Synthesis 2022; 54(17): 3667-3697
DOI: 10.1055/a-1845-3810
review

Recent Advances in C–F Bond Activation of Acyl Fluorides Directed toward Catalytic Transformation by Transition Metals, N-Heterocyclic Carbenes, or Phosphines

Tian Tian
a   Graduate School of Natural Science and Technology, Okayama University, 3-1-1 Tsushimanaka, Kita-ku, Okayama 700-8530, Japan
,
Qiang Chen
a   Graduate School of Natural Science and Technology, Okayama University, 3-1-1 Tsushimanaka, Kita-ku, Okayama 700-8530, Japan
,
Zhiping Li
b   Department of Chemistry, Renmin University of China, Beijing 100872, P. R. of China
,
c   Research Institute for Interdisciplinary Science, Okayama University, 3-1-1 Tsushimanaka, Kita-ku, Okayama 700-8530, Japan
› Author Affiliations
 


Abstract

Numerous studies on the activation of carbon–fluorine bonds have been reported in recent years. For example, acyl fluorides have been utilized as versatile reagents for acylation, arylation, and even fluorination. In this review, we focus on acyl fluorides as compounds with carbon–fluorine bonds, and highlight recent advances in strategies for the activation of their C–F bonds via transition-metal catalysis, N-heterocyclic carbene (NHCs) catalysis, organophosphine catalysis, and classical nucleophilic substitution reactions.

1 Introduction

2 Transition-Metal-Mediated C–F Bond Activation

2.1 Acylation (Carbonyl-Retentive) Coupling Reactions

2.2 Decarbonylative Reactions

2.3 C–F Bond Activation by Other Transition Metals

3 C–F Bond Activation by N-Heterocyclic Carbenes (NHCs)

3.1 NHC-Catalyzed Cycloaddition of Acyl Fluorides

3.2 NHC-Catalyzed Radical Functionalization of Acyl Fluorides

3.3 NHC-Catalyzed Nucleophilic Fluorination of (Hetero)aromatics

4 C–F Bond Activation by Phosphines

4.1 Phosphine-Catalyzed Direct Activation of the C–F Bond of Acyl Fluorides

4.2 Phosphine-Catalyzed Indirect Activation of the C–F Bond of Acyl Fluorides

5 C–F Bond Activation by Classical Nucleophilic Substitution

6 Miscellaneous Examples

7 Summary and Perspective


#

Biographical Sketches

Zoom Image

Tian Tian was born in 1995 in Taiyuan, China. He obtained his M.S. degree in 2020 under the supervision of Prof. Zhiping Li at the Department of Chemistry, Renmin University of China. In 2021, he began his Ph.D. studies under the guidance of Prof. Yasushi Nishihara at the Graduate School of Natural Science and Technology, Okayama University. His current research interests mainly focus on transition-metal-catalyzed activation of C–F bonds of acyl fluorides leading to highly atom economic fluorination via C–F bond reductive elimination.

Zoom Image

Qiang Chen was born in 1991 in Shandong, China. He received his B.S. degree from Qingdao University in 2015 and his M.S. degree in 2018 from the Institute of Chemistry, Chinese Academy of Sciences under the supervision of Prof. Wen-Hua Sun. In 2022, he received his Ph.D. from Okayama University under the supervision of Prof. Yasushi Nishihara. His current research interests are focused on transition-metal-catalyzed cross-coupling reactions of acyl fluorides.

Zoom Image

Zhiping Li studied chemistry at Nanjing University of Science and Technology (B.Sc., 1993) and obtained his Ph.D. at Dalian University of Technology in 1999. After postdoctoral research at Peking University (with Professor Zhenfeng Xi, 1999–2000) and at Hokkaido University (with Professor Tamotsu Takahashi, 2001–2002), he joined Peking University as an assistant professor. In 2004, he moved to McGill University as a postdoctoral fellow with Professor Chao-Jun Li. He started his independent research at Renmin University of China as an associate professor in 2006, and became a professor of chemistry in 2009. His research interests include the development of synthetic methodologies, focusing on iron-catalyzed oxidative C–H bond transformation and selective C–C bond cleavage, the synthesis and applications of organic peroxides, and the synthesis of biologically active natural products.

Zoom Image

Yasushi Nishihara earned his B.S. degree from Hiroshima University in 1992. From 1993 to 1994, he was a research associate at the University of Notre Dame, USA. In 1997 he received his Ph.D. from the Graduate University for Advanced Studies (SOKENDAI) under the supervision of Professor Tamotsu Takahashi. He became an assistant professor at the Tokyo Institute of Technology in 1996, and later moved to Okayama University as an associate professor in 2004 and was promoted to professor in 2010, and has been a professor at the Research Institute for Interdisciplinary Science, Okayama University since 2016. His current research focuses on the development of nickel- and palladium-catalyzed decarbonylative transformations of acyl fluorides, the efficient and selective synthesis of π-conjugated organic molecules for applications in functional materials such as organic field-effect transistors and organic photovoltaics.

1

Introduction

The transformation of C–F bonds represents an important research area because fluorine atoms are present in a wide range of substances in Nature.[1] [2] However, the bond dissociation energy of a C–F bond is quite large and it is also considered to be kinetically inert, hence cleavage of C–F bonds is generally very difficult. Therefore, in order to break through this barrier, methods for cleaving inactive C–F bonds have been developed, which include the use of various transition-metal catalysts and photoredox or electrochemical catalysis. Detailed studies and breakthroughs have resulted in different types of activation methods becoming feasible, especially in the field of transition-metal-catalyzed C–F bond activation. In the C–[M]–F complexes formed through oxidative addition to the C–F bonds, the dissociation energy of the [M]–F bond prior to activation, which is regarded as a thermodynamically favorable process, required an activation energy that was too high, and only a few specific low-valent transition metals could effectively activate the C–F bond.[2,3] In the examples shown on the left-hand side of Scheme [1], the aryl–[M]–F complexes can be formed through oxidative addition to the Ar–F bond.[4] In addition, common alkenyl fluorides can undergo nucleophilic addition in the presence of organometallic reagents, followed by C–F bond cleavage with fluorine elimination.[5]

It has also been reported that C–F bonds are readily activated in gem-difluorocyclopropanes and trifluoromethylated compounds. The former facilitate the formation of [M]–allyl complexes catalyzed by transition metals (e.g., palladium) via ring-opening and C–F bond cleavage,[6] while the latter can undergo single-electron transfer (SET) in photoredox or electrochemical catalytic systems to produce fluoroalkyl radicals.[7] Nevertheless, the types of C–F bonds that can be effectively activated are still relatively limited. Therefore, chemists need to constantly develop new methods of activating C–F bonds while seeking ways to activate classical C–F bonds.

Zoom Image
Scheme 1 Left: Types of activation of C–F bonds. Right: Types of Activation of carboxylic acid derivatives.

Recently, however, as a consequence of the increasing attention from chemists and investigations on acyl fluorides, a new research mode has opened up in the field of C–F bond activation of acyl fluorides. Acyl fluorides, as with acid chlorides, acid anhydrides, esters and amides, are a class of carboxylic acid derivative, and it is well-known that they are inexpensive, readily available, and versatile raw materials in organic synthesis. Over the past few decades, various methods for the synthesis of acyl fluorides have been developed.[8] This is mainly due to the continuous development of different fluorinating reagents, which has resulted in additional procedures for the synthesis of acyl fluorides. At present, methods for the preparation of acyl fluorides are mainly classified as follows: (1) deoxyfluorination and halogen exchange of carboxylic acids and their derivatives, (2) transition-metal-catalyzed fluorocarbonylation of organic halogens, (3) nucleophilic substitution of acid chlorides with fluoride sources, and (4) acyl exchange of carboxylic acid derivatives. At the same time, many common acyl fluorides are commercially available. This not only indicates that the synthetic value of acyl fluorides is constantly improving, but also brings significant convenience to related research.

Meanwhile, acyl chlorides and acid anhydrides are prone to hydrolysis, alcoholysis, and aminolysis. On the other hand, there are some twisted amides reported by Szostak’s group[9] that are more reactive than acyl chlorides and more stable than anhydrides, but in general, esters and amides are considered chemically stable unless they are subjected to harsh conditions for dissolution, since alkoxy and amino groups are not good leaving groups. Detailed studies on transition-metal catalysis have shown that these carboxylic acid derivatives undergo oxidative addition to transition metals (Scheme [1], right), leading to a series of acyl active and/or decarbonylative coupling reactions.[10] [11] [12] [13]

However, research on acyl fluorides, which has taken place over decades, has always remained in the category of classical reactions. The C–F bond activation of acyl fluorides has undergone remarkable development by chemists in recent years. Although acyl fluorides are a type of acyl halide, they are metastable due to the inertness of the C–F bond. Due to the high electronegativity of the fluorine atom, the C–F bond has a relatively large dipole because of its electrostatic rather than covalent nature when compared to acyl chlorides. Consequently, acyl fluorides are much more stable than acyl chlorides.[14] Therefore, the reactivity of carboxylic acid derivatives increases in the following order: acyl chlorides > acid anhydrides > acyl fluorides > esters > amides (except for some twisted amides). Acyl fluorides are not sensitive to humidity, are not prone to solvolysis, and can be isolated by column chromatography, while hydrolysis,[15] alcoholysis,[16] and aminolysis[17] are possible under certain conditions. The versatile nature of acyl fluorides was discovered by chemists decades ago in the construction of C–N, C–C, and C–O bonds.[18] [19] [20] The acylation of amines is an important approach for the construction of C–N bonds by using acyl fluorides. On this basis, they can also be successfully applied to peptide coupling reactions with acid-sensitive N-protecting groups without racemization. Thus, a series of polypeptides was synthesized to overcome the difficulties that were encountered with acyl chlorides (Scheme [2a]).[18] In addition to nucleophilic substitution by organometallic compounds and carbanions, acyl fluorides have been employed to form C–C bonds by Friedel–Crafts reactions and Lewis acid-catalyzed acylation with aromatic hydrocarbons (Scheme [2b]).[19] Similarly, esters can also be synthesized by nucleophilic substitution of acyl fluorides with alcohols to construct C–O bonds (Scheme [2c]).[20]

Zoom Image
Scheme 2 Reactions of acyl fluorides with nucleophiles

Following developments in modern organic chemistry, acyl fluorides have become multifunctional reagents under various conditions, and can be used in organic synthesis as acylating, arylating (alkylating), and fluorinating reagents. To this end, our research group and others have successively reported related reviews,[8] [21] in which the main focus is the coupling reactions of acyl fluorides catalyzed by transition metals. In recent years, however, besides oxidative addition by transition metals, a variety of other methods have been discovered for the activation of the C–F bonds of acyl fluorides (Scheme [3a]). For example, diverse methods for C–F bond activation through the generation of acyl azolium ion intermediates with N-heterocyclic carbenes (NHCs) (Scheme [3b]), the formation of acyl phosphonium intermediates with phosphines (Scheme [3c]), and the synthesis of a series of acyl and hydrocarbyl compounds or fluorides via classical nucleophilic substitution (Scheme [3d]) have been achieved. Hence, this review outlines different organic reactions of acyl fluorides that proceed through C–F bond activation.

