Next Article in Journal
Biodiesel at the Crossroads: A Critical Review
Next Article in Special Issue
Hydrothermal Aging of Pd/LTA Monolithic Catalyst for Complete CH4 Oxidation
Previous Article in Journal
Carbon Quantum Dots (CQDs) Decorated Bi2O3-x Hybrid Photocatalysts with Promising NIR-Light-Driven Photodegradation Activity for AO7
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Organic Transformations with Silver Carbonate as a Key External Base and Oxidant

1
Department of Chemistry, Chungbuk National University, Cheongju 28644, Korea
2
Innovative Therapeutic Research Center, Therapeutics and Biotechnology Division, Korea Research Institute of Chemical Technology, Daejeon 34114, Korea
3
Department of New Drug Discovery and Development, Chungnam National University, Daejeon 34134, Korea
*
Authors to whom correspondence should be addressed.
These authors equally contributed to this work.
Catalysts 2019, 9(12), 1032; https://doi.org/10.3390/catal9121032
Submission received: 11 November 2019 / Revised: 2 December 2019 / Accepted: 4 December 2019 / Published: 6 December 2019
(This article belongs to the Special Issue Noble Metal Catalysts)

Abstract

:
Silver carbonate (Ag2CO3), a common transition metal-based inorganic carbonate, is widely utilized in palladium-catalyzed C–H activations as an oxidant in the redox cycle. Silver carbonate can also act as an external base in the reaction medium, especially in organic solvents with acidic protons. Its superior alkynophilicity and basicity make silver carbonate an ideal catalyst for organic reactions with alkynes, carboxylic acids, and related compounds. This review describes recent reports of silver carbonate-catalyzed and silver carbonate-mediated organic transformations, including cyclizations, cross-couplings, and decarboxylations.

Graphical Abstract

1. Introduction

Organic transformations using transition metal catalysts are essential synthetic techniques in total synthesis, medicinal chemistry, industrial chemistry, and chemical engineering [1]. A variety of transition metal catalysts have been developed and studied for various organic reactions, and they are generally used for selective bond formation and the interconversion of specific organic functional groups. Transition metal-catalyzed reactions can generally be classified into three types based on the role of the metal catalyst. The first class includes those in which the catalytic reaction is based on the oxidation/reduction cycle of the transition metal. As many transition metals can have more than two relatively stable oxidation states, the interconversion between oxidation states of a metal can be utilized to oxidize or reduce target molecules or functionalities. The second class of catalytic reactions are those in which the transition metal serves as a Lewis acid. The final group of organic transformations are those catalyzed by coinage metals [2]. As both Cu(I) and Cu(II) are generally stable states for copper, copper-catalyzed organic transformations have been widely studied. However, silver- and gold-promoted organic reactions are relatively rare.
Silver catalysts are generally less reactive than other transition metal catalysts; therefore, in organic reactions, silver compounds are generally used as cocatalysts or bond activators due to their Lewis acidity [3,4]. In addition, Ag(I) is one of the softest Lewis-acidic metal cations; thus, Ag(I) forms organometallic complexes with π-donor functionalities, such as alkenes, alkynes, allenes, and aromatic rings. Concurrently, Ag(I) generates relatively stable organometallic complexes with n-donor type molecules, for example, ethers, amines, and phosphines.
In the last decade, various silver-catalyzed organic transformations, such as carboxylations, halogenations, C–H bond functionalizations, cycloadditions, and oxidative couplings, have been reported [5,6,7,8]. The reported silver-catalyzed reactions were performed under relatively milder reaction conditions than the reactions by other methodologies, and silver complexes are less expensive than other rare transition metals, such as palladium, rhodium, ruthenium, and platinum.
Among the various silver species, silver carbonate (Ag2CO3) is not typically used as a catalyst or mediator of organic reactions, but it is known as a good oxidant (for example in the Fetizon oxidation) and inorganic base (for example in the Wittig reaction) [9,10]. Scheme 1 describes the general reactions involving Ag2CO3. Ag2CO3 enables the use of Lewis acids with internal alkynes, and readily activates alkynes to nucleophilic attack, providing the corresponding addition products. With terminal alkynes, Ag2CO3 generates silver acetylide intermediates through C–H functionalization, and then various reactions (for example, cross-couplings, and cycloadditions) can be used to convert these intermediates into useful organic frameworks, such as furans [11], pyrroles [12], and nitriles [13]. Additionally, Ag2CO3 is used as an oxidant for single-electron transfer (SET) processes to generate reactive radical species for various organic transformations, for example, the generation of acyl radicals from carboxylic acids. In addition, the basicity of the carbonate moiety can help prepare reactive partners from acidic organic molecules such as carboxylic acids, terminal alkynes, and 1,3-dicarbonyl compounds.
In this review, we summarize the recently reported Ag2CO3-catalyzed and Ag2CO3-mediated organic transformations and small molecule syntheses from 2005 to 2019 based on the reactivity of Ag(I) and the basicity of Ag2CO3. The focus will be on their synthesis and substrate scope as well as the principle reactivity of silver carbonate, and the substrates will also be discussed based on mechanistic clues.

2. Activation of Alkynes Using Silver Carbonate

2.1. Terminal Alkynes: Cross-Coupling Reactions

Oxidative double C–H functionalization/cross-coupling reactions are considered an attractive strategy for the synthesis of various heterocycles. The Ag2CO3-mediated activation of terminal alkynes is a good starting point for this heteroaromatic cyclization. In 2012, Lei and co-workers reported the direct C–H functionalization and oxidative coupling of two C–H bonds as an ideal synthetic strategy for accessing highly substituted furans (Scheme 2) [14]. Although synthetic methods for preparing substituted furans have been reported previously [15], new synthetic strategies are required to achieve selective cross coupling of two more hydrocarbons. The homocoupling of terminal alkynes occurs easily under most oxidative conditions [16]. This report shows that Ag2CO3 can be used as a mediator to induce C–H/C–H functionalization of terminal alkynes and 1,3-dicarbonyl compounds. This reaction affords furans with several functional groups in a facile, one-step manner. Terminal alkyne 1 can be converted to the corresponding silver acetylide intermediate by Ag2CO3, and furan 3 is synthesized by alkylation with 1,3-dicarbonyl compound 2 followed by oxidative radical cyclization.
In the same year, the Lei group reported a Ag-mediated, highly selective C–C/N–H oxidative cross-coupling/cyclization to construct imidazo[1,2-a]pyridines 5 from 2-aminopyridines 4 and terminal alkynes 1 (Scheme 3) [17]. This reaction is also initiated by the generation of a silver acetylide intermediate from the Ag salt and phenylacetylene; then, subsequent Ag-promoted nucleophilic attack and oxidative cyclization form imidazo[1,2-a]pyridine 5 as the product. After the reaction, the spent silver species can be recycled by filtration and treatment with nitric acid and Na2CO3.
As described above, oxidative C–H functionalization/cross-coupling reactions are considered an attractive strategy for the synthesis of various heterocycles. Pyrroles, as 5-membered heterocycles, are also a basic building block of numerous biologically and pharmaceutically important natural compounds [12], and the regiospecific synthesis of pyrroles is an attractive synthetic tool. The Lei group reported that pyrroles 7 containing various functional groups can be effectively synthesized by the C–H/C–H oxidative cross coupling/cyclization of terminal alkynes 1 and β-enamino esters 6 using Ag2CO3 as a mediator (Scheme 4) [18]. The proposed mechanism is shown in Scheme 5. Due to the effects of Ag2CO3, terminal alkyne 1 is primarily converted to silver acetylide 1-I; then, the nucleophilic addition of deprotonated β-enamino ester intermediate 6-I to silver acetylide 1-I delivers oxidative cross-coupling intermediate 6-II via two single-electron oxidations. The coordination of Ag(I) generated internal alkyne 6-III, which is cyclized by an intramolecular nucleophilic addition to provide cyclic intermediate 6-IV, and subsequent aromatization and protodemetallation give highly substituted pyrrole 7 as the product.
In 2014, Bi and co-workers reported an Ag2CO3-catalyzed cross-coupling reaction of propargylic alcohols 8 with isocyanides 9 to give 2,3-allenamides 10 (Scheme 6) [19]. Although the reaction mechanism is not clear, this reaction involves a unique β-C–C coupling and oxygen transposition of propargylic alcohols.
The Agrawal group reported the synthesis of indolizines 12 via a Ag-mediated oxidative C–H functionalization and cyclization of terminal alkynes 1 with pyridine derivatives 11 (Scheme 7) [20]. Indolizine is a basic organic building block for many alkaloids and bioactive molecules, so the development of regioselective synthetic pathways is an attractive topic in organic synthesis [21]. This reaction allows the highly selective synthesis of indolizines from readily available starting materials via a 5-endo-dig cyclization between the silver phenylacetylide generated from alkyne 1 and a chelated intermediate from the deprotonation of 2-pyridine derivative 11. After the reaction, Ag2CO3 can be recycled and reused.
In 2015, Lei and co-workers developed a Ag(I)-mediated selective oxidative cross coupling between C–H and P–H bonds to afford alkynyl (diaryl) phosphine oxides 14 (Scheme 8) [22]. To determine the mechanism of this reaction, the authors performed experiments with silver complexes as the starting materials. Phenylacetylide was mixed with Ph2P(O)H and prepared silver; the reaction only provided only 17% of the desired product, but 60% of the desired product was obtained when an additional 1.0 equiv. of Ag2CO3 was added. When the reaction used Ph2P(O)Ag instead of Ph2P(O)H, only a trace amount of the product was obtained with phenylacetylene. In the presence of an additional 1.0 equiv. of Ag2CO3 in the same reaction, 37% of the product was obtained. With both silver phenylacetylide and Ph2P(O)Ag, the reaction provided 30% of the desired product, and 71% of the product was produced when an additional 1.0 equiv. of Ag2CO3 was added.
Based on the experimental results, a mechanism was proposed as shown in Scheme 9. Silver acetylide 1-I and disubstituted phosphoryl silver species 13-I are formed from silver carbonate, and a cascade involving coordination and cross-coupling reactions via two single-electron oxidations provided alkynyl(diaryl)phosphine oxides 14.

