 Hz, 2H), 7.58 (d, J = 16.0 Hz, 1H), 7.44 (d, J 9713

DOI: 10.1021/acs.joc.7b01808 J. Org. Chem. 2017, 82, 9710−9714

Article

The Journal of Organic Chemistry 7.42 (m, 1H). 13C NMR (101 MHz, DMSO) δ 164.03, 134.78, 133.73, 128.20, 127.72.



(7) Song, C. X.; Chen, P.; Tang, Y. RSC Adv. 2017, 7, 11233−11243.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.joc.7b01808. 1 H and 13C NMR spectra of products (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Xiao-Feng Wu: 0000-0001-6622-3328 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge financial support from NSFC (21472174, 21602201, 21602204), the education department of Zhejiang Province (Y201636555), Zhejiang Sci-Tech University (16062095-Y), and Zhejiang Natural Science Fund for Distinguished Young Scholars (LR16B020002).



REFERENCES

(1) (a) March’s Advanced Organic Chemistry. Reactions, Mechanisms, and Structure; Smith, M. B.; March, J., Eds.; Wiley: New York, 2007. (b) Advanced Organic Chemistry; Carey, F. A.; Sundberg, R. J., Eds.;Springer: New York, 2007. (c) Gooßen, L. J.; Rodriguez, N.; Gooßen, K. Angew. Chem., Int. Ed. 2008, 47, 3100−3120. (2) For selected reviews on carbonylation, see: (a) Kollär, L. Modern Carbonylation Methods; Wiley-VCH: Weinheim, 2008. (b) Barnard, C. F. J. Organometallics 2008, 27, 5402−5422. (c) Brennfuhrer, A.; Neumann, H.; Beller, M. Angew. Chem., Int. Ed. 2009, 48, 4114−4133. (d) Wu, X. F.; Neumann, H.; Beller, M. Chem. Soc. Rev. 2011, 40, 4986−5009. (e) Wu, X.-F.; Neumann, H. ChemCatChem 2012, 4, 447−458. (f) Gabriele, B.; Mancuso, R.; Salerno, G. Eur. J. Org. Chem. 2012, 2012, 6825−6839. (g) Wu, X. F.; Neumann, H.; Beller, M. Chem. Rev. 2013, 113, 1−35. (h) Peng, J.-B.; Qi, X.; Wu, X.-F. ChemSusChem 2016, 9, 2279−2283. (i) Wu, X.-F. RSC Adv. 2016, 6, 83831−83837. (3) For selected reviews on carbonylation using CO surrogates, see: (a) Odell, L. R.; Russo, F.; Larhed, M. Synlett 2012, 23, 685−698. (b) Konishi, H.; Manabe, K. Synlett 2014, 25, 1971−1986. (c) Wu, L.; Liu, Q.; Jackstell, R.; Beller, M. Angew. Chem., Int. Ed. 2014, 53, 6310− 6320. (d) Gautam, P.; Bhanage, B. M. Catal. Sci. Technol. 2015, 5, 4663−4702. (e) Sam, B.; Breit, B.; Krische, M. J. Angew. Chem., Int. Ed. 2015, 54, 3267−3274. (f) Friis, S. D.; Lindhardt, A. T.; Skrydstrup, T. Acc. Chem. Res. 2016, 49, 594−605. (g) Peng, J.-B.; Qi, X.; Wu, X.-F. Synlett 2017, 28, 175−194. (4) (a) Cacchi, S.; Fabrizi, G.; Goggiamani, A. Org. Lett. 2003, 5, 4269−4272. (b) Morimoto, T.; Kakiuchi, K. Angew. Chem., Int. Ed. 2004, 43, 5580−5588. (c) Korsager, S.; Taaning, R. H.; Skrydstrup, T. J. Am. Chem. Soc. 2013, 135, 2891−2894. (d) Hou, J.; Xie, J.-H.; Zhou, Q.-L. Angew. Chem., Int. Ed. 2015, 54, 6302−6305. (e) Qi, X.; Jiang, L.B.; Li, C.-L.; Li, R.; Wu, X.-F. Chem. - Asian J. 2015, 10, 1870−1873. (f) Qi, X.; Li, C.-L.; Jiang, L.-B.; Zhang, W.-Q.; Wu, X.-F. Catal. Sci. Technol. 2016, 6, 3099−3107. (g) Peng, J.-B.; Wu, F.-P.; Li, C.-L.; Qi, X.; Wu, X.-F. Eur. J. Org. Chem. 2017, 2017, 1434−1437. (h) Wu, F.P.; Peng, J.-B.; Meng, L.-S.; Qi, X.; Wu, X.-F. ChemCatChem 2017, 9, 3121−3124. (i) Seo, Y.-S.; Kim, D.-S.; Jun, C.-H. Chem. - Asian J. 2016, 11, 3508−3512. (5) Li, C.-L.; Qi, X.; Wu, X.-F. ChemistrySelect 2016, 1, 1702−1704. (6) Pri-Bar, I.; Buchman, O. J. Org. Chem. 1988, 53, 624−626. 9714

DOI: 10.1021/acs.joc.7b01808 J. Org. Chem. 2017, 82, 9710−9714