Zoom Image
Scheme 3 Activation and transformations of acyl fluorides

# 2

Transition-Metal-Mediated C–F Bond Activation

The development of novel reactions involving transition-metal-catalyzed C–F bond activation has long been one of the research targets in synthetic organic chemistry. The reaction modes involving C–F bond cleavage with transition metals can be classified as follows:[2] (i) oxidative addition of fluorocarbons,[4] (ii) M–C bond formation with HF elimination,[22] (iii) M–C bond formation with fluorosilane elimination,[23] (iv) hydrodefluorination of fluorocarbons with M–F bond formation,[24] (v) nucleophilic attack on fluorocarbons,[25] [26] and (vi) defluorination of fluorocarbons.[27] The observed aspects of these six activation modes are that the C–F bonds are converted into thermodynamically favorable [M]–F, H–F, and Si–F bonds with higher bond energies, assisted by transition metals, and that oxidative addition of C(sp2)–F bonds is the crucial process.

Indeed, transition-metal-catalyzed activation of C(sp2)–F bonds in acyl fluorides has been extensively developed and reported, and it has been found that oxidative addition of acyl fluorides with transition-metal catalysts followed by reactions with suitable nucleophiles facilitates carbonyl-retentive and/or decarbonylative coupling. Based on several systematic reports, the reaction patterns of acyl fluorides are summarized in Scheme [4]. In the acyl couplings shown on the left-hand side of Scheme [4], various reaction modes are thus realized after oxidative addition, such as sequential transmetalation with organometallic reagents, activation of the C–H bond of the aromatic hydrocarbons, or even carbometalative migratory insertion of alkenes. On the other hand, as shown on the right-hand side of Scheme [4], the intermediate complexes can undergo decarbonylation followed by transmetalation with organometallic reagents and activation of the C–H bonds in aromatic hydrocarbons. Next, the various relevant reaction modes shown in Scheme [4] and other activation modes not listed in Scheme [4] will be introduced in detail.

Zoom Image
Scheme 4 Reaction types under transition metal catalysis

Early research dates back more than fifty years, with the transition-metal-catalyzed oxidative addition of acyl fluorides having already been discovered as early as 1967.[28] In that year, Olah’s group realized the intramolecular decarbonylation coupling reaction of acyl fluorides in o-xylene as the solvent using Wilkinson’s catalyst [RhCl(PPh3)3] to synthesize a series of fluoroarenes (Scheme [5a]). This protocol was re-evaluated 15 years later by Ehrenkaufer, who found that the products obtained in the above reaction were actually arenes, not fluoroaromatics (Scheme [5b]).[29] Regardless of whether the products are fluoroarenes or arenes, it is a convincing fact that the rhodium catalytic system can activate the C–F bond of acyl fluorides. However, there had been no speculation or evidence for the oxidative addition of C–F bonds of other acyl fluorides. It was not until a decade later that Fahey and Mahan’s group first demonstrated the oxidative addition of acyl fluorides by stoichiometric reactions using Ni(cod)(PEt3)2 and characterized the isolated organometallic complex. This work opened a new chapter in the transition-metal-catalyzed C–F bond activation of acyl fluorides (Scheme [5c]).[4a]

Zoom Image
Scheme 5 Early studies on transition-metal-catalyzed C–F bond activation of acyl fluorides
2.1

Acylation (Carbonyl-Retentive) Coupling Reactions

2.1.1

Nickel Catalysis

Since low-valent transition metals show great performance in the C–F bond activation of acyl fluorides, nickel catalysts were initially used in this field during the past two decades. Because of the small atomic radius and low electronegativity of fluorine, the occurrence of oxidative addition is very facile. It is considered that catalytic cross-coupling reactions are more likely to occur due to the ease of oxidative addition.[30] In 2004, Rovis and Zhang successfully employed a nickel catalyst and dihydrocarbylzinc reagents as coupling partners for acyl fluorides to synthesize a series of ketones, where the acyl fluorides acted as acylating reagents, accompanied by C–F bond activation of acyl fluorides (Scheme [6]).[31] Under mild conditions and with a wide range of substrates, this reaction was compatible not only with aromatic and aliphatic acyl fluorides, but also with alkyl and arylzinc reagents, which served to broaden the scope of subsequent research.

Until recently, however, nickel-catalyzed acyl coupling reactions of acyl fluorides had not been developed. In 2019, it was reported that the cross-electrophilic coupling reaction proceeded very smoothly due to the favorable process of reducing the electron density of the nickel catalyst. In their report, Shu et al. described the nickel-catalyzed reductive coupling reaction of acyl fluorides with vinyl triflates (Scheme [7]).[32] This was the first report of a nickel-catalyzed cross-electrophilic acylation with acyl fluorides, resulting in the synthesis of a series of enones, which are of significant importance in biologically active compounds with a wide range of activity. Furthermore, a possible reaction mechanism was proposed. Initially, oxidative addition of the vinyl triflate to nickel generates the nickel(II) intermediate A, which is subsequently subjected to single-electron reduction by manganese to form the more active intermediate Ni(I) complex B. Crucially, further oxidative addition by the acyl fluoride leads to the formation of Ni(III) species C, which is followed by reductive elimination to liberate the final enone products. Furthermore, the Ni(I) species formed after reductive elimination is further reduced by manganese to regenerate the initial Ni(0) catalyst.

Zoom Image
Scheme 6 Nickel-catalyzed cross-coupling reaction of acyl fluorides with organozinc reagents
Zoom Image
Scheme 7 Nickel-catalyzed cross-electrophilic coupling reactions of acyl fluorides with alkenyl triflates and the proposed catalytic cycle. a 2,2′:6′,2′′-Terpyridine (15 mol%) was used as the ligand.
Zoom Image
Scheme 8 Synergistic Ni/Cu-catalyzed hydroacylation of alkenes with acyl fluorides

In 2019, Sawamura et al. explored and developed the hydroacylation of alkenes using acyl fluorides and hydrosilanes (Scheme [8]).[33] Using acyl fluorides as the acyl source and hydrosilanes as the hydrogen source, this reaction protocol was achieved by a dual catalyst system combining nickel and copper. The electron-donating, sterically demanding, and rigid properties of the DCYPBz ligand make it effective for the hydroacylation of styrenes under mild conditions, and a rational reaction mechanism has been proposed. Each catalytic cycle operates separately via a synergistic Ni/Cu catalyst system. In the nickel-containing catalytic cycle, the Ni(0) species readily and oxidatively adds to the acyl fluoride to form the acyl–Ni(II)–F intermediate A. Whereas, in the copper catalytic cycle, hydrogen is delivered from the hydrosilane to give the key Cu–H intermediate B. Subsequently, the C–C double bond of the styrene is inserted into the active intermediate B to form the alkylcopper species C, which undergoes transmetalation with the intermediate A to generate the nickel complex D, whilst the copper catalyst is regenerated. Finally, the hydroacylation products are furnished and the nickel catalyst is regenerated by reductive elimination.

Along with the oxidative addition of acyl fluorides, the oxidation state of nickel in the catalytic process is also crucial. In addition to the common Ni(0)/Ni(II) catalytic cycle, the reaction can proceed through Ni(I)/Ni(III), Ni(0)/Ni(I)/Ni(III), and even a single-electron process (radical process) during the Ni(I) catalysis. In 2020, Weix et al. developed a new reaction mode, cross-electrophilic coupling, for the synthesis of ketones catalyzed by nickel (Scheme [9]).[34] This strategy employs N-alkylpyridinium salts as the alkyl source and acyl fluorides as the acyl source to synthesize a series of unsymmetrical ketones, which are often found in the molecular backbones of pharmaceuticals, with good functional-group tolerance. The reaction pathway was designed based on the redox potentials of acyl fluorides and N-alkylpyridinium salts. While N-alkylpyridinium salts rapidly form alkyl radicals by single-electron reduction, other coupling partners, which are usually easily reduced, may also generate radicals, thus reducing the reaction selectivity. Acyl fluorides were selected as good electrophiles for efficient cross-electrophilic coupling by matching with N-alkylpyridinium salts because they exhibit relatively high reduction potentials, are not easily reduced to generate free radicals, and react rapidly with nickel/terpyridine catalysts. Based on such a reaction design, after the N-alkylpyridinium salt is reduced, it undergoes single-electron transfer (SET) to release the alkyl radical B. On the other hand, the nickel complex A, formed after oxidative addition of the acyl fluoride, can rapidly combine with the alkyl radical B to produce unsymmetrical ketones.

Zoom Image
Scheme 9 Nickel-catalyzed cross-electrophilic coupling of acyl fluorides with alkylpyridinium salts. a Dtbbpy was used instead of L1.

Carbamoyl fluorides, which are analogs of acyl fluorides, have recently attracted the attention of scientists because similar reactions can be widely developed with reference to the transition-metal-catalyzed transformations of acyl fluorides. In 2020, Ye et al. developed a nickel-catalyzed enantioselective transformation of carbamoyl fluorides bearing alkene moieties and arylboronic acids accompanied by intramolecular cyclization, and succeeded in constructing a series of γ-lactams with all-carbon quaternary centers in 45–96% yields and 38–97% ee, and with good substrate tolerance (Scheme [10]).[35] Based on the experimental results, a possible reaction mechanism was proposed. Initially, the isomer on the same side as the alkene in the rotamer of the carbamoyl fluoride leads to favorable oxidative addition to nickel, resulting in Ni(II) complex A, followed by rapid intramolecular insertion of the alkene to yield Ni(II) complex B. Next, transmetalation occurs with the arylboronic acid to generate intermediate C, followed by reductive elimination to release the desired product and regenerate the nickel catalyst. Regarding the reaction mechanism, it is reasonable to assume that oxidative addition is initially followed by transmetalation and then the subsequent steps.

Zoom Image
Scheme 10 Nickel-catalyzed intramolecular cyclization and arylation of carbamoyl fluorides tethered with alkene moieties with arylboronic acids

Recently, Schoenebeck et al. ingeniously designed a nickel-catalyzed cross-coupling of N-CF3 carbamoyl fluorides with alkynylsilanes to overcome existing limitations, given the importance of N-CF3 amide compounds in pharmaceutical engineering and the total synthesis of natural products (Scheme [11]).[36] This reaction protocol realized the selective C–F bond activation of carbamoyl fluorides by a commercially available nickel catalyst under mild conditions in order to couple with alkynylsilanes while retaining the N-CF3 amide bonds. The introduction of a C–C triple bond into the molecule gives the products as key synthons, and further derivatization can be used to build a series of N-CF3 alkenylamides, oxindoles, and quinolones that were previously synthetically inaccessible.

Zoom Image
Scheme 11 Nickel-catalyzed cross-coupling of N-CF3 carbamoyl fluorides with alkynylsilanes

# 2.1.2

Palladium Catalysis

Palladium, similar to nickel, is a group 10 metal, and is also one of the key transition metals utilized for the C–F bond activation of acyl fluorides. As palladium has a larger atomic radius and slightly higher electronegativity than nickel, it is more amenable to reductive elimination. In addition, due to the high energy barrier to rotation of the Pd–C bond, palladium is prone to β-hydride elimination.[37]

In the last few years, palladium-catalyzed C–F bond activation of acyl fluorides has been extensively developed by chemists. For example, Ogiwara and Sakai explored and developed palladium-catalyzed cross-coupling reactions utilizing acyl fluorides. The sterically bulky properties of P(t-Bu)3 and the transmetalation ability of fluorine-substituted phenylsilanes have been used to achieve the Hiyama-type acylation, resulting in the successful synthesis of a series of unsymmetrical diaryl ketones (Scheme [12a]).[38] The same research group also achieved the effective syntheses of important diaryl ketones and enones by Suzuki–Miyaura-type coupling of acyl fluorides with boronic acids using a palladium catalyst system (Scheme [12b]). This process also exhibited good functional-group tolerance.[39]

Zoom Image
Scheme 12 (a) Palladium-catalyzed Hiyama-type coupling of acyl fluorides with phenylsilanes. (b) Palladium-catalyzed Suzuki–Miyaura-type coupling of acyl fluorides with boronic acids. a PCy3 was employed instead of P(4-MeOC6H4)3. b 2 equiv each of R′B(OH)2 and KF were used.