2.2. Terminal Alkynes: Cycloaddition Reactions

The ideal method for the synthesis of substituted pyrroles is the cycloaddition of an alkyne and an isocyanide. Isocyanide-alkyne cycloaddition reactions are usually limited to use electron-deficient alkynes, and the yields are low. In 2013, two interesting papers were published on the synthesis of pyrroles via Ag2CO3-catalyzed cycloadditions of isocyanides and terminal alkynes. One is from the Bi group (Scheme 10, Conditions A) [23], and the other is from the Lei group (Scheme 10, Conditions B) [24]. The reaction mechanism is unclear, so these two groups explained the reaction mechanism in different ways. Based on the results of a deuterium-labeling experiment, Bi suggests that the reaction is initiated by the formation of a silver acetylide from terminal alkyne 1 and the Ag(I) catalyst, and this is followed by the 1,1-insertion of isocyanide 9 into the Ag–C bond. Protonolysis and intramolecular cyclization provide a pyrrolenine intermediate, and a 1,5-hydrogen shift gives the pyrrole. On the other hand, Lei suggests that the mechanism involves a click reaction like a Cu(I)-catalyzed azide-alkyne cycloaddition (CuAAC) [25], because they observed the coordination of the silver to the isocyano group by in situ IR spectroscopy of the stoichiometric reaction between ethyl 2-isocyanoacetate and Ag2CO3. Therefore, the Ag2CO3 catalyst generates both a silver acetylide and a silver isocyanide complex, and click cyclization provides substituted pyrroles 15.
In 2014, Bi and co-workers reported the synthesis of pyrroles and indolizines from the cyclization of 2-pyridyl alkynyl carbinols with isocyanides (Scheme 11) [26]. The divergent pathways are described in Scheme 12. 2,4-Disubstituted pyrroles 17a were synthesized from secondary 2-pyridyl alkynyl carbinols 16 by a regioselective [3+2] cycloaddition. Coordination of the pyridine moiety affords coordinated silver acetylide 16-I, and an additional coordination occurs with isocyanide 9 to form intermediate 16-II. [3+2] Cycloaddition affords 2,4-disubstituted pyrrole 16-III, and subsequent oxidation gives pyrrole 17a as the product. The indolizines are generated from tertiary propargyl alcohols 16. The Ag(I) catalyst reacts with terminal alkyne 16 to provide silver acetylide 16-IV, and then a 1,1-insertion of isocyanide 9 into the Ag–C bond affords acetylenic imido intermediate 16-V. Intramolecular rearrangement converts 16-V to 2,3-allenamide 16-VI, and subsequent intramolecular cycloisomerization gives indolizine 17b.

2.3. Terminal Alkynes: Reactions with Azide

In 2013, the Jiao group reported an Ag-catalyzed nitrogenation of alkynes, representing the first direct conversion of alkynes to nitriles via C≡C bond cleavage (Scheme 13) [27]. The reaction was performed with various alkynes as the starting material, including aryl-, heteroaryl-, alkyl- and alkenyl-substituted alkynes. In the proposed reaction mechanism (Scheme 14), Ag(I) simply coordinates to the alkyne, and no silver acetylide is generated. Then, the azide nucleophile adds to the activated alkyne (1-II) to form alkenyl metal complex 1-III. The protodemetallation of 1-III provides vinyl azide 1-IV, and unstable intermediate 1-V is formed by the click reaction of vinyl azide 1-IV. Upon the release of HN3 and CH2N2, intermediate 1-V provides nitrile product 19.
In 2014, the Bi group reported a Ag(I)-catalyzed hydroazidation to provide 2-azidoallyl alcohols 21 from ethynyl carbinols 20 (Scheme 15) [28]. Hydroazidation is one of the most attractive routes for synthesizing vinyl azides [29]. In a study of the effects of residual water in DMSO as the solvent (Scheme 16), dry DMSO provided a mixture of product 23 and TMS-protected product 23-TMS in a 1:3 ratio. When 1.0 equivalents of H2O was added, only the TMS-free product was obtained. When a large amount of water was added, a substantial amount of unreacted starting material 22 remained in the reaction. Based on the results of these reactions, water must play an important role in the reaction.
Based on a deuterium-labeling experiment, the Ag catalyst generates a silver acetylide from ethylnyl carbinol 20 and HN3 by the Ag(I)-catalyzed hydrolysis of TMS-N3 18. HN3 adds to the silver acetylide to give a vinyl silver intermediate, and subsequent protodemetallation gives 2-azidoallyl alcohol 21 as the product.
Unactivated terminal alkynes also provide vinyl azides via a Ag-catalyzed hydroazidation with TMS-N3 (Scheme 17) [30]. Thus, various terminal alkynes with or without hydroxyl groups at the propargyl position can be successfully converted into vinyl azides in the presence of water.
With diynes, the generated vinyl azides can go on to react with the second alkyne moiety via an intramolecular pathway. The Bi group reported a Ag(I)-catalyzed tandem hydroazidation and alkyne-azide cycloaddition of diynes 26 to give 1,5-fused 1,2,3-triazole frameworks (Scheme 18) [31]. As discussed above, terminal alkynes readily provide vinyl azides via Ag-catalyzed hydroazidation. The remaining unreacted alkyne undergoes an intramolecular alkyne-azide 1,3-dipolar cycloaddition to give bicyclic 1,2,3-triazoles 27.
In 2015, Reddy and co-workers developed an efficient pathway for accessing 3-amino quinoline derivatives 30 from readily available 1-(2-aminophenyl) propargyl alcohols 28 (Scheme 19) [32]. The Ag(I) catalyst activates the terminal alkynes to promote the 5-exo-dig cyclization with the amine group. [2+3] Cycloaddition with azide 29 and a C–N bond migration with N2 expulsion results in ring expansion and 6-endo cyclization. Subsequent tautomerization and dehydration afford 4-substituted 3-tosylaminoquinolines 30. This is a rare example of a 6-endo-dig reaction of a terminal alkyne, which tend to undergo 5-exo-dig cyclizations.

