First-Row Transition-Metal-Catalyzed Carbonylative Transformations

4 days ago - In 2014, he received his Ph.D. degree from Lanzhou University under the supervision of Prof. Yong-Qiang Tu. After postdoctoral research w...
159 downloads 4 Views 8MB Size
Review pubs.acs.org/CR

Cite This: Chem. Rev. XXXX, XXX, XXX−XXX

First-Row Transition-Metal-Catalyzed Carbonylative Transformations of Carbon Electrophiles Jin-Bao Peng,† Fu-Peng Wu,† 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 Universität Rostock, Albert-Einstein-Straße 29a, Rostock 18059, Germany



ABSTRACT: The main contributions in the field of first-row transition-metal-catalyzed (base-metal-catalyzed) carbonylative transformations have been summarized and discussed. The contents have been divided according to the electrophiles applied, followed by the different types of nucleophiles. Their reaction mechanisms and applications have been emphatically discussed.

CONTENTS 1. Introduction 1.1. Background of First-Row Transition Metals in Carbonylation Reactions 1.2. Scope of This Review 2. Carbonylation of Organic Halides and Derivatives 2.1. Carbonylation of C(sp2)−X Bonds 2.1.1. Hydroxy-, Alkoxy- and Aminocarbonylations of C(sp2)−X Bonds 2.1.2. Carbonylative Coupling Reactions of C(sp2)−X Bonds 2.2. Carbonylation of C(sp3)−X Bonds 2.2.1. Hydroxy-, Alkoxy- and Aminocarbonylations of C(sp3)−X Bonds 2.2.2. Carbonylative Coupling Reactions of C(sp3)−X 3. Carbonylation of Alkenes and Alkynes 3.1. Hydroformylation of Alkenes 3.2. Carbonylation of Alkynes 3.2.1. Iron-Mediated Carbonylation of Alkynes 3.2.2. Nickel-Mediated Carbonylation of Alkynes 3.2.3. Pauson−Khand Reaction 4. Carbonylation of Aldehydes 4.1. Amidocarbonylation of Aldehydes 4.2. Silylformylation of Aldehydes 5. Carbonylation of Heterocycles 5.1. Alkoxy- and Aminocarboxylations of Heterocycles 5.2. Ring-Expansion Carbonylation of Heterocycles 5.3. Alternating Copolymerization of Epoxides 6. Oxidative Carbonylation Reaction 6.1. Oxidative Carbonylation of C(sp2)−H Bonds 6.2. Oxidative Carbonylation of C(sp3)−H 7. Conclusion

© XXXX American Chemical Society

Author Information Corresponding Author ORCID Notes Biographies Acknowledgments References

A B B C C

AD AD AD AD AD AD AD

C

1. INTRODUCTION Employing CO as a cheap and abundant C1 source, the carbonylation reaction1−15 represents a direct and atomeconomical strategy for the preparation of various carbonylcontaining compounds and their derivatives. After over a halfcentury of developments, the transition-metal-catalyzed carbonylation reaction has experienced impressive progress. This reaction has now emerged as one of the most powerful methods for the synthesis of carbonyl compounds and has been successfully used as a key step in the total synthesis of a number of natural products.16,17 In addition, many carbonylative procedures have already been industrialized. For example, the majority of the world’s acetic acid production is based on carbonylation of methanol (the Monsanto or Cativa process).18,19 Nowadays, research on the development of carbonylation is mainly focused in four directions: (a) developing new, cheap, and reusable catalyst systems for highly efficient carbonylation reactions; (b) developing regio-, chemo-, and stereoselective carbonylation reactions with high selectivity;20 (c) application of unactivated substrates in carbonylation reactions, such as carbonylation of organic chlorides, ethers, and alkanes; and (d) developing new CO surrogates and new

F J J L M M M M P Q Q Q S S S T Z Z Z AB AD

Special Issue: First Row Metals and Catalysis Received: February 1, 2018

A

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Table 1. Electron Configurations and Oxidation States of Base Metals atomic no.

atom

electron configuration

oxidation states

content in Earth’s crust (ppm)

21 22 23 24 25 26 27 28 29 30

Sc Ti V Cr Mn Fe Co Ni Cu Zn

[Ar]3d14s2 [Ar]3d14s2 [Ar]3d34s2 [Ar]3d54s1 [Ar]3d54s2 [Ar]3d64s2 [Ar]3d74s2 [Ar]3d84s2 [Ar]3d104s1 [Ar]3d104s2

1, 2, 3 −2, −1, 1, 2, 3, 4 −3, −1, 1, 2, 3, 4, 5 −4, −2, −1, 1, 2, 3, 4, 5, 6 −3, −2, −1, 1, 2, 3, 4, 5, 6, 7 −4, −2, −1, 1, 2, 3, 4, 5, 6, 7 −3, −1, 1, 2, 3, 4, 5 −2, −1, 1, 2, 3, 4 −2, 1, 2, 3, 4 −2, 1, 2

5 5600 200 100 950 47500 10 70 100 75

Earth’s crust. Although iron exists in a wide range of oxidation states, from −2 to +7, +2 and +3 are the most common oxidation states. Low-valent iron29−36 complexes have been found to be capable of catalyzing a range of reactions, such as cross-coupling reactions and cycloisomerizations. Iron carbonyl complexes are widely used in carbonylation reactions, especially for alkyl halides and alkynes. Common oxidation states of cobalt include +2 and +3, although other oxidation states ranging from −3 to +5 are also known. One of the most significant characteristics of cobalt complexes is their high affinity toward π bonds, including alkenes, alkynes, and allenes. Thus, cobalt is often applied in cycloaddition reactions such as [2 + 2 + 2] cycloadditions37,38 and Pauson−Khand reactions.39 Recently, the utility of cobalt catalysts has been expanded to a number of other types of reactions,40 for example, crosscoupling reactions41 and C−H functionalization reactions.42 Cobalt-mediated carbonylation reactions usually involve the anionic species [Co(CO)4]−, which can be generated from the reduction of various cobalt precursors. Nickel is typically observed in the 0 and +2 oxidation states (+1 and +3 oxidation states are also known). Nickel catalysts are widely used in cyclization reactions43−46 and coupling reactions,47−52 as well as other transformations. Owing to its smaller size compared to palladium, nickel is relatively more nucleophilic and shows higher activity toward carbon−oxygen bonds. Copper can be easily accessed in its 0, +1, +2, and +3 oxidation states, which allows the chemistry of copper to proceed via both singleelectron-transfer pathways and two-electron, bond-forming pathways.53−55

techniques for low-pressure or even CO-gas-free carbonylation reactions.21−25 Although most research on carbonylation reactions is based on catalysis by noble metals (such as Pd, Rh, Ru, and Ir), great achievements have also been made with non-noble-metalcatalyzed and even metal-free carbonylation reactions. In particular, the first-row transition-metal-catalyzed carbonylation has a long history; the earliest investigations into carbonylation reactions were based on first-row transition metals, such as Co and Ni. For example, the first commercialized homogeneous methanol carbonylation for the production of acetic acid was established by BASF in 1955 using a nickel catalyst. 1.1. Background of First-Row Transition Metals in Carbonylation Reactions

The group of first-row transition metals comprises 10 elements that feature the electron configuration of [Ar]3dp4sq (Table 1). The transition metals are known for their wide applications in homogeneous and heterogeneous catalyses. Their varied catalytic activities are ascribed to their multiple oxidation states and abilities in forming complexes. The first-row transition metals utilize 3d and 4s electrons for bonding; thus the chemical environment of the 3d and 4s shells significantly influences their properties, such as magnetic character, variable oxidation states, and catalytic activities. Among the first-row transition metals, zinc has the electronic configuration of [Ar]3d104s2, with a complete d shell. Thus, zinc is, under certain criteria, excluded from the transition metals, and is usually referred to as a post-transition metal.26,27 In fact, with its fully occupied 3d and 4s shells, zinc is chemically similar to magnesium: only one normal oxidation state (+2) is exhibited. Zinc(I) compounds are rarely reported and usually contain the dimeric [Zn2]2+ core. Until now, no examples of zinc-catalyzed cabonylation have been reported. Studies on carbonylation reactions catalyzed by first-row transition metals with half-occupied or less than half-occupied 3d shells (Sc, Ti, V, Cr, and Mn) are comparably rare. Scandium and titanium have been widely used as Lewis acids, but no application of scandium in carbonylation has been reported as yet. Only a few examples of vanadium-, chromium-, and manganese-catalyzed carbonylations have been reported. For example, a vanadium-catalyzed carboxylation of propane into butyric acid was reported in 2007.28 In addition, Cr(CO)6 and Mn2(CO)10 were usually used as CO surrogates in carbonylation reactions. The vast majority of investigations of first-row transitionmetal-catalyzed carbonylation reactions are based on Fe, Co, Ni, and Cu. Iron is widely distributed worldwide, accounting for 4.75% of the Earth’s crust content, and after oxygen, silicon, and aluminum, it is the fourth most common element in the

1.2. Scope of This Review

First-row transition metals have been widely used in organic syntheses. One of the most significant features of the first-row transition metals is that they can participate in both one- and two-electron elementary reactions, while reactions of secondand third-row transition metals usually proceed via two-electron processes. Thus, first-row transition metal catalysis not only provides an economical alternative to noble metals, their unique properties also provide additional opportunities for developing new reactions that are not possible with secondand third-row transition metal catalysis. For example, nickel has been widely applied in cross-coupling reactions involving phenol-derived electrophiles, where palladium catalysts showed lower activity. The carbonylation of unactivated alkyl halides is problematic for palladium catalysts because of the difficulty of the oxidative addition of C(sp3)−X bonds and the competitive β-elimination reactions of alkylpalladium species. However, base metal carbonyl complexes have shown excellent activity in carbonylation reactions of alkyl halides and their analogues. In B

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

addition, the catalytic carbonylation of unactivated alkanes has also been realized with Cu catalysts. The present review aims to summarize the most important advances of first-row transition metals in carbonylative transformations. These results are organized based on their respective reaction types; each subsection is further managed according to the hybridization of the carbon center, i.e., sp2 and sp3. Instead of listing all of the base-metal-catalyzed carbonylation reactions, this review is intended to highlight the strategic applications of first-row transition metals in various carbonylative transformations and discuss the reaction mechanisms of representative reactions.

Scheme 1. Mechanism of Cobalt-Catalyzed Methanol Carbonylation

2. CARBONYLATION OF ORGANIC HALIDES AND DERIVATIVES The synthesis of aromatic acids and their derivatives via transition-metal-catalyzed carbonylation of aromatic halides has been known since the 1920s. Transition-metal-catalyzed carbonylations of C−X bonds followed by attack of nucleophiles such as water, alcohols, and amines, are known as hydroxy-, alkoxy-, and aminocarbonylation reactions, respectively. These reactions are important transformations in organic synthesis, producing a range of valuable carboxylic acids, esters, and amides. Many of these processes have already been applied on industrial scales. For example, the carbonylation of methanol for the synthesis of acetic acid comprises more than 60% of global acetic acid production. In early examples, cobalt and nickel complexes were found to be the most effective catalysts. For example, BASF established the first commercialized acetic acid synthesis via nickelcatalyzed methanol carbonylation in the 1950s. However, extreme conditionshigh reaction temperature (230 °C) and high CO pressure (600 atm)were usually required. Subsequent investigations of methanol carbonylation reactions were focused on the search for new catalytic systems that proceed under milder conditions. Nowadays, Ir catalysts are most often used for the industrial manufacture of acetic acid (the Cativa process). The carbonylation of methanol for the synthesis of acetic acid has been extensively studied and reviewed.56−60 In this review, we will not seek to reiterate the history of methanol carbonylation. However, it is important to discuss the mechanism of the cobalt- and nickel-catalyzed carbonylation since it is a typical representative and it is useful to obtain a general overview of the topic. For the methanol carbonylation, iodine-containing cocatalysts are essential; otherwise the CO insertion would occur at the O−H bond to give methyl formate instead of at the C−O bond to give acetic acid. Thus, the presence of iodine is necessary to convert methanol to methyl iodide. For the cobaltcatalyzed carbonylation (Scheme 1), it is commonly accepted that cobalt precursors react with carbon monoxide and an acidic proton to generate the catalytically active species [HCo(CO)4]. The in situ generated methyl iodide reacts with [HCo(CO)4] to give methylcobalt complex [CH3Co(CO)4]. Then the migratory insertion of CO and recoordination of CO to the cobalt center leads to acylcobalt complex [CH3COCo(CO)4]. Finally, hydrolysis of the acylcobalt complex with water releases the acetic acid and regenerates [HCo(CO)4] for the next catalytic cycle. The acyl cleavage is presumed to take place via nucleophilic attack at the carbonyl carbon of the acyl group. Notably, it was observed that small amounts of H2 increase the catalytic activity, in agreement with

the assumption that the catalytically active species is [HCo(CO)4]. The mechanism of the nickel-catalyzed carbonylation of methanol is depicted in Scheme 2. The nickel-catalyzed Scheme 2. Mechanism for Nickel-Catalyzed Methanol Carbonylation

methanol carbonylation also proceeds via addition of methyl iodide. Nickel tetracarbonyl [Ni(CO)4] is formed from various nickel precursors. Then, the oxidative addition of methyl iodide to the nickel carbonyl forms [CH3Ni(CO)2I]. Coordination and insertion of CO into the C−Ni bond forms the acylnickel complex, which decomposes under reductive elimination to give the corresponding acid halide and regenerate Ni(CO)4. The resulting acetyl iodide is hydrolyzed either by water or alcohols to release acetic acid. Compared to the carbonylation of methanol, the carbonylation of organic halides is relatively facile. Based on the dissociation energies of carbon−halogen bonds, the previously reported relative order of reactivity for oxidative addition of organic halides to an electronically unsaturated metal are as follows: C−I > C−Br ≈ C−OTf ≫ C−OTs/OMs > C−Cl ≫ C−F.61−63 2.1. Carbonylation of C(sp2)−X Bonds

2.1.1. Hydroxy-, Alkoxy- and Aminocarbonylations of C(sp2)−X Bonds. Cassar and Foa developed a nickel-catalyzed hydroxycarbonylation of aryl halides for the synthesis of aromatic acids (Scheme 3).64 In polar aprotic solvent, using nickel carbonyl as the catalyst, and Ca(OH)2 as the base, carbonylation of aromatic halides could be performed even at atmospheric pressure of CO and at a low temperature of 100 C

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

nylnickel anion [Ni(CO)3CN]− was isolated, which was believed to be the key catalytic species. Subsequently, the nickel- and phase-transfer-catalyzed carbonylation reactions were investigated with a series of other organic halides. Aryl iodides, vinyl bromides, and benzyl chlorides were all found to be suitable reactants and produced the corresponding acids in moderate yields (Scheme 5b,c).69−71 A promotion effect of lanthanide reagents was observed in the carbonylation of benzyl chlorides. Iranpoor and Firouzabadi described a nickel-catalyzed carbonylation reaction for the synthesis of thioesters, esters, and amides (Scheme 6).72 With NiCl2 as the catalyst and

Scheme 3. Nickel-Catalyzed Hydroxycarbonylation of Aryl Halides at Atmospheric Pressure of CO

°C. A range of aromatic acids were obtained in good to excellent yields. 1-Chloronaphthalene and 2-chloronaphthalene were also found to be active in this transformation, providing the corresponding naphthoic acids in 95 and 97% yields, respectively. In addition to (hetero)aryl halides, alkenyl and alkyl halides have also been successfully used in carbonylation reactions. Corey and co-workers reported a based-promoted alkoxycarbonylation of alkenyl halides by nickel carbonyl (Scheme 4).65

Scheme 6. Nickel-Catalyzed Synthesis of Thioesters, Esters, and Amides from Aryl Iodides

Scheme 4. Alkoxycarbonylation of Alkenyl Halides by Nickel Carbonyl

Treatment of alkenyl bromides with 6 equiv of nickel carbonyl in alcoholic medium (R2OH) in the presence of the corresponding sodium or potassium alkoxide formed the corresponding esters in moderate to good yields. However, the high toxicity of nickel tetracarbonyl spurred the investigation of other Ni(II) complexes as catalysts. The stoichiometric carbonylation reaction mediated by hexacyanodinickelate(I) was studied by Hashimoto in 1970. By bubbling CO into a water/acetone solution of potassium hexacyanodinickelate and benzyl bromides, a facile carbonylation reaction occurred and provided symmetrical ketones in good yields.66 The reaction of trans-β-bromostyrene gave esters in the presence of alcohol. K4[Ni2(CN)6] is coordinatively unsaturated and easy to disproportionate, and absorbs two molecules of CO to form the dicyanodicarbonylnickel[0] dianion [Ni(CN)2(CO)2]2−, which was believed to be the active intermediate of this reaction.67 Phase-transfer catalysts were found to be valuable promoters of transition-metalcatalyzed carbonylations. In 1985, Alper and co-workers reported a nickel- and phase-transfer-catalyzed carbonylation of allyl halides. Allyl chlorides and bromides were easily carbonylated to acids by cyanonickel(II) catalysts under phasetransfer-catalysis conditions (Scheme 5a).68 The cyanotricarbo-

Cr(CO)6 as a solid source of carbon monoxide, the thiocarbonylation, alkoxycarbonylation, and aminocarbonylation reactions of aryl iodides proceeded effectively in air. A range of varyingly substituted aryl iodides and structurally different thiols, alcohols, and amines were tolerated in this reaction. The corresponding thioesters, esters, and amides were obtained in good to excellent yields. However, aromatic amines and thiols were not suitable for this reaction owing to the direct coupling. Additionally, carbonylation of aromatic bromides failed to give the desired products under these conditions. Mechanistically, oxidative addition of aryl iodides to the in situ generated Ni(0) gives the arylnickel(II) complex 6-A. Coordination and insertion of CO into the C−Ni bond deliver the acylnickel(II) complex 6-B. The nucleophiles attack the acylnickel complex, forming the intermediate 6-C, which undergoes reductive elimination to afford the carbonylated products (Scheme 6b). Transition-metal-catalyzed aminocarbonylation of organic halides is a direct and efficient method for the synthesis of amides. Most of these aminocarbonylation reactions use an amine as the nucleophile. Nitroarenes were reported to be able to be reduced to aryl amines and serve as the nitrogen source under reductive conditions.73 Very recently, Hu and colleagues developed a nickel-catalyzed reductive aminocarbonylation of

Scheme 5. Nickel- and Phase-Transfer-Catalyzed Hydroxycarbonylation of Organic Halides

D

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

aryl halides using nitroarenes as more economical and accessible nitrogen sources (Scheme 7a).74 Using Ni(glyme)Cl2

Scheme 8. Nickel-Catalyzed Synthesis of Isoindolinones from Aryl Iodides, Imines, and CO

Scheme 7. Nickel-Catalyzed Reductive Aminocarbonylation of Aryl Halides with Nitroarenes

as the catalyst, dtbpy (4,4′-di-tert-butyl-2,2′-dipyridyl) as the ligand, Zn powder as the reductant, and Co2(CO)8 as the CO source, the reductive aminocarbonylation proceeded successfully and produced the corresponding amides in moderate to excellent yields. A series of aromatic halides and nitroarenes were tolerated in this reaction. Although a detailed mechanism was not demonstrated, the authors studied the reaction of a number of possible nitrogen-containing intermediates (nitrosobenzene, N-phenylhydroxylamine, azobenzene, and anilines) derived from nitrobenzene under the optimized conditions. NPhenyl hydroxylamine and aniline were found to be the possible intermediates in this reaction. Thus, in the presence of a Zn-activating reagent (TMSCl), the nitroarene was reduced by zinc to form a reduced nitrogen species. The reduced nitrogen containing species reacted with the acylnickel(II) intermediate to produce the amide product upon acidic workup (Scheme 7b). Carbonylation of organic halides would generate the corresponding carboxylic acid halides, which are reactive toward a variety of nucleophiles. Bengali, Arndtsen, and coworkers developed a nickel-catalyzed tandem process for the synthesis of isoindolinones from aryl iodides, imines, and carbon monoxide (Scheme 8).75 The authors suggested that nickel-catalyzed carbonylation of aryl iodides in the presence of chloride anion forms the corresponding acid chlorides, which reacted with imines to generate N-acyl iminium chlorides. Subsequently, nickel-mediated annulation of the latter compounds delivered the cyclized products. A range of substituted isoindolinones were prepared in moderate to high yields. In 2014, Sudalai and co-workers developed a coppercatalyzed protocol for the carbonylative synthesis of carboxylic acid derivatives with sodium cyanide as C1 source (Scheme 9a).76 In the presence of the NiBr2/1,10-phenanthroline catalytic system, carbonylative coupling of aryl halides with phenols, alcohols, and amine nucleophiles provided the corresponding carboxylic acid derivative in high yields. Murahashi reported the first example of Co2(CO)8-catalyzed carbonylative cyclization of aromatic imides in 1955 (Scheme 9b).77 In the presence of a catalytic amount of cobalt, the C−H bond of the imine was activated, followed by CO insertion,

Scheme 9. Carbonylative Synthesis of Amides

leading to the isolation of 2-arylisoindolin-1-ones in good yields. Transition-metal-catalyzed double carbonylation is a special variant of the carbonylation reaction. By incorporating two molecules of carbon monoxide into the parent compounds, a series of α-keto acids, esters, and amides could be synthesized from organic halides.78 Generally, high CO pressure (>50 bar) is usually needed to obtain good selectivity over the monocarbonylation reaction. Several elegant double carbonylation reactions under atmospheric pressure of CO were reported by the palladium catalysis.79 Double carbonylation reactions catalyzed by first-row transition metals have also been studied. To our surprise, cobalt catalysts based on [Co(CO)4]− showed even better reactivity than noble metals in some cases. This superior activity is due to the good nucleophilicity of [Co(CO)4]−. Additionally, the countercation was important as well and better activity can be acheveved with cations having higher charge/radius ratios. Hence catalytic systems based on the combination of [Co(CO)4]− and a Lewis acid have also been developed. Related results were reviewed by Foa and Francalanci.80 The same group performed excellent studies on this topic, for instance reporting a cobalt-catalyzed double carbonylation of aryl and secondary benzyl halides.81,82 The reaction proceeded in alcoholic solvent in the presence of a catalytic amount of Co2(CO)8, providing corresponding acids in moderate yields and selectivity under an atmospheric pressure of CO. Studies on cobalt-catalyzed carbonylation of E

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Sonogashira reactions are typical representatives of transitionmetal-catalyzed carbonylation with CO for the synthesis of various ketones. Generally, carbonylative coupling reactions feature the advantages of high selectivity and broad substrate scope. The transition-metal-catalyzed carbonylative coupling reaction of aryl halides with organostannanes, namely the carbonylative Stille reaction, is a powerful tool for the preparation of aryl ketones.91,92 As alternatives to aryl halides, Kang and colleagues reported the palladium-catalyzed crosscoupling and carbonylative cross-coupling of hypervalent iodonium salts with organostannanes. In 1996, Kang and coworkers reported a copper-catalyzed coupling of iodonium salts with organometallic reagents (Scheme 11a).93 Using CuI as the

aryl and benzyl halides have also been performed in other research groups.83−88 Kashimura’s group developed a cobaltcatalyzed photostimulated carbonylation of o-halogenated benzoic acids in aqueous sodium hydroxide.83 Miura and coworkers succeeded in the cobalt-catalyzed carbonylation of vinyl halides. Using Co2(CO)8 as the catalyst precursor in the presence of methyl iodide and calcium hydroxide at 20 °C, good yields of cinnamic acids were produced from the corresponding vinyl bromides/chlorides. Interestingly, furan2(5H)-ones can be formed from 3-chloroprop-2-enol substrates under identical conditions.86 Mortreux and co-workers reported a cobalt-catalyzed carbonylative synthesis of dimethyl malonate from dichloromethane (DCM).88 Electrochemistry was used for the synthesis of the complex (Co(CO)3PBu3)− in a methanol−methyl formate medium. Under 15 bar of CO at 80 °C, with MeONa as the base, a 75% yield of dimethyl malonate can be produced using MeOH and DCM as the reaction medium. Nickel-catalyzed double carbonylation of halodienes for the synthesis of α-keto acids and α-keto γ-lactones were additionally realized by Amer and co-workers using phase-transfer techniques. The reaction was believed to proceed via the intermediacy of a stable metallacycle.89 In addition, Sun, Xia, and co-workers developed an efficient NHC−Cu−X-based catalyst system for the double carbonylation of aryl iodides and secondary amines. A range of α-keto amides were obtained in good yields and with high selectivity (Scheme 10).90