Subsequently, the same research group extended their studies to palladium-catalyzed cross-coupling reactions of acyl fluorides with other coupling partners. In the event, the C–H bond acylation of heteroaromatic hydrocarbons with acyl fluorides was realized for the first time in order to synthesize a series of diaryl ketones in good yields by using a synergistic palladium/copper catalyst system (Scheme [13]).[40] In this reaction, acyl fluorides maintain a good balance of reactivity and stability among various carboxylic acid derivatives. Meanwhile, it was proved that heterocyclic compounds such as oxazoles, thiazoles, benzoxazoles, benzothiazoles and benzimidazoles were compatible with this reaction system, and that aromatic and aliphatic acyl fluorides were also applicable.

Zoom Image
Scheme 13 Pd/Cu-catalyzed heteroaromatic C–H coupling with acyl fluorides

In addition to the palladium-catalyzed arylation of acyl fluorides, Ogiwara and Sakai also developed the hydrogenation of acyl fluorides by exploiting the good hydrogenation ability of hydrosilanes.[41] As summarized in Table 1, during the initial screening of the reaction conditions, the expected reductive hydrogenation product a as well as a small amount of the decarbonylative hydrogenation product b were detected. Subsequently, when Pd(OAc)2/PCy3 was chosen as the catalyst system for the almost complete elimination of product b, aldehyde a was obtained in a highly selective manner (entry 4). On the other hand, when a bidentate phosphine ligand was employed instead of a monodentate ligand, product b was unexpectedly found to be the main product (entry 7).

Considering the two reaction pathways observed, a possible reaction mechanism was proposed that can realize both reductive hydrogenation products—aldehydes and decarbonylative products—from acyl fluorides (Scheme [14]). Initially, the acyl fluoride undergoes oxidative addition to the Pd(0) catalyst to generate the acyl–Pd(II)–F intermediate A, which then follows a different reaction pathway depending on the ligand used. When the ligand is PCy3, the hydrosilane undergoes a transmetalation process with A to produce acyl–Pd(II)–H intermediate B, which after reductive elimination gives the aldehyde a. On the other hand, when the ligand is DCPE, a fast decarbonylation process forms the aryl–Pd(II)–F intermediate C, which undergoes transmetalation with the hydrosilane to yield aryl–Pd(II)–H species D, reductive elimination of which produces the arene b and regenerates the Pd(0) catalyst.

Table 1 Optimization of the Reaction Conditionsa

Entry

Ligand

Yield of a (%)

Yield of b (%)

Selectivity (a/b)

1

P(4-MeOC6H4)3

53

 9

85:15

2

JohnPhos

17

 1

94:6

3

PCy3

77

 2

97:3

4 b

PCy3

92

<1

99:1

5

dppBz

 7

46

13:87

6

DCPE

 3

82

 4:96

7 b

DCPE

7

84

8:92

a Reaction conditions: Acyl fluoride (0.2 mmol), Et3SiH (0.28 mmol), Pd(OAc)2 (5 mol%), ligand (15 mol% of P), toluene (0.2 mL), 100 °C, 20 h.

b Acyl fluoride (1 mmol), Pd(OAc)2 (2.5 mol%), ligand (7.5 mol% of P), toluene (1 mL).

Zoom Image
Scheme 14 Palladium-catalyzed hydrogenation or decarbonylative hydrogenation of acyl fluorides with hydrosilanes

In view of these results, a reasonable interpretation has been proposed in the corresponding kinetic studies, in which a carbonyl-retentive reaction is favorable via a coordinatively saturated metal center when the Pd/P(PCy3) ratio is 1:3 (Scheme [15a]). When the Pd/P(DCPE ) ratio is 1:2, decarbonylation is facilitated by the presence of a vacant coordination site on the metal center, and the extruded CO makes the intermediate more stable (Scheme [15b]).

Zoom Image
Scheme 15 Possible intermediates using PCy3 and DCPE

Subsequently, Ogiwara and Sakai developed the palladium-catalyzed intermolecular cyclization between acyl fluorides and norbornene derivatives (Scheme [16a]).[42] Optimization of the reaction system led to the development of a Pd(OAc)2/PCy3 catalytic system, and a suitable reaction protocol for (hetero)aroyl and alkenylated acyl fluorides was established. In this transformation, the C–F bond of the acyl fluoride is activated by palladium, followed by activation of the ortho C–H bond accompanied by CO rearrangement during the reaction process, leading to the synthesis of a series of polycyclic ketones in good yields. Very recently, the same research group found that analogous intramolecular cyclization reactions could be achieved with acyl fluorides bearing aryl moieties (Scheme [16b]).[43] Again, activation of the C–H bond at the ortho-position was a critical step in the reaction. Since HF is generated, the reaction proceeds through neutralization by the base to afford a series of fluorenone derivatives, and exhibits high functional-group compatibility.

Zoom Image
Scheme 16 (a) Palladium-catalyzed intermolecular cyclization of acyl fluorides with norbornenes. (b) Palladium-catalyzed intramolecular cyclization of acyl fluorides via intramolecular C–H bond activation.
Zoom Image
Scheme 17 Palladium-catalyzed fluorine exchange using acyl fluorides and anhydrides. a 19F NMR yields. b GC yield.

Furthermore, the same research group has also realized the palladium-catalyzed acyl group exchange between acyl fluorides and acid anhydrides (Scheme [17]).[44] In this reaction, reversible cleavage and formation of the C–F bond were key steps, and new types of acyl fluorides and acid anhydrides were generated in good yields and high selectivity by substituent recombination. This study provides a basis for opening up future research into the synthesis of various acyl fluorides as novel reagents.

Zoom Image
Scheme 18 Palladium-catalyzed decarbonylative trifluoromethylation of acyl fluorides with TMSCF3 and DFT calculations

#
# 2.2

Decarbonylative Reactions

In addition to the reaction modes described above, transition-metal-catalyzed decarbonylation of acyl compounds is also one of the unique classes of synthetic strategies for the construction of C–C and C–heteroatom bonds. Carboxylic acid derivatives, in particular, are key raw materials for various important natural products and pharmaceutical molecules because of their low cost and ready availability. Realization of the transition-metal-catalyzed decarbonylation of carboxylic acid derivatives will undoubtedly greatly expand the scope of electrophiles as reagents for alkylations and arylations.[45] During the development of innovative palladium-catalyzed decarbonylations of acyl fluorides, Schoenebeck’s research group reported pioneering work in 2018.[46] The trifluoromethyl group, which is extremely valuable in the pharmaceutical industry, was selected as the target functional group, whilst acyl fluorides were successfully incorporated into the optimized catalytic system. The use of XantPhos as the ligand enabled trifluoromethylation for the first time, and the presence of fluorine in the substrates allowed for more smoother transmetalation while effectively avoiding excessive transmetalation caused by the introduction of additional fluoride. On the basis of these fundamental findings, a series of trifluoromethylated compounds was obtained with excellent yields from a wide range of substrates, and detailed studies on the mechanism were elucidated (Scheme [18]). DFT calculations on the sequence of decarbonylation and transmetalation steps allowed a reasonable reaction mechanism to be proposed. Initially, oxidative addition of the acyl fluoride to the Pd/XantPhos complex produces the acyl–Pd(II)–F intermediate A, from which two possible pathways were put forward. Namely, for path A, decarbonylation initially occurs from intermediate A to form the aryl–Pd(II)–F intermediate B, which is followed by transmetalation of TESCF3 to generate the Pd(II) complex C. Finally, reductive elimination releases the trifluoromethylated product. In contrast, in path B, transmetalation of TESCF3 occurs first to generate the acyl–Pd(II)–CF3 intermediate B′, which undergoes decarbonylation to form the Pd(II) complex C. A final reductive elimination then gives the desired product. From detailed DFT studies, the Gibbs free energy difference of the transition state of the decarbonylation process in path A was calculated to be 27.3 kcal/mol, and that in path B was 17.4 kcal/mol, which led to the conclusion that path B was a more reasonable reaction pathway. Subsequently, acyl fluorides, a class of carboxylic acid derivatives with moderate stability and reactivity, have recently begun to be widely explored and utilized by chemists in transition-metal-catalyzed decarbonylative transformations. Thus, for such decarbonylative reactions, the similarities and differences between nickel and palladium catalysis are outlined below.

2.2.1

Nickel Catalysis

The construction of C–C and C–heteroatom bonds is an important research topic that has a significant influence on the development of synthetic organic chemistry. In recent years, reaction modes for the C–F bond activation of acyl fluorides catalyzed by transition metals have been continuously innovated and improved. Meanwhile, researchers have continued to explore the nickel-catalyzed decarbonylation of acyl fluorides as a new research field involving the construction of C–C and C–heteroatom bonds.

Zoom Image
Scheme 19 Nickel-catalyzed decarbonylative Suzuki–Miyaura coupling of acyl fluorides with boronic acids without a base

In 2018, Sanford’s group developed a Suzuki–Miyaura coupling reaction via decarbonylation of acyl fluorides. One of the major challenges of the Suzuki–Miyaura coupling is that the addition of an external base is necessary to promote transmetalation, which not only significantly limits the applicability of base-sensitive substrates, but also may promote side reactions following addition of the base. After careful studies, Sanford found that this challenge could be overcome by applying the nickel-catalyzed decarbonylative coupling of acyl fluorides with an array of boronic acids without a base (Scheme [19]).[47] Also, a reasonable mechanism has been proposed based on experimental results. First, oxidative addition of the acyl fluoride to the nickel catalyst forms the acyl–Ni(II)–F intermediate A, which is followed by a smooth decarbonylation process to give aryl–Ni(II)–F intermediate B. Further reaction with the boronic acid coupling partner results in the formation of the Ni(II) intermediate C as a result of a transmetalation process, and the final product is released by reductive elimination along with regeneration of the nickel catalyst.[48]

In addition, Sanford et al. conducted a detailed exploration of the reaction mechanism by utilizing stoichiometric experiments (Scheme [20]). The stoichiometric experiments were carried out using Ni(cod)2 and PCy3 as catalytic systems and benzoyl fluoride and 4-fluorophenylboronic acid as coupling partners. Firstly, the rapid oxidative addition of benzoyl fluoride with stoichiometric amounts of Ni(cod)2 and PCy3 at room temperature within 10 minutes formed Ni(II) intermediate D in 95% yield, which was followed by a decarbonylation process to give Ni(II) intermediate E in 90% yield. Subsequently, although Ni(II) intermediate F, which would be produced by transmetalation between E and 4-fluorophenylboronic acid, could not be detected, the final coupled product was successfully obtained in 90% yield. Therefore, the aryl–Ni(II)–F intermediate E was assumed to exhibit a ‘transmetalation-active’ property that negates the requirement for addition of an external base.