2.4. Internal Alkynes: Synthesis of Heterocyclic Compounds

Although internal alkynes cannot form reactive silver acetylides with Ag(I), where terminal alkynes easily can, internal alkynes can be activated by coordination. In 2005, the Ding group used internal alkynes, (Z)-2-alken-4-ynylphosphonic monoesters 31, as the starting material to access 2H-1,2-oxaphosphorin 2-oxides 32 by a Ag(I)-catalyzed cyclization (Scheme 20) [33]. Compared with using (Z)-2-en-4-ynoic acid to prepare the five- or six-membered ring products [34], this reaction provides high 6-endo-dig regioselectivity. The phosphonyl oxygen regioselectively attacks the Ag(I)-activated C≡C bond in an endo manner to give a vinyl silver intermediate, and subsequent proton transfer and protodemetallation provided 2-ethoxy-2H-1,2-oxaphosphorin 2-oxides under mild conditions. The cyclization of P–OH onto an internal alkyne is a unique reaction, and this platform provides an effective synthetic route for accessing potentially bioactive phosphorus-containing heterocycles.
In 2014, Pan and co-workers reported the construction of substituted indolizines via a Ag(I)-mediated C–H bond functionalization of 2-alkylazaarenes with alkynes (Scheme 21) [35]. We already discussed a Ag(I)-mediated synthesis of indolizines 12 from terminal alkynes 1 and pyridine derivatives 11 (Scheme 7), and the reaction was proposed to involve a 5-endo-dig cyclization between silver phenylacetylide and a chelated intermediate. Here, the authors suggested a reaction mechanism based on the results of a deuterium-labeling experiment. In the case of a terminal alkyne (Scheme 22a), silver acetylide 1-I was initially generated from alkyne 1 and Ag(I), and 2-alkylazaarene 11 was separately deprotonated by KOAc. The nucleophilic attack of anion 11-I by silver acetylide 1-I provides oxidative cross-coupled intermediate 11-II. Subsequent Ag(I)-assisted cycloisomerization and protodemetallation give indolizine product 12. In cases with internal alkynes (Scheme 22b), no silver acetylide intermediate is formed. 2-Alkylazaarene 11 undergoes single-electron oxidation to give radical intermediate 11-VI. Then, single-electron insertion into internal alkyne 33 provides intermediate 11-VII. Further oxidation of intermediate 11-VII by Ag(I) generates carbocation 11-VIII, and indolizine 34 is produced by intramolecular condensation.
A Ag(I)-mediated synthesis of alkynyl (diaryl) phosphine oxides from terminal alkyne with Ph2P(O)H was previously described in Scheme 8. The reaction involves the generation of silver acetylide and an oxidative cross coupling. The Wu group developed a version of this reaction with an internal alkyne, namely, a Ag(I)-catalyzed synthesis of 3-phosphorated coumarins 36 via the radical cyclization of alkynoates 35 and dialkyl H-phosphonates 13 in a highly regioselective manner (Scheme 23) [36]. The reaction was completely suppressed by the presence of stoichiometric TEMPO ((2,2,6,6,-tetramethylpiperidin-1-yl)oxyl), which implies that the reaction proceeds through a radical pathway. With this experimental result in hand, a reaction mechanism was proposed, as shown in Scheme 24. Phosphorus radical 13-I is generated from diethyl H-phosphonate 13 by Ag(I), and subsequent selective addition of the P-radical to the α-position of the C=O bond in the alkynoate gives vinyl radical 35-I. Cyclization of intermediate 35-I generates intermediate 35-II, and subsequent SET from 35-II to Ag(I) releases 3-phosphorated coumarin 36.
In 2014, the Fillion group reported a Ag(I)-catalyzed 5-exo-dig cyclization of propargylic Meldrum’s acid derivatives 37 to give γ-alkylidene butyrolactones 38 (Scheme 25) [37]. The formation of the E/Z-isomers of the product is controlled by the coordination of Ag(I). The alkyne-Ag(I) complex determines the form of an E- or Z-isomer based on the 5-exo-dig cyclization through the attack of the carbonyl-O. Syn-attack of the carbonyl-O and Ag(I)-activated alkyne provides the E-intermediate; on the other hand, anti-attack of the carbonyl-O onto the Ag(I)-coordinated alkyne gives the Z-intermediate. Both intermediates undergo thermally induced cycloreversion to provide the corresponding acylketene intermediates; then, γ-alkylidene butyrolactones 38 are obtained by nucleophilic attack of the solvent.
Indoloquinolines are among the most important skeletons in the design, development, and synthesis of drugs currently on the market [38]. Most synthetic approaches rely on the use of indole or its derivatives as a coupling partner, and a new synthetic pathway for indolo[2,3-b]quinolone has been developed via a cascade radical annulation of o-alkynylthiourea [39]. In 2017, the Patel group reported a microwave-assisted Ag(I)-catalyzed cascade reaction toward indolo[2,3-b]quinolines 41 via the in situ generation of o-alkynylthioureas from 2-(phenylethynyl)anilines 39 with phenyl isothiocyanate 40 (Scheme 26) [40]. The reaction of 2-(phenylethynyl)anilines 39 and phenyl isothiocyanate 40 provides thiourea, followed by desulfurization in the presence of Ag2CO3, to give a carbodiimide intermediate. An intramolecular thermal cyclization drives the carbodiimide intermediate to a carbene-type intermediate; then, indoloquinoline 41 is produced by carbene C−H insertion followed by aromatization.

3. Functionalization of Carboxylic Acid Using Ag2CO3

3.1. Decarboxylation of Carboxylic Acids

In 2009, the Larrosa group reported a Ag(I)-catalyzed protodecarboxylation of ortho-substituted benzoic acids 42 (Scheme 27) [41]. This reaction only proceeds with ortho-substituted benzoic acids although they can have a range of functionalities, such as halogens, NO2, alkoxides, and even unprotected phenols and amines, regardless of the presence of any functional group at positions other than the ortho-position of the benzoic acid. A proposed mechanism for the reaction involves conversion of the Ag(I) species to the silver carboxylate with benzoic acid 42. After decarboxylation of the silver carboxylate, the silver arene intermediate gives corresponding arene 43 by protodemetallation with concomitant generation of a new silver carboxylate to turn over the catalytic cycle (Scheme 28).
In the same year, the Larrosa group reported another Ag(I)-catalyzed protodecarboxylation of various heteroaromatic carboxylic acids 44 (Scheme 29) [42]. This reaction uses acetic acid as an additive. Not only α-heteroaromatic carboxylic acids but also ortho-substituted benzoic acids are converted to silver carboxylates by Ag(I), and the silver carboxylates then undergo decarboxylation. The details of the role of the heteroatoms in the decarboxylation are difficult to predict, but heteroatoms are expected to serve as activators in this reaction. Diacids were also subjected to these conditions, and they provided the corresponding monoacids with complete regioselectivity. In the case of heteroaromatic compounds, the reactivity of the carboxylic acid at the α-position is still controlled by the heteroatoms. In addition to the effect of the α-heteroatom, the conversion to the monoacid proceeds well even if an electron-withdrawing group is located at the ortho-position. Since both acids, 1,2-diacid and 1,4-diacid, underwent decarboxylation with high yields, the number of adjacent carboxylic acids seems to have no effect on the decarboxylation.
The Jafapour group reported a Ag(I)-catalyzed decarboxylation of heteroaromatic compounds (Scheme 30) [43]. In the case of coumarins, the decarboxylation is restricted to carboxylic acids containing α-heteroatoms. After tuning the solvent, even unactivated coumarin-3-carboxylic acids 46 produced corresponding decarboxylated products 47 under mild conditions with wide substrate compatibility.

3.2. Decarboxylative Functionalization and Carboxylations

Aryl halides are useful synthetic intermediates and can be used for a variety of reactions [44]. In general, boron–halogen exchange reactions have attracted attention as a synthetic pathway for accessing aryl halides. Previous common decarboxylative halogenation reactions use very insoluble NaCl and LiCl as chloride sources. In 2010, the Wu group reported a Ag(I)-catalyzed decarboxylative halogenation of benzoic acids in the presence of Cu(II) halides to provide the corresponding aryl halides (Scheme 31) [45]. Cu(II) halides are inexpensive organometallic reagents and readily available halide sources with superior solubility in organic solvents.
Decarboxylation can lead to changes in the functional groups on the starting material as well as the formation of new C−C bonds. Muthusubramanian and co-workers reported a Ag-catalyzed acylation of pyridine N-oxides 49 by α-oxocarboxylic acids 50 in 2014 (Scheme 32) [46]. Through this reaction, various aryl ketones can be synthesized by effectively promoting the acylation at the C-2 position of the pyridine. When this reaction was conducted in the presence of TEMPO, the acylated TEMPO adduct was formed, indicating that the reaction proceeds through an acyl radical intermediate. Based on this result, the reaction may start by generating a silver dicarboxylate intermediate, and then the acyl radical intermediate is generated by the release of CO2 and a Ag(I) carboxylate. A new C−C bond between the acyl radical and pyridine N-oxide 49 was generated by a radical addition reaction, affording acylated pyridine N-oxide 51 as the product.
In 2014, Qi and co-workers reported Ag-catalyzed decarboxylative acylation of arylglyoxylic acids 50 with arylboronic acids 52 (Scheme 33) [47]. This Ag2CO3-catlayzed decarboxylation enables the synthesis of ketones 53 with aromatic rings bearing various functional groups. The most useful feature of this reaction is not only the fact that the arylglyoxylic acid bears a carboxylic acid but also the decarboxylative acylation that occurs when a phenylboronic acid has a carboxylic acid. The reaction pathway may be initiated by a free radical because no product was generated in the presence of TEMPO. Acyl radicals can be prepared by decarboxylation of silver carboxylate, and subsequent radical transformation can provide cross-coupled diaryl ketone 53.
Among the various C1 insertion reactions, CO is the most common acceptor for generating new acyl radicals through carbonylation. Isocyanide is also a radical acceptor for imidoyl radical formation. In recent years, decarboxylations of carboxylic acids using transition metals have contributed significantly to the formation of various C–C bonds and C–heteroatom bonds [48]. The Lei group reported an Ag-mediated oxidative radical decarboxylation-cyclization of α-oxocarboxylates 55 and isocyanides 54 to give 6-acyl phenanthridines 56 (Scheme 34) [49]. In the proposed mechanism (Scheme 35), acyl radical 55-I is formed by oxidative radical decarboxylation in the presence of Ag2CO3 as a catalyst. Then, radical addition to isocyanide 54 provided imidoyl radical 55-II. The intramolecular cyclization of imidoyl radical 55-II then gave cyclohexadienyl radical 55-III, and SET from Ag(I) provides 6-acyl phenanthridine 56 as the product.
In 2016, the Kim group developed an Ag-mediated temperature-controlled acyloxylations and subsequent hydrolysis for hydroxylation (Scheme 36) [50]. At lower temperature (100 °C), various esters 59 were synthesized from 10-bromobenzo[h]quinoline 57. Interestingly, at higher temperature (150 °C), ester products 59 were hydrolyzed to 10-hydroxybenzo[h]quinoline 60. 2-Methylbenzoic acid showed the best conversion in this hydroxylation. The silver species activated the carboxylic acid through silver benzoate formation, and the two-step acyloxylation and hydrolysis were confirmed by NMR spectroscopy.