Scheme 11. Transition-Metal-Catalyzed Carbonylative Stille Reactions

Scheme 10. NHC−Cu−X-Catalyzed Double Carbonylation of Aryl Iodides and Secondary Amines

catalyst, organostannanes and organoboranes reacted with diaryliodonium salts in the presence of atmospheric pressure carbon monoxide and delivered ketones in good yields. Mechanistically, it was presumed that the arylcopper species PhCuIX was formed via the facile oxidative addition of the iodonium salts to the Cu(I) salt. Coordination and insertion of CO to PhCuIX gave an acylcopper species, which underwent transmetalation with RSnBu3 followed by reductive elimination to afford the ketone product. However, only four examples were reported, and substrates with electron-withdrawing substituents were not reported. In addition to copper catalysis, a nickel-catalyzed carbonylative Stille coupling reaction was reported by the same group (Scheme 11b). Employing Ni(acac)2 as the catalyst, ketones were synthesized in 65− 81% yields from their parent compounds under mild conditions.94 Compared to copper and nickel, manganese is little studied as a catalyst for carbonylative coupling reactions. Kang and colleagues developed novel carbonylative Stille reactions of organostannanes and hypervalent iodonium salts with manganese chloride as catalysts (Scheme 11c).95 Under atmospheric pressure of CO, organostannanes and iodonium salts were carbonylatively coupled and produced ketones in moderate to good yields. Iranpoor and Firouzabadi reported a phosphine-free, nickelcatalyzed carbonylative Stille coupling of aryl iodides with Ph3SnCl or Ph3SnOEt (Scheme 12a).96 Using NiBr2 as the catalyst and Cr(CO)6 as a solid CO source, unsymmetrical diaryl ketones were synthesized in good yields. For nickelcatalyzed carbonylative Stille coupling reactions, the mechanism usually constitutes four elementary steps: (I) the oxidative addition of a C−X bond to the active Ni(0) species to generate an arylnickel species, (II) coordination and insertion of CO to give an acylnickel intermediate, (III) transmetalation with an organotin reagent, and (IV) reductive elimination to release the

2.1.2. Carbonylative Coupling Reactions of C(sp2)−X Bonds. Diaryl ketones constitute an important and versatile structure motif that is frequently present in natural products, materials, pharmaceuticals, and other biologically active molecules. Owing to its importance, various synthetic methodologies for the preparation of diaryl,ketones have been developed. Among them, an especially straightforward approach for the synthesis of diaryl,ketones is the CO-based carbonylation reaction. Generally, transition-metal-catalyzed carbonylative coupling reactions for the synthesis of ketones can be classified into two categories: carbonylative crosscoupling reactions and carbonylative homocoupling reactions. Carbonylative cross-coupling reactions involve organic halides R−X and nucleophilic organometallic compounds R′−M, and both symmetrical and unsymmetrical diaryl ketones have been synthesized in this way. Carbonylative homocoupling reactions can be performed either by organometallic nucleophiles R′−M in the presence of an oxidizing reagent or by aryl electrophiles R−X (or Ar2I+) under reductive conditions. 2.1.2.1. Carbonylative Cross-Coupling Reactions of C(sp2)−X Bonds. Three-component coupling reactions of aryl halides/pseudohalides, carbon monoxide, and organometallic reagents, now known as carbonylative coupling reactions, represent some of the most straightforward and efficient procedures for synthesizing various symmetrical and unsymmetrical ketones. The carbonylative Stille, Suzuki, Negishi, and F

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 12. Phosphine-Free Nickel-Catalyzed Carbonylative Stille Reactions

Scheme 13. Iron-Catalyzed Carbonylative Suzuki Reactions

13b).104 Employing FeCl2 as the precatalyst and stoichiometric CHCl3 as the CO source, a broad range of synthetically useful biaryl ketones were synthesized in good to excellent yields. Both aryltrifluoroborates and arylboronic acids were found to be suitable for this reaction. The addition of strongly basic hydroxide played a critical role in the hydrolysis of CHCl3 to release CO. Additionally, this protocol provides an alternative for effective 13C labeling by using the commercially available 13 C-labeled CHCl3. A plausible mechanism for the ironcatalyzed carbonylative Suzuki reaction was proposed. Initially, the hydrolysis of CHCl3 releases CO in the presence of a hydroxide base. The precatalyst FeCl2/FeCl3 reacts with CO to generate an iron carbonyl species Fem(CO)n (13-A), which was believed to be the active catalytic species. Fem(CO)n (13-A) reacted with the arylboronic acid with the assistance of base to generate highly nucleophilic organoiron complex 13-B, which

ketone product and regenerate the Ni(0) catalyst for the next cycle (Scheme 12b). In 2014, the group of Han developed an iron-catalyzed carbonylative Suzuki reaction of aryl iodides (Scheme 13a).97 Using the combination of FeCl2 (4 mol %) and FeCl3 (6 mol %) as the catalyst system, carbonylative Suzuki reactions of various aryl iodides with arylboronic acids proceeded smoothly under atmospheric pressure carbon monoxide using PEG-400 as a green solvent. A range of functional groups, such as hydroxyl, chloro, bromo, and nitrile, were compatible with the reactions. Variously substituted diaryl ketones were obtained in good yields. Chloroform (CHCl3) has recently been used as a safe and inexpensive CO surrogate, and has already been successfully applied in several carbonylative transformations.98−104 In 2016, Han and colleagues reported a general iron-catalyzed carbonylative Suzuki−Miyaura reaction (Scheme G

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 14. Transition-Metal-Nanoparticle-Catalyzed Carbonylative Suzuki Reactions

undergoes intramolecular CO migratory insertion to give organoiron complex 13-C. Subsequently, SNR1-type oxidative addition of the aryl iodide to 13-C occurs to afford intermediate 13-D. Finally, reductive elimination of 13-E delivers the product biaryl ketone and regenerates the catalytically active 13-A under a CO atmosphere (Scheme 13c). Shortly afterward, the same group discovered that transition metal nanoparticles were effective to catalyze the carbonylative Suzuki reaction of aryl iodides with arylboronic acid. Copper nanoparticles (Scheme 14a)105 and in situ generated nickel nanoparticles (Scheme 14b)106 were subsequently reported to catalyze carbonylative Suzuki reactions of aryl iodides with arylboronic acids under ambient CO pressure in poly(ethylene glycol). The development of carbonylative cross-coupling reactions is usually hindered by the competitive direct coupling reactions. The authors found that the use of pivalic acid was effective to suppress the direct Suzuki reactions and facilitate carbonylative Suzuki reactions. In 2008, the group of Chen developed a nickel-catalyzed carbonylative Negishi cross-coupling of enol trifletes and diorganozinc reagents (Scheme 15).107 Utilizing NiCl2 as the

Scheme 16. Copper-Catalyzed Carbonylative Sonogashira Reactions

2.1.2.2. Carbonylative Homocoupling Reactions of C(sp2)−X Bonds. The carbonylative homocoupling of organomercuric halides was reported by Seyferth and Spohn as early as 1968 with stoichiometric dicobalt octacarbonyl. Organomercuric halides RHgX reacted with Co2(CO)8 to form symmetric ketones (Scheme 17a).109,110 Hirota and co-workers discovered that a similar transformation could be realized with Ni(CO)4 (Scheme 17b).111 Scheme 17. Cobalt- and Nickel-Catalyzed Carbonylation of Organomercuric Halides

Scheme 15. Nickel-Catalyzed Carbonylative Negishi Reactions

The group of Knochel reported a cobalt-mediated carbonylative homocoupling of organozinc reagents (Scheme 18).112 Scheme 18. Cobalt-Catalyzed Carbonylation of Organozinc Reagents

catalyst and 4,4′-dimethoxyl-2,2′-bipyridyl as the ligand, both enol triflates and alkenyl iodides were conveniently converted to the corresponding enones in good to excellent yields. The use of polar solvents such as DMSO was found to increase the reaction rate. The addition of lithium or magnesium halides improved the ratio of carbonylative coupling to direct coupling products. Transition-metal-catalyzed coupling of aryl halides (or pseudohalides), CO, and terminal alkynes is now well-known as the carbonylative Sonogashira coupling reaction. However, the majority of the investigations of carbonylative Sonogashira reaction were based on catalysis of palladium. The first coppercatalyzed carbonylative Sonogashira reaction was reported by Bhanage et al. in 2008 (Scheme 16).108 The key to its success was the use of Cu(TMHD)2 [copper bis(2,2,6,6-tetramethyl3,5-heptanedionate)] as the precatalyst since it is highly soluble in organic solvents and stable in air. A series of aliphatic/ aromatic alkynes reacted with aryl iodides to give the ynones in moderate to good yields.

Bubbling CO slowly through a mixture of organozinc reagents and CoBr2 in THF/NMP produced symmetric ketones in moderate to good yields. Functional groups such as chlorides and esters were compatible in this condition. Liu and Lu et al. reported a nickel-catalyzed carbonylation of arylboronic acids for the synthesis of symmetric diaryl ketones (Scheme 19a).113 By employing DMF as an abundant and lowcost CO source, a range of substituted aryl iodides were converted to the corresponding ketones in moderate to excellent yields. The reaction mechanism may involve the following steps: Ni(0) inserts into the formyl C−H bond of DMF to form nickel complex 19-A. Transmetalation of 19-A H

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Recently, Gosmini et al. developed a catalytic procedure for the synthesis of symmetrical diaryl ketones from aryl bromides and ethyl chloroformate (Scheme 22a).118 In the presence of

Scheme 19. Nickel-Catalyzed Carbonylation of Arylboronic Acids

Scheme 22. Cobalt-Catalyzed Synthesis of Diaryl Ketones

catalytic quatities of CoBr2(bpy), the in situ generated arylzinc bromides were converted to the corresponding diaryl ketones in moderate to excellent yields. Ethyl chloroformate acted as a CO surrogate in this reaction. Mechanistic studies revealed that the chloroformate decomposed in the presence of the cobalt catalyst and zinc powder, and reacted with cobalt to generate a cobalt carbonyl intermediate 22-A. Then, transmetalation of 22-A with the arylzinc species formed the intermediate 22-B. After migratory insertion of CO, a second transmetalation took place and produced intermediate 22-D. Finally, reductive elimination released the diaryl ketone and regenerated the cobalt catalyst (Scheme 22b). Electrochemistry119 represents one of the classical synthetic methods where electrons act as clean and efficient redox agents. Electrochemical reduction/oxidation provides a convenient means to generate a desired oxidant state of a metal complex that is active for an organic reaction. Thus, combining electrosynthesis with organometallic catalysis provides versatile and mild processes for selective organic synthesis. Compared to classical organic reactions, where a stoichiometric amount of oxidant or reducing reagent is needed, electrons constitute clean and nonpolluting redox reagents. The electroreduction of Ni(II) or Pd(II) precatalysts offers catalytically active lowvalent nickel and palladium species that are highly active toward oxidative addition of C−X bonds of aryl halides. A series of coupling reactions of organic halides through catalysis by organometallics combined with electrochemical methods were reported.120,121 Troupel and colleagues reported the first electrochemical carbonylation of organic halides for the synthesis of symmetric ketones (Scheme 23a).122,123 Utilizing NiBr2bpy as the precatalyst and Fe(CO)5 as the CO source, symmetrical ketones were obtained from primary benzyl chlorides and alkyl iodides by electroreduction in DMF in an undivided cell. The process is especially well adapted for primary benzyl chlorides and alkyl iodides. The same catalytic system could also be employed in the synthesis of nonsymmetrical aryl benzyl ketones or alkyl aryl ketones (Scheme 23b).124 With 30 mol % NiBr2bpy, benzyl chlorides were successfully coupled with aryl halides and alkyl iodides to form

with an arylboronic acid followed by reductive elimination generates amide intermediate 19-C. Then, the in situ generated Ni(0) inserts into the C−N bond of 19-C to give acylnickel complex 19-D, which releases the diaryl ketone after a second transmetalation and elimination (Scheme 19b). Larhed and colleagues developed a Co2(CO)8-mediated ultrafast carbonylative coupling of aryl halides under microwave irradiation (Scheme 20).114 The reaction was conducted in air and completed in seconds. Several symmetrical diaryl ketones were prepared in moderate to excellent yields. Scheme 20. Cobalt-Catalyzed Carbonylation of Aryl Iodides

The first catalytic carbonylation of aryl halides for the synthesis of symmetric diaryl ketones with metal carbonyl complexes was reported by Brunet (Scheme 21).115−117 In the Scheme 21. Fe(CO)5-Co2(CO)8 Bimetallic Catalysis for the Synthesis of Benzophenones

presence of catalytic amounts of both Fe(CO)5 and Co2(CO)8, substituted iodobenzenes were converted to their corresponding benzophenone derivatives under phase-transfer conditions in atmospheric pressure of carbon monoxide. Although the detailed mechanism was not reported, a synergetic effect of both carbonyl metal complexes (Fe(CO)5 and Co2(CO)8) was observed. I

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

reactions are based on C(sp2)−X bonds. The carbonylation of C(sp3)−X bonds, especially those of unactivated alkyl halides, is relatively challenging owing to the difficulty of the oxidative addition and the competitive β-elimination reactions. Base metal carbonyl complexes show great activity for the carbonylation of alkyl halides. Cobalt seems to be an appropriate alternative for noble metals in these reactions. For example, as early as 1963, Heck and Breslow reported the first usage of Na[Co(CO)4] as a catalyst in the carbonylative formation of esters from alkyl halides.128 Recently, Li and colleagues developed the benzimidazole-based cobalt complex Co(L)2(H2O)2, which was found to be efficient in catalyzing the carbonylation of benzyl chlorides (Scheme 25).129 The electron-rich benzimidazole improved the catalytic activity and the reaction proceeded under atmospheric pressure CO.

Scheme 23. Nickel-Catalyzed Electrochemical Carbonylation Reaction

Scheme 25. Cobalt-Catalyzed Carbonylation of Benzyl Chlorides

The direct carbonylation of benzyl alcohols via cobalt catalysis has also been developed (Scheme 26).130 In the Scheme 26. Cobalt-Catalyzed Carbonylation of Benzyl Alcohols

the corresponding ketones in yields of 43−88%. In addition, the electrochemical carbonylation of organic halides with atmospheric pressure carbon monoxide has also been reported. Fe(CO)5 was found to be a superior CO source to CO gas for electrochemical carbonylation reactions, as higher CO concentrations lead to the formation of unreactive (L)Ni0(CO)2.125,126 Mechanistic studies revealed that the key catalytic species is the transient 16-electron mixed Ni0(CO)bpy complex. Thus, electroreduction of NiBr2bpy in the presence of CO formed two zerovalent nickel complexes: Ni0(CO)2bpy and Ni0bpy. CO exchange between these two nickel complexes generated the active Ni0(CO)bpy. Oxidative addition of the R−X bond to Ni0(CO)bpy and migratory insertion produced an acylnickel(II) intermediate. Finally, the acylnickel(II) complex reacted with the second R−X to give the ketone products (Scheme 23c). Recently, Weix and co-workers investigated the reactivity of the acylnickel(II) complexes in cross-electrophile coupling reactions with a series of carbon electrophiles.127 The catalytic synthesis of alkyl aryl ketones was realized under the catalysis of the NiCl2(dme)/dtbbpy system with zinc as the reducing reagent (Scheme 24).

presence of sodium iodide in PPE (ethyl polyphosphate), Co2(CO)8 catalyzed the carbonylation of benzyl alcohols under CO atmosphere. Both amines and alcohols could be used as the nucleophile and produced the corresponding amides and esters in moderate to good yields. In addition to alkyl halides, alkyl sulfonates have also been successfully applied in carbonylation reactions. In 1991, Fuchikami and colleagues reported a cobalt-catalyzed alkoxycarbonylation of alkyl sulfonates (Scheme 27).131 Using Scheme 27. Cobalt-Catalyzed Alkoxycarbonylation of Alkyl Sulfonates

2.2. Carbonylation of C(sp3)−X Bonds

2.2.1. Hydroxy-, Alkoxy- and Aminocarbonylations of C(sp3)−X Bonds. Despite great advances in the carbonylation of organic halides, most of the electrophiles employed in these Scheme 24. Nickel-Catalyzed Carbonylative Formation of Alkyl Aryl Ketones

Co2(CO)8 as the catalyst, alkyl sulfonates were effectively carbonylated in the presence of NaI under 50 atm of CO in TMU−EtOH solvent mixture to form esters in moderate to good yields (TMU = tetramethylurea). In 1970, Cooke discovered that Na2Fe(CO)4, which was generated from the reduction of Fe(CO)5 by sodium amalgam, J

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

is an effective reagent to convert alkyl bromides to the corresponding aldehydes with an additional carbon (Scheme 28a).132 Treatment of alkyl bromides with stoichiometric

Scheme 29. Manganese-Catalyzed Carboacylation of Alkenes with Alkyl Iodides

Scheme 28. Na2Fe(CO)4-Mediated Carbonylative Conversion of Alkyl Bromides

good diastereoselectivity under mild conditions. A plausible mechanism was proposed by the authors. The active •Mn(CO)5 radical, generated from homolysis of Mn−Mn bond of Mn2(CO)10, reacted with the C(sp3)−I to give an alkyl radical 29-A. A radical cyclization then took place to generate radical 29-B, which reacted with the I−Mn(CO)5 complex and produced the acyl manganese complex 29-C after migratory insertion of CO. Finally, nucleophilic substitution of ethanol to the acyl manganese released the ester. Recoordination of CO to the metal center regenerated the active catalyst for the next cycle. Wu and colleagues reported a carbonylative coupling of alkyl halides with amides for the synthesis of imides via manganese catalysis (Scheme 30a).143 Using Mn2(CO)10 as the catalyst

quantities of in situ generated Na2Fe(CO) 4 gave the corresponding acids in good yields after quenching with acetic acid. However, when secondary alkyl bromides were used under these conditions, only moderate yields were obtained owing to the competing elimination reaction. From the point of view of mechanism, Fe(CO)42− reacted with aliphatic halides to give anionic alkyltetracarbonyliron(0) complexes 28-A. In the presence of CO, migratory insertion of 28-A afforded the acyliron complex 28-B. Protonation of anion 28-B by acetic acid followed by reductive elimination yielded the aldehydes (Scheme 28b). Collman and colleagues systematically investigated the application of Na2Fe(CO)4 for the synthesis of aliphatic ketones133,134 as well as carboxylic acid derivatives135 with different nucleophiles. Halogenation of acyliron complex 28-B followed by nucleophilic attack by water, alcohols, and amines produced carboxylic acids, esters, and amides, respectively (Scheme 28c). Primary and secondary alkyl tosylates also produced the corresponding products in good yields. The Na2Fe(CO)4-mediated carbonylation reaction of aliphatic halides and tosylates features the advantages of high yields and stereospecificity. The main problem with these reactions is the competing elimination of C−X bonds owing to the high basicity of Na2Fe(CO)4. Tertiary alkyl halides were not tolerated, and secondary tosylates gave higher yields than secondary halides. Collman investigated the mechanism of oxidative addition alkyl halides and tosylates to Na2Fe(CO)4136 and the alkyl migration of [RFe(CO)4]−137−139 in detail and reviewed its application in organic synthesis.140 In addition, the Na2Fe(CO)4-mediated synthesis of hemifluorinated ketones using perfluoroalkyl halides was also reported.141 Manganese has been scarcely used in carbonylation reactions. Recently, Alexanian described a manganese-catalyzed carbonylative cascade reaction of alkenes with alkyl iodides (Scheme 29).142 Using commercially available manganese carbonyl as the catalyst, γ-iodo-alkenes reacted with CO and produced cyclopentyl esters in ethanol. A range of carbocyclic and heterocyclic compounds were synthesized with moderate to

Scheme 30. Manganese-Catalyzed Carbonylation of Alkyl Iodides with Amides

and K2CO3 as the base, alkyl iodides were successfully carbonylated with amides as the nucleophiles. A variety of substituted imides were produced in moderate to good yields. Radical scavenging experiments and EPR investigations suggested that an alkyl radical was formed in this reaction. The reaction proceeded via a mechanism similar to that mentioned above (Scheme 30b). K

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

2.2.2. Carbonylative Coupling Reactions of C(sp3)−X. One of the main problems of nickel-catalyzed carbonylation reactions is that CO coordinates very strongly to nickel and leads to coordinatively saturated Ni(CO)4. This limits the oxidative addition of C−X bond to nickel and the migratory insertion of CO. Skrydstrup and co-workers developed a nickel(II)/NN2 pincer complex that shows excellent catalytic activity in the carbonylative Negishi coupling of benzyl bromides with alkylzinc reagents (Scheme 31a).144 The authors

Scheme 32. Cu/Mn Bimetallic Catalyzed Carbonylative Coupling of Alkyl Halides and Arylboronic Esters

Scheme 31. Carbonylative Negishi Coupling Catalyzed by a Nickel(II)/NN2 Pincer Complex

single-electron transfer followed by a CO insertion to generate acylmanganese electrophile 32-B. Reaction of these two metal complexes released the product and regenerated the catalysts (Scheme 32b). Recently, the field of Cu−H-catalyzed reactions has witnessed impressive achievements.146−148 In particular, the Cu−H-catalyzed hydrofunctionalization of alkynes provides a powerful approach to the synthesis of various functionalized alkenes.149−154 The key to these reactions is the addition of Cu−H to the alkyne, thus providing an alkenylcopper species which reacted with numerous electrophiles to give the functionalized alkenes. More recently, Mankad and colleagues demonstrated that the alkenylcopper intermediate can be used in the carbonylative coupling reaction with alkyl halides in the presence of CO (Scheme 33a).155 Using carbene−Cu as the

proposed that the strong tridentate bonding of the pincer to nickel might prevent the bonding of multiple CO molecules to the catalyst center, thus facilitating the carbonylation process. Using a nickel(II) pincer complex (NiL) as the catalyst, a range of benzyl alkyl ketones were synthesized in good yields from the corresponding parent compounds. Notably, the controlled release of CO was also important for this reaction, since the rapid release of CO would result in catalyst inactivation. Mechanistic studies indicated that the reaction proceeded via the intermediacy of a carbon-centered radical. The transmetalation of nickel(II) pincer complex 31-A with alkylzinc reagent produced the alkylnickel complex 31-B, which provided acylnickel complex 31-C after coordination and insertion of CO. 31-C reacted with benzyl bromide via a single-electron transfer to generate Ni(III) complex 31-D and a benzylic radical. The benzylic radical reacted with a second intermediate complex 31-C to give acyl-alkylnickel(III) complex 31-E, which released the product after reductive elimination. The simultaneously generated Ni(I) complex 31-F reacted with 31-D to regenerate catalysts 31-A and 31-C (Scheme 31b). Recently, Mankad and colleagues developed a Cu/Mn bimetallic catalyst system for the carbonylative Suzuki−Miyaura coupling of alkyl halides (Scheme 32a).145 With Cu carbene and Na[Mn(CO)5] complexes as cocatalysts, arylboronic esters reacted with alkyl halides to provide the corresponding ketones under a CO atmosphere. A series of arylboronic acids as well as alkyl halides were tolerated well. Mechanistic studies revealed that the Cu carbene complex reacted with the arylboronic ester via transmetalation to form an organocopper nucleophile 32-A. Meanwhile, the Mn carbonyl activated the alkyl halide by

Scheme 33. Cu-Catalyzed Hydrocarbonylative Coupling of Alkyl Halides with Alkynes

catalyst and polymethylhydrosiloxane (PMHS) as the reducing reagent, terminal alkynes reacted with primary and secondary alkyl iodides and produced unsymmetrical dialkyl ketones. While the reaction of terminal alkynes with tertiary bromides generated allylic alcohols, the reaction of internal alkynes with alkyl iodides also produced allylic alcohols (Scheme 33b). The authors proposed that the reaction proceeds via a two-step L

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

formylation,161−165 which have been reviewed by Kalck in 2009166 and Börner in 2012.156 We will not repeat these reviews here and only briefly discuss the mechanism of cobaltcatalyzed (Scheme 34) and nickel/iron-catalyzed (Scheme 35) hydroformylation reactions.

sequence: First, the Cu-catalyzed hydrocarbonylative coupling of the alkyne with the alkyl halide produced the unsaturated ketone intermediate 33-B. Then Cu-catalyzed 1,2- and 1,4reductions of the unsaturated ketone intermediate generated an allylic alcohol and an alkyl ketone, respectively. Notably, mechanistic studies indicated that a single-electron-transfer (SET) radical carbonylation pathway was involved in reactions featuring primary and secondary alkyl halides. However, for tertiary halides, an acyl halide might be formed via radical atom transfer carbonylation (ATC). In addition, the choice of ligand and the steric effects of the substrates played important roles in the selectivity (Scheme 33c).

Scheme 34. General Mechanism for Cobalt-Catalyzed Hydroformylation of Alkenes

3. CARBONYLATION OF ALKENES AND ALKYNES 3.1. Hydroformylation of Alkenes

Hydroformylation,156 also known as the oxo synthesis or oxo process, is the addition reaction of syngas (a mixture of CO and H2) to alkenes to form aldehydes. A formyl group (CHO) and a hydrogen atom are added to the olefinic double bond under the catalysis of a transition metal. The first hydroformylation was discovered by Otto Roelen in 1938 when he was studying the Fischer−Tropsch reaction.157 Roelen discovered that propanal and diethyl ketone were obtained when ethylene was treated with a high pressure of CO and H2 in the presence of a cobalt catalyst. Since then, this process has undergone continuous growth. Various transition metals neighboring rhodium on the periodic table have shown activity in hydroformylation. Generally, the activities of the unmodified metals are in the order Rh ≫ Co > Ir, Ru > Os > Pt > Pd ≫ Fe > Ni.158 The commonly accepted catalytically active species for the hydroformylation of olefins are the hydridometal complexes [HM(CO)xLy], where L might be CO or an organic ligand. Nowadays, with regard to its large annual production of oxo products, this transformation is considered one of the most important homogeneous catalytic processes in industry. Although various transition metals have been investigated in laboratories for hydroformylation, only Rh and Co are used in industrial processes. In fact, since the 1970s, most applications of hydroformylation have been based on rhodium catalysts. Compared to rhodium, especially for the hydroformylation of propene, cobalt-catalyzed hydroformylation suffers the disadvantages of poor regioselectivity and harsh reaction conditions (high CO pressure and high temperature). However, in addition to its economic advantages (rhodium catalysts are much more expensive than cobalt catalysts), the use of cobalt has some other advantages over rhodium. The recovery of the rhodium catalyst from the produced aldehydes is difficult because of the high instability of hydridorhodium carbonyl as well as the precipitation of insoluble rhodium clusters. Cobalt has turned out to be an optimal catalyst concerning these problems. The isomerization activity of hydridocobalt carbonyl predominantly yields terminal aldehydes even for internal olefins. For the problem of catalyst recycling, various improvements have been established159,160 with the most widely used method being oxidative decobalting. Thus, cobalt catalysts are still used for the synthesis of C10−C16 aldehydes in industrial processes, whereas Rh catalysts are usually used for the hydroformylation of propene. Another advantage of cobalt catalysts is that they are usually robust toward the poisons in the feedstock. Significant efforts have been made by several groups to determine the mechanism of the cobalt-catalyzed hydro-

Scheme 35. General Mechanism for Nickel/Iron-Catalyzed Hydroformylation of Alkenes

3.2. Carbonylation of Alkynes

3.2.1. Iron-Mediated Carbonylation of Alkynes. As early as in 1953, Reppe and Vetter reported their studies on the Fe(CO)5-promoted cyclization of acetylene in alkali aqueous solution for the synthesis of hydroquinone derivatives.167 By the cyclization of two molecules of acetylene and two molecules of carbon monoxide, via the (2 + 1 + 2 + 1) model, hydroquinones can be produced. However, in the case of ethene, under pressure of syngas, propan-1-ol was formed with propionaldehyde as the intermediate. In 1996, Periasamy, Brunet, and co-workers discovered that a reactive iron reagent was generated from the reaction of NaHFe(CO)4 and MeI in THF, which reacted with various substituted alkynes to give the corresponding cyclobutenediones and α,β-unsaturated carboxylic acids after CuCl2 oxidation (Scheme 36).168 Fe(CO)4 (I) or polymeric iron carbonyl species were believed to be the active species. Thus, the α,β-unsaturated carboxylic acids were formed from the reaction of alkynes with hydridoiron carbonyl species, while the cyclobutenedione might result from cyclic intermediate II. Two years later, the authors found that this active iron species could also be produced using the NaHFe(CO)4/ M

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 36. Reaction of Alkynes with NaHFe(CO)4 for the Synthesis of Cyclobutenediones and α,β-Unsaturated Carboxylic Acids

Scheme 38. Preparation of Cyclic Imides

CH2Cl2 system, and the corresponding α,β-unsaturated carboxylic acids were obtained as the major products in moderate yields (Scheme 37a).169 Interestingly, the reaction Scheme 37. Synthesis of Cyclobutenediones and/or α,βUnsaturated Carboxylic Acids with Different Activating Reagents

carbonyl reagents since it can decarbonylate the Fe(CO)5 to produce the coordinatively unsaturated “Fe(CO)4” species.173,174 When the alkynes were treated with Fe(CO)5 in the presence of pyridine-N-oxide (1:1) at 70 °C, the corresponding cyclic anhydrides were obtained in moderate yields after CuCl2·2H2O oxidation. Two years later, the authors realized the synthesis of cyclobutenediones with Fe(CO)5 using inexpensive Me3N+O− as the activating reagent. Interestingly, when an excess of Me3N+O− was used, the corresponding cyclic anhydrides were obtained as the main products in good yields.175 Recently, a regioselective hydrocarbonylation of phenylacetylene for the synthesis of trans-α,β-cinamyl esters has been reported with the Fe(CO)5/ROH/DABCO system.176 The first catalytic carbonylation process with iron as the catalyst was reported by Beller and co-workers in 2009 (Scheme 39a).177 In the presence of catalytic quantities of Fe(CO)5 or Fe3(CO)12, a series of substituted succinimides were successfully synthesized from the reaction of alkynes with amines (or ammonia). Both internal and terminal alkynes were transformed to succinimide derivatives in moderate to good yields. The pressure of CO played an important role in this reaction. When 3-hexyne was treated with ammonia under a CO pressure of 20 bar, an excellent yield of 93% was obtained. However, when a higher CO pressure of 50 bar was used, the yield dropped significantly to 40%. This may have resulted from the lower accessibility of the metal center in a high CO pressure environment. Mechanistically, the precatalyst Fe(CO)5 or Fe3(CO)12 reacts with the amine to generate an active [Fe2(CO)8] species, which reacts with alkynes in the presence of CO to produce the cyclic intermediate 39-A or 39-B. In the presence of excess amine, nucleophilic attack of the amine on one of the carbonyl groups in 39-A gives intermediate 39-C. Finally, intramolecular amidation and reduction produces the succinimide product (Scheme 39b). Substituted maleinimides and succinimides exist widely in a variety of natural products which exhibit diverse biological activities. Beller and co-workers developed a straightforward two-step process for the synthesis of trans-3,4-disubstituted succinimides (Scheme 40a).178 By combining the palladiumcatalyzed Sonogashira reaction and iron-catalyzed double carbonylation of alkynes, a range of trans-3,4-disubstituted

using NaHFe(CO)4/Me3SiCl system produced α,β-unsaturated carboxylic acids at 25 °C, while the corresponding cyclobutenediones became the major products when the reaction was performed at 60 °C (Scheme 37b). Furthermore, the in situ prepared [HFe3(CO)11]− species using Fe(CO)5/NaBH4/ CH3CO2H also reacted with alkynes to give the cyclobutenediones in good yields (Scheme 37c). In subsequent investigations, the authors found that even simple FeCl3 was suitable for the in situ preparation of iron carbonyl species.170 Moreover, the acyloxyferrole complexes II were prepared by the reaction of Fe3(CO)12/Et3N with alkynes in THF, which upon reaction with Br2 in dichloromethane at −78 °C provided the corresponding cyclobutenediones in good yields.171 In 2002, Periasamy and co-workers developed a convenient procedure for the synthesis of cyclic imides via alkyne−iron carbonyl complexes (Scheme 38a).172 Readily available iron carbonyl reagents such as Fe(CO)5 and Fe3(CO)12 were used for the preparation of the active iron carbonyl species Fe3(CO)11, which reacted with alkynes to generate the alkyne−iron carbonyl complex 38-I or 38-II. Subsequently, the alkyne−iron carbonyl complex reacted with excess amine, and produced cyclic imides in moderate to good yields after CuCl2·2H2O oxidation (Scheme 38b). Furthermore, the Noxide R3N+O− has been investigated to activate the iron N

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Sonogashira coupling of 40-1 and 40-2 produced the bisindolylacetylene 40-3. Iron-catalyzed double carbonylation followed by oxidative dehydrogenation produced the corresponding 3,4-bisindolylmaleimide 40-4. Furthermore, the ironcatalyzed double carbonylation was used as a key step in the total synthesis of himanimides A and B (Scheme 40c).179 The palladium cross-coupling aryl bromide 40-5 and 3-phenyl-1propyne delivered the internal alkyne 40-7, which provided himanimide A after iron-catalyzed double carbonylation and oxidative dehydrogenation. Then, Sharpless asymmetric dihydroxylation produced himanimide B with a yield of 90% and 60% enantiomeric excess (ee). Based on their success with iron-catalyzed double aminocarbonylation, Beller et al. investigated the selective monocarbonylation of alkynes to synthesize the α,β-unsaturated amides (Scheme 41a).180 Using Fe3(CO)12 as the precatalyst,

Scheme 39. Iron-Catalyzed Carbonylation for the Synthesis of Succinimides

Scheme 41. Iron-Catalyzed Aminocarbonylation of Alkynes: Synthesis of α,β-Unsaturated Amides

Scheme 40. Application in the Synthesis of Succinimides and Maleinimide Natural Products

treatment of terminal alkynes with amines under a CO pressure of 10 bar produced α,β-unsaturated amides with good linear and (E) selectivities. Higher CO pressure resulted in double carbonylation products. Notably, the use of 1,4-diazabutadiene ligand L improved the chemoselectivity significantly. Internal alkynes also produced the monocarbonylation products in good yields. Two independent reaction pathways were proposed for the monocarbonylation and double carbonylation reactions. In contrast to the concerted double carbonylation mechanism for the succinimide formation, the monocarbonylation proceeds in a stepwise manner. Fe3(CO)12 reacted with amine to give an iron hydride carbonyl cluster, which reacted with the alkyne to give an acylcarbonyliron complex. Subsequently, nucleophilic attack by the amine generated the α,β-unsaturated amide and regenerated the active iron species for the next cycle (Scheme 41b). Shortly afterward, Petricci and co-workers reported a microwave-assisted iron-catalyzed aminocarbonylation of ynamides (Scheme 42).181 Compared to the traditional heating procedure, the microwave-irradiated reaction proceeded under a low pressure of CO (1.3 bar) with a dramatically decreased reaction time (20 min). When MeOH was used as the nucleophile, the (E)-3-amidoacryl ester was obtained.

succinimides were prepared from readily available aryl halides and terminal alkynes. Succinimides could be further transformed to the corresponding maleinimides via the oxidative dehydrogenation by DDQ. The authors used this strategy in the synthesis of the advanced intermediate of arcyriarubins (Scheme 40b). As shown in Scheme 40, 3-alkynylindole 40-2 could be obtained in two steps from 3-bromoindole 40-1. O

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 42. Microwave-Assisted Iron-Catalyzed Aminocarbonylation

Scheme 44. Nickel-Catalyzed Hydrocarboxylation of Alkynes with Formic Acid

Mathur and co-workers developed a Fe(CO)5-catalyzed carbonylation of alkynes with the assistance of photolysis. By photolysis of a solution of alkynes with catalytic quantities of Fe(CO)5 in an alcoholic solvent in the presence of CO, α,βvinylester and γ-alkoxy-γ-lactones could be obtained as the major products depending on the photolysis time (Scheme 43a).182 In 2011, Ogawa and co-workers developed a cobaltScheme 43. Metal Carbonyl Catalyzed Lactonization of Alkynes

anhydride and trifluoromethanesulfonic anhydride were ineffective. DFT calculations also showed that a controlled generation of CO was crucial for this process since fast release of CO would poison the Ni catalyst. Mechanistically, alkyne coordinated to the in situ generated bisphosphine-ligated nickel monocarbonyl complex 44-A to form an alkyne−Ni complex, which reacted with formic acid to generate an alkenyl−Ni species 44-B. Coordination and insertion of CO would generate the intermediate 44-C, which produced a formic anhydride 44-D after reductive elimination. The formic anhydride decomposes to form the acid and release CO (Scheme 44c). In 2008 Ogoshi’s group reported the preparation of a nickeladihydrofuran by oxidative cyclization of an alkyne and an aldehyde with nickel(0).191 By treatment of the nickeladihydrofuran under CO pressure (5 bar), lactones can be produed quantitatively (Scheme 45a). In 2014, Ogoshi and coworkers developed a nickel-catalyzed [2 + 2 + 1] carbonylative cycloaddition reaction of imines and alkynes using phenyl formate as CO source (Scheme 45b).192,193 A series of Nbenzenesulfonyl- and tosyl-substituted imines reacted with alkynes in the presence of 10 mol % Ni(cod)2, 20 mol % PCy3, and 1.5 equiv of phenyl formate to give the corresponding γlactams in good yields. Norbornene was found to be a suitable substrate, and four bicyclic γ-lactams were obtained in good yields under the same conditions. The usage of phenyl formate as a CO source played a key role in this reaction. The concentration of CO was a crucial factor for nickel(0)-catalyzed carbonylative cycloaddition, since too-high CO concentrations facilitate the formation of the catalytically unreactive nickel carbonyl complexes Ni(CO)3L, but the CO concentration needs to be high enough to react with the heteronickelacycle intermediate. Recently, the authors reported a nickel(0)catalyzed [2 + 2 + 1] carbonylative cycloaddition of 1,5- and 1,6-ene-imines with gaseous CO (Scheme 45c).193 In the presence of CO, the catalytically active Ni0 species could be

catalyzed thiolative lactonization of terminal alkynes with thiols (Scheme 43b). Using Co2(CO)8 as the catalyst, terminal alkynes reacted with thiols with incorporation of two molecules of CO and afforded α,β-unsaturated γ-thio-γ-lactones in moderate yields.183 Subsequent studies indicated that the internal alkynes were also suitable substrates for this transformation.184 3.2.2. Nickel-Mediated Carbonylation of Alkynes. In 2015, Zhou and co-workers reported an elegant palladiumcatalyzed hydrocarboxylation of acetylene with formic acid, and acrylic acid was obtained with a turnover number (TON) of up to 350.185 The group of Shi also reported the palladiumcatalyzed hydrocarboxylation of alkenes with formic acid.186,187 In 2016, Shang et al. developed a well-designed nickel-catalyzed hydrocarboxylation of alkynes with formic acid (Scheme 44a). 188 Using Ni(acac) 2 as the catalyst and 1,2-bis(diphenylphosphino)benzene (dppbz) as the ligand, various alkynes were hydrocarboxylated with formic acid in the presence of a catalytic quantity of carboxylic acid anhydride. A broad range of functional groups were compatible with this reaction, and various substituted α,β-unsaturated carboxylic acids were obtained in good yields with high regio- and stereoselectivities. Almost simultaneously, Zhou and co-workers reported their results on this transformation and a high TON of up to 7700 was obtained with di(tert-butylmethylphosphino)benzene as the ligand (Scheme 44b).189 Density functional theory (DFT) calculations190 by Shang et al. revealed that the bisphosphine-ligated nickel monocarbonyl complex 44-A was the active catalytic species for this reaction. Bidentate phosphine ligands such as dppbz were found to promote the coordination of alkyne to the metal center. The acid anhydride as additive also played an important role. Pivalic anhydride, acetic anhydride, and benzoic anhydride were found to be effective, while other anhydrides such as trifluoroacetic P

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

A widely accepted mechanism for the Pauson−Khand reaction was first proposed by Magnus and Principle,203 which has been confirmed by experiments204 and theoretical studies.205 First, the alkyne reacts with Co2(CO)8 to form the isolable alkyne−Co2(CO)6 intermediate 48-A. The olefin coordinates and inserts into the less hindered Co−C bond of the hexacarbonyl complex 48-B to give cyclic cobalt complex 48-C with the loss of one CO ligand. The loss of CO was demonstrated to be important for the overall reaction rate. Tertiary amine N-oxides were usually used as accelerating agents since N-oxides facilitate the generation of free coordination sites at cobalt by oxidative liberation of CO ligands. Then, migratory insertion of a CO group of the cyclic cobalt intermediate 48-C gives intermediate 48-D. Reductive elimination delivers the cyclopentenone product via the intermediate 48-E (Scheme 48). In addition to Co2(CO)8, many other metal carbonyl complexes, such as Mo(CO)6, Fe(CO)5, and W(CO)5, were all reported to facilitate the Pauson−Khand reaction. However, Co2(CO)8 has become the best choice of catalyst, and a broad range of functional groups are tolerated well. Both terminal and internal alkynes are suitable substrates.206−219

Scheme 45. Nickel-Catalyzed [2 + 2 + 1] Carbonylative Cycloaddition Reaction of Imines and Alkynes

4. CARBONYLATION OF ALDEHYDES regenerated from nickel(0) carbonyl complexes Ni(CO)3L. Thus, by heating 1,5- and 1,6-ene-imines with 1 mol % Ni(cod)2/PCy3 in toluene under 0.5 atm of CO, the corresponding tri- and tetracyclic γ-lactams were obtained in excellent yields without stirring. Interestingly, the yield of this reaction dropped dramatically when the reaction was conducted under vigorous stirring, indicating that rapid CO saturation hampers the progress of the catalytic carbonylation. 3.2.3. Pauson−Khand Reaction. The well-known transition-metal-mediated [2 + 2 + 1] cycloaddition of an alkyne, an alkene, and carbon monoxide for the construction of cyclopentenones is termed the Pauson−Khand reaction (PKR). The reaction was first reported by Pauson and co-workers in 1973 (Scheme 46). The readily prepared alkyne−hexacarbonyldicobalt complexes cyclized with norbornene upon heating to form the cyclopentenones in moderate yields.

4.1. Amidocarbonylation of Aldehydes

α-Amino acids, as the key building blocks in peptides and proteins, represent some of the most important structures in chemistry and biology. Efficient synthesis of the optically active amino acid from simple substrates remains a significant challenge even nowadays. From the viewpoint of atom economy, the amidocarbonylation220−222 of aldehydes provides a highly efficient strategy for the construction of these useful compounds. From abundant and cheap starting materials (aldehyde, amide, and carbon monoxide), N-acyl α-amino acid derivatives could be efficiently synthesized under the catalysis of transition metals. The first amidocarbonylation reaction was discovered serendipitously by Wakamatsu when he was investigating the cobalt-catalyzed hydroformylation of acrylonitrile (Scheme 49).223,224 Trace amounts of α-aminobutyric acid were detected in the oxo reactions of acrylonitrile with [HCo(CO)4]. Wakamatsu demonstrated that the α-aminobutyric acid was generated from the reaction of aldehydes and amides with syngas under cobalt catalysis. The reaction was conducted at 110−140 °C with a syngas pressure of 200 bar (CO/H2 = 3/1) and 2 mol % Co2(CO)8 was used as the catalyst precursor. A series of N-acyl α-amino acids were generated in moderate to high yields. Later, Pino et al. further elaborated the general applicability of this transformation and obtained N-acyl αamino acids in high yields.225 The carbonylation of aldehyde in the presence of amide is now known as amidocarbonylation. Numerous natural and non-natural amino acid derivatives can be synthesized with this method from the corresponding aldehydes. The active catalytic species in the cobalt-catalyzed amidocarbonylation is [Co(CO)4]−, which is generated from [Co2(CO)8] in the presence of CO/H2. A proposed mechanism is illustrated in Scheme 50. Initially, the nucleophilic addition of the amide to the aldehyde gave an N-α-hydroxyalkyl amide, which forms the cationic species I after protonation. Reaction of I with [Co(CO)4]− forms the alkylcobalt complex II, which is then transformed into the acylcobalt intermediate III by insertion of CO. Finally,

Scheme 46. First Pauson−Khand Reaction

Since its discovery, the Pauson−Khand reaction has attracted great interest from synthetic chemists. During early investigations, stoichiometric amounts of Co2(CO)8 were usually needed and only strained alkenes such as norbornene provided acceptable yields. Moreover, long reaction times and high reaction temperatures were needed for complete conversion. During the past several decades, considerable efforts have been made to improve the efficiency and generality of this reaction. In addition, asymmetric approaches and hetero-Pauson−Khand reactions have also been established. This useful transformation has emerged as a powerful tool for the construction of cyclopentenone skeletons and has been successfully used as a key step in the total synthesis of complex natural products (Scheme 47).194−202 Q

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 47. Applications of the PKR in the Total Synthesis of Complex Natural Products

Scheme 48. Mechanism of the Pauson−Khand Reaction

Scheme 50. General Mechanism for the Cobalt-Catalyzed Amidocarbonylation

Scheme 49. Discovery of Amidocarbonylation

Scheme 51. Diamidocarbonylation of Primary Amides

hydrolysis of III releases the corresponding N-acyl α-amino acid product and regenerates the catalyst for the next cycle.226 Notably, for cobalt-catalyzed amidocarbonylation, only aldehydes with α-hydrogens and formaldehyde could be used. In 1982, a diamidocarbonylation of primary amides with excess formaldehyde was developed by Chauvin and co-workers, in which N-acylamino diacetic acids were produced in the presence of a controlled amount of water (Scheme 51).227 Ojima and co-workers developed a binary catalyst system for the isomerization−amidocarbonylation of allylic alcohols and oxiranes. With [HRh(CO)(PPh3)3] as the cocatalyst, allylic and

homoallylic alcohols underwent a rearrangement to give aldehydes, and [HCo(CO)4]-catalyzed amidocarbonylation of the in situ generated aldehydes produced the N-acyl amino acids (Scheme 52a).228 On the other hand, in the presence of a Lewis acid such as Ti(OiPr)4, styrene oxide and propene oxide could also be efficiently isomerized to the corresponding aldehydes, which were incorporated into the subsequent amidocarbonylation to give the N-acyl amino acid product (Scheme 52b).229 A regioselective hydroformylation−amidocarbonylation of fluoro olefins was realized with a binary R

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

be [R3Si]+[Co(CO)4]−, which was generated from the reaction of hydrosilane and Co2(CO)8. The Lewis acidic silicon binds the unshared electrons of the carbonyl oxygen, and thus facilitates the nucleophilic attack of [Co(CO)4]− at the carbonyl carbon. The subsequent CO insertion and hydrosilane reduction yielded the formylated products.

Scheme 52. Domino Amidocarbonylation Processes

5. CARBONYLATION OF HETEROCYCLES Carbonylation of epoxides, aziridines, oxazines, lactones, and thietanes234 incorporates carbon monoxide into heterocyclic compounds and provides an efficient method for the synthesis of various high-value compounds. 5.1. Alkoxy- and Aminocarboxylations of Heterocycles

In 1977, Murai and co-workers discovered that treatment of tetrahydrofuran, oxetane, and 1,2-epoxycyclohexane with diethyl(methyl)silane and carbon monoxide in the presence of a catalytic quantity of Co2(CO)8 produced ring-opened, silylprotected hydroxyaldehydes (Scheme 54a,b).235 Murai and coworkers elaborated a cobalt-catalyzed siloxymethylative ring opening of cyclic ethers with the HSiR3/CO/Co2(CO)8 catalytic system.236−239 The reaction proceeded under mild conditions (room temperature and atmospheric pressure of CO) and resulted in high product selectivity. The siloxymethylative ring opening of epoxides led to synthetically useful silyl-protected 1,3-diol derivatives. A broad range of substituted ethylene oxides were tested and resulted in highly regioselective ring opening at the less substituted carbon center. One exception is that the ring opening and CO insertion of styrene oxide occurred at the secondary carbon center. For the reaction of cycloalkene oxides, the stereochemistry of the corresponding products was demonstrated to be trans. A silicon−cobalt complex R′3SiCo(CO)4, generated from the reaction of R′3SiH with Co2(CO)8, was believed to be the reactive catalytic species. The coordination of epoxide oxygen to the silicon atom of R′3SiCo(CO)4 formed a silyloxonium ion. Nucleophilic attack of Co(CO)4− resulted in carbon−oxygen bond cleavage which resulted in alkylcobalt complex 54-A with inversion of configuration. Migratory insertion of CO and reduction of the acylcobalt intermediate with R′3SiH produced the aldehyde 54-

catalyst system. Linear and branched N-acyl α-amino acids were obtained with excellent regioselectivity by using Co2(CO)8 and Co2(CO)8−Rh6(CO)16 as catalysts, respectively (Scheme 52c). Subsequent mechanistic studies showed that the regioselectivity was governed by the hydroformylation step and the catalytic active species for the latter is bimetallic complex CoRh(CO)7.230 4.2. Silylformylation of Aldehydes

In 1978, Murai and colleagues developed a Co2(CO)8/PPh3catalyzed silylformylation of aliphatic aldehydes (Scheme 53a).231 When the aliphatic aldehydes were treated with a 3fold excess of hydrosilane and CO in the presence of Co2(CO)8, 1,2-bis(siloxy)olefins were obtained in moderate yields. The use of PPh3 prevented the hydrosilylation side reaction of the aldehydes. Interestingly, when an excess of aldehyde (3-fold with respect to the hydrosilane) was used, homologated siloxyaldehydes were obtained (Scheme 53b).232 In addition, the same group extended this catalytic system to the silylcarbonylation of ketones. Cyclobutanones reacted with the hydrosilane at 175 °C under CO atmosphere and produced disiloxycyclopentenes (Scheme 53c).233 The reaction proceeded via a tandem formylation/cationic ring expansion process. The active catalyst of these reactions was believed to

Scheme 53. Silylformylation of Aliphatic Aldehydes and Cyclobutanones

S

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 54. Cobalt-Catalyzed Ring-Opening Carbonylation of THF and Epoxides

Scheme 56. Cobalt-Catalyzed Ring-Opening Carbonylation of Chiral Epoxides

resolution of racemic epoxides, this reaction provides a twostep, cobalt-catalyzed reaction for the synthesis of optically active β-hydroxy esters from racemic epoxides. The presence of 3-hydroxypyridine played an important role in this reaction. Interestingly, Denmark and co-workers discovered that the ring-opening carbonylation of simple terminal epoxides can be performed under atmospheric pressure CO and at room temperature with Co2(CO)8 as catalyst without the use of 3hydroxypyridine (Scheme 56b).242 A series of monosubstituted ethylene oxides were converted to the corresponding β-hydroxy esters in modest to good yields. In addition, enantiomerically pure methyl 3-hydroxybutanoate was prepared from chiral propylene oxide with complete retention of configuration. However, the scope of this methodology is somewhat limited, and epoxides bearing carbonyl groups and 1,2-disubstituted epoxide resulted in lower yields. In 2002, the group of Jacobsen developed a cobalt-catalyzed low-pressure carbonylation protocol with 4-(trimethylsilyl)morpholine (Scheme 56c).243 Enantiomerically pure β-hydroxy morpholine amides were effectively obtained under mild conditions. The use of silyl amides effectively activated the epoxides and allowed the reaction to take place under mild conditions. Similar to Watanabe’s work, the amide products were also generated by the catalyst [R2N(SiMe3)2][Co(CO)4], which was effective for the activation of epoxides and ring opening owing to its dual Lewis acidic and strongly nucleophilic character. A series of functional groups, such as ethers, olefins, halides, and esters, were well-suited to the process and the β-hydroxy morpholine amides could be obtained by treatment of the crude product mixture with aqueous acid in >95% purity without chromatography.