Zoom Image
Scheme 20 Mechanistic studies

In 2018, we also developed a decarbonylative ethylation via Suzuki–Miyaura coupling of acyl fluorides (Scheme [21]).[49]

Zoom Image
Scheme 21 Nickel-catalyzed decarbonylative ethylation of acyl fluorides with alkylboranes

The selection of triethylborane and methylboroxine as coupling partners for acyl fluorides allowed the ethylation and methylation of aromatic compounds to be achieved in good yields (Scheme [21] and Equation 1). The borane acts as a Lewis acid as well as a coupling partner, and it is believed that its acidity can accelerate the smooth progress of the transmetalation process. In this ethylation reaction, the addition of an external base, which is usually required in Suzuki­–Miyaura couplings, was unnecessary.

Zoom Image
Equation 1 Nickel-catalyzed decarbonylative methylation of acyl fluorides with methylboroxine

In 2021, we successfully disclosed the nickel-catalyzed Sonogashira–Hagihara-type cross-coupling reaction of acyl fluorides with terminal alkynes via decarbonylation (Scheme [22]).[50] This study not only achieved a copper-free system, but also enabled the construction of C(sp2)–C(sp) bonds for the synthesis of a series of internal alkynes in good yields. Subsequently, relevant explorations of the substrate scope were carried out, including investigations on the electronic effects of the substituents on the acyl fluorides and with different terminal silylethynes. The results indicated that the substrates were all compatible with the developed catalytic system and gave good yields and selectivity.

Zoom Image
Scheme 22 Nickel-catalyzed decarbonylative Sonogashira–Hagihara coupling of acyl fluorides with terminal alkynes

Concurrently, we have reported on the development of nickel-catalyzed decarbonylative C–heteroatom bond-forming reactions using acyl fluorides in addition to C–C bond-forming transformations. Since 2018, we have sequentially developed decarbonylative C–B (Scheme [23a]),[51] C–Sn (Scheme [23b]),[52] C–Si (Scheme [23c]),[53] and C–S (Scheme [23d])[54] bond-forming reactions with acyl fluorides. In these reactions, diboronates as boron sources, Bu3SnTMS as a tin source, silylboronates as silicon sources, and thiols (thiophenols) as sulfur sources can be utilized to construct important heteroatom-containing compounds in good yields. Once again, these protocols prove that acyl fluorides are continuously being developed and applied as versatile coupling reagents.

Zoom Image
Scheme 23 Nickel-catalyzed decarbonylative coupling reactions of acyl fluorides with diboronates, Bu3SnTMS, silylboronates, and thiols (thiophenols) to construct C–B, C–Si, C–Sn, and C–S bonds. a B2neo2 is used instead of B2pin2, then pinacol (4 equiv) and NEt3 (0.5 mL) at room temperature for 1 h.

# 2.2.2

Palladium Catalysis

Besides nickel, palladium catalytic systems have also been explored by many chemists in the past few years for their compatibility with decarbonylative cross-coupling reactions of acyl fluorides. In particular, due to the unique properties of zero-valent palladium, the C–F bond of acyl fluorides is activated as if infused with ‘magical powers’, allowing for the formation of a wide range of C–C and C-heteroatom bonds through decarbonylation. This sub-section focuses on the palladium-catalyzed decarbonylative coupling reactions of acyl fluorides from two perspectives: the construction of C–C and C–heteroatom bonds.

Recently, our group developed a novel protocol for the decarbonylative alkylation of acyl fluorides with a wide variety of alkylboron compounds by utilizing a palladium-catalyzed system (Scheme [24]).[55] This reaction tolerated various functional groups and allowed the construction of a series of compounds containing C(sp2)–C(sp3) bonds, which again greatly expanded the range of substrates containing sp3 carbons. In particular, the realization of long-chain alkylation of arenes is of great value in organic synthesis, and in a detailed search for substrates, it was found that alkylboranes could efficiently introduce various long-chain alkyl groups on the sp2 carbons of acyl fluorides. In addition, it was observed that the electronic effects of acyl fluorides had no significant impact on the reaction, and all the examined substrates were compatible with the palladium catalytic system. In addition, a Pd/DPPP system proved to have better catalytic activity than nickel.

Zoom Image
Scheme 24 Palladium-catalyzed decarbonylative alkylation of acyl fluorides with alkylboranes
Zoom Image
Scheme 25 Pd/Cu-catalyzed decarbonylative alkynylation of acyl fluorides with alkynylsilanes and the proposed mechanism

In the same year, we also developed the decarbonylative alkynylation of acyl fluorides with alkynylsilanes using a dual Pd/Cu catalytic system. This reaction tolerated a wide range of functional groups and was utilized to successfully construct C(sp2)–C(sp) bonds during the preparation of a series of internal alkynes (Scheme [25]).[56] The Si–F bond formation is the driving force, reducing the energy barrier for transmetalation of alkynylsilanes, whilst the synergistic effect caused by the copper catalyst allows the reaction to proceed more smoothly. The reaction mechanism of this alkynylation consists of two catalytic cycles, which are postulated as follows. Firstly, oxidative addition of the acyl fluoride to Pd(0) occurs to form the acyl–Pd(II)–F intermediate A. Concurrently, during the copper catalytic cycle, Si–F bond formation is the key to the reaction with the alkynylsilane, enabling formation of the alkynylcopper intermediate B, which is then incorporated into the palladium catalytic cycle to produce Pd(II) complex C via a transmetalation process with intermediate A. Next, decarbonylation produces intermediate D, which is subjected to reductive elimination to release the final alkynylated products and regenerate the Pd(0) catalyst.

Through mechanistic speculation and analysis, we found that the introduction of copper catalysts can effectively control the selectivity of C–H or C–Si bond cleavage in alkynes. In Scheme 22, we developed the decarbonylative Sonogashira–Hagihara coupling reaction mode for acyl fluorides using a single nickel catalyst, and showed that the coupling reaction occurred efficiently at the C–H bond of alkynylsilanes. The Pd/Cu synergistic catalytic system can be used for efficient coupling reactions at the C–Si bond position of alkynylsilanes. This will undoubtedly provide new research ideas for transition-metal-catalyzed coupling reaction modes for different C(sp)–X bonds.

In 2020, Tobisu and Sakurai made a unique extension of the palladium-catalyzed decarbonylative coupling reaction of acyl fluorides. They used a Pd/XantPhos catalyst as the optimized system and TMSCN as a readily available cyanation reagent to realize the decarbonylative cyanation of acyl fluorides (Scheme [26]).[57] Thus, a series of nitriles was synthesized with good selectivity and high yields, broadening the possibilities of C(sp2/sp3)–C(sp) bond construction.

Zoom Image
Scheme 26 Palladium-catalyzed decarbonylative cyanation of acyl fluorides with TMSCN

Given the importance of fluoroalkyl groups in the field of biopharmaceuticals, in addition to Schoenebeck’s research on the decarbonylative trifluoromethylation of acyl fluorides,[46] Sanford et al. have recently developed a palladium-catalyzed decarbonylation coupling of difluoroacetyl fluoride as the substrate (Scheme [27]).[58] Arylboronic esters were employed as coupling partners because of their superior transmetalation capabilities, and a wide variety of difluoroalkylated arenes were synthesized in excellent yields. Furthermore, stoichiometric experiments were performed for each sub-step of the reaction to demonstrate its feasibility, and possible mechanisms were proposed based on these studies. Initially, difluoroacetyl fluoride undergoes oxidative addition to palladium to yield the acyl–Pd(II)–F intermediate A, and subsequently, a carbonyl de-insertion reaction occurs preferentially on the intermediate A, resulting in the difluoromethyl–Pd(II)–F intermediate B. The presence of the crucial intermediate B in this reaction was detected by 19F NMR measurements. The detection of intermediate B indicates that decarbonylation occurs prior to transmetalation, and also suggests that this intermediate can act as a ‘transmetalation-active’ complex, achieving the coupling reaction without the addition of a base. Next, transmetalation is facilitated between the arylboronate and intermediate B to form the Pd(II) intermediate C, and reductive elimination releases the final difluoromethylation product and regenerates the Pd(0) catalyst.

Zoom Image
Scheme 27 Palladium-catalyzed decarbonylative arylation of difluoroacetyl fluorides with arylboronates

In addition to the described palladium-catalyzed bimolecular decarbonylative coupling of acyl fluorides, we have very recently reported an unprecedented three-component decarbonylative coupling of acyl fluorides with internal alkynes and silylboranes in a synergistic Pd/Cu bicatalytic system, yielding a wide variety of tetrasubstituted alkenylsilanes with good stereoselectivity (Scheme [28]).[59] This protocol proved that acyl fluorides, with their excellent functional-group tolerance, can serve as alternatives to aryl halides as electrophiles. Based on various reference results, a possible reaction mechanism has been proposed as follows. Initially, oxidative addition of the acyl fluoride proceeds on the Pd(0) catalyst to form the acyl–Pd(II)–F intermediate A. Next, the silylborane interacts with the copper catalyst to form the silylcopper intermediate C, followed by insertion of the alkyne to form the alkenylcopper intermediate D, which undergoes transmetalation with intermediate A to produce alkenylpalladium intermediate E. Sequential decarbonylation and CO extrusion then occur to give the Pd(II) complex G. Finally, reductive elimination yields the coupled, tetrasubstituted alkenylsilane derivatives. The assumed reaction mechanism is also consistent with the fact that the ratio of (E)-stereoisomers increases when acyl fluorides with electron-withdrawing groups are employed.

Zoom Image
Scheme 28 Pd/Cu-catalyzed three-component decarbonylative coupling of acyl fluorides with internal alkynes and silylboranes

This isomerization does not originate from the alkenylcopper intermediate D, but is presumably the result of the slow conversion of the alkenylpalladium species E into intermediate F, which also leads to the formation of intermediate E′, via the species H, to finally give the (E)-product.

In the above work, examples of palladium-catalyzed C–C bond-forming reactions have been mainly introduced. Compared to the versatile functions of nickel in the field of C–heteroatom bond construction by decarbonylative coupling of acyl fluorides, related research has been explored in depth under palladium catalysis, but to date, the number of relevant reports are still limited. In 2019, Zhang and co-workers reported a Pd/DCYPB catalytic system that could effectively manipulate the decarbonylative coupling of acyl fluorides in the presence of distannanes to construct a series of arylstannanes; the C–Sn bonds were successfully formed without an exogenous base (Scheme [29]).[60] The reaction protocol was extensively explored with a wide substrate scope and a series of significant derivatization studies were also performed. It was found that a variety of functional groups were compatible, that the electronic effects of substituents had little influence, and most importantly, that the procedure proved to be particularly suitable for the modification of important pharmaceutical molecules due to the good yields obtained.

Zoom Image
Scheme 29 Palladium-catalyzed decarbonylative coupling of acyl fluorides with distannanes

#
# 2.3

C–F Bond Activation by Other Transition Metals

While the C–F bond activation of acyl fluorides by nickel and palladium catalyst systems have been developed and proven to demonstrate good compatibility, other transition metals have also been found to be effective in the C–F bond activation of acyl fluorides. In recent years, scientists have discovered that transition metals such as iridium, copper, rhodium, and cobalt are capable of activating the C–F bond in acyl fluorides, thereby confirming that acyl fluorides are quite effective reagents as alternative and practical carbon or fluorine sources. We next discuss the applications of different transition-metal catalyst systems in the C–F bond activation of acyl fluorides from two aspects: (1) acyl fluorides as a carbon source, and (2) acyl fluorides as a fluorine source.