4. Miscellaneous

Leblanc’s deamination of hydrazides is a useful reaction using Zn and acetic acid [51]. In general, hydrazides are converted into hydrazines instead of generating the corresponding amines from transition metal-catalyzed hydrogenolytic N–N bond cleavage [52]. In 2002, the Cho group reported a Ag(I)-mediated deamination of N-Boc aryl hydrazines 61 to afford N-Boc aryl amines 62 (Scheme 37) [53]. When an electron-donating group was present on the substrate, the corresponding products were obtained in high yields within relatively short reaction time (2 h). When an electron-withdrawing group was present on the substrate, the required reaction time increased to over 48 h (NO2) or 72 h (CO2Me) and the yields were relatively low.
The construction of C−N bonds in heteroaromatic compounds is important in the fields of biological, pharmaceutical, and materials science [54]. Although hydroamination [55] and oxidative amination [56], for the formation of C−N bonds, have been studied in great detail in recent decades, the direct installation of amino groups or their surrogates into aryl or alkyl C–H bonds is still challenging. In 2009, the Chang group reported a Ag-mediated amination of benzoxazoles 63 using formamides 64 or their parent amines (Scheme 38) [57]. The reaction gives 2-aminobenzoxazoles 65 with various functional groups by the regioselective formation of C–N bonds through the direct C(sp2)–H functionalization of heteroarenes.
A possible reaction mechanism is shown in Scheme 39. First, the acid promotes the decarbonylation of formamide 64 to give corresponding secondary amine 64-I. Secondary amine 64-I reacts with protonated benzoxazole 63-I to form 2-amino benzoxazoline intermediate 63-II. Finally, Ag2CO3 facilitates rearomatization to provide 2-aminated benzoxazole 65.
Mukhopadhyay reported a Ag(I)-mediated synthesis of diverse 1,2,4,5-tetrasubstituted imidazoles 68 via the tandem Cα(sp3)–H bond functionalization of primary amines 67 and oxidative C–N cross-coupling reaction (Scheme 40) [58]. Although the reaction required stoichiometric Ag2CO3, the recycled silver reagent provided the product in good yield without any loss of activity.
According to the proposed mechanism (Scheme 41), the reaction is initiated by the generation of imino ketone 66-I from diketone 66 and primary amine 67. Ag(I)-mediated C–H functionalization then gives intermediate 66-II, and the subsequent addition of another equivalent of amine 67 provides diamine intermediate 66-III. Finally, the silver-mediated oxidative cyclization of 66-III gives imidazole product 68.
In 2015, the Youn group reported the silver(I)-mediated synthesis of diverse substituted indoles via the C–H amination of 2-alkenylanilines 69 (Scheme 42) [59]. Interestingly, the products are dependent on the reaction solvents. In DMF, indoles 70 with R1 at the 2-position are selectively obtained; however, mixtures of 2- and 3-substituted indoles are observed when heptane is used as the solvent. To elucidate the reaction mechanism, the reaction was performed in the presence of TEMPO and BHT (butylated hydroxytoluene). In the absence of radical scavengers, the reaction was completed in 30 min, but it took 3 h for the reaction to reach completion with 1.0 equiv. of TEMPO. With 1.0 equivalent of BHT, 80% of the starting material remained unreacted after 24 h. Based on the results of the reactions with radical scavengers, the reaction should proceed through a radical pathway. According to the proposed mechanism, a nitrogen radical cation is generated from substrate 69 with Ag(I) salt by a SET; then, the nitrogen radical undergoes an intramolecular electrophilic addition to the alkene to provide a benzylic radical with the loss of a proton. A benzylic carbocation is generated by a SET from the benzylic radical to Ag(I), and subsequent deprotonation affords desired indole product 70.
The isocyanide group has the similar reactivity as carbenes, making them useful reagents for generating new C–C bonds, and they are used as building blocks in organic synthesis [60]. In addition, isocyanides have reactivities similar to multicomponent reactions, and they are often used for C–H functionalization reactions. In 2015, the Bi group reported a silver-catalyzed cross coupling of isocyanides 72 and active methylene compounds 73 to access various β-aminoenones 74 and tricarbonylmethanes 75 by a radical process (Scheme 43) [61].
To determine the reaction mechanism, both isotope-labeling studies and control experiments with radical scavengers were performed, and the reaction did not proceed in the presence of radical scavengers. Based on the deuterium-labeling experiments, hydrogen atom transfer occurred from the α-hydrogen of methylene compound 72 to the hydrogen on the olefin of product 74. In the 18O-labeling experiments, 18O2 provided a high degree of 18O-incorporated product 75, but H218O did not produce any 18O-incorporated 75. Based on these experimental results, a plausible mechanism was proposed as described in Scheme 44. Ag2CO3 abstracts a proton from 73 due to its basicity, and subsequent oxidation by Ag(I) generates radical intermediate 73-I. Ag2CO3 has a dual role as a base and a one-electron oxidant. At the same time, Ag2CO3 coordinates to isocyanide 72 to form 72-I. The coupling reaction between radical 73-I and silver complex 72-I produces imidoyl radical 72-II. With an aromatic isocyanide, imidoyl radical 72-II can provide imine intermediate 72-IV by abstracting a H atom from the generated AgHCO3 during the one-electron oxidation or by a 1,2-H migration to form stable tricarbonylmethenyl radical 72-III followed by abstraction of an H atom from AgHCO3. Imine 72-IV gives β-aminoenone 74 by tautomerization. With tert-butyl isocyanide, radical 72-II is converted to hydroperoxide 72-V by oxygenation. Hydroperoxide 72-V is reduced by Ag(s) to form oxyanion intermediate 72-VI, and the protonation gives tricarbonylmethane 75 as the product.
Zou and co-workers reported a silver-catalyzed direct C(sp2)–H phosphorylation of indoles in the presence of Mg(NO3)2 leading to phosphoindoles (Scheme 45) [62]. Ag2CO3 is used as an oxidant and Mg(NO3)2 is used as an additive to form a new C–P bond. The reaction occurs by the addition of the electrophilic phosphorous radical to the available C(sp2)–H position of the indole, followed by a SET by Ag(I) and aromatization by deprotonation to give the phosphoindole product.
Recently, the Bi group reported the Ag(I)-catalyzed synthesis of 1,2,4,5-tetrasubstituted pyrroles 80 via the cascade reaction of β-enaminones 78 and isocyanoacetates 79 (Scheme 46) [63]. β-Enaminones are used as building blocks in a variety of reactions and are easily obtained. One of the key features of β-enaminones is their tautomeric equilibrium with β-ketoimines. As β-ketoimines are a minor product, imines are present in this equilibrium system.
To investigate the reaction pathway, a deuterium-labeling experiment was performed using 2.0 equiv. of D2O. As shown in Scheme 47, deuterium-labeled pyrrole 80-D was obtained, and the reaction should involve a proton transfer between the imine-metal complex and D2O. Based on this experimental result, the reaction may be initiated by the deprotonation of isocyanoacetate 79 by basic Ag2CO3 to form an α-metalated isocyanoacetate. Then, [3+2] dipolar cycloaddition with β-ketoimine follows, and the β-ketoimine was derived from β-enaminone 78. The retro-hetero-Michael addition ring-opens the 2-imidazoline (78-I) to give the imidamide intermediate, and then a cascade involving nucleophilic addition, proton transfer, and dehydration provides highly substituted pyrrole 80 as the product.
In 2017, Xu and co-workers developed a silver-catalyzed chemoselective [4+2] annulation of aryl and heteroaryl isocyanides 72 with α-substituted isocyanoacetamides 81 to afford pyridone-fused carbo- and heterocycles 82 (Scheme 48) [64]. To determine the reaction mechanism, the reaction was performed with a 13C labeled isocyanide group in 72, and the NMR spectrum showed the benzo[h]quinolone product 82 with 13C at the 2-position. This result indicated that an α-amidoketenimine is a likely reaction intermediate.
Based on the 13C-labeling experiments, a mechanism was proposed as described in Scheme 49. Ag2CO3 abstracts a proton from isocyanoactamide 81 and coordinates to the isocyanide moiety. Another equivalent of Ag2CO3 coordinates to isocyanide 72, and the subsequent C–C coupling occurs via a nucleophilic attack and elimination to generate amidoketenimine 72-VII. The rearrangement of α-amidoketenimine 72-VII to α-imidoylketene 72-VIII occurs via a 1,3-amino migration. Subsequent 6π electrocyclization and 1,3-proton shift convert 72-VIII to quinolone 82. In this reaction, a catalytic amount of Li2CO3 was used as an additive to promote the regeneration of the Ag2CO3 catalyst.
In 2017, the Xu group developed a formal [3+2] cycloaddition of α-trifluoromethylated methyl isocyanides 84 and aldehydes 83 for the silver-catalyzed divergent synthesis of trifluoromethylated heterocycles (Scheme 50) [65]. With imines or acrylonitriles instead of aldehyde 83, the reaction produces the corresponding imidazoles or pyrrolines as the major products. Ag(I) coordinates to isocyanide 84, then DBU (1,8-diazabicyclo[5.4.0]undec-7-ene) abstracts a proton to form the carbanion nucleophile. A formal [3+2] cycloaddition, which consists of the nucleophilic addition of the carbanion from the isocyanide to aldehyde 83 and cyclization, affords an oxazoline silver complex; then, protodemetallation gives oxazoline 85.
In 2017, Wang and co-workers reported the synthesis of pyrroles via a tandem Ag-catalyzed 1,3-dipolar cycloaddition and BPO (benzoyl peroxide)-mediated oxidative dehydrogenative aromatization (Scheme 51) [66]. Ag2CO3 facilitates the [3+2] cycloaddition of the 1,3-dipolar intermediate generated from compound 86 by deprotonation of the α-position with alkene 87 to form a tetrasubstituted pyrrolidine. Peroxide abstracts a H atom from the α-position of the pyrrolidine, then two oxidations, which each consists of a SET and tautomerization, provide highly substituted pyrrole 88.