C. The final product was produced by the hydrosilylation of 54C (Scheme 54c). In 1994, Watanabe and co-workers reported a Co2(CO)8catalyzed ring-opening carbonylation of epoxides with Nsilylamines (Scheme 55). Primary and secondary N-silylamines Scheme 55. Cobalt-Catalyzed Ring-Opening Carbonylation with N-Silylamines

were found to be effective in the carbonylation of the epoxide reaction to produce β-siloxy amides with high regioselectivity.240 It should be noted that the reaction proceeded under 1 atm of CO at room temperature, while the use of high CO pressure and reaction temperatures resulted in poor selectivity. Oxetane was also suitable in this ring-opening transformation and gave the corresponding γ-siloxy amide in 75% yield. The authors suggested that a [(Me3Si)2NR2R3][Co(CO)4] complex, generated from the reaction of Co2(CO)8 and Nsilylamine, was the catalytically active species. In 1999, Jacobsen et al. developed a regioselective methoxycarbolation of chiral epoxides (Scheme 56a).241 Using Drent’s catalytic system, Co2(CO)8 and 3-hydroxypyridine, as the catalysts, enantiomerically pure terminal epoxides were efficiently transformed to the corresponding β-hydroxy esters with a high degree of configurational retention. The CO insertion occurred selectively into the less substituted C−O bond and provided a linear product. Since the enantiomerically pure epoxides could be obtained by the hydrolytic kinetic

5.2. Ring-Expansion Carbonylation of Heterocycles

The carbonylation of epoxides incorporates carbon monoxide into heterocyclic compounds and provides an efficient method for the synthesis of various hydroxyl carbonyl compounds. For example, ring-expansion carbonylation of epoxides generates βlactones, which exist in a range of natural products (Scheme 57). Furthermore, alternating copolymerization of epoxides with CO provides direct access to the biodegradable poly(3hydroxyalkanoate)s. The carbonylation of epoxides had been explored as early as the 1950s.244,245 Aumann and co-workers reported the first successful catalytic ring-expansion carbonylation of epoxide in the late 1970s.246 Isoprene oxide was converted to an β,γunsaturated lactone in the presence of CO catalyzed by [Rh(cod)Cl]2. However, the generality of this reaction was limited and only one substrate was reported. A breakthrough method for the catalytic epoxide carbonylation was disclosed by T

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 57. Representative Natural Products Containing β-Lactone Groups

good yields. The CO insertion occurred selectively at the unsubstituted C−O bond of the epoxide ring. In the case of 1,2-disubstitued epoxides, B(C6F5)3 was used instead of BF3· Et2O owing to the lower reactivity of the substrates. Interestingly, the stereochemistry indicated a retention of the configuration, which is unusual since cobalt-catalyzed carbonylation of epoxides usually results in configurational inversion. In 2002, the group of Coates designed and prepared a catalyst based on an aluminum−salen complex,251,252 [(salph)Al(THF)2][Co(CO)4] (Scheme 60).253 The basic guidance of

Drent in a patent in 1994. Using Co2(CO)8 in the presence of 3-hydroxypyridine, terminal epoxides were transformed to the corresponding β-lactones with high conversion and selectivity.247 In 1996, Alper developed a Co2(CO)8-catalyzed carbonylation of aziridines for the synthesis of β-lactams (Scheme 58).248 The active catalyst was demonstrated to be the cobalt Scheme 58. Co2(CO)8-Catalyzed Carbonylation of Aziridines

Scheme 60. [(salph)Al(THF)2][Co(CO)4]-Catalyzed Carbonylation of Epoxides

tetracarbonyl anion, which was generated in situ by the reaction of aziridine with Co 2 (CO) 8 . Unlike the [Rh(CO) 2 Cl]2 catalysis,249 this reaction inserts CO into the less substituted carbon of the aziridine with inversion of configuration. The regio- and stereoselectivities resulted from the nucleophilic ring opening of the aziridine by tetracarbonylcobaltate, which proceeds in an SN2 manner. Highly strained trans-7-azabicyclo[4-2-0]octan-8-one derivatives were selectively synthesized in good yields. Simple cobalt-catalyzed carbonylation of epoxides usually suffers from harsh required conditions and limited substrate scope. Great efforts have been made in the development of efficient catalysts with high activity and selectivity. Alper and co-workers developed a regioselective carbonylation of epoxides in 2001, employing [PPN]+[Co(CO)4]− as the catalyst in conjunction with BF3·Et2O (Scheme 59).250 Various monosubstituted epoxides, including aliphatic epoxides, were regioselectively converted to the β-lactones with moderate to

the catalyst design was based on the proposed reaction mechanism (Scheme 61). The Lewis acidic cation coordinates and activates the epoxide, and [Co(CO)4]− attacks the less

Scheme 59. [PPN]+[Co(CO)4]−/BF3·Et2O-Catalyzed Carbonylation of Epoxides

Scheme 61. General Mechanism of [Lewis acid]+[Co(CO)4]−-Catalyzed Carbonylation of Epoxides

U

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

and selectivity in a broad array of epoxide carbonylations. The CO inserts regioselectively between the oxygen and the less substituted carbon of the epoxides. Notably, the identity of the metal and the properties of the ligand in the Lewis acid component have a pronounced effect on the catalytic activity and the substrate scope of carbonylation. In comparison to the Al−Co and Ti−Co systems, the Cr−Co catalyst showed the highest activity with respect to higher yield and lower catalyst loading. For example, the [(OEP)Cr(THF)2][Co(CO)4]catalyzed epoxide carbonylation for the synthesis of β-lactones gave excellent TONs (turnover numbers) up to 10 000 and TOFs (turnover frequencies) up to 1670 h−1. Various functional groups, including alkenes, ethers, esters, and amides, were well tolerated, producing substituted β-lactones in high yields and selectivities. Notably, the carbonylation of epoxides at low CO pressure has also been realized. When [(salph)Cr(THF)2][Co(CO)4] was used as the catalyst, various functionalized β-lactones were synthesized in excellent yields under atmospheric pressure of CO. Coates’s bimetallic catalytic systems showed high catalytic activity for terminal epoxides. However, most of these catalysts are air sensitive; therefore, the preparation and usage of these catalysts needs to be performed in the glovebox. In 2011, Ibrahim and co-workers developed an in situ generated catalytic system for the ring-expansive carbonylation of 1,2-disubstituted meso-epoxides. Using commercially available, air-stable (TPP)CrCl (TPP = tetraphenylporphyrinato) in conjunction with Co2(CO)8 as a catalyst, cyclic and acyclic 1,2-disubstituted meso-epoxides were carbonylated to produce trans-β-lactones in high yields (Scheme 63a).260 Soon after, an asymmetric ringexpansive carbonylation of 1,2-disubstituted meso-epoxides was reported by the same group. Using Jacobsen’s catalyst [(R,R)(salen)Cr]Cl and the D4-symmetric [(+)-porphyrin]CrCl together with Co2(CO)8, moderate levels of asymmetric induction were obtained with up to 56% ee (Scheme 63b, left).261 Coates prepared a series of DABN-based catalysts, such as (R)-63-A and 63-B, which give good to excellent enantioselectivities for the symmetric carbonylation of mesoepoxides. Enantioenriched trans-β-lactones were isolated in moderate yields and with >90% ee (Scheme 63b, right).262−264 In a different approach, Rieger developed the Co2(CO)8/ AlMe3 catalytic system, which was also found to be active for the carbonylation of terminal epoxides.265−267 Via online ATRIR techniques, it was identified that the in situ generated

hindered carbon from the backside of the C−O bond to give a ring-opened species. Migratory insertion of CO to the cobalt− alkyl bond and recoordination of CO to the cobalt center generates an acylcobalt intermediate. Intramolecular attack of the aluminum alkoxide at the acyl carbon produces the βlactone and regenerates the catalyst. [(salph)Al(THF)2][Co(CO)4] showed excellent activity and selectivity with a variety of epoxides. (R)-Propylene oxide, readily available through Jacobsen’s hydrolytic kinetic resolution,254 is converted to (R)β-butyrolactone with >98% retention of configuration. (R)-βButyrolactone is an important precursor for the biodegradable thermoplastic polyester poly((R)-β-hydroxybutyrate). The authors investigated the carbonylative kinetic resolution of racemic trans-2,3-epoxybutane using the chiral cobalt complex [(R,R-salcy)Al(THF)2][Co(CO)4].255 Shortly afterward, the same group prepared and tested a series of [Co(CO)4]−-based bimetallic catalysts, [Cp2Ti(THF)2][Co(CO)4], [(TPP)Cr(THF)2][Co(CO)4], [(OEP)Cr(THF)2][Co(CO)4], and [(salph)Cr(THF)2][Co(CO)4] (Scheme 62a).256−259 All complexes showed good activity Scheme 62. Carbonylation of Epoxides Catalyzed by [Co(CO)4]−-Based Bimetallic Catalysts

Scheme 63. In Situ Generated Catalytic System for Epoxide Carbonylation and Catalysts for Asymmetric Carbonylation

V

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 64. Application of the Ring-Opening Carbonylation of Epoxides in Total Synthesis

cobaltate salt [Me2Al(diglyme)]+[Co(CO)4]− was the active catalyst. Theoretical studies indicated that the Lewis acid influenced both the ring-opening step and the lactoneformation step.268 Additionally, the coordinating ability of the solvent also played an important role in the lactone formation rate.269 In 2011, Beuning, Coates and O’Doherty reported a de novo asymmetric synthesis of both (R)- and (S)-fridamycin E.270 The cobalt-catalyzed epoxide carbonylation was used as a key step for the formation of the tertiary β-hydroxyl carboxylic acids (Scheme 64a). With [(ClTPP)Al(THF)2]+[Co(CO)4]− as the catalyst, enantiomerically pure epoxide 64-A was efficiently carbonylated to give the β-lactone 64-B in 70% yield. Then, hydrolysis of β-lactone 64-B and reductive debenzylation complete the total synthesis of fridamycin E. From the enantiomer of 64-A, the total synthesis of unnatural (S)fridamycin E was also successful. Later, O’Doherty completed the synthesis of vineomycinone B2 methyl ester from the same intermediate 64-B.271 More recently, Coates and O’Doherty used a late-stage cobalt-catalyzed epoxide carbonylation in the total synthesis of tetrahydrolipstatin.272 Enantiomerically pure cis-epoxide 64-C was treated with [(ClTPP)Al(THF)2]+[Co(CO)4]− in the presence of CO at 50 °C, and trans-β-lactone tetrahydrolipstatin was obtained in 80% yield. The total synthesis of tetrahydrolipstatin was accomplished in 10 steps with an overall yield of 31% (Scheme 64b). In 2004, Coates and co-workers discovered that β-lactones could be efficiently converted to succinic anhydrides by a [Lewis acid]+[metal carbonyl]−-catalyzed carbonylation reaction (Scheme 65a).273 Similar to epoxide carbonylation, the reaction started with the coordination and activation of the βlactone by [Lewis acid]+. Nucleophilic attack by [Co(CO)4]− at the β-carbon of the lactone gives alkylcobalt species B with inversion of the configuration. After migratory insertion of CO, intramolecular nucleophilic attack of the aluminum carboxylate to the acyl carbon forms succinic anhydride and regenerates the catalyst (Scheme 65b). The solvent played a profound effect on the rate of the reaction; toluene increased the rate of β-lactone

Scheme 65. Cobalt-Catalyzed Carbonylation of β-Lactones

carbonylation, while donor solvents such as ethers and esters slowed the β-lactone carbonylation. Alkyl ether and olefinic substitution were tolerated in the transformation and produced the corresponding product in high yields. In 2007, a one-pot double carbonylation of the epoxides to form succinic anhydrides was developed by using the cobaltbased bimetallic catalyst [(ClTPP)Al(THF)2]+[Co(CO)4]− as the catalyst (Scheme 66).274 A broad range of substrates were converted directly to the corresponding anhydrides in good to excellent yields. Mechanistic studies by in situ IR spectroscopy revealed that this transformation proceeded in two stages: the carbonylation of epoxide to β-lactone and the carbonylation of W

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

68a).276 Substituted 3-hydroxy-δ-lactones, valuable skeletons of statin drugs,277 were synthesized selectively in moderate to

Scheme 66. Cobalt-Catalyzed Double Carbonylation of Epoxides to Form Succinic Anhydrides

Scheme 68. HCo(CO)4-Catalyzed Carbonylation for the Synthesis of δ-Lactones

β-lactone to succinic anhydride. These two carbonylations occurred sequentially without overlapping, and the carbonylation of β-lactone proceeded after all the epoxide was consumed. 1,2-Disubstituted and enantiomerically pure epoxides were carbonylated to the corresponding succinic anhydrides with excellent retention of configuration. A three-component reaction for the stereospecific construction of 1,3-oxazinane-2,4-diones with epoxides, isocyanates, and carbon monoxide was reported in 2007 (Scheme 67a).275 Using the Al−Co bimetallic catalyst, ring-opening

excellent yields. Screening of the catalysts revealed that [Lewis acid][Co(CO)4] facilitates the formation of β-lactone, while the use of HCo(CO)4 as the catalyst provides δ-lactones exclusively. The carbonylation proceeds with retention of the configuration. Thus, enantiomerically pure (R,R)-68-1 was converted to (R,R)-68-2 with retention of both stereocenters (Scheme 68b). In 2003, Komiya and co-workers reported a carbonylation of thietanes catalyzed by heterodinuclear platinum−cobalt complexes (Scheme 69a).278 When thietane was treated with

Scheme 67. Three-Component Reaction for the Construction of 1,3-Oxazinane-2,4-diones

Scheme 69. Carbonylation of Thietanes

(dppe)MePt-Co(CO)4 at 100 °C in the presence of CO, γthiobutyrolactone was produced in quantitative yield. 2Methylthietane also gives the corresponding γ-thiovalerolactone in 89% yield. The key to the catalytic activity is the presence of a platinum−cobalt bond in 69-A. Mechanistically, coordination of thietane to Pt, with the cleavage of the Pt−Co bond, gives intermediate 69-B. Then the [Co(CO)4]− anion attacks thietane from the less substituted carbon to give a dinuclear complex 69-C. Migratory insertion of CO into the cobalt−carbon bond and recoordination of CO to the cobalt center forms the acylcobalt 69-D. Intramolecular attack of the S atom at the acyl carbon atom generates the γ-thiovalerolactone and releases the catalyst for the next cycle (Scheme 69b).

nucleophilic attack of the Co(CO)4− to the epoxide and insertion of CO generated the key acylcobalt species 67-B, which was demonstrated to be the resting state of the catalytic process. The use of nonpolar, noncoordinating solvent hexane reduced the lactone formation, thus leading to an intermolecular reaction with isocyanates to give the intermediate 67-C, which yielded 1,3-oxazinane-2,4-dione after ring closing (Scheme 67b). Electron-deficient isocyanates gave higher reaction rates and selectivities, and a range of substituted epoxides were tolerated well. In 2007, Coates and co-workers developed a HCo(CO)4catalyzed carbonylation of substituted homoglycidols (Scheme X

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Since oxazolines could be obtained from β-amino alcohols, this transformation provides an alternative strategy for the synthesis of stereopure β-amino acid derivatives from commercially available α-amino acids. The readily prepared [Ph3SiCo(CO)4] reacted with oxazoline in the presence of alcohol and generated oxazolinium cobaltate species 71-A. [Co(CO)4]− attacked the 5-position of the oxazolinium to result in alkylcobalt species 71B with the inversion of the C5 stereocenter. Migratory insertion and recoordination of CO to the cobalt center produced the acylcobalt species 71-C. Ring closing and proton transfer yielded the product via the intermediate 71-D (Scheme 71b). In 2014, Lin, Pan, and Jia discovered a cobalt-catalyzed threecomponent cyclization of CO, imine, and epoxide for the synthesis of substituted 1,3-oxazinan-4-ones, which are valuable intermediates for many important pharmaceutical compounds (Scheme 72a).282 Easily available imines and epoxides were

A novel ring-expanding carbonylation of 2-aryl-2-oxazolines was discovered by Jia and co-workers while they were investigating the copolymerization of 2-oxazolines with CO (Scheme 70).279,280 2-Aryl-substituted 2-oxazolines reacted Scheme 70. BnCo(CO)4-Catalyzed Carbonylation of 2Oxazoline

with CO under the catalysis of BnCo(CO)4 and produced 2aryl-4,5-dihydro-1,3-oxazin-6-ones. The presence of an aromatic substituent at the C2 position of the 2-oxazoline is essential for this reaction, since 2-alkyl-2-oxazoline gives a polymer under the same conditions. Co2(CO)8 combined with AIBN as the catalytic system also showed good catalytic activity. In 2008, Coates investigated this reaction with the [Lewis acid][Co(CO)4] system and found that these complexes were poorly active for the carbonylation of oxazoline. However, the reaction proceeded successfully when they used silylcobalt complex [Ph3SiCo(CO)4], which was readily prepared by the reaction of Ph3SiH and Co2(CO)8, as the precatalyst (Scheme 71a).281 The substitution at the 4-position of the aromatic ring showed a tremendous effect on the yield: a 4-tBu substituted substrate provided an excellent yield. The reaction proceeded stereospecifically with exclusive configurational retention at the C4 atom and inversion at the C5 position of the oxazolines. This stereochemistry could be explained by the backside attack of the Co(CO)4− anion at the C5 position of the oxazolines.

Scheme 72. Cobalt-Catalyzed Three-Component Cyclization of CO, Imines, and Epoxides

Scheme 71. [Ph3SiCo(CO)4]-Catalyzed Carbonylation of 2Oxazoline and its Mechanism

treated with Co2(CO)8 and LiCl under a CO atmosphere to produce 1,3-oxazinan-4-ones. The addition of LiCl played an important role in this transformation; only trace amounts of product could be detected without the addition of LiCl. Ring opening and CO insertion occur selectively at the less hindered carbon of the epoxides. A series of aryl-substituted imines and monosubstituted or 1,1-disubstituted ethylene oxides were tolerated in this reaction. However, styrene oxide and cyclohexene oxide were nonreactive under these conditions. Furthermore, alkyl-substituted imines such as tert-butylidenemethylamine showed poor reactivity and resulted in very low yields of the desired product. Nearly at the same time, Sun and co-workers reported a similar three-component [3 + 2 + 1] cycloaddition reaction using the silylcobalt complex [Ph3SiCo(CO)4] as precatalyst (Scheme 72b).283 The catalytic active species was demonstrated to be HCo(CO)4, which was generated in situ by the alcoholysis of [Ph3SiCo(CO)4]. This catalytic system appears to be more efficient than that used in Y

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

tetrahydrofuran.290,291 In addition to epoxides and aziridines, imines were suitable substrates for carbonylative polymerization reactions. For example, Sun’s group established a cobaltcatalyzed copolymerization of imines and CO for the synthesis of polypeptides.292

the previous work. Both aryl- and alkyl-substituted imines and epoxides were successfully converted to the corresponding 1,3oxazinan-4-ones in high yields. The reaction proceeded via ring opening of epoxides by [HCo(CO)4] followed by migratory insertion of CO to form an acylcobalt species, nucleophilic attack of the imine to the acylcobalt species, and final cyclization (Scheme 72c). One important application of cobalt-catalyzed carbonylation of epoxides is the cobaltcatalyzed hydroformylation of ethylene oxides to produce 3hydroxyaldehydes.284 After Ru3(CO)12-catalyzed hydrogenation of the in situ formed 3-hydroxyaldehydes, 1,3-diols can be produced selectively. This process has been commercially practiced by Shell for the production of 1,3-propylene glycol.

6. OXIDATIVE CARBONYLATION REACTION In recent years, transition-metal-catalyzed C−H functionalization has attracted increasing interest and has witnessed significant development. Compared to the transition-metalcatalyzed carbonylation of aryl halides and sulfonates, the direct carbonylation of C−H bonds eliminates the prefunctionalization of the substrates and thus provides an attractive alternative for the synthesis of various carbonyl compounds. Until recently, although carbonylation of the majority of C−H bonds was achieved employing second- and third-row transition metals,293−304 great success has been obtained by using first-row transition metals such as cobalt, nickel, and copper.

5.3. Alternating Copolymerization of Epoxides

In addition to ring expansion, the cobalt-catalyzed carbonylation reaction is widely used in the so-called alternating copolymerization reaction. This reaction provides an efficient and direct approach to the synthesis of biodegradable poly(3hydroxyalkanoate)s. In 2002, Rieger285 and Osakada286 independently reported their studies on the alternating copolymerization of epoxides with CO (Scheme 73). Using

6.1. Oxidative Carbonylation of C(sp2)−H Bonds

In 2005, Daugulis initially introduced 8-aminoquinoline (Q) as the directing group for the palladium-catalyzed C(sp2)−H and C(sp3)−H bond functionalization.305,306 Since then, this bidentate directing group has been extensively exploited for transition-metal-catalyzed (Pd,307,308 Ru,309−311 Fe,312−315 Ni,316−319 Cu320−324) C−H functionalizations. It was believed that the tridentate, dianionic pincer structure of these auxiliaries stabilized the high-valent transition-metal intermediates, which were suspected to be the key intermediates in the C−H functionalizations (Scheme 74c). In 2014, Daugulis reported

Scheme 73. Cobalt-Catalyzed Alternating Copolymerization of Epoxides and CO

Scheme 74. Cobalt-Catalyzed Carbonylation of sp2 C−H Bonds Drent’s Co2(CO)8/3-hydroxypyridine catalyst system, the propylene oxide reacted with CO and produced poly(3hydroxybutyrate) via an alternating copolymerization process. Enaniomerically pure (S)- and (R)-propylene oxides underwent copolymerization and exclusively produced isotactic polyester with retention of configuration, while atactic polyester was obtained when racemic propylene oxide was used. This observation indicated that the tetracarbonyl cobaltate attacks the less hindered epoxide carbon atom from the backside. Then, the generated alkoxide attacks the second epoxide and forms ester groups intermolecularly. Owing to several chain termination reactions, the molecular weight of polymers obtained using the Co2(CO)8 catalyst was relatively low (Mn < 4000 g/mol). Subsequently, several groups investigated carbonylative polymerization reactions using cobalt carbonyl catalysts. Alper’s group developed a catalyst system based on Co2(CO)8/(6,7dihydro-5,8-dimethyldibenzo[b,f ]-1,10-phenanthroline)/BnBr, which showed good activity for the alternating copolymerization of epoxides and CO. High molecular weight atactic PHB [poly(β-hydroxybutyrate); Mn = 19 400, Mw/Mn = 1.63] was obtained.287 BnCOCo(CO)4, readily prepared by NaCo(CO)4 and phenylacetyl chloride, was found to be an effective catalyst for the alternating copolymerization of N-alkylaziridines with CO to produce poly-β-peptoids. 288 Jia and colleagues discovered that the isolated catalyst CH3COCo(CO)3[P(otolyl)3] was effective for the carbonylative polymerization of epoxides and aziridines.289 This catalyst could be used for the carbonylative copolymerization of N-alkylazetidines and

cobalt-catalyzed oxidative alkyne and alkene annulations with aminoquinoline amides. Ortho-functionalized benzoic acid derivatives were produced by cobalt-catalyzed C−H activation followed by alkyne325 or alkene326 insertion. Based on the success of these annulation reactions, Daugulis reported a cobalt-catalyzed carbonylation of sp2 C−H bonds. With aminoquinoline amide as the directing group, the reactions proceeded in trifluoroethanol at room temperature with Co(acac)2 as the catalyst and Mn(OAc)3·2H2O as the cocatalyst in the presence of CO. Atmospheric oxygen was demonstrated to be the terminal oxidant. Various functional groups including halogen, nitro, cyano, and ester substituents were compatible with the reaction conditions, and high levels of ortho-selectivity were obtained (Scheme 74a).327 The C−H bond of alkene was also tolerated, while the carbonylation of aminoquinoline amides of cinnamic and methacrylic acids delivered the corresponding products in moderate yields Z

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(Scheme 74b). Additionally, the directing group could be easily removed by treatment with methanolic ammonia, thus providing the parent phthalimides in moderate to high yields. The reaction might proceed via an aminoquinoline-directed cobaltation of the C−H bond to give a tridentate Co(III) complex (Scheme 74c). Recently, the same group developed a cobalt-catalyzed sp2 C−H bond carbonylative annulation of aminoquinoline sulfonamides (Scheme 75).328 The reaction proceeded with

intramolecular cyclization produced the desired product (path a, Scheme 76b). On the other hand, CO gas might be generated from the decomposition of DIAD. Coordination and insertion of CO into intermediate 76-A would form intermediate 76-C, which formed the product after reductive elimination (path b, Scheme 76b). Recently, Zhong et al. reported a cobalt-catalyzed carbonylation reaction of benzylamines using picolinamine as a traceless directing group (Scheme 77a).335 N-Benzylamides

Scheme 75. Cobalt-Catalyzed Carbonylation of sp2 C−H Bonds of Aminoquinoline Sulfonamides

Scheme 77. Cobalt-Catalyzed C−H Bond Carbonylation with DEAD as the CO Source

Co(OAc)2 as the precatalyst and Mn(OAc)2 as the co-oxidant. By employing diisopropyl azodicarboxylate (DIAD) as the carbon monoxide source,329 a series of saccharin derivatives were synthesized in moderate to good yields. Most carbonylation reactions use toxic gaseous CO as a cheap carbonyl source. Recently, the developments of green carbonylation based on CO surrogates have attracted more and more attention from organic chemists. Several practical methods have been developed for transition-metal-catalyzed C−H activation/carbonylation reactions. For example, Zhang et al. reported a novel cobalt-catalyzed C−H bond carbonylation of aryl amides with diisopropyl azodicarboxylate (DIAD) as a nontoxic carbonyl source (Scheme 76a).330 Stable Scheme 76. Cobalt-Catalyzed C−H Bond Carbonylation with DIAD as the CO Source were treated with DEAD (diethyl azodicarboxylate) in TFE (trifluoroethanol) with cobalt catalysts, and a variety of Nunprotected isoindolinones were obtained in good yields. Other azodicarboxylates such as diisopropyl azodicarboxylate and dibenzyl azodicarboxylate were also compatible in this process. Additionally, the reaction also occurred when gaseous CO was used as the carbonyl source. A range of functional groups were tolerated in this reaction, Notably, the 4-bromosubstituted substrate was conducted on a gram scale and was used in a formal synthesis of antibiotic natural product (+)-garenoxacin (Scheme 77b). Similar to Zhang’s work, the reaction might proceed via two possible pathways: an esteric radical process or a CO insertion process (Scheme 77c). The use of trifluoroethanol as the solvent played an important role in obtaining the N-unprotected products. In addition to cobalt-catalyzed carbonylation with azodicarboxylates, Koley and co-workers developed a copper-catalyzed ortho-selective transformation that enables the C−H carbonylation of benzamides (Scheme 78a).336 The reaction proceeded with Cu(NO3)2·xH2O as the catalyst and AIBN as a traceless cyanating reagent, whereby AgOAc was found to be the optimal oxidant. Various substituted benzamides were successfully converted to the phthalimide derivatives in good yields. However, only cyanated products were obtained when heteroarene substrates were used under the standard conditions. A site-selective C−H cyanation/intramolecular cyclization mechanism was proposed (Scheme 78b). Recently, a copper-catalyzed reaction of oximes with DIAD for the

and inexpensive Co(OAc)2·4H2O was used as the catalyst with Ag 2CO 3 as the oxidant. In addition to DIAD, other commercially available azodicarboxylates also showed good reactivity as the CO source. According to the mechanistic studies and previous literature,331−334 the authors proposed two plausible reaction pathways, namely esteric radical process and CO insertion process. The in situ generated CoIII coordinated with the aminoquinoline amide and generated intermediate 76A via C−H bond cobaltation. The esteric radical, generated from DIAD, attacked 76-A to form CoIV intermediate 76-B. Subsequent reductive elimination of 76-B followed by an AA

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Finally, reductive elimination produced the coumarin product. The Cp*CoI species was oxidized to regenerate the active CoIII catalyst for the next catalytic cycle (Scheme 79b).

Scheme 78. Copper-Catalyzed C−H Bond Carbonylation with AIBN

6.2. Oxidative Carbonylation of C(sp3)−H

Compared to sp2-hybridized C−H bonds, sp3 C−H bonds are more difficult to functionalize. The groups of Yu341 and Chatani342 independently reported 8-aminoquinoline-directed C(sp3)−H carbonylation of aliphatic amides with palladium and ruthenium, respectively. Cobalt also showed excellent activity in the functionalization of C(sp3)−H bonds. The group of Ge reported the cobalt-catalyzed site-selective intra- and intermolecular amidation reactions of aliphatic amides via a C(sp3)−H bond functionalization.343 Zhang and co-workers developed a cobalt-catalyzed alkynylation/cyclization process for the synthesis of pyrrolidinones from aliphatic amides and terminal alkynes.344 Recently, three different groups (Sundararaju,345 Gaunt,346 and Lei347) have independently developed cobalt-catalyzed intramolecular oxidative C(sp3)−H/N−H carbonylations of aliphatic amides (Scheme 80a). With Co(acac)2 as the catalyst and Ag2CO3 as the oxidant, the sp3 C−H bonds of the α-methyl groups of the propanamides was selectively activated and carbonylated to give succinimides in good to high yields. The reactions proceeded under mild conditions, only atmospheric pressure CO was required, and a range of functional groups were tolerated. Both α-disubstituted and α-monosubstituted propanamides were suitable substrates for this methodology, and the corresponding products were obtained in good yields. Interestingly, the reaction was inert to the α-hydrogen, and simple propanamide was also tolerated. However, other linear aliphatic amides such as n-butanamide, nhexanamide, and 2-ethylbutanamide turned out to be poor substrates and failed to give the desired products. It should be noted that α-phenyl-substituted substrates favored C(sp3)−H activation rather than C(sp2)−H activation, which might result from the greater stability of five-membered cyclocobalt species over six-membered metallocycles. Mechanistic studies showed that the reaction proceeded via a single-electron-transfer process while the cleavage of the C−H bond was the ratelimiting step. The groups of Sundararaju and Lei proposed different reaction mechanisms. According to Lei’s proposed mechanism, initially, coordination of CoII to the substrate gave a two-coordinate CoII species 85-A, which was then oxidized to high-valent CoIII complex 85-B. Then, activation of the C−H bond by CoIII produced cyclic intermediate 85-C. Coordination and insertion of CO generated the cyclic acyl CoIII intermediate 85-D, which produced the product and generated CoI after reductive elimination. Finally, the CoI was oxidized by Ag2CO3 to regenerate CoII for the next catalytic cycle (Scheme 80b). However, Sundararaju and co-workers suggested that a CoIV species might be involved in the reductive elimination step instead of CoIII. Moreover, the 8-aminoquinoline auxiliaries could be removed by treatment with TFA/HCl to give dicarboxylic acids in high yields (Scheme 80c). Directing-group-assisted carbonylative C−H activation has been widely investigated because of its high regioselectivity. The direct carbonylation of unactivated C(sp3)−H bonds is, however, even more challenging. In 2016, Wu and co-workers developed a copper-catalyzed carbonylation reaction of cycloalkanes with amides as coupling partners (Scheme 81a).348 The use of amides as weak nucleophiles turns out to be important since the amides were comparatively stable toward the oxidants. A series of imides were synthesized in good yields by C−H

synthesis of oxime carbonates has been described by Guan and co-workers.337 Coumarin is an important structure and exists in many biologically and pharmacologically active compounds. In 2015, Wang and co-workers developed a Cp*Co(III)-catalyzed carbonylative annulation of 2-alkenylphenols for the synthesis of coumarin derivatives (Scheme 79a).338 Compared to the Scheme 79. Cp*Co(III)-Catalyzed Carbonylative Annulation of 2-Alkenylphenols

similar transformation catalyzed by Pd339 and Rh,340 this methodology proceeded under remarkably mild conditions. Studies of the kinetic isotope effect (KIE) indicated that the C−H bond cleavage is the rate-determining step. The authors suggested that Cp*Co(III)Ln was the active catalytic species, which was generated by the oxidation of the precatalyst. Ligand exchange of CoIII with the 2-vinylphenol followed by concerted cobaltation and deprotonation gave the cyclic cobalt complex 84-B. Subsequently, coordination and migratory insertion of carbon monoxide generated the acyl cobalt intermediate 84-C. AB

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 80. Cobalt-Catalyzed Intramolecular Oxidative C(sp3)−H/N−H Carbonylation

carbonylation occurring selectively at the tertiary C−H bond. However, a mixture of different products was obtained when an open-chain alkane was used in this reaction. Mechanistically, the addition of TEMPO decreased the yields significantly, which indicated that the reaction might proceed via a radical process. Homolytic cleavage of the peroxide generated a tertbutoxy radical, which reacted with the cyclohexane to give cyclohexyl radical 86-A. Then, oxidative addition of CuI produced the CuIII intermediate 86-B. Ligand exchange with amide generated the intermediate 86-C. Subsequently, CO insertion and reductive elimination produced the product and regenerated CuI for the next catalytic cycle (Scheme 81b). Subsequently, the authors expanded the copper-catalyzed C(sp3)−H bond carbonylation to the aminocarbonylation349 and alkoxycarbonylation350 of alkanes with amines and alcohols as nucleophiles, respectively (Scheme 82). Using copper(II)

Scheme 81. Copper-Catalyzed Carbonylation of Cycloalkanes with Amides

Scheme 82. Copper-Catalyzed Aminocarbonylation and Alkoxycarbonylation of Cycloalkanes

carbonylation of cycloalkanes. Interestingly, adamantane was also tolerated and the desired product was obtained with the AC

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

ACKNOWLEDGMENTS The authors are thankful for financial support from the NSFC (21472174, 21772177) and the Zhejiang Natural Science Fund for Distinguished Young Scholars (LR16B020002). The authors also appreciate the general support provided by Prof. Matthias Beller and Prof. Armin Börner at LIKAT. The authors also wish to take this chance to thank the reviewers for their valuable suggestions and Dr. Rian Dewhurst for the language editing, which improved the quality of the manuscript significantly.

1,10-phenanthroline as the catalytic system, the desired aliphatic amides and esters were obtained in moderate to good yields. Reactions with open-chain alkanes produced a mixture of regioisomers. More recently, a copper-catalyzed four-component carbonylative reaction was also developed by the same group.351 By using copper as the catalyst, ethene and other aliphatic alkenes were carbonylated with alcohols and acetonitrile under CO pressure and provided moderate to good yields.

7. CONCLUSION In conclusion, with a focus on mechanistic discussions, the main achievements in first-row transition-metal-catalyzed (basemetal-catalyzed) carbonylative transformations have been summarized and discussed. In general, base metal catalysts have shown unique reactivity and selectivity in the carbonylative transformation of alphatic substrates; for example, the catalytic double carbonylation of benzyl halides and the carbonylation of epoxides have been performed on large scales. However, their activity in the carbonylation of aryl haldies and unsaturated substrates (alkynes, alkenes, imines, etc.) is still not comparable with palladium or rhodium catalysts, and more efforts are needed in this area. With the motto “the first-row transition metal catalysts are not only cheaper than noble metal catalysts”, we hope this review can attract more attention to the field and contribute to the further development of this topic.

REFERENCES (1) Beller, M. Catalytic Carbonylation Reactions; Springer: Berlin, 2006. (2) Kiss, G. Palladium-Catalyzed Reppe Carbonylation. Chem. Rev. 2001, 101, 3435−3456. (3) Bertoux, F.; Monflier, E.; Castanet, Y.; Mortreux, A. Advances in Transition-Metal Catalyzed Hydroxycarbonylation Reactions in Aqueous-Organic Two-Phase System. J. Mol. Catal. A: Chem. 1999, 143, 11−22. (4) Börner, A.; Franke, R. Hydroformylation: Fundamentals, Processes, and Applications in Organic Synthesis; Wiley-VCH: Weinheim, 2016. (5) Beller, M.; Wu, X.-F. Transition Metal Catalyzed Carbonylation Reactions: Carbonylative Activation of C-X Bonds; Springer: Amsterdam, 2013. (6) Kollär , L. Modern Carbonylation Methods; Wiley-VCH: Weinheim, 2008. (7) Barnard, C. F. J. Palladium-Catalyzed Carbonylation-A Reaction Come of Age. Organometallics 2008, 27, 5402−5422. (8) Wu, X.-F.; Neumann, H.; Beller, M. Palladium-Catalyzed Carbonylative Coupling Reactions between Ar-X and Carbon Nucleophiles. Chem. Soc. Rev. 2011, 40, 4986−5009. (9) Li, Y.; Hu, Y.; Wu, X.-F. Non-Noble Metal-Catalysed Carbonylative Transformations. Chem. Soc. Rev. 2018, 47, 172−194. (10) Liu, Q.; Zhang, H.; Lei, A. Oxidative Carbonylation Reactions: Organometallic Compounds (R-M) or Hydrocarbons (R-H) as Nucleophiles. Angew. Chem., Int. Ed. 2011, 50, 10788−10799. (11) Wu, X.-F.; Neumann, H.; Beller, M. Synthesis of Heterocycles via Palladium-Catalyzed Carbonylations. Chem. Rev. 2013, 113, 1−35. (12) Gabriele, B.; Mancuso, R.; Salerno, G. Oxidative Carbonylation as a Powerful Tool for the Direct Synthesis of Carbonylated Heterocycles. Eur. J. Org. Chem. 2012, 2012, 6825−6839. (13) Wu, X.-F.; Fang, X.; Wu, L.; Jackstell, R.; Neumann, H.; Beller, M. Transition-Metal-Catalyzed Carbonylation Reactions of Olefins and Alkynes: A Personal Account. Acc. Chem. Res. 2014, 47, 1041− 1053. (14) Wu, X.-F. Palladium-Catalyzed Carbonylative Transformation of Aryl Chlorides and Aryl Tosylates. RSC Adv. 2016, 6, 83831−83837. (15) Peng, J.-B.; Qi, X.; Wu, X.-F. Visible Light-Induced Carbonylation Reactions with Organic Dyes as the Photosensitizers. ChemSusChem 2016, 9, 2279−2282. (16) Gehrtz, P. H.; Hirschbeck, V.; Ciszek, B.; Fleischer, I. Carbonylations of Alkenes in the Total Synthesis of Natural Compounds. Synthesis 2016, 48, 1573−1596. (17) Bai, Y.; Davis, D. C.; Dai, M. Natural Product Synthesis via Palladium-Catalyzed Carbonylation. J. Org. Chem. 2017, 82, 2319− 2328. (18) Dekleva, T. W.; Forster, D. The Rhodium-Catalyzed Carbonylation of Linear Primary Alcohols. J. Am. Chem. Soc. 1985, 107, 3565−3567. (19) Haynes, A.; Maitlis, P. M.; Morris, G. E.; Sunley, G. J.; Adams, H.; Badger, P. W.; Bowers, C. M.; Cook, D. B.; Elliott, P. I. P.; Ghaffar, T.; Green, H.; Griffin, T. R.; Payne, M.; Pearson, J. M.; Taylor, M. J.; Vickers, P. W.; Watt, R. J. Promotion of Iridium-Catalyzed Methanol Carbonylation: Mechanistic Studies of the Cativa Process. J. Am. Chem. Soc. 2004, 126, 2847−2861.

AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. ORCID

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

The authors declare no competing financial interest. Biographies Jin-Bao Peng was born in 1986 in Gansu, PRC. He obtained his B.S. degree in 2009 from Lanzhou University. In 2014, he received his Ph.D. degree from Lanzhou University under the supervision of Prof. Yong-Qiang Tu. After postdoctoral research with Prof. Suning Wang at Queen’s University, Canada, he joined the Department of Chemistry, Zhejiang Sci-Tech University. His current research interests focus on chemo- and regioselective carbonylation reactions and carbonylation-reaction-based cascade reactions for the synthesis of heterocycles. Fu-Peng Wu was born in PRC. He studied at Hangzhou Normal University, where he obtained his B.S. degree in 2016. Currently he is a master’s degree student working on metal-catalyzed carbonylation at Zhejiang Sci-Tech University. Xiao-Feng Wu was born in China. He studied chemistry at Zhejiang Sci-Tech University, where he obtained his B.S. degree in 2007. In the same year, he moved to University of Rennes 1 (France) and earned his master’s degree in 2009. Then he joined Matthias Beller’s group at the Leibniz Institute for Catalysis (Germany), where he completed his Ph.D. defense in January 2012. Subsequently he began his independent research at LIKAT and ZSTU, where he was promoted to professor in 2013. In March 2017, Xiao-Feng Wu defended his habilitation successfully from University of Rennes 1 (France). XiaoFeng Wu has authored >220 publications in international journals, and is also the editor or author of 10 books. AD

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(20) Peng, J.-B.; Wu, X.-F. Ligand- and Solvent-Controlled Regioand Chemodivergent Carbonylative Reactions. Angew. Chem., Int. Ed. 2018, 57, 1152−1160. (21) Morimoto, T.; Kakiuchi, K. Evolution of Carbonylation Catalysis: No Need for Carbon Monoxide. Angew. Chem., Int. Ed. 2004, 43, 5580−5588. (22) Wu, L.; Liu, Q.; Jackstell, R.; Beller, M. Carbonylations of Alkenes with CO Surrogates. Angew. Chem., Int. Ed. 2014, 53, 6310− 6320. (23) Gautam, P.; Bhanage, B. M. Recent Advances in the Transition Metal Catalyzed Carbonylation of Alkynes, Arenes And Aryl Halides Using CO Surrogates. Catal. Sci. Technol. 2015, 5, 4663−4702. (24) Friis, S. D.; Lindhardt, A. T.; Skrydstrup, T. The Development and Application of Two-Chamber Reactors and Carbon Monoxide Precursors for Safe Carbonylation Reactions. Acc. Chem. Res. 2016, 49, 594−605. (25) Peng, J.-B.; Qi, X.; Wu, X.-F. Recent Achievements in Carbonylation Reactions: A Personal Account. Synlett 2017, 28, 175−194. (26) Wu, X.-F. Non-Redox-Metal-Catalyzed Redox Reactions: Zinc Catalysts. Chem. - Asian J. 2012, 7, 2502−2509. (27) Wu, X.-F.; Neumann, H. Zinc-Catalyzed Organic Synthesis: CC, C-N, C-O Bond Formation Reactions. Adv. Synth. Catal. 2012, 354, 3141−3160. (28) Kirillova, M. V.; da Silva, J. A. L.; da Silva, J. J. R. F.; Palavra, A. F.; Pombeiro, A. J. L. Highly Efficient Direct Carboxylation of Propane into Butyric Acids Catalyzed by Vanadium Complexes. Adv. Synth. Catal. 2007, 349, 1765−1774. (29) Bolm, C.; Legros, J.; Le Paih, J.; Zani, L. Iron-Catalyzed Reactions in Organic Synthesis. Chem. Rev. 2004, 104, 6217−6254. (30) Enthaler, S.; Junge, K.; Beller, M. Sustainable Metal Catalysis with Iron: From Rust to a Rising Star? Angew. Chem., Int. Ed. 2008, 47, 3317−3321. (31) Correa, A.; Garcia Mancheno, O.; Bolm, C. Iron-Catalysed Carbon-Heteroatom and Heteroatom-Heteroatom Bond Forming Processes. Chem. Soc. Rev. 2008, 37, 1108−1117. (32) Sherry, B. D.; Fürstner, A. The Promise and Challenge of IronCatalyzed Cross Coupling. Acc. Chem. Res. 2008, 41, 1500−1511. (33) Iron Catalysis in Organic Chemistry: Reactions and Applications; Plietker, B., Ed.; Wiley: New York, 2008. (34) Iron Catalysis: Fundamentals and Applications; Plietker, B., Ed.; Springer-Verlag: Berlin, 2011. (35) Riener, K.; Haslinger, S.; Raba, A.; Högerl, M. P.; Cokoja, M.; Herrmann, W. A.; Kühn, F. E. Chemistry of Iron N-Heterocyclic Carbene Complexes: Syntheses, Structures, Reactivities, and Catalytic Applications. Chem. Rev. 2014, 114, 5215−5272. (36) Bauer, I.; Knölker, H.-J. Iron Catalysis in Organic Synthesis. Chem. Rev. 2015, 115, 3170−3387. (37) Chopade, P. R.; Louie, J. [2 + 2 + 2] Cycloaddition Reactions Catalyzed by Transition Metal Complexes. Adv. Synth. Catal. 2006, 348, 2307−2327. (38) Domínguez, G.; Pérez-Castells, J. Recent Advances in [2 + 2 + 2] Cycloaddition Reactions. Chem. Soc. Rev. 2011, 40, 3430−3444. (39) Khand, I. U.; Knox, G. R.; Pauson, P. L.; Watts, W. E.; Foreman, M. I. Organocobalt Complexes. Part II. Reaction of Acetylenehexacarbonyldicobalt Complexes, (R1C2R2)Co2(CO)6, With Norbornene and Its Derivatives. J. Chem. Soc., Perkin Trans. 1 1973, 1, 977−981. (40) Gandeepan, P.; Cheng, C.-H. Cobalt Catalysis Involving π Components in Organic Synthesis. Acc. Chem. Res. 2015, 48, 1194− 1206. (41) Cahiez, G.; Moyeux, A. Cobalt-Catalyzed Cross-Coupling Reactions. Chem. Rev. 2010, 110, 1435−1462. (42) Kommagalla, Y.; Chatani, N. Cobalt(II)-Catalyzed C-H Functionalization Using an N,N’-Bidentate Directing Group. Coord. Chem. Rev. 2017, 350, 117−135. (43) Montgomery, J. Nickel-Catalyzed Cyclizations, Couplings, and Cycloadditions Involving Three Reactive Components. Acc. Chem. Res. 2000, 33, 467−473.

(44) Standley, E. A.; Tasker, S. Z.; Jensen, K. L.; Jamison, T. F. Nickel Catalysis: Synergy between Method Development and Total Synthesis. Acc. Chem. Res. 2015, 48, 1503−1514. (45) Kurahashi, T.; Matsubara, S. Nickel-Catalyzed Reactions Directed Toward the Formation of Heterocycles. Acc. Chem. Res. 2015, 48, 1703−1716. (46) Thakur, A.; Louie, J. Advances in Nickel-Catalyzed Cycloaddition Reactions To Construct Carbocycles and Heterocycles. Acc. Chem. Res. 2015, 48, 2354−2365. (47) Rosen, B. M.; Quasdorf, K. W.; Wilson, D. A.; Zhang, N.; Resmerita, A.-M.; Garg, N. K.; Percec, V. Nickel-Catalyzed CrossCouplings Involving Carbon-Oxygen Bonds. Chem. Rev. 2011, 111, 1346−1416. (48) Yamaguchi, J.; Muto, K.; Itami, K. Recent Progress in NickelCatalyzed Biaryl Coupling. Eur. J. Org. Chem. 2013, 2013, 19−30. (49) Han, F.-S. Transition-Metal-Catalyzed Suzuki−Miyaura CrossCoupling Reactions: A Remarkable Advance from Palladium to Nickel Catalysts. Chem. Soc. Rev. 2013, 42, 5270−5298. (50) Tobisu, M.; Chatani, N. Cross-Couplings Using Aryl Ethers via C−O Bond Activation Enabled by Nickel Catalysts. Acc. Chem. Res. 2015, 48, 1717−1726. (51) Jackson, E. P.; Malik, H. A.; Sormunen, G. J.; Baxter, R. D.; Liu, P.; Wang, H.; Shareef, A.-R.; Montgomery, J. Mechanistic Basis for Regioselection and Regiodivergence in Nickel- Catalyzed Reductive Couplings. Acc. Chem. Res. 2015, 48, 1736−1745. (52) Tollefson, E. J.; Hanna, L. E.; Jarvo, E. R. Stereospecific NickelCatalyzed Cross-Coupling Reactions of Benzylic Ethers and Esters. Acc. Chem. Res. 2015, 48, 2344−2353. (53) Zhang, C.; Tang, C.; Jiao, N. Recent Advances in CopperCatalyzed Dehydrogenative Functionalization via a Single Electron Transfer (SET) Process. Chem. Soc. Rev. 2012, 41, 3464−3484. (54) Allen, S. E.; Walvoord, R. R.; Padilla-Salinas, R.; Kozlowski, M. C. Aerobic Copper-Catalyzed Organic Reactions. Chem. Rev. 2013, 113, 6234−6458. (55) Guo, X.-X.; Gu, D.-W.; Wu, Z.; Zhang, W. Copper-Catalyzed C−H Functionalization Reactions: Efficient Synthesis of Heterocycles. Chem. Rev. 2015, 115, 1622−1651. (56) Dekleva, T. W.; Forster, D. Mechanistic Aspects of TransitionMetal-Catalyzed Alcohol Carbonylations. Adv. Catal. 1986, 34, 81− 130. (57) Howard, M. J.; Jones, M. D.; Roberts, M. S.; Taylor, S. A. C1 to Acetyls: Catalysis and Process. Catal. Catal. Today 1993, 18, 325−354. (58) Maitlis, P. M.; Haynes, A.; Sunley, G. J.; Howard, M. J. Methanol Carbonylation Revisited: Thirty Years on. J. Chem. Soc., Dalton Trans. 1996, 2187−2196. (59) Thomas, C. M.; Süss-Fink, G. Ligand Effects in the RhodiumCatalyzed Carbonylation Of Methanol. Coord. Chem. Rev. 2003, 243, 125−142. (60) Torrence, P. Carbonylations: Synthesis of Acetic Acid and Acetic Acid Anhydride from Methanol. In Applied Homogeneous Catalysis with Organometallic Compounds, 2nd ed.; Cornils, B., Herrmann, W. A., Eds.; Wiley: Weinheim, 2002; Vol. 1, p 104. (61) Kalvet, I.; Magnin, G.; Schoenebeck, F. Rapid RoomTemperature, Chemoselective Csp2− Csp2 Coupling of Poly(Pseudo) Halogenated Arenes Enabled by Palladium(I) Catalysis in Air. Angew. Chem., Int. Ed. 2017, 56, 1581−1585. (62) Mesganaw, T.; Garg, N. K. Ni- and Fe-Catalyzed CrossCoupling Reactions of Phenol Derivatives. Org. Process Res. Dev. 2013, 17, 29−39. (63) So, C. M.; Lau, C. P.; Chan, A. S. C.; Kwong, F. Y. Suzuki− Miyaura Coupling of Aryl Tosylates Catalyzed by an Array of Indolyl Phosphine−Palladium Catalysts. J. Org. Chem. 2008, 73, 7731−7734. (64) Cassar, L.; Foa, M. Nickel-Catalyzed Carbonylation of Aromatic Halides at Atmospheric Pressure of Carbon Monoxide. J. Organomet. Chem. 1973, 51, 381−393. (65) Corey, E. J.; Hegedus, L. S. Base-Catalyzed Carboxylation of Organic Halides by Nickel Carbonyl in Protic Media. J. Am. Chem. Soc. 1969, 91, 1233−1234. AE

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(66) Hashimoto, I.; Tsuruta, N.; Ryang, M.; Tsutsumi, S. Reaction of Potassium Hexacyanodinickelate(I) with Organic Halides. J. Org. Chem. 1970, 35, 3748−3752. (67) del Rosario, R.; Stuhl, L. S. Concerning the Mechanism of A Cyanometallate Mediated Carbonylation Reaction. Tetrahedron Lett. 1982, 23, 3999−4002. (68) Joo, F.; Alper, H. Cyanonickel(II) Complexes as Catalysts for the Phase-Transfer-Catalyzed Carbonylation of Allyl halides. The Isolation of the Catalytically Active Species [Ni(CO)3CN−]. Organometallics 1985, 4, 1775−1778. (69) Amer, I.; Alper, H. Nickel Cyanide and Phase-TransferCatalyzed Carbonylation of Aryl Iodides in The Absence of Light. J. Org. Chem. 1988, 53, 5147−5148. (70) Amer, I.; Alper, H. Lanthanide-Promoted and Nickel CyanideCatalyzed Carbonylation Reactions under Phase-Transfer Conditions. J. Am. Chem. Soc. 1989, 111, 927−930. (71) Alper, H.; Amer, I.; Vasapollo, G. Stereospecific Nickel and Phase Transfer Catalyzed Carbonylation of Vinyl Bromides and Chlorides. Tetrahedron Lett. 1989, 30, 2615−2616. (72) Iranpoor, N.; Firouzabadi, H.; Etemadi-Davan, E.; Nematollahi, A.; Firouzi, H. R. A Novel Nickel-Catalyzed Synthesis of Thioesters, Esters and Amides from Aryl Iodides in the Presence of Chromium Hexacarbonyl. New J. Chem. 2015, 39, 6445−6452. (73) Fang, X.; Jackstell, R.; Beller, M. Selective Palladium-Catalyzed Aminocarbonylation of Olefins with Aromatic Amines and Nitroarenes. Angew. Chem., Int. Ed. 2013, 52, 14089−14093. (74) Cheung, C. W.; Ploeger, M. L.; Hu, X. Amide Synthesis via Nickel-Catalysed Reductive Aminocarbonylation of Aryl Halides with Nitroarenes. Chem. Sci. 2018, 9, 655−659. (75) Tjutrins, J.; Shao, J. L.; Yempally, V.; Bengali, A. A.; Arndtsen, B. A. A. Nickel-Based, Tandem Catalytic Approach to Isoindolinones from Imines, Aryl Iodides, and CO. Organometallics 2015, 34, 1802− 1805. (76) Prasad, P. K.; Sudalai, A. Copper(I) Bromide-Catalyzed Carbonylative Coupling of Aryl Halides with Phenols, Alcohols and Amines Using Sodium Cyanide as C1 Source: A Synthesis of Carboxylic Acid Derivatives. Adv. Synth. Catal. 2014, 356 (10), 2231− 2238. (77) Murahashi, S. Synthesis of Phthalimidines from Schiff Bases and Carbon Monoxide. J. Am. Chem. Soc. 1955, 77, 6403−6404. (78) De Risi, C.; Pollini, G. P.; Zanirato, V. Recent Developments in General Methodologies for the Synthesis of α-Ketoamides. Chem. Rev. 2016, 116, 3241−3305. (79) Uozumi, Y.; Arii, T.; Watanabe, T. Double Carbonylation of Aryl Iodides with Primary Amines under Atmospheric Pressure Conditions Using the Pd/PPh3/DABCO/THF System. J. Org. Chem. 2001, 66, 5272−5274. (80) Foa, M.; Francalanci, F. Recent Developments in CobaltCatalyzed Carbonylation. J. Mol. Catal. 1987, 41, 89−107. (81) Francalanci, F.; Bencini, E.; Gardano, A.; Vincenti, M.; Foa, M. Cobalt-Catalyzed Low Pressure Double Carbonylation of Aryl and Secondary Benzyl Halides. J. Organomet. Chem. 1986, 301, C27−C30. (82) Francalanci, F.; Gardano, A.; Abis, L.; Foa, M. Formation and Reactivity of the New Complex (CNCH2Co(COOCH3)(CO)3)−. J. Organomet. Chem. 1983, 251, C5−C8. (83) Kashimura, T.; Kudo, K.; Mori, S.; Sugita, N. Cobalt CarbonylCatalyzed Double-Carbonylation of O-Halogenated Benzoic Acids under Photostimulation. Chem. Lett. 1986, 15, 483−486. (84) Li, G.-X.; Li, L.; Huang, H.-M.; Cai, H.-Q. One Pot Formation of Catalyst and Double Carbonylation of Benzyl Chloride. J. Mol. Catal. A: Chem. 2003, 193, 97−102. (85) Miura, M.; Itoh, K.; Nomura, M. Cobalt(II) Chloride Catalyzed Normal Pressure Carbonylation of Aryl Halides. J. Mol. Catal. 1988, 48, 11−13. (86) Miura, M.; Okuro, K.; Hattori, A.; Nomura, M. Carbonylation of Vinyl Halides with Carbonylcobalt. J. Chem. Soc., Perkin Trans. 1 1989, 73−76.

(87) Monflier, E.; Mortreux, A.; Petit, F. A Convenient Synthesis of Benzylpyruvic Acid: the Double Carbonylation of Phenethyl Bromide. Appl. Catal., A 1993, 102, 53−67. (88) Suisse, P.; Pellegrini, S.; Castanet, Y.; Mortreux, A.; Lecolier, S. Efficient Dimethyl Malonate Synthesis by Methoxycarbonylation of Dichloromethane Catalysed by Electrogenerated (Co(CO)3PBu3)− Species. J. Chem. Soc., Chem. Commun. 1995, 847−848. (89) Younis, K.; Amer, I. Nickel-Catalyzed Double Carbonylation of Halodienes: A Possible New Mechanism for the Double Carbonylation Reaction. Organometallics 1994, 13, 3120−3126. (90) Liu, J.; Zhang, R.; Wang, S.; Sun, W.; Xia, C. A General and Efficient Copper Catalyst for the Double Carbonylation Reaction. Org. Lett. 2009, 11, 1321−1324. (91) Goure, W. F.; Wright, M. E.; Davis, P. D.; Labadie, S. S.; Stille, J. K. Palladium-Catalyzed Cross-Coupling of Vinyl Iodides with Organostannanes: Synthesis of Unsymmetrical Divinyl Ketones. J. Am. Chem. Soc. 1984, 106, 6417−6422. (92) Crisp, G. T.; Scott, W. J.; Stille, J. K. Palladium-Catalyzed Carbonylative Coupling of Vinyl Triflates with Organostannanes. A Total Synthesis of (±)Δ9(12)-Capnellene. J. Am. Chem. Soc. 1984, 106, 7500−7506. (93) Kang, S.-K.; Yamaguchi, T.; Kim, T.-H.; Ho, P.-S. CopperCatalyzed Cross-Coupling and Carbonylative Cross-Coupling of Organostannanes and Organoboranes with Hypervalent Iodine Compounds. J. Org. Chem. 1996, 61, 9082−9083. (94) Kang, S. K.; Ryu, H. C.; Lee, S. W. Ni(acac)2-Catalyzed CrossCoupling and Carbonylative Cross-Coupling of Organostannanes with Hypervalent Iodonium Salts. J. Chem. Soc., Perkin Trans. 1 1999, 19, 2661−2663. (95) Kang, S.-K.; Kim, W.-Y.; Lee, Y.-T.; Ahn, S.-K.; Kim, J.-C. Manganese Chloride-Catalyzed Cross-Coupling and Carbonylative Cross-Coupling of Organostannanes with Iodonium Salts. Tetrahedron Lett. 1998, 39, 2131−2132. (96) Iranpoor, N.; Firouzabadi, H.; Etemadi-Davan, E. PhosphineFree NiBr2-Catalyzed Synthesis of Unsymmetrical Diaryl Ketones via Carbonylative Cross-Coupling of Aryl Iodides with Ph3SnX (X= Cl, OEt). J. Organomet. Chem. 2015, 794, 282−287. (97) Zhong, Y.; Han, W. Iron-Catalyzed Carbonylative Suzuki Reactions under Atmospheric Pressure of Carbon Monoxide. Chem. Commun. 2014, 50, 3874−3877. (98) Grushin, V. V.; Alper, H. Novel Palladium-Catalyzed Carbonylation of Organic Halides by Chloroform and Alkali. Organometallics 1993, 12, 3846−3850. (99) Liu, X. L.; Li, B.; Gu, Z. H. Palladium-Catalyzed Heck-type Domino Cyclization and Carboxylation to Synthesize Carboxylic Acids by Utilizing Chloroform as the Carbon Monoxide Source. J. Org. Chem. 2015, 80, 7547−7554. (100) Liu, X.; Gu, Z. Pd-Catalyzed Heck Cyclization and in situ Hydrocarboxylation or Hydromethenylation via a Hydrogen Borrowing Strategy. Org. Chem. Front. 2015, 2, 778−782. (101) Gockel, S. N.; Hull, K. L. Chloroform as A Carbon Monoxide Precursor: In or Ex Situ Generation of CO for Pd-Catalyzed Aminocarbonylations. Org. Lett. 2015, 17, 3236−3239. (102) Li, Z. Y.; Wang, L. Palladium-Catalyzed Aminocarbonylation Reaction to Access 1,3,4-Oxadiazoles Using Chloroform as The Carbon Monoxide Source. Adv. Synth. Catal. 2015, 357, 3469−3473. (103) Sun, G.; Lei, M.; Hu, L. A. Facile and Efficient Method for the Synthesis of Alkynone by Carbonylative Sonogashira Coupling Using CHCl3 as the CO Source. RSC Adv. 2016, 6, 28442−28446. (104) Zhao, H. Y.; Du, H. G.; Yuan, X. R.; Wang, T. J.; Han, W. IronCatalyzed Carbonylation of Aryl Halides with Arylborons Using Stoichiometric Chloroform as the Carbon Monoxide Source. Green Chem. 2016, 18, 5782−5787. (105) Cheng, L.; Zhong, Y.; Ni, Z.; Du, H.; Jin, F.; Rong, Q.; Han, W. Copper-Catalyzed Carbonylative Suzuki Coupling of Aryl Iodides with Arylboronic Acids under Ambient Pressure of Carbon Monoxide. RSC Adv. 2014, 4, 44312−44316. (106) Zhong, Y.; Gong, X.; Zhu, X.; Ni, Z.; Wang, H.; Fu, J.; Han, W. In situ Generated Nickel Nanoparticle-Catalyzed Carbonylative Suzuki AF

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Reactions of Aryl Iodides with Arylboronic Acids at Ambient CO Pressure in Poly(Ethylene Glycol). RSC Adv. 2014, 4, 63216−63220. (107) Wang, Q.; Chen, C. Nickel-Catalyzed Carbonylative Negishi Cross-Coupling Reactions. Tetrahedron Lett. 2008, 49, 2916−2921. (108) Tambade, P. J.; Patil, Y. P.; Nandurkar, N. S.; Bhanage, B. M. Copper-Catalyzed, Palladium-Free Carbonylative Sonogashira Coupling Reaction of Aliphatic and Aromatic Alkynes with Iodoaryls. Synlett 2008, 2008, 886−888. (109) Seyferth, D.; Spohn, R. J. Novel, Mercurial-Based Symmetrical Ketone Synthesis. J. Am. Chem. Soc. 1968, 90, 540−541. (110) Seyferth, D.; Spohn, R. J. Reaction of Organomercuric Halides with Dicobalt Octacarbonyl. A New Ketone Synthesis. J. Am. Chem. Soc. 1969, 91, 3037−3044. (111) Hirota, Y.; Ryang, M.; Tsutsumi, S. The Reaction of Organomercuric Compounds with Nickel Carbonyl. Tetrahedron Lett. 1971, 12 (19), 1531−1534. (112) Devasagayaraj, A.; Knochel, P. Preparation of Polyfunctional Ketones by A Cobalt (II) Mediated Carbonylation of Organozinc Reagents. Tetrahedron Lett. 1995, 36, 8411−8414. (113) Li, Y.; Tu, D. H.; Wang, B.; Lu, J. Y.; Wang, Y. Y.; Liu, Z. T.; Liu, Z. W.; Lu, J. Nickel-Catalyzed Carbonylation of Arylboronic Acids with DMF as a CO Source. Org. Chem. Front. 2017, 4, 569−572. (114) Enquist, P.-A.; Nilsson, P.; Larhed, M. Ultrafast Chemistry: Cobalt Carbonyl-Mediated Synthesis of Diaryl Ketones under Microwave Irradiation. Org. Lett. 2003, 5, 4875−4878. (115) Brunet, J.-J.; Taillefer, M. Bimetallic Catalysis. A New Method for Carbonylation of Aryl Iodides under Mild Conditions. J. Organomet. Chem. 1989, 361, C1−C4. (116) Brunet, J.-J.; Taillefer, M. Bimetallic Catalysis: Synthesis of Benzophenones Through Carbonylation of Aryl Iodides Catalysed By Fe(CO)5-Co2(CO)8 Systems. J. Organomet. Chem. 1990, 384, 193− 197. (117) Brunet, J.-J.; El Zaizi, A. Synthesis of Benzophenone by Carbonylation of Iodobenzene with [Bu4nN][HFe(CO)4]. J. Organomet. Chem. 1995, 486, 275−277. (118) Rérat, A.; Michon, C.; Agbossou-Niedercorn, F.; Gosmini, C. Synthesis of Symmetrical Diaryl Ketones by Cobalt-Catalyzed Reaction of Arylzinc Reagents with Ethyl Chloroformate. Eur. J. Org. Chem. 2016, 2016, 4554−4560. (119) Yan, M.; Kawamata, Y.; Baran, P. S. Synthetic Organic Electrochemical Methods since 2000: On the Verge of a Renaissance. Chem. Rev. 2017, 117, 13230−13319. (120) Nédélec, J.-Y.; Périchon, J.; Troupel, M. Organic Electroreductive Coupling Reactions Using Transition Metal Complexes as Catalysts. Top. Curr. Chem. 1997, 185, 141−173. (121) Duñach, E.; Franco, D.; Olivero, S. Carbon-Carbon Bond Formation with Electrogenerated Nickel and Palladium Complexes. Eur. J. Org. Chem. 2003, 2003, 1605−1622. (122) Oçafrain, M.; Devaud, M.; Troupel, M.; Perichon, J. New Electrochemical Synthesis of Ketones from Organic Halides and Carbon Monoxide. J. Chem. Soc., Chem. Commun. 1995, 2331−2332. (123) Dolhem, E.; Oçafrain, M.; Nédélec, J. Y.; Troupel, M. Nickel Catalyzed Electrosynthesis of Ketones from Organic Halides and Metal Carbonyls. Tetrahedron 1997, 53, 17089−17096. (124) Dolhem, E.; Barhdadi, R.; Folest, J.; Nédélec, J.; Troupel, M. Nickel Catalysed Electrosynthesis of Ketones from Organic Halides and Iron Pentacarbonyl. Part 2: Unsymmetrical Ketones. Tetrahedron 2001, 57, 525−529. (125) Oçafrain, M.; Devaud, M.; Nédélec, J. Y.; Troupel, M. Electrochemical Generation of A Nickel−Carbonyl Complex, Catalyst for the Electroreductive Coupling of Organic Halides and Carbon Monoxide into Ketones. J. Organomet. Chem. 1998, 560, 103−107. (126) Oçafrain, M.; Dolhem, E.; Nédélec, J. Y.; Troupel, M. Nickel− Bipyridine Catalysed Electrosynthesis of Ketones from Organic Halides and Carbon Monoxide: Kinetic and Mechanistic Investigations. J. Organomet. Chem. 1998, 571, 37−42. (127) Wotal, A. C.; Ribson, R. D.; Weix, D. J. Stoichiometric Reactions of Acylnickel (II) Complexes with Electrophiles and the Catalytic Synthesis of Ketones. Organometallics 2014, 33, 5874−5881.

(128) Heck, R. F.; Breslow, D. S. Carboxyalkylation Reactions Catalyzed by Cobalt Carbonylate Ion. J. Am. Chem. Soc. 1963, 85, 2779−2782. (129) She, M.; Xiao, D.; Yin, B.; Yang, Z.; Liu, P.; Li, J.; Shi, Z. An Efficiently Cobalt-Catalyzed Carbonylative Approach to Phenylacetic Acid Derivatives. Tetrahedron 2013, 69, 7264−7268. (130) Imamoto, T.; Kusumoto, T.; Yokoyama, M. The Conversion of Benzyl Alcohols into Phenylacetic Acid Derivatives by Cobalt Carbonyl Catalyzed Carbonylation. Bull. Chem. Soc. Jpn. 1982, 55, 643−644. (131) Urata, H.; Goto, D.; Fuchikami, T. Carbonylation of Alkyl Sulfonates Catalyzed by Cobalt Complexes. Tetrahedron Lett. 1991, 32, 3091−3094. (132) Cooke, M. P., Jr. Facile Conversion of Alkyl Bromides into Aldehydes Using Sodium Tetracarbonylferrate(-II). J. Am. Chem. Soc. 1970, 92, 6080−6082. (133) Siegl, W. O.; Collman, J. P. Acyl and Alkyl Tetracarbonylferrate (0) Complexes and Intermediates in the Synthesis of Aldehydes and Ketones. J. Am. Chem. Soc. 1972, 94, 2516−2518. (134) Collman, J. P.; Winter, S. R.; Clark, D. R. Selective Syntheses of Aliphatic Ketones Using Sodium Tetracarbonylferrate (-II). J. Am. Chem. Soc. 1972, 94, 1788−1789. (135) Collman, J. P.; Winter, S. R.; Komoto, R. G. Selective Syntheses of Aliphatic Carboxylic Acids, Esters, and Amides Using Sodium Tetracarbonylferrate(-II). J. Am. Chem. Soc. 1973, 95, 249− 250. (136) Collman, J. P.; Finke, R. G.; Cawse, J. N.; Brauman, J. I. Oxidative-Addition Reactions of the Disodium Tetracarbonylferrate Supernucleophile. J. Am. Chem. Soc. 1977, 99, 2515−2526. (137) Collman, J. P.; Cawse, J. N.; Brauman, J. I. Role of Ion Pairing in Reactions of Metal Carbonyl Anions. I. Cation-Assisted Alkyl-Acyl Migratory Insertions. J. Am. Chem. Soc. 1972, 94, 5905−5906. (138) Collman, J. P.; Rothrock, R. K.; Finke, R. G.; Rose-Munch, F. Metal Promoted Alkyl Migration in A Bimetallic Complex. J. Am. Chem. Soc. 1977, 99, 7381−7383. (139) Collman, J. P.; Finke, R. G.; Cawse, J. N.; Brauman, J. I. Lewis Acid Catalyzed [RFe(CO)4]-Alkyl Migration Reactions. A Mechanistic Investigation. J. Am. Chem. Soc. 1978, 100, 4766−4772. (140) Collman, J. P. Disodium Tetracarbonylferrate, a Transition Metal Analog of a Grignard Reagent. Acc. Chem. Res. 1975, 8, 342− 347. (141) Collman, J. P.; Hoffman, N. W. Synthesis of Hemifluorinated Ketones Using Disodium Tetracarbonylferrate(II). J. Am. Chem. Soc. 1973, 95, 2689−2691. (142) McMahon, C. M.; Renn, M. S.; Alexanian, E. J. ManganeseCatalyzed Carboacylations of Alkenes with Alkyl Iodides. Org. Lett. 2016, 18, 4148−4150. (143) Li, Y.; Zhu, F.; Wang, Z.; Rabeah, J.; Brückner, A.; Wu, X.-F. Practical and General Manganese-Catalyzed Carbonylative Coupling of Alkyl Iodides with Amides. ChemCatChem 2017, 9, 915−919. (144) Andersen, T. L.; Donslund, A. S.; Neumann, K. T.; Skrydstrup, T. Carbonylative Coupling of Alkyl Zinc Reagents with Benzyl Bromides Catalyzed by a Nickel/NN2 Pincer Ligand Complex. Angew. Chem., Int. Ed. 2018, 57, 800−804. (145) Pye, D. R.; Cheng, L.-J.; Mankad, N. P. Cu/Mn Bimetallic Catalysis Enables Carbonylative Suzuki−Miyaura Coupling with Unactivated Alkyl Electrophiles. Chem. Sci. 2017, 8, 4750−4755. (146) Deutsch, C.; Krause, N.; Lipshutz, B. H. CuH-Catalyzed Reactions. Chem. Rev. 2008, 108, 2916−2927. (147) Suess, A. M.; Lalic, G. Copper-Catalyzed Hydrofunctionalization of Alkynes. Synlett 2016, 27, 1165−1174. (148) Jordan, A. J.; Lalic, G.; Sadighi, J. P. Coinage Metal Hydrides: Synthesis, Characterization, and Reactivity. Chem. Rev. 2016, 116, 8318−8372. (149) Fujihara, T.; Xu, T.; Semba, K.; Terao, J.; Tsuji, Y. CopperCatalyzed Hydrocarboxylation of Alkynes Using Carbon Dioxide and Hydrosilanes. Angew. Chem., Int. Ed. 2011, 50, 523−527. AG

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(150) Jang, W. J.; Lee, W. L.; Moon, J. H.; Lee, J. Y.; Yun, J. CopperCatalyzed Hydrocarboxylation of Alkynes Using Carbon Dioxide and Hydrosilanes. Org. Lett. 2016, 18, 1390−1393. (151) Shi, S.-L.; Buchwald, S. L. Copper-Catalyzed Selective Hydroamination Reactions of Alkynes. Nat. Chem. 2015, 7, 38−44. (152) Uehling, M. R.; Rucker, R. P.; Lalic, G. Catalytic AntiMarkovnikov Hydrobromination of Alkynes. J. Am. Chem. Soc. 2014, 136, 8799−8803. (153) Mailig, M.; Hazra, A.; Armstrong, M. K.; Lalic, G. Catalytic Anti-Markovnikov Hydroallylation of Terminal and Functionalized Internal Alkynes: Synthesis of Skipped Dienes and Trisubstituted Alkenes. J. Am. Chem. Soc. 2017, 139, 6969−6977. (154) Cheng, L.-J.; Mankad, N. P. Cu-Catalyzed Hydrocarbonylative C−C Coupling of Terminal Alkynes with Alkyl Iodides. J. Am. Chem. Soc. 2017, 139, 10200−10203. (155) Cheng, L.-J.; Islam, S. M.; Mankad, N. P. Synthesis of Allylic Alcohols via Cu-Catalyzed Hydrocarbonylative Coupling of Alkynes with Alkyl Halides. J. Am. Chem. Soc. 2018, 140, 1159−1164. (156) Franke, R.; Selent, D.; Börner, A. Applied Hydroformylation. Chem. Rev. 2012, 112, 5675−5732. (157) Roelen, O. Ger. Patent DE 849548, 1938. U.S. Patent 2,327,066, 1943. (158) Pruchnik, F. P. Organometallic Chemistry of Transition Elements; Plenum Press: New York, 1990; p 691. (159) Beller, M.; Cornils, B.; Frohning, C. D.; Kohlpaintner, C. W. Progress in Hydroformylation and Carbonylation. J. Mol. Catal. A: Chem. 1995, 104, 17−85. (160) Frohning, C. D.; Kohlpaintner, C. W. In Applied Homogeneous Catalysis with Organometallic Compounds; Cornils, B., Herrmann, W. A., Eds.; Wiley-VCH: Weinheim, 1996; Vol. 1; pp 29−104. (161) Wender, I.; Sternberg, H. W.; Orchin, M. Evidence for Cobalt Hydrocarbonyl as the Hydroformylation Catalyst. J. Am. Chem. Soc. 1953, 75, 3041−3042. (162) Orchin, M.; Kirch, L.; Goldfarb, I. Evidence for the Presence of Cobalt Hydrocarbonyl under Conditions of the Oxo Reaction. J. Am. Chem. Soc. 1956, 78, 5450−5451. (163) Wender, I.; Metlin, S.; Ergun, S.; Sternberg, H. W.; Greenfield, H. Kinetics and Mechanism of the Hydroformylation Reaction. The Effect of Olefin Structure on Rate. J. Am. Chem. Soc. 1956, 78, 5401− 5405. (164) Kirch, L.; Orchin, M. The Intermediate Cobalt HydrocarbonylOlefin Complex in The Oxo Reaction1. J. Am. Chem. Soc. 1958, 80, 4428−4429. (165) Kirch, L.; Orchin, M. On the Mechanism for the Oxo Reaction. J. Am. Chem. Soc. 1959, 81, 3597−3599. (166) Hebrard, F.; Kalck, P. Cobalt-Catalyzed Hydroformylation of Alkenes: Generation and Recycling of the Carbonyl Species, and Catalytic Cycle. Chem. Rev. 2009, 109, 4272−4282. (167) Reppe, W.; Vetter, H. Carbonylierung VI. Synthesen mit Metallcarbonylwasserstoffen. Liebigs Ann. Chem. 1953, 582, 133−161. (168) Periasamy, M.; Radhakrishnan, U.; Brunet, J. J.; Chauvin, R.; El-Zaizi, A. W. Double Carbonylation of Alkynes Using NaHFe(CO)4. Chem. Commun. 1996, 13, 1499−1500. (169) Periasamy, M.; Rameshkumar, C.; Radhakrishnan, U.; Brunet, J. J. New Convenient One-Pot Methods of Conversion of Alkynes to Cyclobutenediones or α,β-Unsaturated Carboxylic Acids Using Novel Reactive Iron Carbonyl Reagents. J. Org. Chem. 1998, 63, 4930−4935. (170) Rameshkumar, C.; Periasamy, M. Reactive Iron Carbonyl Species via Reduction of FeCl3 with NaBH4 in the Presence of CO: Conversion of 1-Alkynes to Benzoquinones and Cyclobutenediones. Organometallics 2000, 19, 2400−2402. (171) Periasamy, M.; Mukkanti, A.; Raj, D. S. Novel Synthesis of Acyloxyferrole Complexes from Alkynes and Their Conversion to Cyclobutenediones. Organometallics 2004, 23, 619−621. (172) Periasamy, M.; Rameshkumar, C.; Mukkanti, A. Conversion of Alkynes to Cyclic Imides and Anhydrides Using Reactive Iron Carbonyls Prepared from Fe(CO)5 and Fe3(CO)12. J. Organomet. Chem. 2002, 649, 209−213.

(173) Alper, H.; Edward, J. T. Reactions of Iron Pentacarbonyl with Compounds Containing the N-O Linkage. Can. J. Chem. 1970, 48, 1543−1552. (174) Elzinga, J.; Hogeveen, H. Intermediate Complex in the Reduction of Trimethylamine Oxide by Pentacarbonyl Iron: its Use as Catalyst in the Addition of Carbon Tetrachloride to Carbon-Carbon Double Bonds. J. Chem. Soc., Chem. Commun. 1977, 20, 705−706. (175) Periasamy, M.; Mukkanti, A.; Raj, D. S. Synthesis of Cyclobutenediones and Anhydrides from Alkynes Using the Fe(CO)5/Me3NO Reagent System. Organometallics 2004, 23, 6323− 6326. (176) Iranpoor, N.; Firouzabadi, H.; Riazi, A.; Pedrood, K. Regioselective Hydrocarbonylation of Phenylacetylene to α,β-Unsaturated Esters and Thioesters with Fe(CO)5 and Mo(CO)6. J. Organomet. Chem. 2016, 822, 67−73. (177) Driller, K. M.; Klein, H.; Jackstell, R.; Beller, M. Iron-Catalyzed Carbonylation: Selective and Efficient Synthesis of Succinimides. Angew. Chem., Int. Ed. 2009, 48, 6041−6044. (178) Prateeptongkum, S.; Driller, K. M.; Jackstell, R.; Spannenberg, A.; Beller, M. Efficient Synthesis of Biologically Interesting 3, 4-DiarylSubstituted Succinimides and Maleimides: Application of IronCatalyzed Carbonylations. Chem. - Eur. J. 2010, 16, 9606−9615. (179) Prateeptongkum, S.; Driller, K. M.; Jackstell, R.; Beller, M. Iron-Catalyzed Carbonylation as a Key Step in the Short and Efficient Syntheses of Himanimide A and B. Chem. - Asian J. 2010, 5, 2173− 2176. (180) Driller, K. M.; Prateeptongkum, S.; Jackstell, R.; Beller, M. A General and Selective Iron-Catalyzed Aminocarbonylation of Alkynes: Synthesis of Acryl-and Cinnamides. Angew. Chem., Int. Ed. 2011, 50, 537−541. (181) Pizzetti, M.; Russo, A.; Petricci, E. Microwave-Assisted Aminocarbonylation of Ynamides by Using Catalytic [Fe3(CO)12] at Low Pressures of Carbon Monoxide. Chem. - Eur. J. 2011, 17, 4523− 4528. (182) Mathur, P.; Joshi, R. K.; Jha, B.; Singh, A. K.; Mobin, S. M. Towards the Catalytic Formation of α, β-Vinylesters and Alkoxy Substituted γ-Lactones. J. Organomet. Chem. 2010, 695, 2687−2694. (183) Higuchi, Y.; Atobe, S.; Tanaka, M.; Kamiya, I.; Yamamoto, T.; Nomoto, A.; Sonoda, M.; Ogawa, A. Cobalt-Catalyzed Thiolative Lactonization of Alkynes with Double CO Incorporation. Organometallics 2011, 30, 4539−4543. (184) Higuchi, Y.; Higashimae, S.; Tamai, T.; Ogawa, A. A Highly Selective Cobalt-Catalyzed Carbonylative Cyclization of Internal Alkynes with Carbon Monoxide and Organic Thiols. Tetrahedron 2013, 69, 11197−11202. (185) Hou, J.; Xie, J.; Zhou, Q. Palladium-Catalyzed Hydrocarboxylation of Alkynes with Formic Acid. Angew. Chem., Int. Ed. 2015, 54, 6302−6305. (186) Dai, J.; Ren, W.; Wang, H.; Shi, Y. A Facile Approach to βAmino Acid Derivatives via Palladium-Catalyzed Hydrocarboxylation of Enimides with Formic Acid. Org. Biomol. Chem. 2015, 13, 8429− 8432. (187) Wang, Y.; Ren, W.; Shi, Y. An Atom-Economic Approach to Carboxylic Acids via Pd-Catalyzed Direct Addition of Formic Acid to Olefins with Acetic Anhydride as a Co-Catalyst. Org. Biomol. Chem. 2015, 13, 8416−8419. (188) Fu, M.; Shang, R.; Cheng, W.; Fu, Y. Nickel-Catalyzed RegioAnd Stereoselective Hydrocarboxylation of Alkynes with Formic Acid through Catalytic CO Recycling. ACS Catal. 2016, 6, 2501−2505. (189) Hou, J.; Yuan, M.; Xie, J.; Zhou, Q. Nickel-Catalyzed Hydrocarboxylation of Alkynes with Formic Acid. Green Chem. 2016, 18, 2981−2984. (190) Jiang, J.; Fu, M.; Li, C.; Shang, R.; Fu, Y. Theoretical Investigation on Nickel-Catalyzed Hydrocarboxylation of Alkynes Employing Formic Acid. Organometallics 2017, 36, 2818−2825. (191) Ogoshi, S.; Arai, T.; Ohashi, M.; Kurosawa, H. Nickeladihydrofuran. Key Intermediate for Nickel-Catalyzed Reaction of Alkyne and Aldehyde. Chem. Chem. Commun. 2008, 1347−1349. AH

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(192) Hoshimoto, Y.; Ohata, T.; Sasaoka, Y.; Ohashi, M.; Ogoshi, S. Nickel (0)-Catalyzed [2 + 2+1] Carbonylative Cycloaddition of Imines and Alkynes or Norbornene Leading to γ-lactams. J. Am. Chem. Soc. 2014, 136, 15877−15880. (193) Hoshimoto, Y.; Ashida, K.; Sasaoka, Y.; Kumar, R.; Kamikawa, K.; Verdaguer, X.; Riera, A.; Ohashi, M.; Ogoshi, S. Efficient Synthesis of Polycyclic γ-Lactams by Catalytic Carbonylation of Ene-Imines via Nickelacycle Intermediates. Angew. Chem., Int. Ed. 2017, 56, 8206− 8210. (194) Xiao, Q.; Ren, W.-W.; Chen, Z.-X.; Sun, T.-W.; Li, Y.; Ye, Q.D.; Gong, J.-X.; Meng, F.-K.; You, L.; Liu, Y.-F.; Zhao, M.-Z.; Xu, L.M.; Shan, Z.-H.; Tang, Y.-F.; Chen, J.-H.; Yang, Z.; Shi, Y. Diastereoselective Total Synthesis of (±)-Schindilactone A. Angew. Chem., Int. Ed. 2011, 50, 7373−7377. (195) Yang, Y.; Fu, X.; Chen, J.; Zhai, H. Total Synthesis of (−)-Jiadifenin. Angew. Chem., Int. Ed. 2012, 51, 9825−9828. (196) Liu, Q.; Yue, G.-Z.; Wu, N.; Lin, G.; Li, Y. Z.; Quan, J.-M.; Li, C.-C.; Wang, G.-X.; Yang, Z. Total Synthesis of (±)-Pentalenolactone A Methyl Ester. Angew. Chem., Int. Ed. 2012, 51, 12072−12076. (197) Huang, J.; Fang, L.-C.; Long, R.; Shi, L.-L.; Shen, H.-J.; Li, C.C.; Yang, Z. Asymmetric Total Synthesis of (+)-Fusarisetin A via the Intramolecular Pauson−Khand Reaction. Org. Lett. 2013, 15, 4018− 4021. (198) Jørgensen, L.; McKerrall, S. J.; Kuttruff, C. A.; Ungeheuer, F.; Felding, J.; Baran, P. S. 14-Step Synthesis of (+)-Ingenol from (+)-3Carene. Science 2013, 341, 878−882. (199) McKerrall, S. J.; Jørgensen, L.; Kuttruff, C. A.; Ungeheuer, F.; Baran, P. S. Development of a Concise Synthesis of (+)-Ingenol. J. Am. Chem. Soc. 2014, 136, 5799−5810. (200) You, L.; Liang, X.-T.; Xu, L.-M.; Wang, Y.-F.; Zhang, J.-J.; Su, Q.; Li, Y.-H.; Zhang, B.; Yang, S.-L.; Chen, J.-H.; Yang, Z. Asymmetric Total Synthesis of Propindilactone G. J. Am. Chem. Soc. 2015, 137, 10120−10123. (201) Lv, C.; Yan, X.-H.; Tu, Q.; Di, Y.-T.; Yuan, C.-M.; Fang, X.; Ben-David, Y.; Xia, L.; Gong, J.-X.; Shen, Y.-M.; Yang, Z.; Hao, X.-J. Isolation and Asymmetric Total Synthesis of Perforanoid A. Angew. Chem., Int. Ed. 2016, 55, 7539−7543. (202) Chuang, K. V.; Xu, C.; Reisman, S. E. A 15-Step Synthesis of (+)-Ryanodol. Science 2016, 353, 912−915. (203) Magnus, P.; Principe, L. M. Origins Of 1,2- And 1,3Stereoselectivity in Dicobaltoctacarbonyl Alkene-Alkyne Cyclizations for the Synthesis of Substituted Bicyclo [3.3.0] Octenones. Tetrahedron Lett. 1985, 26, 4851−4854. (204) Gimbert, Y.; Lesage, D.; Milet, A.; Fournier, F.; Greene, A. E.; Tabet, J.-C. On Early Events in the Pauson−Khand Reaction. Org. Lett. 2003, 5, 4073−4075. (205) Yamanaka, M.; Nakamura, E. Density Functional Studies on the Pauson−Khand Reaction. J. Am. Chem. Soc. 2001, 123, 1703− 1708. (206) Brummond, K. M.; Kent, J. L. Recent Advances in the Pauson−Khand Reaction and Related [2 + 2+1] Cycloadditions. Tetrahedron 2000, 56, 3263−3283. (207) Gibson, S. E.; Stevenazzi, A. The Pauson−Khand Reaction: the Catalytic Age Is Here! Angew. Chem., Int. Ed. 2003, 42, 1800−1810. (208) Blanco-Urgoiti, J.; Añorbe, L.; Pérez-Serrano, L.; Domínguez, G.; Pérez-Castells, J. The Pauson−Khand Reaction, a Powerful Synthetic Tool for the Synthesis of Complex Molecules. Chem. Soc. Rev. 2004, 33, 32−42. (209) Gibson, S. E.; Mainolfi, N. The Intermolecular Pauson−Khand Reaction. Angew. Chem., Int. Ed. 2005, 44, 3022−3037. (210) Laschat, S.; Becheanu, A.; Bell, T.; Baro, A. Regioselectivity, Stereoselectivity and Catalysis in Intermolecular Pauson-Khand Reactions: Teaching an Old Dog New Tricks. Synlett 2005, 2005, 2547−2570. (211) Shibata, T. Recent Advances in the Catalytic Pauson−KhandType Reaction. Adv. Synth. Catal. 2006, 348, 2328−2336. (212) Park, J. H.; Chang, K.-M.; Chung, Y. K. Catalytic Pauson− Khand-Type Reactions and Related Carbonylative Cycloaddition Reactions. Coord. Chem. Rev. 2009, 253, 2461−2480.

(213) Lee, H.-W.; Kwong, F.-Y. A Decade of Advancements in Pauson−Khand-Type Reactions. Eur. J. Org. Chem. 2010, 2010, 789− 811. (214) Kitagaki, S.; Inagaki, F.; Mukai, C. [2 + 2+1] Cyclization of Allenes. Chem. Soc. Rev. 2014, 43, 2956−2978. (215) Röse, P.; Hilt, G. Cobalt-Catalysed Bond Formation Reactions; Part 2. Synthesis 2016, 48, 463−492. (216) Shi, L.; Yang, Z. Exploring the Complexity-Generating Features of the Pauson−Khand Reaction from a Synthetic Perspective. Eur. J. Org. Chem. 2016, 2016, 2356−2368. (217) Simeonov, S. P.; Nunes, J. P. M.; Guerra, K.; Kurteva, V. B.; Afonso, C. A. M. Synthesis of Chiral Cyclopentenones. Chem. Rev. 2016, 116, 5744−5893. (218) Jeong, N. The Pauson−Khand Reaction. In Comprehensive Organic Synthesis II, 2nd ed.; Knochel, P., Molander, G. A., Eds.; Elsevier: Amsterdam, 2014; Vol. 5, Chapter 5.24, p 1106. (219) Ricker, J. D.; Geary, L. M. Recent Advances in the Pauson− Khand Reaction. Top. Catal. 2017, 60, 609−619. (220) Ojima, I. Homogeneous Mixed-Metal Catalyst Systems: New and Effective Routes to N-Acyl-Α-Amino Acids via Carbonylations. J. Mol. Catal. 1986, 37, 25−44. (221) Ojima, I. New Aspects of Carbonylations Catalyzed by Transition Metal Complexes. Chem. Rev. 1988, 88, 1011−1030. (222) Beller, M.; Eckert, M. AmidocarbonylationAn Efficient Route to Amino Acid Derivatives. Angew. Chem., Int. Ed. 2000, 39, 1010−1027. (223) Wakamatsu, H.; Uda, J.; Yamakami, N. Synthesis of N-Acyl Acids by a Carbonylation Reaction. J. Chem. Soc. D 1971, 23, 1540− 1540. (224) Wakamatsu, H.; Uda, J.; Yamakami, N. Ger. Patent DEB 2115985, 1971. (225) Parnaud, J.-J.; Campari, G.; Pino, P. Some Aspects of the Catalytic Synthesis of N-Acyl-α-Aminoacids by Carbonylation of Aldehydes in the Presence of Amides. J. Mol. Catal. 1979, 6, 341−350. (226) Magnus, P.; Slater, M. Studies on the Hydrocarboxylation of N-Acetylimines, Enamines and Allylamines. Tetrahedron Lett. 1987, 28, 2829−2832. (227) Stern, R.; Reffet, D.; Hirschauer, A.; Commereuc, D.; Chauvin, Y. New Synthesis of N-Acyliminodiacetic Acid by Condensation of Amides with Formaldehyde in the Presence of Carbon Monoxide Catalyzed by Dicobalt Octacarbonyl. Synth. Commun. 1982, 12, 1111− 1114. (228) Hirai, K.; Takahashi, Y.; Ojima, I. Direct Conversion of Allylic Alcohols into N-Acyl-Α-Amino Acids by Catalytic Amidocarbonylation by Means of Homogeneous Binary Systems. Tetrahedron Lett. 1982, 23, 2491−2494. (229) Ojima, I.; Hirai, K.; Fujita, M.; Fuchikami, T. New Synthetic Route to N-Acyl-α-Amino Acids via Amidocarbonylation by Means of Homogeneous Binary Catalyst Systems. J. Organomet. Chem. 1985, 279, 203−214. (230) Ojima, I.; Okabe, M.; Kato, K.; Kwon, H. B.; Horváth, I. T. Homogeneous Catalysis of Mixed-Metal Systems. Highly Regioselective Hydroformylation-Amidocarbonylation of a Fluoro Olefin Catalyzed by Co-Rh Mixed-Metal Systems. Observation of CoRh(CO)7 Catalysis. J. Am. Chem. Soc. 1988, 110, 150−157. (231) Seki, Y.; Murai, S.; Sonoda, N. The Co2(CO)8/Ph3P-Catalyzed Reaction of Aldehydes with Hydrosilane and Carbon Monoxide. Angew. Chem., Int. Ed. Engl. 1978, 17, 119−120. (232) Murai, S.; Kato, T.; Sonoda, N.; Seki, Y.; Kawamoto, K. Catalytic Conversion of Aldehydes into Higher α-Siloxy Aldehydes by Hydrosilane and Carbon Monoxide. Angew. Chem., Int. Ed. Engl. 1979, 18, 393−394. (233) Chatani, N.; Furukawa, H.; Kato, T.; Murai, S.; Sonoda, N. Catalytic Incorporation of Carbon Monoxide into A Ketonic Carbon. Conversion of Cyclobutanones to Disiloxycyclopentenes with Hydrosilane and Carbon Monoxide in the Presence of Cobalt Carbony. J. Am. Chem. Soc. 1984, 106, 430−432. AI

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(234) Khumtaveeporn, K.; Alper, H. Transition Metal Mediated Carbonylative Ring Expansion of Heterocyclic Compounds. Acc. Chem. Res. 1995, 28, 414−422. (235) Seki, Y.; Murai, S.; Yamamoto, I.; Sonoda, N. Co2(CO)8 Catalyzed Reactions of Cyclic Ethers with Hydrosilanes and Carbon Monoxide. Angew. Chem., Int. Ed. Engl. 1977, 16, 789−789. (236) Murai, S.; Sonoda, N. Catalytic reactions with Hydrosilane and Carbon Monoxide [New synthetic methods (30)]. Angew. Chem., Int. Ed. Engl. 1979, 18, 837−846. (237) Murai, S.; Seki, Y. Silylcobalt Carbonyl and New Catalytic Reactions. J. Mol. Catal. 1987, 41, 197−207. (238) Murai, T.; Kato, S.; Murai, S.; Toki, T.; Suzuki, S.; Sonoda, N. Oxymethylative Opening of Oxiranes Leading to 1,3-Diol Derivatives by Cobalt Carbonyl Catalyzed Reaction with A Hydrosilane and Carbon Monoxide. J. Am. Chem. Soc. 1984, 106, 6093−6095. (239) Murai, T.; Yasui, E.; Kato, S.; Hatayama, Y.; Suzuki, S.; Yamasaki, Y.; Sonoda, N.; Kurosawa, H.; Kawasaki, Y.; Murai, S. Cobalt Carbonyl-Catalyzed Reactions of Cyclic Ethers with A Hydrosilane and Carbon Monoxide. A New Synthetic Reaction Equivalent to Nucleophilic Oxymethylation. J. Am. Chem. Soc. 1989, 111, 7938−7946. (240) Watanabe, Y.; Nishiyama, K.; Zhang, K.; Okuda, F.; Kondo, T.; Tsuji, Y. Co2(CO)8-Catalyzed Ring-Opening Carbonylation of Cyclic Ethers Using N-Silylamines. Bull. Chem. Soc. Jpn. 1994, 67, 879−882. (241) Hinterding, K.; Jacobsen, E. N. Regioselective Carbomethoxylation of Chiral Epoxides: A New Route to Enantiomerically Pure ΒHydroxy Esters. J. Org. Chem. 1999, 64, 2164−2165. (242) Denmark, S. E.; Ahmad, M. Carbonylative Ring Opening of Terminal Epoxides at Atmospheric Pressure. J. Org. Chem. 2007, 72, 9630−9634. (243) Goodman, S. N.; Jacobsen, E. N. Enantiopure β-Hydroxy Morpholine Amides from Terminal Epoxides by Carbonylation at 1 atm. Angew. Chem., Int. Ed. 2002, 41, 4703−4705. (244) Reppe, W.; Kroper, H.; Pistor, H. J.; Weissbarth, O. Carbonylierung IV. Einwirkung von Kohlenoxyd und Wasser auf Cyclische Ä ther. Justus Liebigs Ann. Chem. 1953, 582, 87−116. (245) Murahashi, S.; Horiie, S. The Reaction of Azobenzene and Carbon Monoxide. J. Am. Chem. Soc. 1956, 78, 4816−4816. (246) Aumann, R.; Ring, H. δ-Lactones by Carbonylation of Vinyloxiranes. Angew. Chem., Int. Ed. Engl. 1977, 16, 50−50. (247) Drent, E.; Kragtwijk, E. Eur. Pat. Appl. EP 577206, 1994; Chem. Abstr. 1994, 120, 191517c. (248) Piotti, M. E.; Alper, H. Inversion of Stereochemistry in the Co2(CO)8-Catalyzed Carbonylation of Aziridines to Β-Lactams. The First Synthesis of Highly Strained Trans-Bicyclic Β-Lactams. J. Am. Chem. Soc. 1996, 118, 111−116. (249) Calet, S.; Urso, F.; Alper, H. Enantiospecific and Stereospecific Rhodium (I)-Catalyzed Carbonylation and Ring Expansion of Aziridines. Asymmetric Synthesis of β-Lactams and the Kinetic Resolution of Aziridines. J. Am. Chem. Soc. 1989, 111, 931−934. (250) Lee, T. L.; Thomas, P. J.; Alper, H. Synthesis of β-Lactones by the Regioselective, Cobalt and Lewis Acid Catalyzed Carbonylation of Simple and Functionalized Epoxides. J. Org. Chem. 2001, 66, 5424− 5426. (251) Sigman, M. S.; Jacobsen, E. N. Enantioselective Addition of Hydrogen Cyanide to Imines Catalyzed by a Chiral (Salen)Al(III) Complex. J. Am. Chem. Soc. 1998, 120, 5315−5316. (252) Evans, D. A.; Janey, J. M.; Magomedov, N.; Tedrow, J. S. Chiral Salen−Aluminum Complexes as Catalysts for Enantioselective Aldol Reactions of Aldehydes and 5-Alkoxyoxazoles: An Efficient Approach to the Asymmetric Synthesis of syn and anti β-Hydroxy-αamino Acid Derivatives. Angew. Chem., Int. Ed. 2001, 40, 1884−1888. (253) Getzler, Y. D. Y. L.; Mahadevan, V.; Lobkovsky, E. B.; Coates, G. W. Synthesis of β-lactones: A Highly Active and Selective Catalyst for Epoxide Carbonylation. J. Am. Chem. Soc. 2002, 124, 1174−1175. (254) Tokunaga, M.; Larrow, J. F.; Kakiuchi, F.; Jacobsen, E. N. Asymmetric Catalysis with Water: Efficient Kinetic Resolution of Terminal Epoxides by Means of Catalytic Hydrolysis. Science 1997, 277, 936−938.

(255) Getzler, Y. D. Y. L.; Mahadevan, V.; Lobkovsky, E. B.; Coates, G. W. Stereochemistry of Epoxide Carbonylation Using Bimetallic Lewis Acid/Metal Carbonyl Complexes. Pure Appl. Chem. 2004, 76, 557−588. (256) Mahadevan, V.; Getzler, Y. D. Y. L.; Coates, G. W. [Lewis Acid]+[Co(CO)4]− Complexes: A Versatile Class of Catalysts for Carbonylative Ring Expansion of Epoxides And Aziridines. Angew. Chem., Int. Ed. 2002, 41, 2781−2784. (257) Schmidt, J. A. R.; Mahadevan, V.; Getzler, Y. D. Y. L.; Coates, G. W. A Readily Synthesized and Highly Active Epoxide Carbonylation Catalyst Based on a Chromium Porphyrin Framework: Expanding the Range of Available β-Lactones. Org. Lett. 2004, 6, 373−376. (258) Schmidt, J. A. R.; Lobkovsky, E. B.; Coates, G. W. Chromium (III) Octaethylporphyrinato Tetracarbonylcobaltate: A Highly Active, Selective, and Versatile Catalyst for Epoxide Carbonylation. J. Am. Chem. Soc. 2005, 127, 11426−11435. (259) Kramer, J. W.; Lobkovsky, E. B.; Coates, G. W. Practical βLactone Synthesis: Epoxide Carbonylation at 1 atm. Org. Lett. 2006, 8, 3709−3712. (260) Ganji, P.; Doyle, D. J.; Ibrahim, H. In Situ Generation of the Coates Catalyst: A Practical and Versatile Catalytic System for the Carbonylation of meso-Epoxides. Org. Lett. 2011, 13, 3142−3145. (261) Ganji, P.; Ibrahim, H. The First Asymmetric Ring-Expansion Carbonylation of Meso-Epoxides. Chem. Commun. 2012, 48, 10138− 10140. (262) Mulzer, M.; Ellis, W. C.; Lobkovsky, E. B.; Coates, G. W. Enantioenriched β-Lactone and Aldol-Type Products from Regiodivergent Carbonylation of Racemic cis-Epoxides. Chem. Sci. 2014, 5, 1928−1933. (263) Mulzer, M.; Lamb, J. R.; Nelson, Z.; Coates, G. W. Carbonylative Enantioselective Meso-Desymmetrization of cis-Epoxides to Trans-β-Lactones: Effect of Salen-Ligand Electronic Variation on Enantioselectivity. Chem. Commun. 2014, 50, 9842−9845. (264) Mulzer, M.; Coates, G. W. Carbonylation of cis-Disubstituted Epoxides to trans-β-Lactones: Catalysts Displaying Steric and Contrasteric Regioselectivity. J. Org. Chem. 2014, 79, 11851−11862. (265) Molnar, F.; Luinstra, G. A.; Allmendinger, M.; Rieger, B. Multisite Catalysis: A Mechanistic Study of β-Lactone Synthesis from Epoxides and CO-Insights into a Difficult Case of Homogeneous Catalysis. Chem. - Eur. J. 2003, 9, 1273−1280. (266) Allmendinger, M.; Eberhardt, R.; Luinstra, G. A.; Molnar, F.; Rieger, B. Cobaltcarbonyl ComplexesTunable Catalysts for the Carbonylation of Epoxides. Z. Anorg. Allg. Chem. 2003, 629, 1347− 1352. (267) Allmendinger, M.; Zintl, M.; Eberhardt, R.; Luinstra, G. A.; Molnar, F.; Rieger, B. Online ATR-IR Investigations And Mechanistic Understanding of the Carbonylation of Epoxides−the Selective Synthesis of Lactones or Polyesters from Epoxides and CO. J. Organomet. Chem. 2004, 689, 971−979. (268) Stirling, A.; Iannuzzi, M.; Parrinello, M.; Molnar, F.; Bernhart, V.; Luinstra, G. A. β-Lactone Synthesis from Epoxide and CO: Reaction Mechanism Revisited. Organometallics 2005, 24, 2533−2537. (269) Church, T. L.; Getzler, Y. D. Y. L.; Coates, G. W. The Mechanism of Epoxide Carbonylation by [Lewis Acid]+[Co(CO)4]− Catalysts. J. Am. Chem. Soc. 2006, 128, 10125−10133. (270) Chen, Q.; Mulzer, M.; Shi, P.; Beuning, P. J.; Coates, G. W.; O’Doherty, G. A. De Novo Asymmetric Synthesis of Fridamycin E. Org. Lett. 2011, 13, 6592−6595. (271) Chen, Q.; Zhong, Y.; O’Doherty, G. A. Convergent De novo Synthesis of Vineomycinone B2 Methyl Ester. Chem. Commun. 2013, 49, 6806−6808. (272) Mulzer, M.; Tiegs, B. J.; Wang, Y.; Coates, G. W.; O’Doherty, G. A. Total Synthesis of Tetrahydrolipstatin and Stereoisomers via a Highly Regio- and Diastereoselective Carbonylation of Epoxyhomoallylic Alcohols. J. Am. Chem. Soc. 2014, 136, 10814−10820. (273) Getzler, Y. D. Y. L.; Kundnani, V.; Lobkovsky, E. B.; Coates, G. W. Catalytic Carbonylation of β-lactones to Succinic Anhydrides. J. Am. Chem. Soc. 2004, 126, 6842−6843. AJ

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(274) Rowley, J. M.; Lobkovsky, E. B.; Coates, G. W. Catalytic Double Carbonylation of Epoxides to Succinic Anhydrides: Catalyst Discovery, Reaction Scope, and Mechanism. J. Am. Chem. Soc. 2007, 129, 4948−4960. (275) Church, T. L.; Byrne, C. M.; Lobkovsky, E. B.; Coates, G. W. A New Multicomponent Reaction Catalyzed by a [Lewis Acid]+[Co(CO)4]− Catalyst: Stereospecific Synthesis of 1,3-Oxazinane-2,4diones from Epoxides, Isocyanates, and CO. J. Am. Chem. Soc. 2007, 129, 8156−8162. (276) Kramer, J. W.; Joh, D. Y.; Coates, G. W. Carbonylation of Epoxides to Substituted 3-Hydroxy-δ-Lactones. Org. Lett. 2007, 9, 5581−5583. (277) de Lorenzo, F.; Feher, M.; Martin, J.; Collot-Teixeira, S.; Dotsenko, O.; McGregor, J. L. Statin Therapy - Evidence Beyond Lipid Lowering Contributing to Plaque Stability. Curr. Med. Chem. 2006, 13, 3385−3393. (278) Furuya, M.; Tsutsuminai, S.; Nagasawa, H.; Komine, N.; Hirano, M.; Komiya, S. Catalytic Synthesis of Thiobutyrolactones via CO Insertion into the C−S Bond of Thietanes in the Presence of a Heterodinuclear Organoplatinum−Cobalt Complex. Chem. Commun. 2003, 2046−2047. (279) Xu, H.; Jia, L. Novel Cobalt-Catalyzed Carbonylation of 2Aryl-2-oxazolines. Org. Lett. 2003, 5, 1575−1577. (280) Xu, H.; Gladding, J. A.; Jia, L. Stereochemistry of CobaltCatalyzed Carbonylation of 2-Oxazolines. Inorg. Chim. Acta 2004, 357, 4024−4028. (281) Byrne, C. M.; Church, T. L.; Kramer, J. W.; Coates, G. W. Catalytic Synthesis of β3-Amino Acid Derivatives from α-Amino Acids. Angew. Chem., Int. Ed. 2008, 47, 3979−3983. (282) Zhang, Y.; Ji, J.; Zhang, X.; Lin, S.; Pan, Q.; Jia, L. CobaltCatalyzed Cyclization of Carbon Monoxide, Imine, and Epoxide. Org. Lett. 2014, 16, 2130−2133. (283) Liu, L.; Sun, H. [HCo(CO)4]− Catalyzed Three-component Cycloaddition of Epoxides, Imines, and Carbon Monoxide: Facile Construction of 1,3-Oxazinan-4-ones. Angew. Chem., Int. Ed. 2014, 53, 9865−9869. (284) Slaugh, L. H.; Weider, P. R.; Arhancet, J. P.; Lin, J.-J. (Shell). Process for Making 1,3-Diols and 3-Hydroxyaldehydes. PCT Int. Appl. WO9418149, 1994. (285) Allmendinger, M.; Eberhardt, R.; Luinstra, G.; Rieger, B. The Cobalt-Catalyzed Alternating Copolymerization of Epoxides and Carbon Monoxide: A Novel Approach to Polyesters. J. Am. Chem. Soc. 2002, 124, 5646−5647. (286) Takeuchi, D.; Sakaguchi, Y.; Osakada, K. Alternating Copolymerization of Propylene Oxide with Carbon Monoxide Catalyzed by Co Complex and Co/Ru Complexes. J. Polym. Sci., Part A: Polym. Chem. 2002, 40, 4530−4537. (287) Lee, J. T.; Alper, H. Alternating Copolymerization of Propylene Oxide and Carbon Monoxide to form Aliphatic Polyesters. Macromolecules 2004, 37, 2417−2421. (288) Jia, L.; Sun, H.; Shay, J. T.; Allgeier, A. M.; Hanton, S. D. Living Alternating Copolymerization of N-Alkylaziridines and Carbon Monoxide as a Route for Synthesis of Poly-β-peptoids. J. Am. Chem. Soc. 2002, 124, 7282−7283. (289) Liu, G.; Jia, L. Design of Catalytic Carbonylative Polymerizations of Heterocycles. Synthesis of Polyesters and Amphiphilic Poly(amide-block-ester)s. J. Am. Chem. Soc. 2004, 126, 14716−14717. (290) Liu, G.; Jia, L. Cobalt-Catalyzed Carbonylative Copolymerization of N-Alkylazetidines and Tetrahydrofuran. Angew. Chem., Int. Ed. 2006, 45, 129−131. (291) Chai, J.; Liu, G.; Chaicharoen, K.; Wesdemiotis, C.; Jia, L. Cobalt-Catalyzed Carbonylative Polymerization of Azetidines. Macromolecules 2008, 41, 8980−8985. (292) Sun, H.; Zhang, J.; Liu, Q.; Yu, L.; Zhao, J. Metal-Catalyzed Copolymerization of Imines and CO: A Non-Amino Acid Route to Polypeptides. Angew. Chem., Int. Ed. 2007, 46, 6068−6072. (293) Ohashi, S.; Sakaguchi, S.; Ishii, Y. Carboxylation of Anisole Derivatives with CO and O 2 Catalyzed by Pd(Oac) 2 and Molybdovanadophosphates. Chem. Commun. 2005, 2005, 486−488.

(294) Giri, R.; Yu, J.-Q. Synthesis of 1,2-and 1,3-Dicarboxylic Acids via Pd(II)-Catalyzed Carboxylation of Aryl and Vinyl C−H Bonds. J. Am. Chem. Soc. 2008, 130, 14082−14083. (295) Giri, R.; Lam, J. K.; Yu, J.-Q. Synthetic Applications of Pd(II)Catalyzed C-H Carboxylation and Mechanistic Insights: Expedient Routes to Anthranilic Acids, Oxazolinones, and Quinazolinones. J. Am. Chem. Soc. 2010, 132, 686−693. (296) Houlden, C. E.; Hutchby, M.; Bailey, C. D.; Ford, J. G.; Tyler, S. N. G.; Gagné, M. R.; Lloyd-Jones, G. C.; Booker-Milburn, K. I. Room-Temperature Palladium-Catalyzed C-H Activation: orthoCarbonylation of Aniline Derivatives. Angew. Chem., Int. Ed. 2009, 48, 1830−1833. (297) Guan, Z.-H.; Chen, M.; Ren, Z.-H. Palladium-Catalyzed Regioselective Carbonylation of C−H Bonds of N-Alkyl Anilines for Synthesis of Isatoic Anhydrides. J. Am. Chem. Soc. 2012, 134, 17490− 17493. (298) Lang, R.; Shi, L.; Li, D.; Xia, C.; Li, F. A General Method for Palladium-Catalyzed Direct Carbonylation of Indole with Alcohol and Phenol. Org. Lett. 2012, 14, 4130−4133. (299) Zhang, H.; Liu, D.; Chen, C.; Liu, C.; Lei, A. PalladiumCatalyzed Regioselective Aerobic Oxidative C-H/N-H Carbonylation of Heteroarenes under Base-Free Conditions. Chem. - Eur. J. 2011, 17, 9581−9585. (300) López, B.; Rodriguez, A.; Santos, D.; Albert, J.; Ariza, X.; Garcia, J.; Granell, J. Preparation of Benzolactams by Pd(II)-Catalyzed Carbonylation of N-Unprotected Arylethylamines. Chem. Commun. 2011, 47, 1054−1056. (301) Wen, J.; Tang, S.; Zhang, F.; Shi, R.; Lei, A. Palladium/Copper Co-catalyzed Oxidative C−H/C−H Carbonylation of Diphenylamines: A Way to Access Acridones. Org. Lett. 2017, 19, 94−97. (302) Guan, Z.-H.; Ren, Z.-H.; Spinella, S. M.; Yu, S.; Liang, Y.-M.; Zhang, X. Rhodium-Catalyzed Direct Oxidative Carbonylation of Aromatic C− H Bond with CO and Alcohols. J. Am. Chem. Soc. 2009, 131, 729−733. (303) Du, Y.; Hyster, T. K.; Rovis, T. Rhodium(III)-catalyzed oxidative Carbonylation of Benzamides with Carbon Monoxide. Chem. Commun. 2011, 47, 12074−12076. (304) Inoue, S.; Shiota, H.; Fukumoto, Y.; Chatani, N. RutheniumCatalyzed Carbonylation at Ortho C−H Bonds in Aromatic Amides Leading to Phthalimides: C-H Bond Activation Utilizing a Bidentate System. J. Am. Chem. Soc. 2009, 131, 6898−6899. (305) Zaitsev, V. G.; Shabashov, D.; Daugulis, O. Highly Regioselective Arylation of sp3 C−H Bonds Catalyzed by Palladium Acetate. J. Am. Chem. Soc. 2005, 127, 13154−13155. (306) Shabashov, D.; Daugulis, O. Auxiliary-Assisted PalladiumCatalyzed Arylation and Alkylation of sp2 and sp3 Carbon−Hydrogen Bonds. J. Am. Chem. Soc. 2010, 132, 3965−3972. (307) Zhang, S.-Y.; He, G.; Nack, W. A.; Zhao, Y.; Li, Q.; Chen, G. Palladium-Catalyzed Picolinamide-Directed Alkylation of Unactivated C(sp3)−H Bonds with Alkyl Iodides. J. Am. Chem. Soc. 2013, 135, 2124−2127. (308) He, G.; Chen, G. A Practical Strategy for the Structural Diversification of Aliphatic Scaffolds through the Palladium-Catalyzed Picolinamide-Directed Remote Functionalization of Unactivated C(sp3)H Bonds. Angew. Chem., Int. Ed. 2011, 50, 5192−5196. (309) Aihara, Y.; Chatani, N. Ruthenium-Catalyzed Direct Arylation of C−H Bonds in Aromatic Amides Containing A Bidentate Directing Group: Significant Electronic Effects on Arylation. Chem. Sci. 2013, 4, 664−670. (310) Rouquet, G.; Chatani, N. Ruthenium-Catalyzed ortho-C−H Bond Alkylation of Aromatic Amides with α, β-Unsaturated Ketones via Bidentate-Chelation Assistance. Chem. Sci. 2013, 4, 2201−2208. (311) Shibata, K.; Chatani, N. Rhodium-Catalyzed Alkylation of C− H Bonds in Aromatic Amides with α, β-Unsaturated Esters. Org. Lett. 2014, 16, 5148−5151. (312) Shang, R.; Ilies, L.; Matsumoto, A.; Nakamura, E. β-Arylation of Carboxamides via Iron-Catalyzed C(sp3)−H Bond Activation. J. Am. Chem. Soc. 2013, 135, 6030−6032. AK

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(313) Asako, S.; Ilies, L.; Nakamura, E. Iron-Catalyzed OrthoAllylation of Aromatic Carboxamides with Allyl Ethers. J. Am. Chem. Soc. 2013, 135, 17755−17757. (314) Gu, Q.; Al Mamari, H. H.; Graczyk, K.; Diers, E.; Ackermann, L. Iron-Catalyzed C(sp2)−H and C(sp3)−H Arylation by Triazole Assistance. Angew. Chem., Int. Ed. 2014, 53, 3868−3871. (315) Monks, B. M.; Fruchey, E. R.; Cook, S. P. Iron-Catalyzed C(sp2)−H Alkylation of Carboxamides with Primary Electrophiles. Angew. Chem., Int. Ed. 2014, 53, 11065−11069. (316) Aihara, Y.; Chatani, N. Nickel-Catalyzed Direct Arylation of C(sp3)−H Bonds in Aliphatic Amides via Bidentate-Chelation Assistance. J. Am. Chem. Soc. 2014, 136, 898−901. (317) Li, M.; Dong, J.; Huang, X.; Li, K.; Wu, Q.; Song, F.; You, J. Nickel-Catalyzed Chelation-Assisted Direct Arylation of Unactivated C(sp3)−H Bonds with Aryl Halides. Chem. Commun. 2014, 50, 3944− 3946. (318) Song, W.; Lackner, S.; Ackermann, L. Nickel-Catalyzed C−H Alkylations: Direct Secondary Alkylations and Trifluoroethylations of Arenes. Angew. Chem., Int. Ed. 2014, 53, 2477−2480. (319) Wu, X.; Zhao, Y.; Ge, H. Nickel-Catalyzed Site-Selective Alkylation of Unactivated C(sp3)−H Bonds. J. Am. Chem. Soc. 2014, 136, 1789−1792. (320) Nishino, M.; Hirano, K.; Satoh, T.; Miura, M. CopperMediated C−H/C−H Biaryl Coupling of Benzoic Acid Derivatives and 1,3-Azoles. Angew. Chem., Int. Ed. 2013, 52, 4457−4461. (321) Tran, L. D.; Popov, I.; Daugulis, O. Copper-Promoted Sulfenylation of sp2 C−H Bonds. J. Am. Chem. Soc. 2012, 134, 18237− 18240. (322) Tran, L. D.; Roane, J.; Daugulis, O. Directed Amination of Non-Acidic Arene C−H Bonds by a Copper−Silver Catalytic System. Angew. Chem., Int. Ed. 2013, 52, 6043−6046. (323) Truong, T.; Klimovica, K.; Daugulis, O. Copper-Catalyzed, Directing Group-Assisted Fluorination of Arene and Heteroarene C− H Bonds. J. Am. Chem. Soc. 2013, 135, 9342−9345. (324) Wang, Z.; Ni, J.; Kuninobu, Y.; Kanai, M. Copper-Catalyzed Intramolecular C(sp3) − H and C(sp2) − H Amidation by Oxidative Cyclization. Angew. Chem., Int. Ed. 2014, 53, 3496−3499. (325) Grigorjeva, L.; Daugulis, O. Cobalt−Catalyzed, Aminoquinoline−Directed C (sp2) − H Bond Alkenylation by Alkynes. Angew. Chem., Int. Ed. 2014, 53, 10209−10212. (326) Grigorjeva, L.; Daugulis, O. Cobalt-Catalyzed, Aminoquinoline-Directed Coupling of sp2 C−H Bonds with Alkenes. Org. Lett. 2014, 16, 4684−4687. (327) Grigorjeva, L.; Daugulis, O. Cobalt-Catalyzed Direct Carbonylation of Aminoquinoline Benzamides. Org. Lett. 2014, 16, 4688− 4690. (328) Nguyen, T. T.; Grigorjeva, L.; Daugulis, O. AminoquinolineDirected, Cobalt-Catalyzed Carbonylation of Sulfonamide sp 2 C−H Bonds. Chem. Commun. 2017, 53, 5136−5138. (329) Yu, W.-Y.; Sit, W. N.; Lai, K.-M.; Zhou, Z.; Chan, A. S. C. Palladium-Catalyzed Oxidative Ethoxycarbonylation of Aromatic C− H Bond with Diethyl Azodicarboxylate. J. Am. Chem. Soc. 2008, 130, 3304−3306. (330) Ni, J.; Li, J.; Fan, Z.; Zhang, A. Cobalt-Catalyzed Carbonylation of C(sp2)−H Bonds with Azodicarboxylate as the Carbonyl Source. Org. Lett. 2016, 18, 5960−5963. (331) Hao, X.-Q.; Du, C.; Zhu, X.; Li, P.-X.; Zhang, J.-H.; Niu, J.-L.; Song, M.-P. Cobalt (II)-Catalyzed Decarboxylative C−H Activation/ Annulation Cascades: Regioselective Access to Isoquinolones and Isoindolinones. Org. Lett. 2016, 18, 3610−3613. (332) Landge, V. G.; Jaiswal, G.; Balaraman, E. Cobalt-Catalyzed Bisalkynylation of Amides via Double C−H Bond Activation. Org. Lett. 2016, 18, 812−815. (333) Huang, Y.; Li, G.; Huang, J.; You, J. Palladium-Catalyzed Direct Ortho-C−H Ethoxycarboxylation of Anilides at Room Temperature. Org. Chem. Front. 2014, 1, 347−350. (334) Xu, N.; Li, D.; Zhang, Y.; Wang, L. Palladium-Catalyzed Direct Ortho-Ethoxycarbonylation of Azobenzenes and Azoxybenzenes with Diethyl Azodicarboxylate. Org. Biomol. Chem. 2015, 13, 9083−9092.

(335) Ling, F.; Ai, C.; Lv, Y.; Zhong, W. Traceless Directing Group Assisted Cobalt-Catalyzed C−H Carbonylation of Benzylamines. Adv. Synth. Catal. 2017, 359, 3707−3712. (336) Khan, B.; Khan, A. A.; Kant, R.; Koley, D. Directing GroupAssisted Copper(II)-Catalyzed Ortho-Carbonylation to Benzamide Using 2,2’-Azobisisobutyronitrile (AIBN). Adv. Synth. Catal. 2016, 358, 3753−3758. (337) Usman, M.; Ren, Z.-H.; Wang, Y.-Y.; Guan, Z.-H. A CopperCatalyzed Reaction of Oximes with Diisopropyl Azodicarboxylate: An Alternative Method for the Synthesis of Oxime Carbonates. Org. Biomol. Chem. 2017, 15, 1091−1095. (338) Liu, X.-G.; Zhang, S.-S.; Jiang, C.-Y.; Wu, J.-Q.; Li, Q.; Wang, H. Cp* Co (III)-Catalyzed Annulations of 2-Alkenylphenols with CO: Mild Access to Coumarin Derivatives. Org. Lett. 2015, 17, 5404−5407. (339) Ferguson, J.; Zeng, F.; Alper, H. Synthesis Of Coumarins via Pd-Catalyzed Oxidative Cyclocarbonylation of 2-Vinylphenols. Org. Lett. 2012, 14, 5602−5605. (340) Seoane, A.; Casanova, N.; Quiñones, N.; Mascareñas, J. L.; Gulías, M. Straightforward Assembly of Benzoxepines by Means of a Rhodium(III)-Catalyzed C−H Functionalization of o-Vinylphenols. J. Am. Chem. Soc. 2014, 136, 834−837. (341) Yoo, E. J.; Wasa, M.; Yu, J.-Q. Pd(II)-Catalyzed Carbonylation of C(sp3)−H Bonds: A New Entry to 1,4-Dicarbonyl Compounds. J. Am. Chem. Soc. 2010, 132, 17378−17380. (342) Hasegawa, N.; Charra, V.; Inoue, S.; Fukumoto, Y.; Chatani, N. Highly Regioselective Carbonylation of Unactivated C(sp3)−H Bonds by Ruthenium Carbonyl. J. Am. Chem. Soc. 2011, 133, 8070−8073. (343) Wu, X.; Yang, K.; Zhao, Y.; Sun, H.; Li, G.; Ge, H. CobaltCatalysed Site-Selective Intra- and Intermolecular Dehydrogenative Amination of Unactivated sp3 Carbons. Nat. Commun. 2015, 6, 6462. (344) Zhang, J.; Chen, H.; Lin, C.; Liu, Z.; Wang, C.; Zhang, Y. Cobalt-Catalyzed Cyclization of Aliphatic Amides and Terminal Alkynes with Silver-Cocatalyst. J. Am. Chem. Soc. 2015, 137, 12990− 12996. (345) Barsu, N.; Bolli, S. K.; Sundararaju, B. Cobalt Catalyzed Carbonylation of Unactivated C(sp3)−H Bonds. Chem. Sci. 2017, 8, 2431−2435. (346) Williamson, P.; Galvan, A.; Gaunt, M. J. Cobalt-Catalysed C− H Carbonylative Cyclisation of Aliphatic Amides. Chem. Sci. 2017, 8, 2588−2591. (347) Zeng, Li.; Tang, S.; Wang, D.; Deng, Y.; Chen, J.-L.; Lee, J.-F.; Lei, A. Cobalt-Catalyzed Intramolecular Oxidative C(sp3)−H/N−H Carbonylation of Aliphatic Amides. Org. Lett. 2017, 19, 2170−2173. (348) Li, Y.; Dong, K.; Zhu, F.; Wang, Z.; Wu, X.-F. CopperCatalyzed Carbonylative Coupling of Cycloalkanes and Amides. Angew. Chem., Int. Ed. 2016, 55, 7227−7230. (349) Li, Y.; Zhu, F.; Wang, Z.; Wu, X.-F. Copper-Catalyzed Carbonylative Synthesis of Aliphatic Amides from Alkanes and Primary Amines via C(sp3)−H Bond Activation. ACS Catal. 2016, 6, 5561−5564. (350) Li, Y.; Wang, C.; Zhu, F.; Wang, Z.; Dixneuf, P. H.; Wu, X.-F. Copper-Catalyzed Alkoxycarbonylation of Alkanes with Alcohols. ChemSusChem 2017, 10, 1341−1345. (351) Li, Y.; Zhu, F.; Wang, Z.; Wu, X.-F. A Copper-Catalyzed Carbonylative Four-Component Reaction of Ethene and Aliphatic Olefins. Chem. Commun. 2018, 54, 1984−1987.

AL

DOI: 10.1021/acs.chemrev.8b00068 Chem. Rev. XXXX, XXX, XXX−XXX