2.3.1

Acyl Fluorides as a Carbon Source

While investigating arylation and alkylation using acyl fluorides in various nickel and palladium-catalyzed systems, in 2017, Riant et al. developed a three-component coupling reaction utilizing acyl fluorides in a copper-catalyzed system (Scheme [30]).[61] In this reaction, acyl fluorides and B2pin2 were employed as acylating and boronating reagents, respectively, to combine with allenes to afford a series of deconjugated enones bearing quaternary carbon centers in good yields and with a wide substrate scope. Follow-up mechanistic studies suggest that acyl fluorides, being a suitable electrophile, can rapidly capture the allylcopper intermediate to form a six-membered transition state, which proves to be the key step.

Zoom Image
Scheme 30 Cu-catalyzed three-component acylborylation of allenes with acyl fluorides and B2pin2

Based on the results of various additional experiments, a possible reaction mechanism was given (Scheme [31]). Initially, Cu(OAc)2 is reduced by TMSONa and dppf to form Cu(I) species A. Additionally, TMSONa reacts with B2pin2 to produce the borate B, which then reacts with A to afford the [Cu]–Bpin intermediate D. The allene substrate inserts into intermediate D to give the allylcopper intermediate E. The acyl fluoride and intermediate E then form a highly organized six-membered transition state F, which eventually releases the difunctionalized product and regenerates the active Cu(I) catalyst A.

Zoom Image
Scheme 31 A proposed mechanism for the reaction shown in Scheme [30]
Zoom Image
Scheme 32 Iridium-catalyzed decarbonylative coupling of acyl fluorides with arenes, and the possible transition state or intermediate

In 2018, Tobisu and co-workers developed the iridium-catalyzed cross-coupling reaction of acyl fluorides via C–H bond activation of arenes. Employing Ir/BrettPhos as a catalyst system, cross-coupling involving C–H bond cleavage of arenes and decarbonylation of acyl fluorides has been successfully demonstrated to construct various important aromatic frameworks (Scheme [32]).[62] A feasible reaction mechanism based on experimental results has been put forward. It is thought that the fluorine atom in acyl fluorides acts as a key component and promotes concerted metalation–deprotonation (CMD) of the benzene ring by the Ir catalyst, which facilitates the formation of transition state A. A related description of the C–H bond activation of nitrogen-containing heterocycles also applies the concept of this reaction. It is believed that addition of the Ir–aryl bond can proceed smoothly to the C=N bond of the heteroarene to form intermediate B, and the final product is obtained through aromatization.

Very recently, Su et al. developed the rhodium-catalyzed ortho-C–H functionalization of (hetero)aromatics with 8-quinolylamine as a bidentate directing group that proceeded via the in situ preparation of a series of acyl fluorides from the corresponding carboxylic acids (Scheme [33]).[63] In addition, the substrate scope, which included a wide variety of pharmaceutical molecules, was extensively explored and a possible reaction mechanism was deduced (Scheme [34]).

Zoom Image
Scheme 33 Rhodium-catalyzed decarbonylative arylation of acyl fluorides, formed in situ from carboxylic acids, with arenes tethered with a directing group

Initially, the bidentate directing group (8-aminoquinolylamido) coordinates to the rhodium center, which simultaneously activates the C–F bond of the in situ generated acyl fluoride during the process of oxidative addition to form Rh(III) complex A, from which a decarbonylation process occurs to form the Rh(III) complex C through CO extrusion. Subsequently, intramolecular C–H bond activation in the intermediate C and neutralization of HF by the base present in the system generated Rh(III) intermediate D. Finally, reductive elimination releases the product and regenerates the Rh(I) catalyst.

Zoom Image
Scheme 34 A proposed mechanism for the reaction shown in Scheme [33]

# 2.3.2

Acyl Fluorides as a Fluorine Source

Fluorine has attracted considerable attention from the scientific community, and research on organofluorine chemistry has become of increasing important to chemists engaged in organic synthesis.[64] Notably, fluorine atoms are found in 30–40% of pesticides and in 20% of pharmaceutical molecules.[65] Due to the unique properties of fluorine atoms, their introduction into organic molecules can improve lipophilicity and solubility, and reduce polarity and intermolecular interactions.[66] Therefore, organofluorine compounds have been studied in a wide variety of fields, including science and technology. In this research environment, the installation of fluorine atoms into organic molecules has attracted significant attention.[65f] [67] Currently, there are two common routes to fluorination: nucleophilic and electrophilic fluorination. However, nucleophilic fluorinating reagents (fluoride salts) require harsh reaction conditions, while electrophilic fluorinating reagents are generally expensive. Each of these drawbacks makes it difficult to achieve major breakthroughs in fluorination reactions. As such, scientists have been constantly trying to develop new fluorinating reagents to overcome the existing problems. In this context, acyl fluorides can be regarded as a new type of inexpensive and readily available fluorinating reagent, and reactions using acyl fluorides as the fluorine source are discussed below.

In 2010, Doyle’s group developed a dual catalyst system combining a chiral salen cobalt complex and an amine, and used benzoyl fluoride as a fluorinating reagent to achieve the asymmetric nucleophilic fluorination of epoxides (Scheme [35a]).[68] To date, the known asymmetric nucleophilic ring-opening reactions of epoxides usually take place in the presence of HF-containing reagents, such as Olah’s reagent. Although such reagents are considered to be a class of nucleophilic fluorinating reagents that can remarkably enable the asymmetric ring-opening of epoxides, these reactions in conjunction with chiral Lewis acids suffer from competing background pathways and catalyst inhibition. In this reaction, benzoyl fluoride was used to form a new HF-containing reagent, which completely overcomes the conventional challenges described above to efficiently achieve the asymmetric fluorination of epoxides.

Zoom Image
Scheme 35 (a) Cobalt-catalyzed asymmetric ring-opening fluorination of epoxides with benzoyl fluoride. (b) Cobalt-catalyzed asymmetric ring-opening fluorination of aziridines with benzoyl fluoride.

Subsequently, in 2013, the same research group utilized the above strategy to realize asymmetric fluorination with ring-opening of aziridines, and successfully prepared a series of β-fluoroamines through the desymmetrization of meso-aziridines (Scheme [35b]).[69] In this protocol, the same chiral (salen)Co catalyst was combined with a Ti(IV) complex as a Lewis acid to generate analogous HF-containing reagents from benzoyl fluoride and HFIP, achieving asymmetric fluorination of N-picolinamide aziridines. A possible mechanism has been proposed for these asymmetric fluorination reactions (Scheme [36]). Firstly, benzoyl fluoride and HFIP react to form hexafluoroisopropyl benzoate B, whilst the chiral cobalt catalyst captures a molecule of HF to give the cobalt complex C. Meanwhile, the Lewis acid interacts with the aziridine (containing picolinamide as a protecting group) to form the complex D, which reacts with Co complex C to give the fluorinated product. Simultaneously, both the cobalt catalyst A and the Lewis acid are regenerated.[70]

Zoom Image
Scheme 36 A proposed mechanism for the reaction shown in Scheme [35b]

#
#
# 3

C–F Bond Activation by N-Heterocyclic Carbenes (NHCs)

The pioneering studies of Bertrand[71] and Arduengo[72] on stabilized carbene chemistry has provided a wealth of research interest and numerous possibilities for subsequent related work. From the advent of polar inversions of aldehydes, such as the benzoin condensation[73] and the Stetter reaction,[74] to the later Breslow polar inversion intermediates[75] and their extensive applications, carbenes have played an important role in modern synthetic organic chemistry. Due to the unique properties of N-heterocyclic carbenes (NHCs), such as their ability to function as neutral two-electron ligands and their steric bulkiness, the proportion of related research on this topic has increased. In addition to their use as important transition-metal ligands in the field of catalysis, synthetic organic reactions using NHCs as organocatalysts emerged in 2004, and they now attract significant attention as new catalysts with practical value. NHCs can form discrete reactive intermediates with many basic organic molecules, for example, acyl anions, homoenolates and enolate equivalents, which are usually generated by oxidation state reorganization, and undergo a series of diverse transformations.[76]

In recent years, activation of the C–F bonds of acyl fluorides catalyzed by NHCs has started to be explored. As a result of previous studies, it is known that the C–F bond of acyl fluorides can be readily activated in NHC-catalyzed systems to form acyl azolium intermediates, which can then undergo three reaction modes: (1) cycloaddition reactions, (2) single-electron transfer in a photoredox catalytic system, and (3) nucleophilic aromatic fluorination (Scheme [37]). These three reaction modes will be introduced in the following sub-sections.

Zoom Image
Scheme 37 The three reaction modes of the C–F bonds in acyl fluorides catalyzed by N-heterocyclic carbenes (NHCs)
3.1

NHC-Catalyzed Cycloaddition of Acyl Fluorides

Lupton’s group pioneered the establishment of the research field of NHC-catalyzed C–F bond activation of acyl fluorides. For the past 15 years, this group has focused on the cycloaddition of α,β-unsaturated acyl fluorides with enols catalyzed by NHCs, and has discovered that the acyl azolium intermediates formed from acyl fluorides in NHC catalysis have various functions. In 2009, they employed a series of disubstituted silyl enol ethers as conjugate acceptors in the first conjugate addition with acyl azolium intermediates formed from acyl fluorides and an NHC to synthesize various dihydropyranones, which are known as medicinal agents and are useful building blocks in organic synthesis (Scheme [38a]).[77] Subsequently, in 2011, they used a series of silyl dienol ethers and α,β-unsaturated acyl fluorides to achieve synergistic [4+2] cycloaddition under the catalysis of an NHC, synthesizing a variety of 1,3-cyclohexadienes in good yields and with excellent diastereoselectivity in the decarboxylation step (Scheme [38b]).[78] This synergistic approach makes the reaction more inclined toward the endo-orientation mode, and the protocol expands on ideas and directions for further research of nucleophilic organocatalysts. Subsequent mechanistic studies revealed that the final product is obtained by isomerization via a retro-aldol/aldol sequence, lactonization, and then decarboxylation. The specific sequence is shown in Scheme [39]. Initially, reaction of the α,β-unsaturated acyl fluoride with the silyl dienol ether is catalyzed by the NHC to form α,β-unsaturated acyl azolium species A and TMS–F is removed to produce enol intermediate B. Next, intermediate B undergoes [4+2] cycloaddition with intermediate A to generate cyclohexene intermediate C. Next, intramolecular nucleophilic addition forms β-lactone D and regenerates the NHC catalyst, and finally decarboxylation affords the cyclohexadiene product.

Zoom Image
Scheme 38 (a) NHC-catalyzed cycloaddition of α,β-unsaturated acyl fluorides with silyl enol ethers. (b) NHC-catalyzed cycloaddition and decarboxylation of α,β-unsaturated acyl fluoride tethered conjugated dienes with silyl enol ethers. (c) NHC-catalyzed cycloaddition of α,β-unsaturated acyl fluorides with conjugated silyl enol ethers. (d) NHC-catalyzed cycloaddition and decarboxylation of α,β-unsaturated acid fluorides with silyl enol ethers.