5. Conclusions

In this review, we summarized the recent studies and developments in Ag2CO3-catalyzed/mediated organic transformations. Silver carbonates were employed as either external bases for activating acidic molecules or an external oxidant for turning over catalytic cycles. Notably, heteroaromatic compounds can be elegantly synthesized using silver carbonate to combine heteroatom-containing fragments and more than two carbon skeletons through alkyne activation. At the same time, silver carboxylate can be generated from the simple reaction between silver carbonates and various carboxylic acids. The following cross-couplings and decarboxylative reactions were investigated.
Despite the substantial progress and important advances described in these reports, there is still significant room for improvement at the present level. For example, alkene-related substrates cannot be efficiently activated by silver carbonate. The utilization of an alkene could expand the possible synthetic targets of these silver carbonate-catalyzed reactions. Further detailed studies on the transmetallation from silver to other transition/main group metals should be conducted. In addition, detailed mechanism studies, including the structural determination of various intermediates should be performed. We hope that this review will inspire the development of various silver carbonate-catalyzed organic reactions and other related methodologies.

Author Contributions

K.Y., D.G.J., and H.-E.L. contributed equally to this work. K.Y., D.G.J. and M.K. designed and conceptualized the original project. K.Y., D.G.J., and H.J.K. categorized target reactions. H.-E.L. and C.K. built the mechanism discussion. All authors wrote the manuscript together.

Funding

This project was supported by the Basic Science Research Program (2019R1A2C4070584) and the Science Research Center (2016R1A5A1009405) through the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT. K. Yoo was supported by NRF Global Ph.D. Fellowship program (NRF-2017H1A2A1045495).

Acknowledgments

Sung Min Kang is gratefully acknowledged for discussion.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Beller, M.; Bolm, C. Transition Metals for Organic Synthesis: Buidling Blocks and Find Chemicals, 2nd rev. and enl. ed.; WILEY-VCH: Weinheim, Germany, 2004; ISBN 3527306137. [Google Scholar]
  2. Patil, N.T.; Yamamoto, Y. Coinage Metal-Assisted Synthesis of Heterocycles. Chem. Rev. 2008, 108, 3395–3442. [Google Scholar] [CrossRef] [PubMed]
  3. Zheng, Q.Z.; Jiao, N. Ag-catalyzed C-H/C-C bond functionalization. Chem. Soc. Rev. 2016, 45, 4590–4627. [Google Scholar] [CrossRef] [PubMed]
  4. Weibel, J.M.; Blanc, A.; Pale, P. Ag-mediated reactions: Coupling and heterocyclization reactions. Chem. Rev. 2008, 108, 3149–3173. [Google Scholar] [CrossRef] [PubMed]
  5. Halbes-Letinois, U.; Weibel, J.M.; Pale, P. The organic chemistry of silver acetylides. Chem. Soc. Rev. 2007, 36, 759–769. [Google Scholar] [CrossRef]
  6. Harmata, M. Silver in Organic Chemistry; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2010; ISBN 9780470466117. [Google Scholar]
  7. Naodovic, M.; Yamamoto, H. Asymmetric silver-catalyzed reactions. Chem. Rev. 2008, 108, 3132–3148. [Google Scholar] [CrossRef]
  8. Sekine, K.; Yamada, T. Silver-catalyzed carboxylation. Chem. Soc. Rev. 2016, 45, 4524–4532. [Google Scholar] [CrossRef] [Green Version]
  9. Fetizon, M.; Balogh, V.; Golfier, M. Oxidations with silver carbonate/celite. V. Oxidations of phenols and related compounds. J. Org. Chem. 1971, 36, 1339–1341. [Google Scholar] [CrossRef]
  10. Jedinak, L.; Rush, L.; Lee, M.; Hesek, D.; Fisher, J.F.; Boggess, B.; Noll, B.C.; Mobashery, S. Use of Silver Carbonate in the Wittig Reaction. J. Org. Chem. 2013, 78, 12224–12228. [Google Scholar] [CrossRef] [Green Version]
  11. Boto, A.; Alvares, L. Furan and Its Derivatives. In Heterocycles in Natural Product Synthesis; Majumdar, K.C., Chattopadhyay, S.K., Eds.; Wiley-VCH: Weinheim, Germany, 2011; pp. 97–152. ISBN 3527634894. [Google Scholar]
  12. Young, I.S.; Thornton, P.D.; Thompson, A. Synthesis of natural products containing the pyrrolic ring. Nat. Prod. Rep. 2010, 27, 1801–1839. [Google Scholar] [CrossRef]
  13. Fleming, F.F.; Yao, L.; Ravikumar, P.C.; Funk, L.; Shook, B.C. Nitrile-Containing Pharmaceuticals: Efficacious Roles of the Nitrile Pharmacophore. J. Med. Chem. 2010, 53, 7902–7917. [Google Scholar] [CrossRef] [Green Version]
  14. He, C.; Guo, S.; Ke, J.; Hao, J.; Xu, H.; Chen, H.; Lei, A. Silver-mediated oxidative C-H/C-H functionalization: A strategy to construct polysubstituted furans. J. Am. Chem. Soc. 2012, 134, 5766–5769. [Google Scholar] [CrossRef] [PubMed]
  15. Moran, W.J.; Rodríguez, A. Metal-catalyzed furan synthesis. A review. Org. Prep. Proced. Int. 2012, 44, 103–130. [Google Scholar] [CrossRef] [Green Version]
  16. Hay, A. Oxidative Coupling of Acetylenes. J. Org. Chem. 1960, 25, 1275–1276. [Google Scholar] [CrossRef]
  17. He, C.; Hao, J.; Xu, H.; Mo, Y.; Liu, H.; Han, J.; Lei, A. Heteroaromatic imidazo[1,2-a]pyridines synthesis from C-H/N-H oxidative cross-coupling/cyclization. Chem. Commun. 2012, 48, 11073–11075. [Google Scholar] [CrossRef] [PubMed]
  18. Ke, J.; He, C.; Liu, H.; Li, M.; Lei, A. Oxidative cross-coupling/cyclization to build polysubstituted pyrroles from terminal alkynes and β-enamino esters. Chem. Commun. 2013, 49, 7549–7551. [Google Scholar] [CrossRef]
  19. Liu, J.; Liu, Z.; Wu, N.; Liao, P.; Bi, X. Silver-catalyzed cross-coupling of propargylic alcohols with isocyanides: An atom-economical synthesis of 2,3-ALLENAMIDES. Chem. Eur. J. 2014, 20, 2154–2158. [Google Scholar] [CrossRef]
  20. Pandya, A.N.; Fletcher, J.T.; Villa, E.M.; Agrawal, D.K. Silver-mediated synthesis of indolizines via oxidative C-H functionalization and 5-endo-dig cyclization. Tetrahedron Lett. 2014, 55, 6922–6924. [Google Scholar] [CrossRef] [Green Version]
  21. Sadowski, B.; Klajn, J.; Gryko, D.T. Recent advances in the synthesis of indolizines and their π-expanded analogues. Org. Biomol. Chem. 2016, 14, 7804–7828. [Google Scholar] [CrossRef] [Green Version]
  22. Wang, T.; Chen, S.; Shao, A.; Gao, M.; Huang, Y.; Lei, A. Silver-Mediated Selective Oxidative Cross-Coupling between C-H/P-H: A Strategy to Construct Alkynyl(diaryl)phosphine Oxide. Org. Lett. 2015, 17, 118–121. [Google Scholar] [CrossRef]
  23. Liu, J.; Fang, Z.; Zhang, Q.; Liu, Q.; Bi, X. Silver-catalyzed isocyanide-alkyne cycloaddition: A general and practical method to oligosubstituted pyrroles. Angew. Chem. Int. Ed. 2013, 52, 6953–6957. [Google Scholar] [CrossRef]
  24. Gao, M.; He, C.; Chen, H.; Bai, R.; Cheng, B.; Lei, A. Synthesis of pyrroles by click reaction: Silver-catalyzed cycloaddition of terminal alkynes with isocyanides. Angew. Chem. Int. Ed. 2013, 52, 6958–6961. [Google Scholar] [CrossRef] [PubMed]
  25. Liang, L.; Astruc, D. The copper(I)-catalyzed alkyne-azide cycloaddition (CuAAC) “click” reaction and its applications. An overview. Coord. Chem. Rev. 2011, 255, 2933–2945. [Google Scholar] [CrossRef]
  26. Meng, X.; Liao, P.; Liu, J.; Bi, X. Silver-catalyzed cyclization of 2-pyridyl alkynyl carbinols with isocyanides: Divergent synthesis of indolizines and pyrroles. Chem. Commun. 2014, 50, 11837–11839. [Google Scholar] [CrossRef] [PubMed]
  27. Shen, T.; Wang, T.; Qin, C.; Jiao, N. Silver-catalyzed nitrogenation of alkynes: A direct approach to nitriles through C≡C bond cleavage. Angew. Chem. Int. Ed. 2013, 52, 6677–6680. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, Z.; Liu, J.; Zhang, L.; Liao, P.; Song, J.; Bi, X. Silver(I)-catalyzed hydroazidation of ethynyl carbinols: Synthesis of 2-azidoallyl alcohols. Angew. Chem. Int. Ed. 2014, 53, 5305–5309. [Google Scholar] [CrossRef] [PubMed]
  29. Barnert, K. The Chemistry of Vinyl, Allenyl, and Ethynyl Azides. In Organic Azides: Syntheses and Applications; Bräse, S., Banert, K., Eds.; John Wiley & Sons Ltd.: Chichester, UK, 2010; pp. 115–156. ISBN 0470682523. [Google Scholar]
  30. Liu, Z.; Liao, P.; Bi, X. General silver-catalyzed hydroazidation of terminal alkynes by combining TMS-N3 and H2O: Synthesis of vinyl azides. Org. Lett. 2014, 16, 3668–3671. [Google Scholar] [CrossRef]
  31. Ning, Y.; Wu, N.; Yu, H.; Liao, P.; Li, X.; Bi, X. Silver-catalyzed tandem hydroazidation/alkyne-azide cycloaddition of diynes with TMS-N3: An easy access to 1,5-fused 1,2,3-triazole frameworks. Org. Lett. 2015, 17, 2198–2201. [Google Scholar] [CrossRef]
  32. Kumar, Y.K.; Kumar, G.R.; Reddy, T.J.; Sridhar, B.; Reddy, M.S. Synthesis of 3-Sulfonylamino Quinolines from 1-(2-Aminophenyl) Propargyl Alcohols through a Ag(I)-Catalyzed Hydroamination, (2+3) Cycloaddition, and an Unusual Strain-Driven Ring Expansion. Org. Lett. 2015, 17, 2226–2229. [Google Scholar] [CrossRef]
  33. Peng, A.-Y.; Ding, Y.-X. Synthesis of 2H-1,2-Oxaphosphorin 2-Oxides via Ag2CO3-Catalyzed Cyclization of (Z)-2-Alken-4-ynylphosphonic Monoesters. Org. Lett. 2005, 7, 3299–3301. [Google Scholar] [CrossRef]
  34. Anastasia, L.; Xu, C.; Negishi, E.I. Catalytic and selective conversion of (Z)-2-en-4-ynoic acids to either 2H-pyran-2-ones in the presence of ZnBr2 or (Z)-5-alkylidenefuran-2(5H)-ones in the presence of Ag2CO3. Tetrahedron Lett. 2002, 43, 5673–5676. [Google Scholar] [CrossRef]
  35. Tan, X.C.; Liang, Y.; Bao, F.P.; Wang, H.S.; Pan, Y.M. Silver-mediated C-H bond functionalization: One-pot to construct substituted indolizines from 2-alkylazaarenes with alkynes. Tetrahedron 2014, 70, 6717–6722. [Google Scholar] [CrossRef]
  36. Mi, X.; Wang, C.; Huang, M.; Zhang, J.; Wu, Y.; Wu, Y. Silver-catalyzed synthesis of 3-phosphorated coumarins via radical cyclization of alkynoates and dialkyl H -phosphonates. Org. Lett. 2014, 16, 3356–3359. [Google Scholar] [CrossRef] [PubMed]
  37. Ahmar, S.F.E. Expedient Synthesis of Complex γ-Butyrolactones from 5-(1-Arylalkylidene) Meldrum’s Acids via Sequential Conjugate Alkynylation/Ag(I)-Catalyzed Lactonization. Org. Lett. 2014, 16, 5748–5751. [Google Scholar] [CrossRef] [PubMed]
  38. Suresh Kumar, E.; Etukala, J.; Ablordeppey, S. Indolo[3,2-b]quinolines: Synthesis, Biological Evaluation and Structure Activity-Relationships. Mini-Rev. Med. Chem. 2008, 8, 538–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Du, W.; Curran, D.P. Synthesis of Carbocyclic and Heterocyclic Fused Quinolines by Cascade Radical Annulations of Unsaturated N-Aryl Thiocarbamates, Thioamides, and Thioureas. Org. Lett. 2003, 5, 1765–1768. [Google Scholar] [CrossRef]
  40. Ali, W.; Dahiya, A.; Pandey, R.; Alam, T.; Patel, B.K. Microwave-Assisted Cascade Strategy for the Synthesis of Indolo[2,3-b]quinolines from 2-(Phenylethynyl)anilines and Aryl Isothiocynates. J. Org. Chem. 2017, 82, 2089–2096. [Google Scholar] [CrossRef]
  41. Cornella, J.; Sanchez, C.; Banawa, D.; Larrosa, I. Silver-catalysed protodecarboxylation of ortho-substituted benzoic acids. Chem. Commun. 2009, 7176–7178. [Google Scholar] [CrossRef]
  42. Lu, P.; Sanchez, C.; Cornella, J.; Larrosa, I. Silver-catalyzed protodecarboxylation of heteroaromatic carboxylic acids. Org. Lett. 2009, 11, 5710–5713. [Google Scholar] [CrossRef]
  43. Jafarpour, F.; Jalalimanesh, N.; Olia, M.B.A.; Kashani, A.O. Silver-catalyzed facile decarboxylation of coumarin-3-carboxylic acids. Tetrahedron 2010, 66, 9508–9511. [Google Scholar] [CrossRef]