Furthermore, in 2018, Lupton et al. found that this unique extension of the cycloaddition reaction yielded different cyclized frameworks depending on the substrate. In this reaction, a key dienyl acyl azolium intermediate was prepared from dienyl acyl fluorides. When combined with enols, the reaction system was found to undergo direct [4+2] cycloaddition without decarboxylation, resulting in the synthesis of a series of bi-, tri-, and tetracyclic β-lactones with high stereoselectivity (Scheme [38c]).[79] Lupton has also reported the NHC-catalyzed difunctionalization of silyl enol ethers with α,β-unsaturated acyl fluorides. This protocol realized the first β-lactonization via cyclization and enabled construction of a series of indene skeletons through decarboxylation (Scheme [38d]).[80]

In fact, α,β-unsaturated acyl azoliums generated by NHCs exhibit versatile electrophilicity. In 2013, Lupton and co-workers extended the NHC-catalyzed cycloaddition of α,β-unsaturated acyl azoliums by replacing silyl enol ethers, which have been widely used as enol donors, with cyclopropyl silyl ethers (Scheme [40]).[81] The development of a new Claisen rearrangement reaction with Ireland–Coates hybrid structures has enabled the synthesis of a series of multifunctionalized β-butyrolactone-fused cyclopentanes with good diastereoselectivity. The reaction mechanism assumed from the experimental data is as follows. First, the NHC-catalyzed reaction of the α,β-unsaturated acyl fluoride with the cyclopropylsilyl ether forms α,β-unsaturated acyl azolium A and enolate intermediate B accompanied with elimination of TMS–F. Next, nucleophilic addition of B to A produces the hemiacetal intermediate C, which is followed by a turnover-limiting Ireland–Coates Claisen rearrangement to yield intermediate D. Finally, aldol cyclization and lactonization take place to afford the final product.

Zoom Image
Scheme 39 A proposed mechanism for the reaction shown in Scheme [38b]

On the basis of these results, Lupton recently conducted a more comprehensive study of the reaction mode of α,β-unsaturated acyl azoliums with nucleophiles. As a result, important fused butyrolactone or lactam skeletons were synthesized, with high enantioselectivity and without using exogenous oxidants and/or bases, by applying 1,3-diketones, nitrile esters, or hydrazones in the NHC catalytic system, together with acyl azoliums or acyl azolium enolates (Schemes 41a–d).[82] This reaction design has generality and realized excellent enantioselectivity, mainly due to its incompatibility with Breslow intermediates.

Zoom Image
Scheme 40 NHC-catalyzed cycloaddition of α,β-unsaturated acyl fluorides with cyclopropyl silyl ethers
Zoom Image
Scheme 41 NHC-catalyzed cycloaddition of α,β-unsaturated acyl fluorides with 1,3-diketones, nitrile esters, or hydrazones

A possible mechanism for this transformation is shown in Scheme [42]. First, an active NHC·C4 species is formed in the presence of KHMDS and reacts with the α,β-unsaturated acyl fluoride to give the acyl azolium intermediate A, which further reacts with the in situ formed HMDS (B) to generate the silyl amide anion C, accompanied by elimination of TMS–F. The anion C serves as a base, and deprotonation of the 1,3-diketone results in the formation of enolate D, which undergoes Michael addition with the acyl azolium species A to form the intermediate E. This then smoothly isomerizes into F. Finally, intramolecular cyclization of F yields the product and regenerates the NHC catalyst.

Zoom Image
Scheme 42 A proposed mechanism for the reaction shown in Scheme [41a]

# 3.2

NHC-Catalyzed Radical Functionalization of Acyl Fluorides

Furthermore, the acyl azolium intermediates generated by NHC catalysis can be combined with a photoredox catalytic system to form ketyl radicals through a novel single-electron reduction process, which enables subsequent coupling with a wide variety of radicals. In 2020, Studer et al. developed a radical three-component acyltrifluoromethylation of alkenes by using an NHC/photoredox dual catalyst system (Scheme [43a]).[83] This reaction utilizes acyl fluorides as the acyl source and sodium trifluoromethanesulfonate as the trifluoromethyl source to prepare a series of important β-trifluoromethyl-α-substituted ketones with excellent functional-group tolerance under mild conditions. Acyl azolium intermediates formed by NHC catalysis and acyl fluorides can be reduced by a single electron in a photoredox system to produce ketyl radicals, which are completely different from those produced by oxidation of Breslow intermediates. Moreover, this study has established a redox-neutral system that requires neither an external oxidant nor a reductant, in which single-electron reduction produces the ketyl radical, and single-electron oxidation forms the trifluoromethyl radicals, thereby opening a new avenue for NHC-catalyzed C–F activation of acyl fluorides.

Zoom Image
Scheme 43 (a) NHC/photoredox dual-catalyzed difunctionalization of alkenes with acyl fluorides and sodium trifluoromethanesulfonate (CF3SO2Na). (b) NHCs/photoredox/sulfonate tri-catalyzed acylation of alkenes with acyl fluorides.

Very recently, Studer’s group developed a tri-catalyst system that included NHC catalysis, photoredox catalysis, and sulfinate catalysis, using acyl fluorides as extremely reliable substrates (Scheme [43b]).[84] The acyl azolium intermediate formed by the NHC catalyst can generate ketyl radicals in this reaction system. When an alkene was added to the reaction system, α-acylation of the alkene unexpectedly proceeded to yield a series of α-substituted vinyl ketones. Notably, this is in contrast to the existing acylation of styrene derivatives, which usually affords β-acylation products. It was also found in this reaction mode that the sulfinyl radical formed by the single-electron oxidation of sulfinate can act as a traceless species, which can be captured by the ketyl radical after addition to the alkene to form a radical intermediate. The specific process is shown in Scheme [44]. Firstly, the photoredox catalyst generates the excited state A in the presence of light, and then sulfinic acid anion C is subjected to single-electron oxidation to form the sulfinyl radical D, while photoredox catalyst A is reduced to the intermediate B. Subsequently, the addition of sulfinyl radical D to the alkene occurs to form the radical E, while in the presence of the NHC catalyst, the acyl azolium intermediate F generated from the acyl fluoride is subjected to single-electron reduction by the photoredox catalyst B to produce ketyl radical G. The generated radical G couples with the radical E to give the precursor H and regenerates the NHC catalyst. Finally, sulfinic acid is eliminated to give the α-substituted vinyl ketone product.

Zoom Image
Scheme 44 A proposed mechanism for the reaction shown in Scheme [43b]

Furthermore, Studer et al. have also developed a direct C–H acylation with acyl fluorides by utilizing the synergistic effect of an NHC and photocatalysis (Scheme [45a]).[85] Because this method exhibits good functional group compatibility and diastereoselectivity, a variety of benzylic acylated products, which are often used as the building blocks of important natural products or pharmaceutical molecules, can also be synthesized. The following possibilities were proposed for the mechanism of this reaction. The photoredox catalyst becomes the excited state A upon photoexcitation, and single-electron oxidation of the substrate forms the radical cation C, which reduces the photoredox catalyst to give B. Meanwhile, deprotonation of the resulting radical cation C gives rise to the benzyl radical D. The acyl azolium intermediate E, formed from reaction of the acyl fluoride with the NHC catalyst, is reduced by photocatalyst B via single-electron transfer to generate the ketyl radical F and regenerate the photoredox catalyst. The generated radicals D and F then undergo coupling to yield the final product and regenerate the NHC catalyst.

Zoom Image
Scheme 45 (a) NHC/photoredox dual-catalyzed acylation of benzylic C–H bonds with acyl fluorides. (b) NHCs/photoredox dual-catalyzed ring-opening and 1,3-difunctionalization of cyclopropanes with acyl fluorides.

Recently, the same research group also reported the manipulation of ketyl radicals using a dual NHC/photoredox catalyst system and acyl azolium intermediates formed from acyl fluorides in a single-electron reduction strategy. During this procedure a cyclopropyl radical cation is generated by single-electron oxidation of a cyclopropane, which is then ring-opened by nucleophilic attack. Subsequent radical–radical coupling then gives the γ-aroyloxy ketone products in good to excellent yields with excellent functional group tolerance (Scheme [45b]).[86] The mechanistic elucidation also confirmed that the nucleophile attacking the cyclopropane ring is the aromatic acid anion formed after decarboxylation of the carbonate intermediate generated from the acyl fluoride and carbonate ions.

Very recently, Studer’s group once again broadened the range of applications of acyl fluorides. They utilized their previously reported NHC/photoredox catalyst system to manipulate aromatic acyl fluorides as bifunctional reagents to achieve the 2,3-fluoroaroylation of benzofurans or indoles accompanied by dearomatization.[87] Based on a series of mechanistic studies, they believe that this reaction involves an unusual radical–radical cation coupling process. A possible reaction mechanism has been given as follows. Initially, irradiation of PC (photoredox catalyst) generates the excited state catalyst PC*, the strong oxidative properties of which results in single-electron oxidation to benzofuran (E1/2= 1.28 V vs SCE), forming radical cation intermediate A and a PC radical anion [E1/2(P/P•–) = –1.37 V vs SCE]. At the same time, the NHC catalyst reacts with acyl fluoride to form acyl azolium ion intermediate B (E1/2 = –1.29 V vs SCE), and single-electron reduction of this intermediate B generates ketyl radical intermediate C and regenerates PC. The radical C then couples with the radical cation B to form intermediate D, which is followed by nucleophilic addition of the fluoride anion to generate the final product and regenerate the NHC catalyst (Scheme [46]).

Zoom Image
Scheme 46 Dearomatizing 2,3-fluoroaroylation of benzofurans with aroyl fluorides as bifunctional reagents for both C–C and C–F bond formation via cooperative NHC/photoredox catalysis

# 3.3

NHC-Catalyzed Nucleophilic Fluorination of (Hetero)aromatics

By employing NHC catalyst systems, acyl azolium intermediates formed from acyl fluorides may not only undergo subsequent cycloaddition or single-electron reduction to form ketyl radicals, but serve as a potential fluorine source for aromatic radicals. In 2015, Sanford et al. developed an NHC-catalyzed nucleophilic fluorination of electron-deficient (hetero)aryl chlorides by employing acyl fluorides as the fluorine source (Scheme [47]).[88] This protocol enables nucleophilic fluorination at room temperature, and this novel fluorination mode is more broadly applicable to methods requiring anhydrous fluorinating agents.

Zoom Image
Scheme 47 NHC-catalyzed SNAr fluorination of aryl chlorides with acyl fluorides. a PhCOF is used. b 4-MeOC6H4COF is used. c Acyl fluoride (4 equiv) is used.

#
# 4

C–F Bond Activation by Phosphines

Organophosphorus compounds have played an important role in organic synthesis. For example, the phosphine-mediated Wittig[89] and Mitsunobu[90] reactions have been widely utilized in synthetic organic chemistry. To date, the activation and formation of various chemical bonds catalyzed by phosphines have been explored and developed by chemists. For example, in 1968, the Morita–Baylis–Hillman reaction catalyzed by a phosphine was discovered, in which the key intermediate, a Horner zwitterion, turned out to be amphiphilic.[91]

Zoom Image
Scheme 48 Direct and indirect activation of C–F bonds in acyl fluorides

Besides, a number of synthetic organic reactions using organophosphorus compounds as catalysts have been developed in the past few decades.[92] In recent years, the activation of C–F bonds catalyzed by phosphines has gained interest from scientists, especially transformations of the C–F bonds in acyl fluorides, which have been explored intensively, leading to several breakthroughs. Currently, relevant reports mainly focus on two types of reaction. One is the phosphine-catalyzed direct activation of the C–F bond of acyl fluorides, and the other is the indirect activation of alkenylphosphine intermediates formed by nucleophilic addition of phosphines to electron-deficient alkynes (Scheme [48]). Herein, several examples of these two types of reaction will be presented.

4.1

Phosphine-Catalyzed Direct Activation of the C–F Bond of Acyl Fluorides

The direct activation of C–F bonds by phosphines has been pioneered by our research group. We found that organophosphine compounds can act as a Lewis base to form onium ions with acyl fluorides for subsequent transformations. Among them, the activation of Lewis bases for inert carboxylic acid derivatives, such as amides, was reported by Szostak’s group in 2017,[93] whilst for acyl fluorides, a few years ago, we discovered two triarylphosphine-catalyzed reactions that occurred via activation of a C–F bond in acyl fluorides and an inert C–O bond. In the first protocol, a series of acyl fluorides was accidentally defluoromethoxylated using PPh3 as the catalyst and methyl ethers were utilized as a methoxy source to synthesize an array of methyl benzoates with good functional group tolerance (Scheme [49a]).[94] In the second protocol, tris(2,4,6-trimethoxyphenyl)phosphine was utilized as both a catalyst and a methoxylating agent to again afford a series of methyl benzoates in good yields (Scheme [49b]).[95] After extensive studies, it was concluded that the phosphine forms an acyl phosphonium fluoride from the acyl fluoride, and that regiospecific cleavage of the C–O bond promoted by the acyl phosphonium fluoride is a key step in the reaction. Hence, a possible mechanism for the former reaction is shown in Scheme [49c]. Phosphine reacts first with the acyl fluoride to form acyl phosphonium fluoride A, while the methyl ether and TBAT form the key silicate B, which can lead to regiospecific cleavage of the C–O bond to produce silicate [Ph3FSiOMe] (C). Finally, the electrophilic carbonyl group undergoes nucleophilic attack of the methoxide ion of the silicate C to deliver the methoxylated product and regenerates PPh3.

Zoom Image
Scheme 49 (a) Phosphine-catalyzed alkoxylation of acyl fluorides. (b) Phosphine-catalyzed methoxylation of acyl fluorides. (c) A proposed mechanism for the reaction shown in Scheme 49a.

# 4.2

Phosphine-Catalyzed Indirect Activation of the C–F Bond of Acyl Fluorides

Zoom Image
Scheme 50 (a) Phosphine-catalyzed carbofluorination of electron-deficient alkynes with acyl fluorides. (b) A proposed mechanism. (c) Phosphine-catalyzed three-component anti-Michael addition of electron-deficient alkynes with acyl fluorides and silyl enol ethers.

Tobisu et al. have performed pioneering research in the field of phosphine-catalyzed indirect C–F bond activation of acyl fluorides. In 2020, they achieved an unprecedented phosphine-catalyzed carbofluorination of alkynes, in which acyl fluorides served as acylating and fluorinating reagents to construct a series of alkenyl fluorides in excellent yields and with notable functional-group tolerance (Scheme [50a]).[96] A series of subsequent mechanistic studies and DFT calculations revealed that an unprecedented ligand coupling on phosphine phosphorus constructs the C–F bond according to the mechanism shown in Scheme [50b]. First, nucleophilic addition of tricyclohexylphosphine to the electron-deficient alkyne results in the formation of carbanion intermediate A, which can then nucleophilically attack the acyl fluoride through ready activation of the C–F bond to form fluorophosphorane B, bearing an electron-withdrawing group in an equatorial position, which promotes ligand coupling and releases the carbofluorinated product.

Recently, the same research group developed a novel PCy3-catalyzed three-component coupling reaction of acyl fluorides, silyl enol ethers, and alkynes, based on previous research. This strategy provides for the synthesis of a variety of 1,3-dienes with good substrate scope and high yields (Scheme [50c]).[97] In-depth mechanistic studies and DFT calculations indicated that the initial reaction is nucleophilic addition of the phosphine to the electron-deficient alkyne to form a carbanion intermediate. The crucial formation of a fluorophosphorane intermediate B via nucleophilic acyl substitution of acyl fluorides is also key. A sequential and unusual ligand metathesis then leads to polarity-mismatched bonding, resulting in anti-Michael-addition-type products.


#
# 5

C–F Bond Activation by Classical Nucleo­philic Substitution

Nucleophilic substitution of carboxylic acid derivatives is one of the most classic strategies for transformative applications. Conventional nucleophilic substitution reactions using acyl fluorides can also be regarded as one of the earliest methods for such transformations. Among the research trends in modern organic chemistry, the study of classical nucleophilic substitution toward acyl fluorides is still popular, because the electron density of the carbon atom in the carbonyl group is lower due to the very strong electronegativity of the fluorine atom, resulting in an increase in electrophilicity. Although fluoride ions are not a good leaving group for nucleophilic substitution reactions, carbon centers with low localized electron density are still more likely to be attacked by nucleophiles.[98] In addition, the classical ‘addition–elimination’ mechanism model fully interprets the pathway of this reaction (Scheme [51]).

Zoom Image
Scheme 51 A classical nucleophilic substitution reaction of acyl fluorides via an ‘addition–elimination’ sequence

In 2020, Maruoka et al. succeeded in achieving the one-pot intermolecular nucleophilic substitution of amides with various nitrogen, oxygen, and carbon nucleophiles catalyzed by copper, by using a strategy to generate acyl fluorides in situ with Selectfluor.[99] Furthermore, this reaction was applied to the synthesis of multifunctionalized peptides, which greatly increases the synthetic value of this protocol (Scheme [52a]).

Zoom Image
Scheme 52 (a) Nucleophilic substitution of acyl fluorides formed in situ from amides with amines, alcohols or thiophenes. (b) Nucleophilic substitution of acyl fluorides formed in situ from carboxylic acids with amines and alcohols.

Very recently, Cobb et al. utilized pentafluoropyridine (PFP) as a deoxyfluorinating agent to form acyl fluorides in situ from carboxylic acids, and then constructed a series of amides and esters in good to excellent yields via intermolecular nucleophilic substitution with amines or alcohols as nucleophiles (Scheme [52b]).[100]


# 6

Miscellaneous Examples

In this section, some unique methods involving C–F bond activation of acyl fluorides are introduced as miscellaneous examples. In 2020, a base-promoted decarboxylative coupling reaction of acyl fluorides with polyfluorobenzoates was developed by our group.[101] In this reaction system, polyfluorobenzoates play an unprecedented role in the synthesis of a series of unsymmetrical ketones. The method exhibits broad functional-group tolerance and proceeds under extremely practical and environmentally benign conditions (Scheme [53]). In this reaction, the acyl fluoride reacts with polyfluorobenzoate to form the mixed anhydride A, from which three possible pathways are possible. In path A, acyl radical B and ester radical C are generated by homolytic cleavage of a C–O bond in anhydride A. Subsequently, rapid decarboxylation of unstable radical C produces the polyfluoroaryl radical D. Finally, this radical recombines with the acyl radical B to give the coupled product. In path B, the intramolecular reaction of the formed mixed anhydride A proceeds directly to yield the final unsymmetrical ketones, accompanied by decarboxylation through the four-membered transition state E. Both paths A and B have been ruled out. In path C, spontaneous decarboxylation of the polyfluorobenzoate forms a highly reactive phenyl potassium intermediate F, which nucleophilically attacks the acyl fluoride to give the final product.

Zoom Image
Scheme 53 Base-promoted decarboxylative coupling of acyl fluorides and polyfluorobenzoates
Zoom Image
Scheme 54 FLUOLEAD®-mediated deoxytrifluoromethylation of acyl fluorides

In 2020, Shibata’s group reported a small-molecule-promoted transformation of acyl fluorides. This protocol involves deoxytrifluoromethylation of acyl fluorides with a combination of FLUOLEAD® ((4-(tert-butyl)-2,6-dimethylphenyl)trifluoro-λ4-sulfane) and Olah’s reagent (Scheme [54]).[102] In the postulated reaction mechanism, FLUOLEAD® and Olah’s reagent form complex A through hydrogen bonding, which nucleophilically attacks the carbonyl group of the acyl fluoride to form intermediate C. Then, in the presence of HF, intermediate C is further activated to yield the final trifluoromethylated product.

Recently, during the course of developing Sm/SmI2 as a new activator of carboxylic acid derivatives, An and co-workers found that acyl fluorides were compatible with such a system (Scheme [55]).[103] Single-electron transfer of acyl fluorides would generate ketyl radicals, which are then deuterated by D2O to give a wide variety of α,α-dideuterated alcohols with a broad substrate scope. In particular, this strategy was used to achieve the deuteration of a series of compounds that could potentially be used as drugs and agrochemicals, and it was found that a high incorporation of deuterium could be accomplished under various conditions. This new transformation is expected to play an important role in future research on the development of life sciences, pharmaceutical engineering, and agrochemicals.

Zoom Image
Scheme 55 SmI2-assisted defluorination and bis-deuteration of acyl fluorides with deuterium oxide

In 2021, Ogoshi’s group developed a CsF-catalyzed fluoroacylation of tetrafluoroethylene (TFE). Tetrafluoroethylene can react smoothly at a low pressure of 1.5 atmospheres, and a series of pentafluoroethyl ketones was produced via this method with good functional-group compatibility.[104] The authors also conducted related studies on side reactions and found that ester C, bearing two pentafluoroethyl groups, was formed as a transient intermediate during the reaction. Intermediate C is present only at low temperatures, and on raising the temperature gave the desired pentafluoroethyl ketone product. Following a mechanistic study, the authors deduced a possible mechanism, which is divided into two parts (Cycles 1 and 2), representing formation of the pentafluoroethyl ketone and ester C, respectively. In Cycle 1, CsF first reacts with TFE to form CsC2F5 intermediate A, which then reacts with the acyl fluoride to give the desired pentafluoroethyl ketone product and regenerates CsF. In Cycle 2, on the other hand, pentafluoroethyl ketone, due to its strong electrophilicity, reacts with CsC2F5 intermediate A to form intermediate B, which then reacts with the acyl fluoride to form ester C and releases CsF. Subsequent reaction of CsF with TFE produces intermediate A to complete the catalytic cycle. At low temperatures, Cycle 2 can be viewed as a minor reaction (Scheme [56]).

Zoom Image
Scheme 56 CsF-catalyzed fluoroacylation of tetrafluoroethylene using acyl fluorides for the synthesis of pentafluoroethyl ketones

# 7

Summary and Perspective

In conclusion, this review has summarized synthetic strategies for activating the C–F bond in acyl fluorides over the past two decades and the versatile use of acyl fluorides as novel reagents for acylation, arylation (or alkylation), and fluorination. In addition, the four most common modes for activation of the C–F bond of acyl fluorides are (1) transition-metal catalysis, (2) N-heterocyclic carbene (NHC) catalysis, (3) phosphine catalysis, and (4) classical nucleophilic substitution. Initially, catalytic activation procedures were introduced in detail, and then various synthetic methods for the formation of C–C, C–F, C–O, C–Si, C–S, and C–Sn bonds were discussed. Among them, transition-metal-catalyzed cross-coupling reactions using acyl fluorides, especially decarbonylative couplings, will be more powerful in advanced research due to their advantages such as the formation of ‘transmetalation-active’ intermediates without exogenous bases. At present, however, acyl fluorides as effective electrophilic coupling partners under transition-metal catalysis are mainly limited to aromatic or activated acyl fluorides, while inactive aliphatic acyl fluorides still show significant limitations and low efficiency. This is mainly due to the low stability of the corresponding acyl fluorides, the simultaneous occurrence of side reactions, and the inability of catalysts to control the selectivity of the reaction. Further research is expected to overcome these related challenges and extend the appeal of reactions using acyl fluorides.

Compared to other carbonyl electrophiles such as acyl chlorides and acid anhydrides, acyl fluorides exhibit unique advantages in NHC and NHC/photoredox synergistic catalysis. For example, in NHC catalytic systems, when α,β-unsaturated acyl fluorides are used as the substrates, acyl azolium intermediates can be formed without the addition of oxidants and/or bases. With other carbonyl electrophiles, however, a base is often required to achieve this reaction efficiently. Furthermore, when combined with a photocatalytic oxidation system, acyl azolium intermediates are advantageous in that they can readily produce ketyl radical intermediates that are quite different from those produced by the oxidation of Breslow intermediates via single-electron reduction.

Moreover, with the renaissance of radical chemistry, the development of reactions employing activation by single-electron transfer of acyl fluorides is anticipated to become a new research direction in the future. In addition, the combination of this process with various other techniques, including photoredox catalysis, electrochemical systems, microwave-assisted systems, continuous-flow systems, quantum-dot catalysts, and polymer materials systems, is expected to lead to more extensive research and useful future applications.


#
#

Conflict of Interest

The authors declare no conflict of interest.


Corresponding Author

Yasushi Nishihara
Research Institute for Interdisciplinary Science, Okayama University
3-1-1 Tsushimanaka, Kita-ku, Okayama 700-8530
Japan   

Publication History

Received: 31 March 2022

Accepted after revision: 06 May 2022

Accepted Manuscript online:
06 May 2022

Article published online:
30 June 2022

© 2022. Thieme. All rights reserved

Georg Thieme Verlag KG
Rüdigerstraße 14, 70469 Stuttgart, Germany


Zoom Image
Zoom Image
Zoom Image
Zoom Image
Zoom Image
Scheme 1 Left: Types of activation of C–F bonds. Right: Types of Activation of carboxylic acid derivatives.
Zoom Image
Scheme 2 Reactions of acyl fluorides with nucleophiles
Zoom Image
Scheme 3 Activation and transformations of acyl fluorides
Zoom Image
Scheme 4 Reaction types under transition metal catalysis
Zoom Image
Scheme 5 Early studies on transition-metal-catalyzed C–F bond activation of acyl fluorides
Zoom Image
Scheme 6 Nickel-catalyzed cross-coupling reaction of acyl fluorides with organozinc reagents
Zoom Image
Scheme 7 Nickel-catalyzed cross-electrophilic coupling reactions of acyl fluorides with alkenyl triflates and the proposed catalytic cycle. a 2,2′:6′,2′′-Terpyridine (15 mol%) was used as the ligand.
Zoom Image
Scheme 8 Synergistic Ni/Cu-catalyzed hydroacylation of alkenes with acyl fluorides
Zoom Image
Scheme 9 Nickel-catalyzed cross-electrophilic coupling of acyl fluorides with alkylpyridinium salts. a Dtbbpy was used instead of L1.
Zoom Image
Scheme 10 Nickel-catalyzed intramolecular cyclization and arylation of carbamoyl fluorides tethered with alkene moieties with arylboronic acids
Zoom Image
Scheme 11 Nickel-catalyzed cross-coupling of N-CF3 carbamoyl fluorides with alkynylsilanes
Zoom Image
Scheme 12 (a) Palladium-catalyzed Hiyama-type coupling of acyl fluorides with phenylsilanes. (b) Palladium-catalyzed Suzuki–Miyaura-type coupling of acyl fluorides with boronic acids. a PCy3 was employed instead of P(4-MeOC6H4)3. b 2 equiv each of R′B(OH)2 and KF were used.
Zoom Image
Scheme 13 Pd/Cu-catalyzed heteroaromatic C–H coupling with acyl fluorides
Zoom Image
Scheme 14 Palladium-catalyzed hydrogenation or decarbonylative hydrogenation of acyl fluorides with hydrosilanes
Zoom Image
Scheme 15 Possible intermediates using PCy3 and DCPE
Zoom Image
Scheme 16 (a) Palladium-catalyzed intermolecular cyclization of acyl fluorides with norbornenes. (b) Palladium-catalyzed intramolecular cyclization of acyl fluorides via intramolecular C–H bond activation.
Zoom Image
Scheme 17 Palladium-catalyzed fluorine exchange using acyl fluorides and anhydrides. a 19F NMR yields. b GC yield.
Zoom Image
Scheme 18 Palladium-catalyzed decarbonylative trifluoromethylation of acyl fluorides with TMSCF3 and DFT calculations
Zoom Image
Scheme 19 Nickel-catalyzed decarbonylative Suzuki–Miyaura coupling of acyl fluorides with boronic acids without a base
Zoom Image
Scheme 20 Mechanistic studies
Zoom Image
Scheme 21 Nickel-catalyzed decarbonylative ethylation of acyl fluorides with alkylboranes
Zoom Image
Equation 1 Nickel-catalyzed decarbonylative methylation of acyl fluorides with methylboroxine
Zoom Image
Scheme 22 Nickel-catalyzed decarbonylative Sonogashira–Hagihara coupling of acyl fluorides with terminal alkynes
Zoom Image
Scheme 23 Nickel-catalyzed decarbonylative coupling reactions of acyl fluorides with diboronates, Bu3SnTMS, silylboronates, and thiols (thiophenols) to construct C–B, C–Si, C–Sn, and C–S bonds. a B2neo2 is used instead of B2pin2, then pinacol (4 equiv) and NEt3 (0.5 mL) at room temperature for 1 h.
Zoom Image
Scheme 24 Palladium-catalyzed decarbonylative alkylation of acyl fluorides with alkylboranes
Zoom Image
Scheme 25 Pd/Cu-catalyzed decarbonylative alkynylation of acyl fluorides with alkynylsilanes and the proposed mechanism
Zoom Image
Scheme 26 Palladium-catalyzed decarbonylative cyanation of acyl fluorides with TMSCN
Zoom Image
Scheme 27 Palladium-catalyzed decarbonylative arylation of difluoroacetyl fluorides with arylboronates
Zoom Image
Scheme 28 Pd/Cu-catalyzed three-component decarbonylative coupling of acyl fluorides with internal alkynes and silylboranes
Zoom Image
Scheme 29 Palladium-catalyzed decarbonylative coupling of acyl fluorides with distannanes
Zoom Image
Scheme 30 Cu-catalyzed three-component acylborylation of allenes with acyl fluorides and B2pin2
Zoom Image
Scheme 31 A proposed mechanism for the reaction shown in Scheme [30]
Zoom Image
Scheme 32 Iridium-catalyzed decarbonylative coupling of acyl fluorides with arenes, and the possible transition state or intermediate
Zoom Image
Scheme 33 Rhodium-catalyzed decarbonylative arylation of acyl fluorides, formed in situ from carboxylic acids, with arenes tethered with a directing group
Zoom Image
Scheme 34 A proposed mechanism for the reaction shown in Scheme [33]
Zoom Image
Scheme 35 (a) Cobalt-catalyzed asymmetric ring-opening fluorination of epoxides with benzoyl fluoride. (b) Cobalt-catalyzed asymmetric ring-opening fluorination of aziridines with benzoyl fluoride.
Zoom Image
Scheme 36 A proposed mechanism for the reaction shown in Scheme [35b]
Zoom Image
Scheme 37 The three reaction modes of the C–F bonds in acyl fluorides catalyzed by N-heterocyclic carbenes (NHCs)
Zoom Image
Scheme 38 (a) NHC-catalyzed cycloaddition of α,β-unsaturated acyl fluorides with silyl enol ethers. (b) NHC-catalyzed cycloaddition and decarboxylation of α,β-unsaturated acyl fluoride tethered conjugated dienes with silyl enol ethers. (c) NHC-catalyzed cycloaddition of α,β-unsaturated acyl fluorides with conjugated silyl enol ethers. (d) NHC-catalyzed cycloaddition and decarboxylation of α,β-unsaturated acid fluorides with silyl enol ethers.
Zoom Image
Scheme 39 A proposed mechanism for the reaction shown in Scheme [38b]
Zoom Image
Scheme 40 NHC-catalyzed cycloaddition of α,β-unsaturated acyl fluorides with cyclopropyl silyl ethers
Zoom Image
Scheme 41 NHC-catalyzed cycloaddition of α,β-unsaturated acyl fluorides with 1,3-diketones, nitrile esters, or hydrazones
Zoom Image
Scheme 42 A proposed mechanism for the reaction shown in Scheme [41a]
Zoom Image
Scheme 43 (a) NHC/photoredox dual-catalyzed difunctionalization of alkenes with acyl fluorides and sodium trifluoromethanesulfonate (CF3SO2Na). (b) NHCs/photoredox/sulfonate tri-catalyzed acylation of alkenes with acyl fluorides.
Zoom Image
Scheme 44 A proposed mechanism for the reaction shown in Scheme [43b]
Zoom Image
Scheme 45 (a) NHC/photoredox dual-catalyzed acylation of benzylic C–H bonds with acyl fluorides. (b) NHCs/photoredox dual-catalyzed ring-opening and 1,3-difunctionalization of cyclopropanes with acyl fluorides.
Zoom Image
Scheme 46 Dearomatizing 2,3-fluoroaroylation of benzofurans with aroyl fluorides as bifunctional reagents for both C–C and C–F bond formation via cooperative NHC/photoredox catalysis
Zoom Image
Scheme 47 NHC-catalyzed SNAr fluorination of aryl chlorides with acyl fluorides. a PhCOF is used. b 4-MeOC6H4COF is used. c Acyl fluoride (4 equiv) is used.
Zoom Image
Scheme 48 Direct and indirect activation of C–F bonds in acyl fluorides
Zoom Image
Scheme 49 (a) Phosphine-catalyzed alkoxylation of acyl fluorides. (b) Phosphine-catalyzed methoxylation of acyl fluorides. (c) A proposed mechanism for the reaction shown in Scheme 49a.
Zoom Image
Scheme 50 (a) Phosphine-catalyzed carbofluorination of electron-deficient alkynes with acyl fluorides. (b) A proposed mechanism. (c) Phosphine-catalyzed three-component anti-Michael addition of electron-deficient alkynes with acyl fluorides and silyl enol ethers.
Zoom Image
Scheme 51 A classical nucleophilic substitution reaction of acyl fluorides via an ‘addition–elimination’ sequence
Zoom Image
Scheme 52 (a) Nucleophilic substitution of acyl fluorides formed in situ from amides with amines, alcohols or thiophenes. (b) Nucleophilic substitution of acyl fluorides formed in situ from carboxylic acids with amines and alcohols.
Zoom Image
Scheme 53 Base-promoted decarboxylative coupling of acyl fluorides and polyfluorobenzoates
Zoom Image
Scheme 54 FLUOLEAD®-mediated deoxytrifluoromethylation of acyl fluorides
Zoom Image
Scheme 55 SmI2-assisted defluorination and bis-deuteration of acyl fluorides with deuterium oxide
Zoom Image
Scheme 56 CsF-catalyzed fluoroacylation of tetrafluoroethylene using acyl fluorides for the synthesis of pentafluoroethyl ketones