  44. Hartwig, J.F. Transition Metal Catalyzed Synthesis of Arylamines and Aryl Ethers from Aryl Halides and Triflates: Scope and Mechanism. Angew. Chem. Int. Ed. 1998, 37, 2046–2067. [Google Scholar] [CrossRef]
  45. Luo, Y.; Pan, X.; Wu, J. Silver-catalyzed decarboxylative halogenation of carboxylic acids. Tetrahedron Lett. 2010, 51, 6646–6648. [Google Scholar] [CrossRef]
  46. Suresh, R.; Kumaran, R.S.; Senthilkumar, V.; Muthusubramanian, S. Silver catalyzed decarboxylative acylation of pyridine-N-oxides using α-oxocarboxylic acids. RSC Adv. 2014, 4, 31685–31688. [Google Scholar] [CrossRef]
  47. Cheng, K.; Zhao, B.; Qi, C. Silver-catalyzed decarboxylative acylation of arylglyoxylic acids with arylboronic acids. RSC Adv. 2014, 4, 48698–48702. [Google Scholar] [CrossRef]
  48. Shang, R.; Liu, L. Transition metal-catalyzed decarboxylative cross-coupling reactions. Sci. China Chem. 2011, 54, 1670–1687. [Google Scholar] [CrossRef]
  49. Liu, J.; Fan, C.; Yin, H.; Qin, C.; Zhang, G.; Zhang, X.; Yi, H.; Lei, A. Synthesis of 6-acyl phenanthridines by oxidative radical decarboxylation-cyclization of α-oxocarboxylates and isocyanides. Chem. Commun. 2014, 50, 2145–2147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Yoo, K.; Park, H.; Kim, S.; Kim, M. Temperature-controlled acyloxylations and hydroxylations of bromoarene by a silver salt. Tetrahedron Lett. 2016, 57, 781–783. [Google Scholar] [CrossRef]
  51. Leblanc, Y.; Fitzsimmons, B.J. [4+2] Cycloaddition reaction of bis (trichloroethyl) azodicarboxylate and glycals: Preparation of a C1-C1 2-amino disaccharide. Tetrahedron Lett. 1989, 30, 2889–2892. [Google Scholar] [CrossRef]
  52. Trimble, L.A.; Vederas, J.C. Amination of chiral enolates by dialkyl azodiformates. Synthesis of .alpha.-hydrazino acids and .alpha.-amino acids. J. Am. Chem. Soc. 1986, 108, 6397–6399. [Google Scholar] [CrossRef]
  53. Lee, K.S.; Lim, Y.K.; Cho, C.G. Ag+ mediated deaminations of N-Boc aryl hydrazines for the efficient synthesis of N-Boc aryl amines. Tetrahedron Lett. 2002, 43, 7463–7464. [Google Scholar] [CrossRef]
  54. Hili, R.; Yudin, A.K. Making carbon-nitrogen bonds in biological and chemical synthesis. Nat. Chem. Biol. 2006, 2, 284–287. [Google Scholar] [CrossRef]
  55. Müller, T.E.; Hultzsch, K.C.; Yus, M.; Foubelo, F.; Tada, M. Hydroamination: Direct Addition of Amines to Alkenes and Alkynes. Chem. Rev. 2008, 108, 3795–3892. [Google Scholar] [CrossRef] [PubMed]
  56. Louillat, M.L.; Patureau, F.W. Oxidative C-H amination reactions. Chem. Soc. Rev. 2014, 43, 901–910. [Google Scholar] [CrossRef] [PubMed]
  57. Cho, S.H.; Kim, J.Y.; Lee, S.Y.; Chang, S. Silver-mediated direct amination of benzoxazoles: Tuning the amino group source from formamides to parent amines. Angew. Chem. Int. Ed. 2009, 48, 9127–9130. [Google Scholar] [CrossRef] [PubMed]
  58. Sarkar, R.; Mukhopadhyay, C. Silver-mediated Cα(sp3)-H functionalization of primary amines: An oxidative C-N coupling strategy for the synthesis of two different types of 1,2,4,5-tetrasubstituted imidazoles. Eur. J. Org. Chem. 2015, 2015, 1246–1256. [Google Scholar] [CrossRef]
  59. Youn, S.W.; Ko, T.Y.; Jang, M.J.; Jang, S.S. Silver(I)-mediated C-H amination of 2-alkenylanilines: Unique solvent-dependent migratory aptitude. Adv. Synth. Catal. 2015, 357, 227–234. [Google Scholar] [CrossRef]
  60. Qiu, G.; Ding, Q.; Wu, J. Recent advances in isocyanide insertion chemistry. Chem. Soc. Rev. 2013, 42, 5257–5269. [Google Scholar] [CrossRef]
  61. Liu, J.; Liu, Z.; Liao, P.; Zhang, L.; Tu, T.; Bi, X. Silver-Catalyzed Cross-Coupling of Isocyanides and Active Methylene Compounds by a Radical Process. Angew. Chem. Int. Ed. 2015, 54, 10618–10622. [Google Scholar] [CrossRef]
  62. Sun, W.B.; Xue, J.F.; Zhang, G.Y.; Zeng, R.S.; An, L.T.; Zhang, P.Z.; Zou, J.P. Silver-Catalyzed Direct Csp2-H Phosphorylation of Indoles Leading to Phosphoindoles. Adv. Synth. Catal. 2016, 358, 1753–1758. [Google Scholar] [CrossRef]
  63. Fang, G.; Liu, J.; Fu, J.; Liu, Q.; Bi, X. Silver-Catalyzed Cascade Reaction of β-Enaminones and Isocyanoacetates to Construct Functionalized Pyrroles. Org. Lett. 2017, 19, 1346–1349. [Google Scholar] [CrossRef]
  64. Hu, Z.; Dong, J.; Men, Y.; Lin, Z.; Cai, J.; Xu, X. Silver-Catalyzed Chemoselective [4+2] Annulation of Two Isocyanides: A General Route to Pyridone-Fused Carbo- and Heterocycles. Angew. Chem. Int. Ed. 2017, 56, 1805–1809. [Google Scholar] [CrossRef]
  65. Zhang, X.; Wang, X.; Gao, Y.; Xu, X. Silver-catalyzed formal [3+2]-cycloaddition of α-trifluoromethylated methyl isocyanides: A facile stereoselective synthesis of CF3-substituted heterocycles. Chem. Commun. 2017, 53, 2427–2430. [Google Scholar] [CrossRef] [PubMed]
  66. Liu, Y.; Hu, H.; Wang, X.; Zhi, S.; Kan, Y.; Wang, C. Synthesis of Pyrrole via a Silver-Catalyzed 1,3-Dipolar Cycloaddition/Oxidative Dehydrogenative Aromatization Tandem Reaction. J. Org. Chem. 2017, 82, 4194–4202. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. The role of Ag(I) in transformations of alkynes and carboxylic acids.
Scheme 1. The role of Ag(I) in transformations of alkynes and carboxylic acids.
Catalysts 09 01032 sch001
Scheme 2. Silver-mediated oxidative C–H/C–H functionalization process to afford highly substituted furans.
Scheme 2. Silver-mediated oxidative C–H/C–H functionalization process to afford highly substituted furans.
Catalysts 09 01032 sch002
Scheme 3. Ag-mediated C–H/N–H oxidative cross-coupling/cyclization.
Scheme 3. Ag-mediated C–H/N–H oxidative cross-coupling/cyclization.
Catalysts 09 01032 sch003
Scheme 4. Synthesis of polysubstituted pyrroles via Ag-mediated oxidative cyclization.
Scheme 4. Synthesis of polysubstituted pyrroles via Ag-mediated oxidative cyclization.
Catalysts 09 01032 sch004
Scheme 5. Proposed mechanism for Ag-mediated pyrrole synthesis. Chem. Commun. 2013, 49, 7549–7551, doi:10.1039/c3cc43682a—Reproduced by permission of The Royal Society of Chemistry.
Scheme 5. Proposed mechanism for Ag-mediated pyrrole synthesis. Chem. Commun. 2013, 49, 7549–7551, doi:10.1039/c3cc43682a—Reproduced by permission of The Royal Society of Chemistry.
Catalysts 09 01032 sch005
Scheme 6. Ag-catalyzed cross coupling of propargylic alcohols with isocyanides.
Scheme 6. Ag-catalyzed cross coupling of propargylic alcohols with isocyanides.
Catalysts 09 01032 sch006
Scheme 7. Ag-mediated oxidative C–H functionalization to synthesize indolizines.
Scheme 7. Ag-mediated oxidative C–H functionalization to synthesize indolizines.
Catalysts 09 01032 sch007
Scheme 8. Ag-mediated selective oxidative cross coupling between C–H and P–H. Adapted with permission from Org. Lett. 2015, 17, 118–121, doi:10.1021/ol503341t. Copyright (2019) American Chemical Society.
Scheme 8. Ag-mediated selective oxidative cross coupling between C–H and P–H. Adapted with permission from Org. Lett. 2015, 17, 118–121, doi:10.1021/ol503341t. Copyright (2019) American Chemical Society.
Catalysts 09 01032 sch008
Scheme 9. Proposed mechanism of the Ag-mediated oxidative cross coupling between C–H and P–H.
Scheme 9. Proposed mechanism of the Ag-mediated oxidative cross coupling between C–H and P–H.
Catalysts 09 01032 sch009
Scheme 10. Ag-catalyzed cycloaddition of alkynes and isocyanides.
Scheme 10. Ag-catalyzed cycloaddition of alkynes and isocyanides.
Catalysts 09 01032 sch010
Scheme 11. Ag-catalyzed cyclization of 2-pyridyl alkynyl carbinols with isocyanides.
Scheme 11. Ag-catalyzed cyclization of 2-pyridyl alkynyl carbinols with isocyanides.
Catalysts 09 01032 sch011
Scheme 12. Proposed mechanism for Ag-catalyzed cyclization of 2-pyridyl alkynyl carbinols with isocyanides. Chem. Commun. 2014, 50, 11837–11839, doi:10.1039/c4cc04905e—Reproduced by permission of The Royal Society of Chemistry.
Scheme 12. Proposed mechanism for Ag-catalyzed cyclization of 2-pyridyl alkynyl carbinols with isocyanides. Chem. Commun. 2014, 50, 11837–11839, doi:10.1039/c4cc04905e—Reproduced by permission of The Royal Society of Chemistry.
Catalysts 09 01032 sch012
Scheme 13. Ag-catalyzed nitrogenation of alkynes.
Scheme 13. Ag-catalyzed nitrogenation of alkynes.
Catalysts 09 01032 sch013
Scheme 14. Proposed mechanism of the Ag-catalyzed nitrogenation.
Scheme 14. Proposed mechanism of the Ag-catalyzed nitrogenation.
Catalysts 09 01032 sch014
Scheme 15. Ag-catalyzed hydroazidation of ethynyl carbinols.
Scheme 15. Ag-catalyzed hydroazidation of ethynyl carbinols.
Catalysts 09 01032 sch015
Scheme 16. Effect of water as an additive in the Ag-catalyzed hydroazidation.
Scheme 16. Effect of water as an additive in the Ag-catalyzed hydroazidation.
Catalysts 09 01032 sch016
Scheme 17. Ag-catalyzed hydroazidation of propargyl compounds. Adapted with permission from Org. Lett. 2014, 16, 3668–3671, doi:10.1021/ol501661k. Copyright (2019) American Chemical Society.
Scheme 17. Ag-catalyzed hydroazidation of propargyl compounds. Adapted with permission from Org. Lett. 2014, 16, 3668–3671, doi:10.1021/ol501661k. Copyright (2019) American Chemical Society.
Catalysts 09 01032 sch017
Scheme 18. Ag-catalyzed tandem hydroazidation/alkyne-azide cycloaddition of diynes with TMS-N3. Adapted with permission from Org. Lett. 2015, 17, 2198–2201, doi:10.1021/acs.orglett.5b00784. Copyright (2019) American Chemical Society.
Scheme 18. Ag-catalyzed tandem hydroazidation/alkyne-azide cycloaddition of diynes with TMS-N3. Adapted with permission from Org. Lett. 2015, 17, 2198–2201, doi:10.1021/acs.orglett.5b00784. Copyright (2019) American Chemical Society.
Catalysts 09 01032 sch018
Scheme 19. Ag-catalyzed hydroamination, [2+3] cycloaddition and ring expansion.
Scheme 19. Ag-catalyzed hydroamination, [2+3] cycloaddition and ring expansion.
Catalysts 09 01032 sch019
Scheme 20. Ag-catalyzed cyclization of (Z)-2-alken-4-ynyl-phosphonic monoesters. Adapted with permission from Org. Lett. 2005, 7, 3299–3301, doi:10.1021/OL051126+. Copyright (2019) American Chemical Society.
Scheme 20. Ag-catalyzed cyclization of (Z)-2-alken-4-ynyl-phosphonic monoesters. Adapted with permission from Org. Lett. 2005, 7, 3299–3301, doi:10.1021/OL051126+. Copyright (2019) American Chemical Society.
Catalysts 09 01032 sch020
Scheme 21. Ag-mediated one-pot synthesis of indolizines from alkynes and 2-alkylazaarenes.
Scheme 21. Ag-mediated one-pot synthesis of indolizines from alkynes and 2-alkylazaarenes.
Catalysts 09 01032 sch021
Scheme 22. Proposed mechanisms for Ag-mediated indolizine synthesis with terminal and internal alkynes.
Scheme 22. Proposed mechanisms for Ag-mediated indolizine synthesis with terminal and internal alkynes.
Catalysts 09 01032 sch022
Scheme 23. Ag-catalyzed phosphorus carbocyclization of alkynoates. Adapted with permission from Org. Lett. 2014, 16, 3356–3359, doi:10.1021/ol5013839. Copyright (2019) American Chemical Society.
Scheme 23. Ag-catalyzed phosphorus carbocyclization of alkynoates. Adapted with permission from Org. Lett. 2014, 16, 3356–3359, doi:10.1021/ol5013839. Copyright (2019) American Chemical Society.
Catalysts 09 01032 sch023
Scheme 24. Proposed mechanism for the Ag-catalyzed phosphorus carbocyclization.
Scheme 24. Proposed mechanism for the Ag-catalyzed phosphorus carbocyclization.
Catalysts 09 01032 sch024
Scheme 25. Ag-catalyzed formation of alkylidene γ-butyrolactones.
Scheme 25. Ag-catalyzed formation of alkylidene γ-butyrolactones.
Catalysts 09 01032 sch025
Scheme 26. Ag-catalyzed cascade cyclization by the reaction of 2-(phenylethynyl)aniline and isothiocyanate. Adapted with permission from J. Org. Chem. 2017, 82, 2089–2096, doi:10.1021/acs.joc.6b02912. Copyright (2019) American Chemical Society.
Scheme 26. Ag-catalyzed cascade cyclization by the reaction of 2-(phenylethynyl)aniline and isothiocyanate. Adapted with permission from J. Org. Chem. 2017, 82, 2089–2096, doi:10.1021/acs.joc.6b02912. Copyright (2019) American Chemical Society.
Catalysts 09 01032 sch026
Scheme 27. Ag-catalyzed decarboxylation of ortho-substituted benzoic acids.
Scheme 27. Ag-catalyzed decarboxylation of ortho-substituted benzoic acids.
Catalysts 09 01032 sch027
Scheme 28. Proposed mechanism for decarboxylation of ortho-substituted benzoic acids.
Scheme 28. Proposed mechanism for decarboxylation of ortho-substituted benzoic acids.
Catalysts 09 01032 sch028
Scheme 29. Protodecarboxylation of heteroaromatic carboxylic acids.
Scheme 29. Protodecarboxylation of heteroaromatic carboxylic acids.
Catalysts 09 01032 sch029
Scheme 30. Ag-catalyzed protodecarboxylation of coumarin-3-carboxylic acid.
Scheme 30. Ag-catalyzed protodecarboxylation of coumarin-3-carboxylic acid.
Catalysts 09 01032 sch030
Scheme 31. Ag-catalyzed decarboxylative halogenation of carboxylic acids.
Scheme 31. Ag-catalyzed decarboxylative halogenation of carboxylic acids.
Catalysts 09 01032 sch031
Scheme 32. Ag-catalyzed decarboxylative acylation of pyridine N-oxides.
Scheme 32. Ag-catalyzed decarboxylative acylation of pyridine N-oxides.
Catalysts 09 01032 sch032
Scheme 33. Ag-catalyzed cross coupling of arylboronic acids with arylglyoxylic acids.
Scheme 33. Ag-catalyzed cross coupling of arylboronic acids with arylglyoxylic acids.
Catalysts 09 01032 sch033
Scheme 34. Oxidative radical decarboxylation/cyclization of α-oxocarboxylates and isocyanides. Chem. Commun. 2014, 50, 2145–2147, doi:10.1039/c3cc49026b—Reproduced by permission of The Royal Society of Chemistry.
Scheme 34. Oxidative radical decarboxylation/cyclization of α-oxocarboxylates and isocyanides. Chem. Commun. 2014, 50, 2145–2147, doi:10.1039/c3cc49026b—Reproduced by permission of The Royal Society of Chemistry.
Catalysts 09 01032 sch034
Scheme 35. Proposed mechanism of the Ag-catalyzed oxidative radical decarboxylation/cyclization. Chem. Commun. 2014, 50, 2145–2147, doi:10.1039/c3cc49026b—Reproduced by permission of The Royal Society of Chemistry.
Scheme 35. Proposed mechanism of the Ag-catalyzed oxidative radical decarboxylation/cyclization. Chem. Commun. 2014, 50, 2145–2147, doi:10.1039/c3cc49026b—Reproduced by permission of The Royal Society of Chemistry.
Catalysts 09 01032 sch035
Scheme 36. Temperature-controlled acyloxylation and hydroxylation of bromoarene.
Scheme 36. Temperature-controlled acyloxylation and hydroxylation of bromoarene.
Catalysts 09 01032 sch036
Scheme 37. Ag-mediated deamination of N-Boc aryl hydrazines.
Scheme 37. Ag-mediated deamination of N-Boc aryl hydrazines.
Catalysts 09 01032 sch037
Scheme 38. Ag-mediated direct amination of benzoxazoles.
Scheme 38. Ag-mediated direct amination of benzoxazoles.
Catalysts 09 01032 sch038
Scheme 39. Proposed mechanism for the Ag-mediated direct amination of benzoxazoles.
Scheme 39. Proposed mechanism for the Ag-mediated direct amination of benzoxazoles.
Catalysts 09 01032 sch039
Scheme 40. Ag-mediated C–N coupling for the synthesis of imidazoles.
Scheme 40. Ag-mediated C–N coupling for the synthesis of imidazoles.
Catalysts 09 01032 sch040
Scheme 41. Proposed mechanism for the Ag-mediated C-N coupling to afford imidazoles.
Scheme 41. Proposed mechanism for the Ag-mediated C-N coupling to afford imidazoles.
Catalysts 09 01032 sch041
Scheme 42. Ag-mediated oxidative C–H amination of 2-alkenylanilines.
Scheme 42. Ag-mediated oxidative C–H amination of 2-alkenylanilines.
Catalysts 09 01032 sch042
Scheme 43. Ag-catalyzed cross coupling of isocyanides and activated methylene compounds.
Scheme 43. Ag-catalyzed cross coupling of isocyanides and activated methylene compounds.
Catalysts 09 01032 sch043
Scheme 44. Proposed mechanism for Ag-catalyzed cross-couplings with isocyanide.
Scheme 44. Proposed mechanism for Ag-catalyzed cross-couplings with isocyanide.
Catalysts 09 01032 sch044
Scheme 45. Ag-catalyzed direct Csp2–H phosphorylation of 2- and 3-substituted indoles.
Scheme 45. Ag-catalyzed direct Csp2–H phosphorylation of 2- and 3-substituted indoles.
Catalysts 09 01032 sch045
Scheme 46. Construction of pyrroles via the cascade reaction of β-enaminones and isocyanoacetates.
Scheme 46. Construction of pyrroles via the cascade reaction of β-enaminones and isocyanoacetates.
Catalysts 09 01032 sch046
Scheme 47. Mechanistic investigation of the Ag-mediated pyrrole synthesis. Adapted with permission from Org. Lett. 2017, 19, 1346–1349, doi:10.1021/acs.orglett.7b00201. Copyright (2019) American Chemical Society.
Scheme 47. Mechanistic investigation of the Ag-mediated pyrrole synthesis. Adapted with permission from Org. Lett. 2017, 19, 1346–1349, doi:10.1021/acs.orglett.7b00201. Copyright (2019) American Chemical Society.
Catalysts 09 01032 sch047
Scheme 48. Ag-catalyzed cycloaddition reactions of isocyanoacetamides.
Scheme 48. Ag-catalyzed cycloaddition reactions of isocyanoacetamides.
Catalysts 09 01032 sch048
Scheme 49. Proposed mechanism for the cycloaddition of isocyanoacetamide.
Scheme 49. Proposed mechanism for the cycloaddition of isocyanoacetamide.
Catalysts 09 01032 sch049
Scheme 50. Ag-catalyzed cycloaddition of α-trifluoromethylated methyl isocyanides.
Scheme 50. Ag-catalyzed cycloaddition of α-trifluoromethylated methyl isocyanides.
Catalysts 09 01032 sch050
Scheme 51. Synthesis of pyrrole via a Ag-catalyzed tandem 1,3-dipolar cycloaddition/oxidative dehydrogenative aromatization reaction. Adapted with permission from J. Org. Chem. 2017, 82, 4194-4202, doi:10.1021/acs.joc.7b00180. Copyright (2019) American Chemical Society.
Scheme 51. Synthesis of pyrrole via a Ag-catalyzed tandem 1,3-dipolar cycloaddition/oxidative dehydrogenative aromatization reaction. Adapted with permission from J. Org. Chem. 2017, 82, 4194-4202, doi:10.1021/acs.joc.7b00180. Copyright (2019) American Chemical Society.
Catalysts 09 01032 sch051

Share and Cite

MDPI and ACS Style

Yoo, K.; Jwa, D.G.; Lee, H.-E.; Kim, H.J.; Kim, C.; Kim, M. Recent Organic Transformations with Silver Carbonate as a Key External Base and Oxidant. Catalysts 2019, 9, 1032. https://doi.org/10.3390/catal9121032

AMA Style

Yoo K, Jwa DG, Lee H-E, Kim HJ, Kim C, Kim M. Recent Organic Transformations with Silver Carbonate as a Key External Base and Oxidant. Catalysts. 2019; 9(12):1032. https://doi.org/10.3390/catal9121032

Chicago/Turabian Style

Yoo, Kwangho, Dong Gyun Jwa, Ha-Eun Lee, Hyun Jin Kim, Cheoljae Kim, and Min Kim. 2019. "Recent Organic Transformations with Silver Carbonate as a Key External Base and Oxidant" Catalysts 9, no. 12: 1032. https://doi.org/10.3390/catal9121032

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop