Next Article in Journal
Plasma-Assisted CO2 Conversion to Methanol in Energy Systems: Parameter Optimization and Synergistic Effects
Previous Article in Journal
Catalytic and Non-Catalytic Co-Gasification of Biomass and Plastic Wastes for Energy Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

PdII Catalysis: Recent Advances in the Intramolecular Wacker-Type Reaction of Alkenols and Related Domino Reactions

Institut de Chimie Moléculaire de Reims, UMR 7312, CNRS-Université de Reims Champagne-Ardenne, B.P. 1039, CEDEX 2, 51687 Reims, France
Catalysts 2025, 15(9), 845; https://doi.org/10.3390/catal15090845
Submission received: 15 July 2025 / Revised: 19 August 2025 / Accepted: 25 August 2025 / Published: 2 September 2025
(This article belongs to the Special Issue Feature Review Papers on Catalysis in Organic and Polymer Chemistry)

Abstract

This review surveys the last twenty years of the PdII-catalysed oxaheterocyclisation of alkenes bearing a hydroxylated tether as well as the plausible subsequent in situ reactions, the intramolecular C–O bond formation being the first step of a domino process involving, among others, the Heck reaction, etherification, esterification and lactonisation. Versatile intermediates usable for the total synthesis of natural products have been thus produced. The proposed reaction mechanisms are highlighted with, as far as possible, personal comments.

Graphical Abstract

1. Introduction

The discovery in 1894 by Francis C. Phillips of the PdCl2-mediated oxidation of ethylene to acetaldehyde [1,2] sank into oblivion until Jürgen Smidt and co-workers from Wacker Chemie AG reported, more than a half-century later, a process using an acidic aqueous solution of catalytic amounts of PdCl2 and CuCl2 or CuCl under oxygen [3,4,5], leading to an industrial process [6]. The process involves the addition of water on the η2-PdII ethylene intermediate, a key step which has led to long controversy [7,8]. The reaction produces Pd0 species; regeneration of the catalyst is assumed with CuCl2, which is easily reoxidised with oxygen (Scheme 1) [5]. In 1960, Moiseev and co-workers suggested benzoquinone as the reoxidant of the Pd0 species [9].
The name “Wacker process” has been subsequently attributed to the method which has been used, under usually modified conditions, for the oxidation of terminal alkenes into the corresponding methyl ketones [10,11,12,13] or aldehydes [14,15] with application in the synthesis of natural products [16,17]. Convenient laboratory methods were proposed in 1964 by Selwitz and Clement, who performed the efficient oxidation of 1-dodecene into 2-dodecanone in aqueous DMF using catalytic PdCl2 with either catalytic CuCl2 and a flow of oxygen, or stoichiometric benzoquinone [18]. Then, PdCl2-catalysed Wacker reactions in aqueous DMF were performed by McQuillin and Parker in 1974 with either catalytic CuCl under oxygen or benzoquinone [19], and by Tsuji’s team in 1976 with stoichiometric CuCl and oxygen [20,21]. In 2001, Wong and Yick disclosed an approach towards the synthesis of eudesmanolides, using for one of the steps the PdCl2/CuCl/O2/aq. DMF procedure, that they inadequately named the “Wacker–Tsuji reaction” [22]. This name was, a few years later, used for reactions under different experimental conditions [23,24], leading to the inappropriate attribution of the name “Wacker–Tsuji (or Tsuji–Wacker) reaction (or oxidation)” to various Wacker-type procedures even in a book devoted to reaction names [25]. Other reoxidant systems have been used [12,13,16,26,27].
Alcohols may also form a C–O bond via PdII-catalysed inter- or intramolecular reaction with a C=C bond. In 2005, we highlighted such Wacker-type reactions [28], some of them being documented in other reviews [29,30,31]. We herein present the intramolecular oxidative cyclisations of alkenols mediated by PdII catalysts, and the related domino reactions published over the last twenty years [32,33]. The intramolecular Wacker-type cyclisation may proceed via anti nucleophilic attack [34,35], affording intermediate 2a or 2a’, which may evolve following two pathways: β-H elimination, providing O-heterocycle 2b or 2b’, respectively, while reaction with other reagents, an alkene for an example, may arise, leading to the corresponding Heck-type product 2c or 2c’, respectively (Scheme 2a). Regeneration of the active catalyst from the Pd0 species produced in the course of these reactions is assumed by the oxidant. In fact, like the addition of water [36,37,38], the anti/syn selectivity of the hydroxyl attack may depend on the nature of the Pd catalyst [39]. Moreover, intermediates 2a and 2a’ may arise via the formation of the palladium alcoholate followed by syn oxypalladation of the C=C bond (Scheme 2b). Whatever the mechanism may be, these annelations [40,41] are mediated by the PdII/Pd0 redox system. The PdII-catalysed reactions of alkenols in the presence of PhI(OAc)2 are discarded from the present review because they involve PdIV intermediates [42,43,44].
The PdII-catalysed oxaheterocyclisations of substrates having both an allylic leaving group and a hydroxylated tether also implicate Wacker-type intermediates. Corresponding studies disclosed up to mid-2009 have been previously reviewed [45], and consequently, will not be included here.

2. Intramolecular Wacker-Type Reaction

2.1. Formation of Dihydro- or Tetrahydrofurans

Pd(OCOCF3)2 or Pd(OAc)2 as the catalyst with oxygen as the oxidant mediated the oxidative cyclisation of primary alkenols 3a, 3b, 3c [46], 3d [47] and 3d’ [48] (Scheme 3a–d). Unsaturated fused bicyclic or spiro compounds were isolated depending on the length of the alcohol tether. The competing formation of aldehydes 3b1 and 3c1 (Scheme 3b,c) leads us to suspect that the cyclisation proceeded via syn oxypalladation. Experiments with stereospecifically deuterated substrates, carried out by the Stoltz team, agree with such a process [46]. The team of Kanai and Shibasaki disclosed the total synthesis of (±)-garsubellin A using 3e1 obtained from oxidative cyclisation of 3e mediated with a large excess of both Na2PdCl4 and t-BuOOH (Scheme 3e) [49].
Under PdCl2/CuCl/O2/aq. DMF conditions [19], the secondary alcohol 4a provided the bridged product 4a1 in high yield (Scheme 4a) [50]. Oxidation of α-bisabolol, which is a component in cosmetic, fragrance and therapeutic formulations, has been carried out under Pd(OAc)2 catalysis and oxygen pressure in aqueous methanol (Scheme 4b); the main reaction occurred at the level of the acyclic olefinic bond, with the Wacker-type process arising with low efficiency, leading to little amounts of a five-membered O-heterocycle [51].

2.2. Formation of Dihydro- or Tetrahydropyrans

While catalytic PdCl2 with stoichiometric CuCl2 and O2 in aq. DMF performed the oxidation of 5a into 5a1, which is intermediate in the total synthesis of cardinalin 3 (Scheme 5a) [52], a modified Wacker process based on hydroperoxides as the reoxidant in AcOH [53] was used to obtain (−)-alstonerine and (−)-alstophylline from 5b and 5b’, respectively (Scheme 5b) [24,54]. We assume that the reaction conditions firstly led to hydrolysis of the trimethylsilyloxy group of 5b and 5b’ [55,56,57].
Under PdCl2 catalysis in aq. DMA (Scheme 5c) instead of Pd(OAc)2 catalysis in aq. MeOH (Scheme 5b), the oxidation of α-bisabolol provided a fair yield of the six-membered O-heterocycle [51]. The Wacker reaction of ene-diol 5d under chiral conditions afforded the 6-endo cyclised product 5d1 with ee up to 82% (Scheme 5d) [58,59].

2.3. Formation of Dihydro-1,4-Oxazines

A variety of oxidants was experimented with for the PdCl2(MeCN)2-catalysed oxidative cyclisation in THF of aza-alkenol 6a into 6a1, the best result arising with O2 as the sole oxidant (Scheme 6a) [60]. Addition of catalytic CuCl2, or use of stoichiometric amounts of Cu(OAc)2, benzoquinone, H2O2 or t-BuOOH instead of O2 led to lower yields. With molecular oxygen as the sole oxidant, the Wacker-type reactions of terminal olefins were initially reported either in water [61] or aqueous DMA [62,63] in 1998 and 2006, respectively, using more elaborated PdII catalysts or O2 pressure up to 30 atm. Broggini’s team proposed that cyclised intermediate 6bA undergoes β-H-elimination, affording HPdClLn and the exo-methylene compound 6bB, the latter undergoing PdII-catalysed isomerisation [64,65,66,67] giving 6a1 (Scheme 6b, path a). The authors assumed that HPdClLn suffers reductive elimination leading to Pd0Ln, and a subsequent reaction with O2, giving η2-peroxopalladium complex 6bC [68,69,70]. Next, acidic opening of the peroxocycle followed by a second protonation regenerates the catalyst. In fact, it is never clear if the regeneration of the active PdII catalyst occurs from interaction of O2 with Pd0 or with HPdClLn [71,72,73,74]. Indeed, insertion of O2 into the H–Pd bond of the latter could directly afford HOOPdClLn [75,76,77] (Scheme 6b, path b). Instead of the formation of 6bB, we suggest the formation of a Pd-complex 6bD followed by successive addition/elimination of HPdClLn, delivering 6a1 (Scheme 6b, path c).

2.4. Formation of a Tetrahydrooxocine

The oxidative cyclisation of β-citronellol (7a) into eight-membered O-heterocycle 7a1 effectively arose under Pd(OAc)2-catalysis with H2O2 as the oxidant (Scheme 7a) [78]. Switching from Pd(OAc)2 to Pd(OCOCF3)2, Pd(acac)2 or PdCl2 was detrimental to the process. The catalytic cycle assumed by da Silva and Villarreal involved PdIV intermediates 7bA and 7bB (Scheme 7b, path a). We rather suspect the formation of 7bC (path b), which undergoes β-H elimination, delivering 7a1 and HPdCl; the reaction of the latter with H2O2 and AcOH regenerates the catalyst possibly via the formation of HOOPdOAc [26].

2.5. Formation of Dihydropyranones or Furanones

Dihydro-pyranyl and -furanyl cores have been synthesised from oxidative cyclisation of β-hydroxyenones 8a and α-hydroxyenones 8b, respectively (Scheme 8a,b) [79]. The five-membered ring has been exclusively formed from diol 8c, although both cyclisation pathways were possible (Scheme 8c) [79]. Gouverneur and co-workers pointed out that furan-3(2H)-ones are found in many natural products, one of them being bullatenone (Scheme 8b).

3. Intramolecular Wacker-Type Reaction/Dehydration

The PdCl2/CuCl2-catalysed reaction of alkenyl diols 9a under oxygen at room temperature unexpectedly provided furans 9a1 (Scheme 9a) via oxidative cyclisation followed by dehydration [80]. Although Reddy’s team claimed for a “Tsuji–Wacker-type cyclization”, the proposed mechanism corresponded to syn oxypalladation (see Scheme 2). Indeed, the authors assumed the formation of (η2-en)palladium alcoholate 9bA (which was incorrectly named π-allyl palladium) (Scheme 9b). Then, intramolecular olefin insertion affords 9bB, which undergoes β-H elimination, leading, according to the authors, to HCl, Pd0 and 9bC (path a), subsequent isomerisation of the latter to 9bD followed by spontaneous dehydration, giving 9a1. We suggest the formation of 9bD via η2-olefin hydridopalladium complex 9bE and η1-palladium chloride 9bF (Scheme 9b, path b).

4. Intramolecular Wacker-Type Reaction/Nucleophilic Addition

2-Allylphenols are particularly prone to Pd-catalysed dialkoxylation via domino reactions, which firstly involve isomerisation of the substrates into the corresponding 2-propenylphenols [65,81,82,83]. That urged Sigman’s team to study the enantioselective Wacker-type reaction of tertiary alcohol 10a in the presence of exogenous nucleophiles. Using (S)-iPrQuinox as the chiral ligand with catalytic amounts of both PdCl2(MeCN)2 and CuCl2 under oxygen atmosphere, reaction with alcohols or water afforded dialkoxylation or hydroalkoxylation products 10a1 (Scheme 10a) [84], while reaction with N-methylindoles or N-methylpyrrole provided adducts 10b1 and 10b2, respectively (Scheme 10b) [85], these reactions occurring with excellent enantioselectivities and fair to high diastereoselectivities. The process is compatible with the use of primary alcohols 10c (Scheme 10c) [84]. The domino reaction would proceed as depicted in Scheme 10d but detailed experimental investigations highlight rapid ligand exchange between palladium and copper, leading the authors to assume that “copper is involved in productive product formation, not just catalyst turnover” [86]. Coordination of the substrate to the catalyst affords palladium phenolate 10dA, which undergoes intramolecular nucleophilic addition of the hydroxyl group, leading to palladacycle 10dB. The latter evolves towards quinone methide intermediate 10dC, which suffers addition of the nucleophile delivering the product. The proposal of 10dC as an intermediate agrees with the formation of 10e1 from the reaction of 10a with propenyl ether 10e via a domino process involving a formal Diels–Alder reaction (Scheme 10e) [84].

5. Intramolecular Wacker-Type Reaction/Hydroalkoxylation or Hydroaryloxylation

The domino intramolecular Wacker-type reaction/hydroalkoxylation of ene-diols 11a, 11b and 11b’ provided ketals 11a1, 11b1 and 11b2 used for the total synthesis of (+)-buergerinin F (Scheme 11a) [87] and (−)-saliniketals A and B (Scheme 11b) [88,89]; the latter, inhibiting ornithine decarboxylase induction, may have a role in the chemoprevention of cancer [90]. The procedure led to 11c1 from hydroxyl phenol 11c, with 11c1 possessing the tetracyclic core of integrastatins (Scheme 11c) [91], which display HIV-1 integrase inhibitory activities [92]. The formation of ketal 11c1 from the heterocyclisation intermediate 11dA obtained from 11c could arise via hydroxyl addition, leading to hydridopalladium 11dB and then reductive elimination of Pd0 (Scheme 11d) [91,93].

6. Intramolecular Wacker-Type Reaction/Dehydrogenation (Dehydration)/Palladation/Oxidation

The enantioselective total synthesis of (−)-halenaquinone, which possesses antibiotic activities [94], has been carried out using intermediate 12a1 obtained from the Pd(OCOCF3)2-catalysed oxidative heterocyclisation of primary alcohol 12a in aerobic DMSO containing molecular sieves (Scheme 12a) [95]. The absence of molecular sieves led to 12a2, which provided 12a1 under oxidation with 2-iodoxybenzoic acid in CF3CO2H. According to Carter and co-workers, who tentatively explained the formation of both 12a1 and 12a2 from the Wacker product 12bA, oxidative dehydrogenation of the latter affords 12bB, which reacts with the catalyst, leading to η1-palladium complex 12bC (Scheme 12b). Next, oxidative cleavage of the C–Pd bond gives alcohol 12bD, which undergoes oxidation [96,97], providing 12a1. The authors proposed that 12a2 was formed from addition of adventitious water to 12bA, leading to hemiketal 12bE, the latter undergoing oxidative dehydrogenation and then dehydration. The oxidation steps noted “[O]” were not commented on by the authors; we suspect some participation of the Pd(OCOCF3)2/DMSO/O2 association [67].

7. Intra/Intermolecular Wacker-Type Reaction

Treatment of secondary or tertiary homoallylic alcohols 13a with catalytic amounts of both PdCl2 and CuCl2 at 80 °C in MeNO2 containing aq. t-BuOOH led to γ-lactones 13a1 (Scheme 13a) [98]. Use of a bidentate ligand may increase the reaction efficiency. A labeling experiment with H2O18 led Wu, Jiang and co-workers to retain two plausible catalytic cycles (Scheme 13b, paths a and b; the proposed catalytic scheme by the authors has been somewhat modified, in particular path b) which, however, did not fully explain the labeling result. Both catalytic cycles implicated Pd alcoholate 13ba as an intermediate. According to path a, 13bA undergoes intramolecular insertion of the O–Pd bond into the C=C bond, leading 13bB, which suffers β-H elimination, giving Pd0 and dihydrofuran 13bC. Coordination of the latter to PdCl2 triggers addition of water, leading to 13bD. Next, β-H elimination affords the corresponding enol, which tautomerises to the product. Path b involves addition of water to 13bA, resulting in the formation of palladacycle 13bE. Subsequent reductive elimination of palladium affords hydroxyl enol 13bF, which provides the corresponding aldehyde and finally the lactone.

8. Cyclisation/Chlorination or Esterification

The nature of the products issued from intramolecular reaction of aza-alkenol 14a under PdCl2(MeCN)2)2 (cat.)/CuCl2 conditions was highly dependent on the solvent, the chloro derivative 14a1 being produced in THF while formic ester 14a2 or acetic ester 14a3 were competitively formed in DMF or DMA, respectively (Scheme 14a) [99]. According to G. Broggini [100], the heterobimetallic Pd/Cu complex 14bA (Scheme 14b) undergoes reductive palladium elimination, liberating either 14a1 or the iminium intermediate 14bB [101,102,103], suggested in a previous paper of the team [104], subsequent work-up giving 14a2 or 14a3 (Scheme 14b). Comparison with another Broggini’s report using O2 and a catalytic amount of CuCl2 (see Scheme 6a) [60] leads us to suspect that the surprising reactivity is due to the over-stoichiometric amount of CuCl2, which favors the formation of the heterobimetallic complex inhibiting the β-H elimination. In fact, a similar complex was postulated as an intermediate of the PdCl2/CuCl2-catalysed chlorocyclisation of aminoalkenitols, which provided compounds corresponding to 14a1 [105].
The PdII-catalysed reaction of ene-diol 14c at 60 °C in AcOH containing over-stoichiometric amounts of CuCl2 and AcONa provided a mixture of chlorinated tetrahydropyran and tetrahydrofuran derivatives—14c1 and 14c2—due to exo- and endo-trig heterocyclisation, respectively (Scheme 14c) [106]. Using BuLi and LiCl in THF at room temperature instead of AcONa in AcOH at 60 °C selectively led to 14c2.

9. Cyclisation/Alkoxylation

A Wacker procedure was used to transform alkenyl polyols into functionalised 2,5-dioxabicyclo [2.2.1]heptanes, which are plausible intermediates of the total synthesis of natural products (Scheme 15a–c) [106,107,108,109,110], such as (+)-varitriol, which has significant activity against a variety of tumors [111]. The catalytic cycle occurs via intramolecular heterocyclisation of η2-palladium complex 15dA, leading to η1-palladium complex 15dB (Scheme 15d). Subsequent formation of palladacycle 15dC is followed by reductive elimination of Pd0, delivering the product.

10. Cyclisation/Dehydrogenation or Isomerisation

The behaviour of the PdII-catalysed reaction of 4-octen-1,8-diol 16a depended on the nature of both the Pd salt and oxidant (Scheme 16a) [112]. BQ as the oxidant provided five-membered heterocycles, either alcohol 16a1 with Pd(OAc)2 or aldehyde 16a2 with PdCl2, while the latter, associated to CuCl2, led to six-membered heterocycle 16a3. According to Tan and co-workers, the η1-Pd complex 16bA obtained via anti-oxypalladation of 16a undergoes β-H elimination, giving 16bB. Then, de-coordination of HPdII releases 16a1 (Scheme 16b, path a) while re-addition leads to 16bC. The authors assumed that redox relay provides 16bD, and subsequent elimination of HPdII, liberating 16a2 (path b). We rather propose that successive elimination/additions of HPdII lead to enol 16bE and then to 16a2 (path c). Indeed, 16bE as an intermediate agrees with the deuterium migration occurring when 16c was subjected to the PdCl2/BQ process (Scheme 16c). The formation of 16d1 from hydroxy-acid, -ester or -amide 16d (Scheme 16d) also involved double bond migration. For the formation of 16a3, we propose that the 6-endo-trig heterocyclisation intermediate 16bF undergoes intramolecular reaction, leading successively to palladaycle 16bG and hydridopalladium complex 16bH, and subsequent reductive elimination of Pd0, delivering the product (Scheme 16b, path d). The dependence of the reactivity on the anion of the Pd catalyst as well as on the oxidant is unclear.

11. Cyclisation with β-Heterounit Elimination

As pointed out in the introduction, the present chapter will only highlight studies published from mid-2009, except when previous reports are beneficial to the discussion.

11.1. β-OH Elimination

11.1.1. With 1,3-Chirality Transfer

After a first report on the synthesis of stereospecific formation of tetrahydro- and 3,6-dihydro [2H]pyran rings from intramolecular oxypalladation of unichiral [113,114] 2-ene-1,7-ols and 4-ene-1,3-ols in 2005 [115], Uenishi’s team extended the dehydrative cyclisation to various ene-diols [50,116,117,118,119,120], those published after 2009 [45] being depicted in Scheme 17.
The primary/secondary diols (R)-17a and (R)-17a’ underwent 5-exo-trig heterocyclisation from an Si face attack, leading to 17a1 and 17a’1, respectively (Scheme 17a) [121,122]. The relationship between the configuration of the new chiral center and that of the allylic alcohol was also observed for 7-exo-trig heterocyclisation but with lower dependence (Scheme 17b) [119]. The efficiency of the 1,3-chirality transfer has been exploited for the synthesis of C-nucleosides (Scheme 17c) [123] and (+)-goniocin (Scheme 17d) [124].
According to above results and those previously reported [45,116,117,118,119,120], the stereoselectivity depends on the configuration of the allylic hydroxyl. Uenishi’s team assumed that the reaction of E-ene-diols “occurs through the hydroxy-group-induced facioselective formation of the π-complex I, hydroxy ligand exchange leading to another π-complex II in equilibrium, an intramoleular syn oxypalladation leading to all-syn intermediate III, and syn elimination as the final step” (Scheme 17e) [119]. As pointed out by the team, the mechanism is more complicated in the case of Z-ene-diols [117]. Density functional calculations led Aponick, Ess and co-workers to disagree with Uenishi’s mechanism and to assume the anti-addition/anti-elimination depicted in Scheme 18 [125] for the cyclisation of 18a previously documented by Uenishi’s team [117]. According to the proposal, coordination of 18a to PdII is followed by anti-addition, leading to 18B via transition state 18A. Next, anti-elimination via 18C delivers 18a1. Subsequent studies from Uenishi and co-workers using poly-hydroxylated substrates (see Scheme 17c for an example) led the authors to note that they “were unable to account for the current results” according to that anti-oxypalladation and anti-elimination mechanism [123].

11.1.2. Without Chirality Transfer

Zawisza, who reported with Sinou the dependence of the 6-endo-trig versus 5-exo-trig reaction selectivity of Wacker-type cyclisation of bis-hydroxy allylic alcohols on both substitution and solvent [126], has subsequently studied the reactivity of tertiary alcohols 19a in THF, CH2Cl2 and Et2O (Scheme 19) [127]. Substrates with R = Et or CH=CH2 afforded a quasi-1:1 mixture of corresponding 19a1 and 19a2 while the 19a1/19a2 ratio strongly differed from 1:1 when R = CMe=CH2 with selectivity vary depending on the solvent, 19a1 being the main product in THF and Et2O, while reaction in CH2Cl2 mainly afforded 19a2. Such a dependence of diastereoelectivity on the solvent and size of the tertiary carbon remains unclear. The cyclisation of (E)- and (Z)-ene-triols 19b produced a mixture of diastereoisomers 19b1 and 19b2 in ratios independent of the stereochemistry of the C=C bond (Scheme 19b) [106].
The total synthesis of biologically active (±)-chelonin A has been performed via functionalised morpholine 20a1 obtained from intramolecular diastereoselective cyclisation of N-tethered alkenol 20a (Scheme 20) [128].
Dehydrative cyclisation of chiral θ-hydroxy-α,β,γ,δ-unsaturated dienols 21a and 21a’ effectively occurred, leading to a 3:1 mixture of O-heterocycles 21a1 and 21a2 (Scheme 21a) [129]. According Uenishi and co-workers, the absence of 1,5-chirality transfer arises from the formation of π-palladium conformers 21bA, 21bB and 21bC, 21bA leading to 21a1 while 21bB or 21bC affords 21a2 (Scheme 21b). Equilibrium between the three conformers favors the formation of the more stable cis isomer.
The approach to the synthesis of yessotoxin (Scheme 22a), which is a toxin isolated from the digestive glands of the scallop Patinopecten yessoensis, [130] urged to carry out the dehydrative cyclisation of bis-hydroxy allylic alcohols 22b, 22c, 22c’ and 22d to produce polyheterocycles (Scheme 22b–d) [131,132,133,134]; in particular, the tricyclic fused polyether core 22d1. The latter, which has the AB ring system of yessotoxin, arose from an iterative process starting with the dehydrative cyclisation of 22d.

11.2. β-OCOR Elimination

The total synthesis of muconin has been performed using six-membered heterocycles 23a1 obtained from diastereoselective 6-exo-trig heterocyclisation of 23a (Scheme 23a) [135]. The partial deprotection of the MOM group in the course of the annelation was caused by in situ-produced HCl and 2-naphthoic acid. The diasteroselectivity may be explained by the preponderance of the chair-like transition state 23bA, which displays less steric hindrance than 23bB (Scheme 23b) [135,136]. Cores of natural products have also been obtained from Wacker-type cyclisation of dihydroxy acetate 23c (Scheme 23c) [106].

11.3. β-OR Elimination

Other plausible intermediates of natural products have been synthesised from Wacker-type cyclisation of allylethers 24a (Scheme 24a) [106]. Cyclisation with elimination of the methoxyether unit of 24b arose, giving 1:1 mixtures of spiroketals 24b1 and 24b2, which have been used for total synthesis of acortatarin A (Scheme 24b) [137]. This reaction differs somewhat from the topic of this review because the cyclisation occurs from in situ-produced hemiketal 24cA (Scheme 24c) [138].
The formation of the seven-membered heterocycle 25a1 from protected ene-diol 25a under PdCl2(PhCN)2 catalysis in the presence of 2-propanol (Scheme 25) [132] would firstly imply the hydrolysis of the silylether [56,57].

12. Cyclisation/Intramolecular Heck Reaction

The reaction of dialkenyl alcohol 26a in the presence of catalytic amounts of Pd(OCOCF3)2 and a chiral ligand provided the bridged bicyclic ether 26a3 with ee up to 97% besides small amounts of Wacker-type O-heterocycles 26a1 and 26a2 (Scheme 26a) [139]. The products arose via PdII intermediate 26bA, which suffers intramolecular nucleophilic anti-addition, leading to σ-complex 26bB (Scheme 26b). The latter evolves via either β-H elimination, delivering 26a1, which may isomerise into 26a2 under the reaction conditions, or Heck-type addition giving 26bC which undergoes β-H elimination leading to 26a3. The domino etherification/Heck reaction also occurred from diene-alcohols 26c, leading to tricyclic products 26c1 (Scheme 26c) [140].

13. Cyclisation/Intermolecular Heck Reaction

Semmelhack conditions [141] have been used for the reaction of ene-diol 27a with methyl acrylate, leading to a mixture of 27a1 and 27a2 (Scheme 27a) [142], 27a2 being apparently formed from Pd-catalysed cyclisation of 27a1 [143]. The cyclisation of 27b and 27b’ in the presence of allyl bromide afforded heteroallylation products 27b1 and 27b2, respectively (Scheme 27b) [144]. After formation of the cyclisation intermediate 27cA, coordination of allyl bromide is followed by insertion of the Pd–C bond into the C=C bond to afford 27cB, which undergoes β-Br elimination to release the product and active PdII species (Scheme 27c). The use of deuterated allyl bromide (Scheme 27c) and the absence of reaction under Pd2(dba)3 catalysis agree with the proposal. The method has been used for the total synthesis of citalopram (Scheme 27d) [144], which may be prescribed for the treatment of depression [145].

14. Cyclisation/Dehydrogenative Heck Reaction

The Pd(OAc)2/ethylnicotinate association efficiently catalysed the intramolecular oxidative oxyarylation of hydroxyalkenes 28a, which bear an unactivated (hetero)arene (Scheme 28a) [146]. The formation of tricyclic compound 28a1 was greatly depreciated in the absence of the ligand. Ethylnicotinate was more efficient than other pyridine ligands; such superiority was previously observed for other Pd-catalysed annelations and could be due to its electronic properties [147]. After formation of Wacker intermediate 28bA, concerted metalation-deprotonation [148,149] provides palladacycle 28bB, which undergoes reductive elimination, delivering the tricyclic product (Scheme 28b). A PdII/PdIV catalytic cycle was previously assumed for a similar domino reaction performed with Ph(IOAc)2 as the oxidant [150].

15. Cyclisation/Alkoxy- or Hydroxycarbonylation

Looking for the total synthesis of polycavernoside A, White’s team observed that the PdII-catalysed cyclisation/alkoxycarbonylation of 29a arose in 61% yield while the expected heterocycles were not obtained from 29a’ and 29a” (Scheme 29a) [151]. Attempts to explain the determining role of the arrangement of the substituents in the reactivity were unsuccessful. Difficulties for such reactions were previously encountered, 29b requiring over-stoichiometric quantities of Pd(OAc)2 to efficiently afford 29b1 (Scheme 29b) [152]. Finally, this approach towards polycavernoside A was abandoned. In contrast, catalytic conditions successfully performed the cyclisation/alkoxycarbonylation of (i) 29c, which led to 29c1, the latter being tentatively used for the total synthesis of phorboxazole A (Scheme 29c) [153,154], which exhibits some in vitro antifungal and cytostatic activities [155]; and (ii) 29d, which led to 29d1 (Scheme 29d) [155], the latter providing an intermediate [156] previously used for the synthesis of cyanolide A [157], which exhibits cytotoxic properties [158].
Aza-alkenol 14a, which above led to 14a1, 14a2 and 14a3 (Scheme 14a), provided 14a4 under alkoxycarbonylation conditions (Scheme 29e) [99]. Similar conditions except the change of MeOH for AcOH and addition of AcONa afforded tetrahydropyranyl acetic acids 29f1 and 29g1 from 29f and 29g, respectively (Scheme 29f,g) [159], 29f1 being a civet cat compound and a perfume material while 29g1 was used to synthesise diospongin A [159], which exhibits slight inhibited bone resorption [160].
The above PdII-catalysed domino reactions involve CO coordination to a Wacker intermediate, giving 29hA, which undergoes insertion of CO into the C–Pd bond, leading to acylpalladium species 29hB (Scheme 29h). In the presence of ROH, the exchange of the ligand provides 29hC, which releases the ester via reductive elimination of Pd0 (path a) [161,162], while we suspect that anhydride 29hD is formed in the presence of AcOH/AcONa, work-up leading to the carboxylic acid (path b). 29hC has also been trapped via a two-step reaction consisting in in situ addition of a hydride source or an organometallic species, leading to an aldehyde or a ketone, respectively [163].

16. Cyclisation/Alkoxycarbonylative Lactonisation

In 1984, Semmelhack’s team disclosed the synthesis of pyranes fused with a γ-lactone from 5-hexen-1,4-diols, carbon dioxide and stoichiometric amounts of Pd(OAc)2 [164], while the next year, Yoshida’s team proposed a PdCl2-catalysed procedure leading to tetrahydrofuranolactones [165]. These seminal papers have been followed by a number of reports leading to bicyclic lactones from Pd-catalysed reactions of ene-diols under CO atmosphere [28]. For the last twenty years, the corresponding research mainly concerned the formation of synthons being plausible intermediates of the total synthesis of natural products [166].
After an approach to the synthesis of plakortone B using the stoichiometric Pd(OAc)2 procedure [167], the target natural product, which has potential interest for treatment of cardiac abnormalities [168], was synthesised by Semmelhack’s team from diol 30a, catalytic PdCl2 and over-stoichiometric amounts of both CuCl2 and AcONa in AcOH under CO atmosphere (Scheme 30a) [169]. Subsequently to a short communication in 2002 [170], Kitching’s team detailed in 2010 the synthesis with the same methodology of plakortones C, D, F and racemic E [171]; an example depicted in Scheme 30b concerns the synthon 30b1 used for the preparation of plakortone D and obtained from ene-triol 30b.
For an approach to the synthesis of micrandilactone A, which was traditionally used for the treatment of traumatic injuries [172], Yang and co-workers firstly observed that the Pd(OAc)2-catalysed reaction of 31a in the presence of CuCl2 and CO caused complete decomposition of the substrate [173]. Consequently, they performed the reaction with addition of a C2-symmetric thiourea (Scheme 31a) [174], which was successfully employed in previous carbonylative reactions [175,176]. The formation of the expected tricyclic compound 31a1 was accompanied by that of chloro bicyclic compound 31a2 due to some competitive participation of the internal double bond in the process. Consequently, the corresponding epoxide 31a’ was used instead of 31a, providing a high yield of 31a3, which would be the FGH ring of micrandilactone A. In fact, the latter is not a synthon for the synthesis of micrandilactone A because the previous published structure was erroneous [172].
Looking for the synthesis of crisamicin A, which exhibits in vitro activity against Gram-positive bacteria [177], Yang’s team firstly studied the cyclisation/carbonylation of 31b [178]. Under experimental conditions similar to those used above, the alkoxycarbonylative lactonisation arose, leading to expected heterocycles 31b1 and 31b2, but 31b3 and 31b4 were also formed (Scheme 31b). To preclude the side products, the reaction was carried out with additives. Interestingly, propylene oxide precluded the formation of 31b3 in removing Cl in situ generated from the regeneration of active PdII species, while NH4OAc prevented the formation of 31b4, the presence of both propylene oxide and NH4OAc, allowing the selective synthesis of 31b1 and 31b2. A plausible mechanism showed in Scheme 31c somewhat differs from that proposed by the authors and corrects the structure of some intermediates. The coordination of 31b to the catalyst affords η2-complex 31cA, which undergoes intramolecular nucleophilic addition, giving 31cB. The authors assumed the subsequent formation of palladacycle 31cC following insertion of CO to provide 31cD and then reductive elimination of Pd0, delivering 31b1 and 31b2 (path a). We suspect that 31cD is rather provided via 31cF obtained from CO insertion into the Pd–Cl bond of 31cB (path b). The authors assumed that “due to the existence of an active β-hydrogen”, 31cC could afford hydridopalladium 31cE, and subsequent reductive elimination of Pd0, giving 31b3. In fact, 31cE could be directly obtained from 31cB (path c). The authors envisioned that ionisation of 31cA affords η3-allyl complex 31cB, which provides 31b4 via addition of Cl (path d). We propose the formation of 31b4 via addition of Cl to 31cA (path e). The improved experimental conditions were effective for the alkoxycarbonylative lactonisation of 31d, affording 31d1, which was employed for the synthesis of crisamicin A (Scheme 31d). The preparation of a range of fused pyran-γ-lactones with the same method was then documented [179], the most interesting aspect of this report being the high yield of the reaction of 31e, for which the allylic hydrogen was replaced with a methyl group (Scheme 31e).
McErlean and Nesbitt prepared bicyclic systems (Scheme 32) which have been used to prepare antipodal compounds of marine natural products [180] and were intermediates [181] of the synthesis of kumausallene [182,183].
Macrolides have been synthesised from ene-diols 33a and 33b (Scheme 33a,b) by Dai and co-workers, who applied the procedure to the total synthesis of 9-demethylneopeltolide (Scheme 33c) [184].
After a first report in 1991 on the oxycarbonylation of enitols [185], Gracza remained strongly involved in the topic [28,166]. A preliminary communication on the total synthesis of goniothalesdiol was followed by a detailed report, with key intermediate 34a1 being obtained from ene-triol 34a (Scheme 34a) [186,187]. Then, tetrahydrofuranolactone 34b1, useable for the synthesis of jaspine B [188], was attained from amino-ene-diol 34b (Scheme 34b) [189] while Szolcsányi’s team from the same university prepared a mixture of fused tetrahydropyranyllactones 34c1 and 34c2 which could be used for the synthesis of decarestrictine L (Scheme 34c) [190].
Gracza and Kapitán studied the stereocontrol of the oxycarbonylation of 4-benzyloxy-hepta-1,6-diene-3,5-diols, a few results being highlighted in Scheme 34d,e [191]. Treatment of di-(allylic alcohol) 34d under Pd(OAc)2 catalysis and CO atmosphere in the presence of over-stoichiometric amounts of both CuCl2 and AcONa in AcOH provided a mixture of 34d1 and 34d2 at 25 °C while the former was selectively produced at 65 °C. Under asymmetric conditions with the Pd(OCOCF3)2/benzoquinone association, the diasteroisomer 34e led to either 34e1 or 34e2 depending on the chiral ligand, the reaction occurring, however, with low enantioselectivity.
After observing that the efficiency of the kinetic resolution of pent-4-ene-1,3-diol in AcOH depended on the Pd salt (Scheme 34f) [192], the reaction was carried out in ionic liquids, giving the starting diol with up to 60% ee and lactone 34g1 with up to 79% ee (Scheme 34g) [193]. The long reaction time was considerably reduced under microwave irradiation without a decrease in the enantioselectivity.
Then, reactions were performed with iron pentacarbonyl as the CO source, possibly leading to improved yields as depicted with the cyclocarbonylation of ene-diol 34h and ene-triol 34i, which provided the bis-tetrahydrofuran subunit of neopallavicinin (Scheme 34h) [194] and the tetrahydrofuranolactone core of pyrenolide D (Scheme 34i) [195,196], respectively. With ene-tetraol 34j as the substrate, the method brought a mixture of 34j1, 34j2 and 34j3, the latter not being produced in increasing the amount of Fe(CO)5 (Scheme 34j) or using CO atmosphere under historical conditions [197]. The Fe(CO)5 procedure also mediated the domino reaction of ene-triol 34k, providing 34k1 (Scheme 34k) [197], which was subsequently used for the asymmetric synthesis of sauropunols A–D [198], and N-protected amino diol 34l, leading to 34l1, which was an intermediate of the asymmetric synthesis of jaspine B (Scheme 34l) [199].
Moreover, Gracza’s team designated experimental conditions to perform the Fe(CO)5-carbonylative reaction in a flow reactor, effectively producing multi-grams of furanolactones [199,200,201].

17. Conclusions and Perspectives

This survey has shown the versatility of the Pd-catalysed heterocyclisation of alkenols. Indeed, the intramolecular Wacker-type process may be the first step of a domino reaction leading to building blocks of the total synthesis of biologically active natural compounds. In most cases, the reactions efficiently occur under mild experimental conditions and are characterised by regio- and stereoselectivities as well as functional group tolerance and a broad substrate scope. Consequently, the process is a valuable tool for the bench organic chemists.
At the present time, the above domino reactions have been seldom carried out under asymmetric conditions. The large set of available chiral ligands is an opportunity for new efficient enantioselective constructions of compounds having the scaffold similar to that of natural compounds, the aim being the discovery of drugs with improved or novel pharmaceutical properties. Thus, it is likely that new applications, based on the above procedures and providing important heterocyclic architectures, will appear in the near future.

Funding

This research received no external funding.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Phillips, F.C. Researches upon the phenomena of oxidation and chemical properties of gases. Am. Chem. J. 1894, 16, 255–277. [Google Scholar]
  2. Phillips, F.C. Untersuchungen über die chemischen eigenschaften von gasen. I. Mitteilung. Erscheinungen bei der oxydation von wasserstoff und kohlenwasserstoffen. Z. Anorg. Chem. 1894, 6, 213–228. [Google Scholar] [CrossRef]
  3. Smidt, J.; Hafner, W.; Jira, R.; Sedlmeier, J.; Sieber, R.; Rüttinger, R.; Kojer, H. Katalytische umsetzungen von olefinen an platinmetall-verbindungen. Das consortium-verfahren zur herstellung von acetaldehyde. Angew. Chem. 1959, 71, 176–182. [Google Scholar] [CrossRef]
  4. Hafner, W.; Jira, R.; Sedlmeier, J.; Smidt, J. Über die reaktionen von olefinen mit wässrigen lösungen with palladiumsalzen. Chem. Ber. 1962, 95, 1575–1581. [Google Scholar] [CrossRef]
  5. Smidt, J.; Hafner, W.; Jira, R.; Sedlmeier, J.; Sieber, R.; Ruttinger, R.; Kojer, H. The oxidation of olefins with palladium chloride catalysts. Angew. Chem. Int. Ed. 1962, 1, 80–88. [Google Scholar] [CrossRef]
  6. Jira, R. Acetaldehyde from Ethylene—A Retrospective on the Discovery of the Wacker Process. Angew. Chem. Int. Ed. 2009, 48, 9034–9037. [Google Scholar] [CrossRef]
  7. Keith, J.A.; Henry, P.M. The mechanism of the Wacker reaction: A tale of two hydroxypalladations. Angew. Chem. Int. Ed. 2009, 48, 9038–9049. [Google Scholar] [CrossRef]
  8. Stirling, A.; Nair, N.N.; Lledós, A.; Ujaque, G. Challenges in modelling homogeneous catalysis: New answers from ab initio molecular dynamics to the controversy over the Wacker process. Chem. Soc. Rev. 2014, 43, 4940–4952. [Google Scholar] [CrossRef]
  9. Moiseev, I.I.; Vargaftik, M.N.; Syrkin, Y.K. Oxidation reactions of olefins. Dokl. Akad. Nauk SSSR 1960, 130, 820–823. [Google Scholar]
  10. Tsuji, J. Synthetic applications of the palladium-catalyzed oxidation of olefins to ketones. Synthesis 1984, 1984, 369–384. [Google Scholar] [CrossRef]
  11. Takacs, J.M.; Jiang, X.-T. The Wacker reaction and related alkene oxidation reaction. Curr. Org. Chem. 2003, 7, 369–396. [Google Scholar] [CrossRef]
  12. Cornell, C.N.; Sigman, M.S. Recent progress in Wacker oxidations: Moving toward molecular oxygen as the sole oxidant. Inorg. Chem. 2007, 46, 1903–1909. [Google Scholar] [CrossRef] [PubMed]
  13. Jose, J.; Jose, D.E.; Mathew, T.V. Advances in Wacker oxidation: From palladium to first-row transition metal catalysts. J. Organomet. Chem. 2025, 1032, 123625. [Google Scholar] [CrossRef]
  14. Muzart, J. Aldehydes from Pd-catalysed oxidation of terminal olefins. Tetrahedron 2007, 63, 7505–7521. [Google Scholar] [CrossRef]
  15. Muzart, J. Progress in the synthesis of aldehydes from Pd-catalyzed Wacker-type reactions of terminal olefins. Tetrahedron 2021, 87, 132024. [Google Scholar] [CrossRef]
  16. Fernandes, R.A.; Jha, A.K.; Kumar, P. Recent advances in Wacker oxidation: From conventional to modern variants and applications. Catal. Sci. Technol. 2020, 10, 7448–7470. [Google Scholar] [CrossRef]
  17. Mhaldar, S.N.; Gaikwad, M.M.; Tilve, S.G. Wacker Oxidation: A Powerful Tool for Natural Product Synthesis. Asian J. Org. Chem. 2025, 14, e202500117. [Google Scholar] [CrossRef]
  18. Clement, W.H.; Selwitz, C.M. Improved procedures for converting higher α-olefins to methyl ketones with palladium chloride. J. Org. Chem. 1964, 29, 241–243. [Google Scholar] [CrossRef]
  19. McQuillin, F.J.; Parker, D.G. Complexing of terpenes with transition metals. Part IV. A comparison of the reactions of (+)-3,7-dimethylocta-1,6-diene with palladium(II) and with mercury(II). J. Chem. Soc. Perkin Trans. 1 1974, 1974, 809–815. [Google Scholar] [CrossRef]
  20. Tsuji, J.; Shimizu, I.; Yamamoto, K. Convenient general synthetic method for 1,4- and 1,5-diketones by palladium catalyzed oxidation of α-allyl and α-3-butenyl ketones. Tetrahedron Lett. 1976, 34, 2975–2976. [Google Scholar] [CrossRef]
  21. Tsuji, J.; Nagashima, H.; Nemoto, H. A general method for the preparation of methyl ketones from terminal olefins: 2-decanone. In Organic Syntheses; Semmelhack, M.F., Ed.; Wiley: Hoboken, NJ, USA, 1984; Volume 62, pp. 9–11. [Google Scholar]
  22. Yick, C.-Y.; Wong, H.N.C. Regiospecific substitution of the carbon–boron bond of tris(4-methylfuran-3-yl)boroxine: A model ring C→BC→ABC approach towards eudesmanolides. Tetrahedron 2001, 57, 6935–6940. [Google Scholar] [CrossRef]
  23. Thadani, A.N.; Rawal, V.H. Multifunctional palladium catalysis. 2. Tandem haloallylation followed by Wacker-Tsuji oxidation or Sonogashira cross-coupling. Org. Lett. 2002, 4, 4321–4323. [Google Scholar] [CrossRef] [PubMed]
  24. Liao, X.; Zhou, H.; Yu, J.; Cook, J.M. An improved total synthesis of (+)-macroline and alstonerine as well as the formal total synthesis of (−)-talcarpine and (−)-anhydromacrosalhine-methine. J. Org. Chem. 2006, 71, 8884–8890. [Google Scholar] [CrossRef]
  25. Li, J.J. Wacker-Tsuji oxidation. In Name Reactions for Functional Group Transformations; Li, J.J., Corey, E.J., Eds.; Wiley-Interscience: Hoboken, NJ, USA, 2007; pp. 309–326. [Google Scholar]
  26. Roussel, M.; Mimoun, H. Palladium-catalyzed oxidation of terminal olefins to methyl ketones by hydrogen peroxide. J. Org. Chem. 1980, 45, 5387–5390. [Google Scholar] [CrossRef]
  27. Nishimura, T.; Kakiuchi, N.; Onoue, T.; Ohe, K.; Uemura, S. Palladium(II)-catalyzed oxidation of terminal alkenes to methyl ketones using molecular oxygen. J. Chem. Soc. Perkin Trans. 1 2000, 2000, 1915–1918. [Google Scholar] [CrossRef]
  28. Muzart, J. Palladium-catalysed reactions of alcohols. Part C: Formation of ether linkages. Tetrahedron 2005, 61, 5955–6008. [Google Scholar] [CrossRef]
  29. Zeni, G.; Larock, R.C. Synthesis of heterocycles via palladium π-olefin and π-alkyne chemistry. Chem. Rev. 2004, 104, 2285–2309. [Google Scholar] [CrossRef]
  30. Sharma, G.V.M.; Krishna, P.R. Carbohydrates: From ‘chirons’ to new glycosubstances. Curr. Org. Chem. 2004, 8, 1187–1209. [Google Scholar] [CrossRef]
  31. Doháňošová, J.; Gracza, T. Asymmetric palladium-catalysed intramolecular Wacker-type cyclisations of unsaturated alcohols and amino alcohols. Molecules 2013, 18, 6173–6192. [Google Scholar] [CrossRef]
  32. Beccalli, E.M.; Broggini, G.; Martinelli, M.; Sottocornola, S. C-C, C-O, C-N bond formation on sp2 carbon by Pd(II)-catalyzed reactions involving oxidant agents. Chem. Rev. 2007, 107, 5318–5365. [Google Scholar] [CrossRef]
  33. Butt, N.A.; Zhang, W. Synthesis of heterocycles via palladium-catalyzed Wacker-type oxidative cyclization reactions of hydroxy- and amino-alkenes. Top. Heterocycl. Chem. 2013, 32, 77–108. [Google Scholar]
  34. Hosokawa, T.; Yamanaka, T.; Murahashi, S.-I. Palladium(II)-catalysed asymmetric acetalization of alkenes. J. Chem. Soc. Chem. Commun. 1993, 1993, 117–119. [Google Scholar] [CrossRef]
  35. Hosokawa, T.; Yamanaka, T.; Itotani, M.; Murahashi, S.-I. Palladium(II)-catalyzed asymmetric acetalization of alkenes. J. Org. Chem. 1995, 60, 6159–6167. [Google Scholar] [CrossRef]
  36. Henry, P.M. Palladium Catalyzed Oxidation of Hydrocarbons; Reidel: Dordrecht, The Netherlands, 1980; pp. 41–84. [Google Scholar]
  37. ten Brink, G.-J.; Arends, I.W.C.E.; Papadogianakis, G.; Sheldon, R.A. Catalytic conversions in water: Part 13. Aerobic oxidation of olefins to methyl ketones catalysed by a water-soluble palladium complex—Mechanistic investigations. Appl. Catal. A 2000, 194–195, 435–442. [Google Scholar] [CrossRef]
  38. Nelson, D.J.; Li, R.; Brammer, C. Correlation of relative rates of PdCl2 oxidation of functionalized acyclic alkenes versus alkene ionization potentials, HOMOs, and LUMOs. J. Am. Chem. Soc. 2001, 123, 1564–1568. [Google Scholar] [CrossRef]
  39. Hayashi, T.; Yamasaki, K.; Mimura, M.; Uozumi, Y. Deuterium-labeling studies establishing stereochemistry at the oxypalladation step in Wacker-type oxidative cyclization of an o-allylphenol. J. Am. Chem. Soc. 2004, 126, 3036–3037. [Google Scholar] [CrossRef]
  40. Sakakibara, T.; Tanaka, Y.; Yamasaki, S.-I. Palladium-promoted aromatic annelation with acetylene dicarboxylates. Chem. Lett. 1986, 15, 797–800. [Google Scholar] [CrossRef]
  41. Wu, G.; Rheingold, A.L.; Geib, S.J.; Heck, R.F. Palladium-catalyzed annelation of aryl iodides with diphenylacetylene. Organometallics 1987, 6, 1941–1946. [Google Scholar] [CrossRef]
  42. Wang, W.; Wang, F.; Shi, M. Bis(NHC)-palladium(II) complex-catalyzed dioxygenation of alkenes. Organometallics 2010, 29, 928–933. [Google Scholar] [CrossRef]
  43. Takenaka, K.; Dhagea, Y.D.; Sasai, H. Enantioselective Pd(II)–Pd(IV) catalysis utilizing a SPRIX ligand: Efficient construction of chiral 3-oxy-tetrahydrofurans. Chem. Commun. 2013, 49, 11224–11226. [Google Scholar] [CrossRef]
  44. Giofrè, S.; Molteni, L.; Nava, D.; Lo Presti, L.; Beccalli, E.M. Enantio- and regioselective palladium(II)-catalyzed dioxygenation of (aza-)alkenols. Angew. Chem. Int. Ed. 2021, 60, 21723–21727. [Google Scholar] [CrossRef]
  45. Muzart, J. Pd0- and PdII-catalyzed oxaheterocyclization of substrates having both an allylic leaving group and a hydroxylated tether. J. Mol. Catal. A Chem. 2010, 319, 1–29. [Google Scholar] [CrossRef]
  46. Trend, R.M.; Ramtohul, Y.K.; Stoltz, B.M. Oxidative cyclizations in a nonpolar solvent using molecular oxygen and studies on the stereochemistry of oxypalladation. J. Am. Chem. Soc. 2005, 127, 17778–17788. [Google Scholar] [CrossRef]
  47. Ebner, D.C.; Novak, Z.; Stoltz, B.M. Oxidative kinetic resolution-Claisen rearrangement sequence to enantioenriched arylcycloalkenes. Synlett 2006, 2006, 3533–3539. [Google Scholar] [CrossRef]
  48. Jin, L.; Nemoto, T.; Nakamura, H.; Hamada, Y. Pd-catalyzed asymmetric allylic alkylation of 2-substituted cycloalkenyl carbonates using a chiral diaminophosphine oxide: (S,RP)-Ph-DIAPHOX. Tetrahedron Asymmetry 2008, 19, 1106–1113. [Google Scholar] [CrossRef]
  49. Kuramochi, A.; Usuda, H.; Yamatsugu, K.; Kanai, M.; Shibasaki, M. Total synthesis of (±)-garsubellin A. J. Am. Chem. Soc. 2005, 127, 14200–14201. [Google Scholar] [CrossRef]
  50. Kawai, N.; Hande, S.M.; Uenishi, J. Stereoselective synthesis of (−)-diospongins A and B and their stereoisomers at C-5. Tetrahedron 2007, 63, 9049–9056. [Google Scholar] [CrossRef]
  51. Parreira, L.A.; Gonçalves, D.C.; Menini, L.; Gusevskaya, E.V. Aerobic oxidation of naturally occurring -bisabolol catalyzed by palladium(II) salts as sole catalysts. Appl. Catal. A 2016, 524, 126–133. [Google Scholar] [CrossRef]
  52. Govender, S.; Mmutlane, E.M.; van Otterlo, W.A.L.; de Koning, C.B. Bidirectional racemic synthesis of the biologically active quinone cardinalin 3. Org. Biomol. Chem. 2007, 5, 2433–2440. [Google Scholar] [CrossRef]
  53. Tsuji, J.; Nagashima, H.; Hori, K. A new preparative method for 1, 3-dicarbonyl compounds by the regioselective oxidation of α, β-unsaturated carbonyl compounds, catalyzed by PdCl2 using hydroperoxides as the reoxidant of Pd0. Chem. Lett. 1980, 9, 257–260. [Google Scholar] [CrossRef]
  54. Liao, X.; Zhou, H.; Wearing, X.Z.; Ma, J.; Cook, J.M. The first regiospecific, enantiospecific total synthesis of 6-oxoalstophylline and an improved total synthesis of alstonerine and alstophylline as well as the bisindole alkaloid macralstonine. Org. Lett. 2005, 7, 3501–3504. [Google Scholar] [CrossRef] [PubMed]
  55. Cort, A.D. A simple and convenient method for the cleavage of silylethers. Synth. Commun. 1990, 20, 757–760. [Google Scholar] [CrossRef]
  56. Lipshutz, B.H.; Pollart, D.; Monforte, J.; Kotsuki, H. Pd(II)-catalyzed acetal/ketal hydrolysis/exchange reactions. Tetrahedron Lett. 1985, 26, 705–708. [Google Scholar] [CrossRef]
  57. Cormier, J.F. Removal of TBDMS protecting groups from carbohydrates using catalytic transfer hydrogenation. Tetrahedron Lett. 1991, 32, 187–188. [Google Scholar] [CrossRef]
  58. Takenaka, K.; Nakatsuka, S.; Tsujihara, T.; Koranne, P.S.; Sasai, H. Divergent synthesis of chiral spiro (isoxazole–isoxazoline) hybrid ligands. Tetrahedron Asymmetry 2008, 19, 2492–2496. [Google Scholar] [CrossRef]
  59. Takenaka, K.; Nagano, T.; Takizawa, S.; Sasai, H. Asymmetric synthesis of chiral spiro bis(isoxazoline) and spiro (isoxazole–isoxazoline) ligands. Tetrahedron Asymmetry 2010, 21, 379–381. [Google Scholar] [CrossRef]
  60. Broggini, G.; Beccalli, E.M.; Borsini, E.; Fasana, A.; Zecchi, G. Intramolecular oxidative Pd(II)-catalyzed alkoxylation of 3-aza-5-alkenols with O2 as sole oxidant: Mild conditions for the synthesis of 1,4-oxazine derivatives. Synlett 2011, 2011, 227–230. [Google Scholar] [CrossRef]
  61. ten Brink, G.-J.; Arends, I.W.C.E.; Papadogianakis, G.; Sheldon, R.A. Catalytic conversions in water. Part 10. Aerobic oxidation of terminal olefins to methyl ketones catalysed by water soluble palladium complexes. Chem. Commun. 1998, 1998, 2359–2360. [Google Scholar] [CrossRef]
  62. Mitsudome, T.; Umetani, T.; Nosaka, N.; Mori, K.; Mizugaki, T.; Ebitani, K.; Kaneda, K. Convenient and efficient Pd-catalyzed regioselective oxyfunctionalization of terminal olefins by using molecular oxygen as sole reoxidant. Angew. Chem. Int. Ed. 2006, 45, 481–485. [Google Scholar] [CrossRef]
  63. Cornell, C.N.; Sigman, M.S. Discovery of a practical direct O2-coupled Wacker Oxidation with Pd[(-)-sparteine]Cl2. Org. Lett. 2006, 8, 4117–4120. [Google Scholar] [CrossRef] [PubMed]
  64. Gross, J.L. A concise stereospecific synthesis of repinotan (BAY × 3702). Tetrahedron Lett. 2003, 44, 8563–8565. [Google Scholar] [CrossRef]
  65. Thiery, E.; Chevrin, C.; Le Bras, J.; Hakarat, D.; Muzart, J. Mechanism insights into the palladiumII-catalyzed hydroalkoxylation of 2-allylphenols. J. Org. Chem. 2007, 72, 1859–1862. [Google Scholar] [CrossRef]
  66. Fan, J.; Wan, C.; Wang, Q.; Gao, L.; Zheng, X.; Wang, Z. Palladium catalyzed isomerization of alkenes: A pronounced influence of an o-phenol hydroxyl group. Org. Biomol. Chem. 2009, 7, 3168–3172. [Google Scholar] [CrossRef]
  67. Nishiwaki, N.; Kamimura, R.; Shono, K.; Kawakami, T.; Nakayama, K.; Nishino, K.; Nakayama, T.; Takahashi, K.; Nakamura, A.; Hosokawa, T. Efficient double bond migration of allylbenzenes catalyzed by Pd(OAc)2–HFIP system with unique substituent effect. Tetrahedron Lett. 2010, 51, 3590–3592. [Google Scholar] [CrossRef]
  68. Landis, C.R.; Morales, C.M.; Stahl, S.S. Insights into the spin-forbidden reaction between L2Pd0 and molecular oxygen. J. Am. Chem. Soc. 2004, 126, 16302–16303. [Google Scholar] [CrossRef]
  69. Gligorich, K.M.; Sigman, M.S. Recent advancements and challenges of palladiumII-catalyzed oxidation reactions with molecular oxygen as the sole oxidant. Chem. Commun. 2009, 2009, 3854–3867. [Google Scholar] [CrossRef]
  70. Scheuermann, M.L.; Goldberg, K.I. Reactions of Pd and Pt complexes with molecular oxygen. Chem. Eur. J. 2014, 20, 14556–14568. [Google Scholar] [CrossRef]
  71. Gligorich, K.M.; Sigman, M.S. Mechanistic questions about the reaction of molecular oxygen with palladium in oxidase catalysis. Angew. Chem. Int. Ed. 2006, 118, 6612–6615. [Google Scholar] [CrossRef] [PubMed]
  72. Muzart, J. Molecular oxygen to regenerate PdII active species. Chem. Asian J. 2006, 1, 508–515. [Google Scholar] [CrossRef] [PubMed]
  73. Keith, J.M.; Goddard, W.A., III; Oxgaard, J. Pd-Mediated activation of molecular oxygen: Pd(0) versus direct insertion. J. Am. Chem. Soc. 2007, 129, 10361–10369. [Google Scholar] [CrossRef]
  74. Konnick, M.M.; Stahl, S.S. Reaction of molecular oxygen with a PdII-hydride to produce a PdII-hydroperoxide: Experimental evidence for an HX-reductive-elimination pathway. J. Am. Chem. Soc. 2008, 130, 5753–5762. [Google Scholar] [CrossRef]
  75. Muzart, J.; Pète, J.-P. Dehydrogenation of cyclohexanones catalyzed by palladiumII trifluoroacetate. J. Mol. Catal. 1982, 15, 373–376. [Google Scholar] [CrossRef]
  76. Keith, J.M.; Muller, R.P.; Kemp, R.A.; Goldberg, K.I.; Goddard, W.A., III; Oxgaard, J. Mechanism of direct molecular oxygen insertion in a palladium(II)-hydride bond. Inorg. Chem. 2006, 45, 9631–9633. [Google Scholar] [CrossRef]
  77. Chowdhury, S.; Rivalta, I.; Russo, N.; Sicilia, E. Theoretical investigation of the mechanism of acid-catalyzed oxygenation of a Pd(II)-hydride to produce a Pd(II)-hydroperoxide. J. Chem. Theory Comput. 2008, 4, 1283–1292. [Google Scholar] [CrossRef] [PubMed]
  78. da Silva, M.J.; Villarreal, J.A. Novel palladium-catalyzed oxidative intramolecular cyclization of β-citronellol with H2O2: A green and selective process to synthesize oxocine. Catal. Lett. 2017, 147, 1646–1653. [Google Scholar] [CrossRef]
  79. Reiter, M.; Turner, H.; Mills-Webb, R.; Gouverneur, V. Palladium-catalyzed oxidative cyclizations: Synthesis of dihydropyranones and furanones. J. Org. Chem. 2005, 70, 8478–8485. [Google Scholar] [CrossRef]
  80. Masal, D.P.; Choudhury, R.; Singh, A.; Reddy, S. Ready access to densely substituted furans using Tsuji-Wacker-type cyclization. J. Org. Chem. 2022, 87, 556–568. [Google Scholar] [CrossRef]
  81. Chevrin, C.; Le Bras, J.; Hénin, F.; Muzart, J. One-pot catalytic synthesis of 2-(1,2-dihydroxypropyl)-phenol derivatives from 2-allylphenols in aqueous media. Synthesis 2005, 2005, 2615–2618. [Google Scholar] [CrossRef]
  82. Schultz, M.J.; Sigman, M.S. Palladium(II)-catalyzed aerobic dialkoxylation of styrenes: A profound influence of an o-phenol. J. Am. Chem. Soc. 2006, 128, 1460–1461. [Google Scholar] [CrossRef]
  83. Zhang, Y.; Sigman, M.S. Palladium(II)-catalyzed enantioselective aerobic dialkoxylation of 2-propenyl phenols: A pronounced effect of copper additives on enantioselectivity. J. Am. Chem. Soc. 2007, 129, 3076–3077. [Google Scholar] [CrossRef]
  84. Jensen, K.H.; Pathak, T.P.; Zhang, Y.; Sigman, M.S. Palladium-catalyzed enantioselective addition of two distinct nucleophiles across alkenes capable of quinone methide formation. J. Am. Chem. Soc. 2009, 131, 17074–17075. [Google Scholar] [CrossRef]
  85. Pathak, T.P.; Gligorich, K.M.; Welm, B.E.; Sigman, M.S. Synthesis and preliminary biological studies of 3-substituted indoles accessed by a palladium-catalyzed enantioselective alkene difunctionalization reaction. J. Am. Chem. Soc. 2010, 132, 7870–7871. [Google Scholar] [CrossRef] [PubMed]
  86. Jensen, K.H.; Webb, J.D.; Sigman, M.S. Advancing the mechanistic understanding of an enantioselective palladium-catalyzed alkene difunctionalization reaction. J. Am. Chem. Soc. 2010, 132, 17471–17482. [Google Scholar] [CrossRef]
  87. Shiina, I.; Kawakita, Y.; Ibuka, R.; Yokoyama, K.; Tamai, Y. Total synthesis of buergerinin F via effective construction of the asymmetric quaternary carbons using an enantioselective aldol reaction. Chem. Commun. 2005, 2005, 4062–4064. [Google Scholar] [CrossRef] [PubMed]
  88. Paterson, I.; Razzak, M.; Anderson, E.A. Total synthesis of (−)-saliniketals A and B. Org. Lett. 2008, 10, 3295–3298. [Google Scholar] [CrossRef]
  89. Yadav, J.S.; Hossain, S.S.; Madhu, M.; Mohapatra, D.K. Formal total synthesis of (−)-saliniketals. J. Org. Chem. 2009, 74, 8822–8825. [Google Scholar] [CrossRef]
  90. Williams, P.G.; Asolkar, R.N.; Kondratyuk, T.; Pezzuto, J.M.; Jensen, P.R.; Fenical, W. Saliniketals A and B, bicyclic polyketides from the marine actinomycete Salinispora arenicola. J. Nat. Prod. 2007, 70, 83–88. [Google Scholar] [CrossRef]
  91. Tadross, P.M.; Bugga, P.; Stoltz, B.M. A rapid and convergent synthesis of the integrastatin core. Org. Biomol. Chem. 2011, 9, 5354–5357. [Google Scholar] [CrossRef]
  92. Singh, S.B.; Zink, D.L.; Quamina, D.S.; Pelaez, F.; Teran, A.; Felock, P.; Hazuda, D.J. Integrastatins: Structure and HIV-1 integrase inhibitory activities of two novel racemic tetracyclic aromatic heterocycles produced by two fungal species. Tetrahedron Lett. 2002, 43, 2351–2354. [Google Scholar] [CrossRef]
  93. Balija, A.M.; Stowers, K.J.; Schultz, M.J.; Sigman, M.S. Pd(II)-catalyzed conversion of styrene derivatives to acetals: Impact of (−)-sparteine on regioselectivity. Org. Lett. 2006, 8, 1121–1124. [Google Scholar] [CrossRef]
  94. Roll, D.M.; Scheuer, P.J.; Matsumoto, G.K.; Clardy, J. Halenaquinone, a pentacyclic polyketide from a marine sponge. J. Am. Chem. Soc. 1983, 105, 6177–6178. [Google Scholar] [CrossRef]
  95. Goswami, S.; Harada, K.; El-Mansy, M.F.; Lingampally, R.; Carter, R.G. Enantioselective synthesis of (−)-halenaquinone. Angew. Chem. Int. Ed. 2018, 57, 9117–9121. [Google Scholar] [CrossRef]
  96. Muzart, J. Palladium-catalysed oxidation of primary and secondary alcohols. Tetrahedron 2003, 59, 5789–5816. [Google Scholar] [CrossRef]
  97. Muzart, J. Oxidation adjacent to oxygen of alcohols catalyzed by palladium/dimethyl sulfoxide. In Comprehensive Organic Synthesis, 2nd ed.; Molander, G.A., Knochel, P., Eds.; Elsevier: Oxford, UK, 2014; Volume 7, pp. 295–301. [Google Scholar]
  98. Zheng, M.; Chen, P.; Huang, L.; Wu, W.; Jiang, H. Nucleo-palladation-triggering alkene functionalization: A route to γ-lactones. Org. Lett. 2017, 19, 5756–5759. [Google Scholar] [CrossRef]
  99. Gazzola, S.; Beccalli, E.M.; Borelli, T.; Castellano, C.; Diamante, D.; Broggini, G. Selective 7-endo-cyclization of 3-aza-5-alkenols through oxidative Pd(II)-catalyzed olefin oxyarylation. Synlett 2018, 29, 503–508. [Google Scholar]
  100. Broggini, G. (Università dell’Insubria, Como, Italy). Personal communication, 2025.
  101. Muzart, J. N,N-Dimethylformamide: Much more than a solvent. Tetrahedron 2009, 65, 8313–8323. [Google Scholar] [CrossRef]
  102. Le Bras, J.; Muzart, J. N,N-Dimethylformamide and N,N-dimethylacetamide as carbon, hydrogen, nitrogen and/or oxygen sources. In Solvents as Reagents in Organic Synthesis, Reactions and Applications; Wu, X.-F., Ed.; Wiley-VCH: Weinheim, Germany, 2017; pp. 199–314. [Google Scholar]
  103. Le Bras, J.; Muzart, J. Recent uses of N,N-dimethylformamide and N,N-dimethylacetamide as reagents. Molecules 2018, 23, 1939. [Google Scholar] [CrossRef] [PubMed]
  104. Broggini, G.; Barbera, V.; Beccalli, E.M.; Borsini, E.; Galli, S.; Lanza, G.; Zecchi, G. Palladium(II)/copper halide/solvent combination for selective intramolecular domino reactions of indolecarboxylic acid allylamides: An unprecedented arylation/esterification sequence. Adv. Synth. Catal. 2012, 354, 159–170. [Google Scholar] [CrossRef]
  105. Szolcsányi, P.; Gracza, T. PdCl2/CuCl2-catalysed chlorocyclisation of sugar-derived aminoalkenitols in the synthesis of new iminohexitols. Tetrahedron 2006, 62, 8498–8502. [Google Scholar] [CrossRef]
  106. Palík, M.; Kožíšek, J.; Koóš, P.; Gracza, T. Palladium-catalysed cyclisation of alkenols: Synthesis of oxaheterocycles as core intermediates of natural compounds. Beilstein J. Org. Chem. 2014, 10, 2077–2086. [Google Scholar] [CrossRef]
  107. Babjak, M.; Remeň, L.; Karlubíková, O.; Gracza, T. Synthesis of tetrahydrofurans through a novel pseudo-meso-trick. Synlett 2005, 2005, 1609–1611. [Google Scholar] [CrossRef]
  108. Babjak, M.; Remeň, L.; Szolcsányi, P.; Zálupský, P.; Mikloš, D.; Gracza, T. Novel bicyclisation of unsaturated polyols in PdCl2-CuCl2-AcOH catalytic system. J. Organomet. Chem. 2006, 691, 928–940. [Google Scholar] [CrossRef]
  109. Palík, M.; Karlubíková, O.; Lásiková, A.; Kožíšek, J.; Gracza, T. Total synthesis of (+)-varitriol from dimethyl L-tartrate. Eur. J. Org. Chem. 2009, 2009, 709–715. [Google Scholar] [CrossRef]
  110. Malmstrøm, J.; Christophersen, C.; Barrero, A.F.; Oltra, J.E.; Justicia, J.; Rosales, A. Bioactive metabolites from a marine-derived strain of the fungus Em11ericella variecolor. J. Nat. Prod. 2002, 65, 364–367. [Google Scholar] [CrossRef]
  111. Palík, M.; Karlubíková, O.; Lackovičová, D.; Lásiková, A.; Gracza, T. Formal synthesis of (+)-varitriol. Application of Pd(II)/Cu(II)-catalysed bicyclisation of unsaturated polyols. Tetrahedron 2010, 66, 5244–5249. [Google Scholar] [CrossRef]
  112. Brooks, J.L.; Xu, L.; Wiest, O.; Tan, D.S. Diastereoselective synthesis of highly substituted tetrahydrofurans by Pd-catalyzed tandem oxidative cyclization-redox relay reactions controlled by intramolecular hydrogen bonding. J. Org. Chem. 2017, 82, 57–75. [Google Scholar] [CrossRef]
  113. Gal, J. Problems of Stereochemical Nomenclature and Terminology. 1. The Homochiral Controversy. Its Nature and Origins, and a Proposed Solution. Enantiomer 1998, 3, 263–273. [Google Scholar]
  114. Gal, J. On the meaning and use of homochiral. J. Chromatogr. A 1998, 829, 417–418. [Google Scholar] [CrossRef]
  115. Uenishi, J.; Ohmi, M.; Ueda, A. PdII-catalyzed stereospecific formation of tetrahydro- and 3,6-dihydro [2H]pyran rings: 1,3-chirality transfer by intramolecular oxypalladation reaction. Tetrahedron Asymmetry 2005, 16, 1299–1303. [Google Scholar] [CrossRef]
  116. Uenishi, J.; Ohmi, M. Total synthesis of (−)-laulimalide; Pd-catalyzed stereospecific ring construction of substituted 3,6-dihydro [2H]pyran. Angew. Chem. Int. Ed. 2005, 44, 2756–2760. [Google Scholar] [CrossRef] [PubMed]
  117. Kawi, N.; Lagrange, J.-M.; Ohmi, M.; Uenishi, J. Palladium-catalyzed stereospecific synthesis of 2,6-disubstituted tetrahydropyrans: 1,3-chirality transfer by an intramolecular oxypalladation reaction. J. Org. Chem. 2006, 71, 4530–4537. [Google Scholar] [CrossRef]
  118. Kawai, N.; Lagrange, J.-M.; Uenishi, J. Stereochemistry and construction of tetrasubstituted chiral carbon centers in intramolecular Pd-catalyzed 1,3-chirality transfer reactions. Eur. J. Org. Chem. 2007, 2007, 2808–2814. [Google Scholar] [CrossRef]
  119. Uenishi, J.; Vikhe, Y.S.; Kawai, N. Stereochemistry and mechanistic study of intramolecular PdII-catalyzed oxypalladation and 1,3-chirality-transfer reactions. Chem. Asian J. 2008, 3, 473–484. [Google Scholar] [CrossRef]
  120. Hande, S.M.; Uenishi, J. Total synthesis of aspergillide B and structural discrepancy of aspergillide A. Tetrahedron Lett. 2009, 50, 189–192. [Google Scholar] [CrossRef]
  121. Uenishi, J.; Vikhe, Y.S. A short access to chiral non-racemic oxa- and azaheterocycles by cross-metathesis and Pd-catalyzed cyclization sequence. Heterocycles 2010, 80, 1463–1469. [Google Scholar] [CrossRef] [PubMed]
  122. Ida, A.; Kitao, K.; Hoshiya, N.; Uenishi, J. Hydroxy group directed stereochemistry in oxypalladation of chiral allylic alcohol. Tetrahedron Lett. 2015, 56, 1956–1959. [Google Scholar] [CrossRef]
  123. Murata, Y.; Uenishi, J. Stereochemistry of PdII-catalyzed THF ring formation of ε-hydroxy allylic alcohols and synthesis of 2,3,5-trisubstituted and 2,3,4,5-tetrasubstituted tetrahydrofurans. J. Org. Chem. 2016, 81, 7471–7485. [Google Scholar] [CrossRef] [PubMed]
  124. Suzuki, A.; Sasaki, M.; Nakagishi, T.; Ueda, T.; Hoshiya, N.; Uenishi, J. Construction of iterative tetrahydrofuran ring units and total synthesis of (+)-goniocin. Org. Lett. 2016, 18, 2248–2251. [Google Scholar] [CrossRef]
  125. Ghebreghiorgis, T.; Kirk, B.H.; Aponick, A.; Ess, D.H. Multiple mechanisms in Pd(II)-catalyzed SN2′ reactions of allylic alcohols. J. Org. Chem. 2013, 78, 7664–7673. [Google Scholar] [CrossRef]
  126. Zawisza, A.; Sinou, D. An unexpected palladium-catalyzed cyclization of bis-hydroxy allylic alcohols to dioxabicyclo [2.2.2]octanes. Tetrahedron Lett. 2006, 47, 3271–3274. [Google Scholar] [CrossRef]
  127. Olszewska, B.; Kryczka, B.; Zawisza, A. Synthesis of 2,5-substituted tetrahydrofurans by Pd(II)-catalyzed cyclization. Lett. Org. Chem. 2012, 9, 563–567. [Google Scholar] [CrossRef]
  128. Borah, M.; Borthakur, U.; Saikia, A.K. Diastereoselective synthesis of substituted morpholines from N-tethered alkenols: Total synthesis of (±)-chelonin A. J. Org. Chem. 2017, 82, 1330–1339. [Google Scholar] [CrossRef] [PubMed]
  129. Ida, A.; Hoshiya, N.; Uenishi, J. Investigation of Pd(II)-catalyzed cyclization of chiral θ-hydroxy-α,β,γ,δ-unsaturated dienol. Heterocycles 2015, 90, 1082–1093. [Google Scholar]
  130. Yasumoto, T. The chemistry and biological function of natural marine toxins. Chem. Rec. 2001, 1, 228–242. [Google Scholar] [CrossRef] [PubMed]
  131. Yokoyama, H.; Nishida, K.; Togawa, T.; Yamagami, M.; Miyazawa, M.; Hirai, Y. Stereoselective synthesis of the KJ ring system of yessotoxin by Pd(II)-catalyzed cyclization. Tetrahedron Lett. 2016, 57, 4379–4381. [Google Scholar] [CrossRef]
  132. Miyazawa, M.; Yano, Y.; Hayashi, C.; Okuno, M.; Hara, K.; Hirai, Y.; Yokoyama, H. Efficient synthesis of seven-membered cyclic ethers using Pd(II) catalyst. Heterocycles 2017, 94, 1885–1894. [Google Scholar] [CrossRef]
  133. Yokoyama, H.; Nakayama, S.; Murase, M.; Miyazawa, M.; Yamaguchi, S.; Hirai, Y. Pd(II)-catalyzed cyclization to ether and its application to the synthesis of the trans-fused polyether core. Heterocycles 2009, 77, 211–215. [Google Scholar] [CrossRef]
  134. Yokoyama, H.; Kusumoto, Y.; Sumiyoshi, K.; Miyazawa, M.; Hirai, Y. Synthetic studies of yessotoxin: Iterative synthesis of the AB ring system via Pd(II)-catalyzed cyclization of alcohol. Heterocycles 2014, 89, 353–358. [Google Scholar] [CrossRef]
  135. Sugimoto, M.; Nosho, S.; Hikosaka, G.; Hattori, Y.; Umezawa, K.; Kawamura, A.; Makabe, H. Synthesis of (+)-muconin via diastereoselective oxypalladation. J. Org. Chem. 2021, 86, 4859–4866. [Google Scholar] [CrossRef]
  136. Furuhata, S.; Hattori, Y.; Okajima, M.; Konno, H.; Abe, M.; Miyoshi, H.; Goto, T.; Makabe, H. Synthesis of pyranicin and its deoxygenated analogues and their inhibitory action with bovine heart mitochondrial complex I. Tetrahedron 2008, 64, 7695–7703. [Google Scholar] [CrossRef]
  137. Borrero, N.V.; Aponick, A. Total synthesis of acortatarin A using a Pd(II)-catalyzed spiroketalization strategy. J. Org. Chem. 2012, 77, 8410–8416. [Google Scholar] [CrossRef] [PubMed]
  138. Awasaguchi, K.; Miyazawa, M.; Uoya, I.; Inoue, K.; Nakamura, K.; Yokoyama, H.; Kakuda, H.; Hirai, Y. Synthesis of spiro C-arylglycoriboside via Pd(II)-catalyzed spirocyclization. Synlett 2010, 2010, 2392–2396. [Google Scholar]
  139. Koranne, P.S.; Tsujihara, T.; Arai, M.A.; Bajracharya, G.B.; Suzuki, T.; Onitsuka, K.; Sasai, H. Design and synthesis of chiral hybrid spiro (isoxazole–isoxazoline) ligands. Tetrahedron Asymmetry 2007, 18, 919–923. [Google Scholar] [CrossRef]
  140. Lankri, D.; Mostinski, Y.; Tsvelikhovsky, D. Palladium-catalyzed cascade assembly of tricyclic spiroethers from diene-alcohol precursors. J. Org. Chem. 2017, 82, 9452–9463. [Google Scholar] [CrossRef]
  141. Semmelhack, M.F.; Epa, W.R. Catalytic tandem oxy-palladation and vinylation. Tetrahedron Lett. 1993, 34, 7205–7208. [Google Scholar] [CrossRef]
  142. Lásiková, A.; Doháňošová, J.; Hlavínová, L.; Toffano, M.; Vo-Thanh, G.; Kožíšek, J.; Gracza, T. Domino reaction: Pd(II)-catalyzed cyclization of unsaturated polyols and cross-coupling. Tetrahedron Asymmetry 2012, 23, 818–827. [Google Scholar] [CrossRef]
  143. Abbineni, C.; Sasmal, P.K.; Mukkanti, K.; Iqbal, J. Synthesis of the C1-C12 fragment of amphidinolide T. Tetrahedron Lett. 2007, 48, 4259–4262. [Google Scholar] [CrossRef]
  144. Hewitt, J.F.M.; Williams, L.; Aggarwal, P.; Smith, C.D.; France, D.J. Palladium-catalyzed heteroallylation of unactivated alkenes—Synthesis of citalopram. Chem. Sci. 2013, 4, 3538–3543. [Google Scholar] [CrossRef]
  145. Keller, M.B. Citalopram therapy for depression: A review of 10 years of European experience and data from U.S. clinical trials. J. Clin. Psychiatry 2000, 61, 896–908. [Google Scholar] [CrossRef]
  146. Zhu, R.; Buchwald, S.L. Combined oxypalladation/C-H functionalization: Palladium(II)-catalyzed intramolecular oxidative oxyarylation of hydroxyalkenes. Angew. Chem. Int. Ed. 2012, 51, 1926–1929. [Google Scholar] [CrossRef]
  147. Ferreira, E.M.; Stoltz, B.M. Catalytic C-H bond functionalization with palladium(II):  aerobic oxidative annulations of indoles. J. Am. Chem. Soc. 2003, 125, 9578–9579. [Google Scholar] [CrossRef]
  148. Lafrance, M.; Rowley, C.N.; Woo, T.K.; Fagnou, K. Catalytic intermolecular direct arylation of perfluorobenzenes. J. Am. Chem. Soc. 2006, 128, 8754–8756. [Google Scholar] [CrossRef]
  149. Le Bras, J.; Muzart, J. Intermolecular dehydrogenative Heck reactions. Chem. Rev. 2011, 111, 1170–1214. [Google Scholar] [CrossRef]
  150. Matsuura, B.S.; Condie, A.G.; Buff, R.C.; Karahalis, G.J.; Stephenson, C.R.J. Intercepting Wacker intermediates with arenes: C–H functionalization and dearomatization. Org. Lett. 2011, 13, 6320–6323. [Google Scholar] [CrossRef] [PubMed]
  151. Blakemore, P.R.; Browder, C.C.; Hong, J.; Lincoln, C.M.; Nagornyy, P.A.; Robarge, L.A.; Wardrop, D.J.; White, J.D. Total synthesis of polycavernoside A, a lethal toxin of the red alga Polycavernosa tsudai. J. Org. Chem. 2005, 70, 5449–5460. [Google Scholar] [CrossRef] [PubMed]
  152. White, J.D.; Kranemann, C.L.; Kuntiyong, P. Intramolecular palladium(II)-mediated alkoxy carbonylation as a route to functionalized tetrahydropyrans. Synthesis of the C9-C32 segment of phorboxazole A. Org. Lett. 2001, 3, 4003–4006. [Google Scholar] [CrossRef]
  153. White, J.D.; Kuntiyong, P.; Lee, T.H. Total synthesis of phorboxazole A. 1. Preparation of four subunits. Org. Lett. 2006, 8, 6039–6042. [Google Scholar] [CrossRef]
  154. White, J.D.; Lee, T.H.; Kuntiyong, P. Total synthesis of phorboxazole A. 2. Assembly of subunits and completion of the synthesis. Org. Lett. 2006, 8, 6043–6046. [Google Scholar] [CrossRef] [PubMed]
  155. Searle, P.A.; Molinski, T.F. Phorboxazoles A and B: Potent cytostatic macrolides from marine sponge Phorbas sp. J. Am. Chem. Soc. 1995, 117, 8126–8131. [Google Scholar] [CrossRef]
  156. Yang, Z.; Xie, X.; Jing, P.; Zhao, G.; Zheng, J.; Zhao, C.; She, X. Total synthesis of cyanolide A. Org. Biomol. Chem. 2011, 9, 984–986. [Google Scholar] [CrossRef]
  157. Hong, J.Y.; Kim, H.S. Total synthesis of cyanolide A and confirmation of its absolute configuration. Org. Lett. 2010, 12, 2880–2883. [Google Scholar] [CrossRef]
  158. Erickson, K.L.; Gustafson, K.R.; Pannell, L.K.; Beutler, J.A.; Boyd, M.R. New dimeric macrolide glycosides from the marine sponge Myriastra clavosa. J. Nat. Prod. 2002, 65, 1303–1306. [Google Scholar] [CrossRef] [PubMed]
  159. Karlubíková, O.; Babjak, M.; Gracza, T. Tetrahydropyran synthesis by palladium(II)-catalysed hydroxycarbonylation of hexenols: Synthesis of (±)-diospongin A and (+)-civet cat compound. Tetrahedron 2011, 67, 4980–4987. [Google Scholar] [CrossRef]
  160. Yin, J.; Kouda, K.; Yasuhiro, T.; Tran, Q.L.; Miyahara, T.; Chen, Y.; Kadota, S. New diarylheptanoids from the rhizomes of Dioscorea spongiosa and their antiosteoporotic activity. Planta Med. 2004, 70, 54–58. [Google Scholar] [PubMed]
  161. Hegedus, L.S.; Allen, G.F.; Olsen, D.J. Palladium-assisted cyclization-insertion reactions. Synthesis of functionalized heterocycles. J. Am. Chem. Soc. 1980, 102, 3583–3587. [Google Scholar] [CrossRef]
  162. Semmelhack, M.F.; Bodurow, C. Intramolecular alkoxypalladation/carbonylation of alkenes. J. Am. Chem. Soc. 1984, 106, 1496–1498. [Google Scholar] [CrossRef]
  163. Ambrosini, L.M.; Cernak, T.A.; Lambert, T.H. Development of oxidative formylation and ketonylation reactions. Synthesis 2010, 2010, 870–881. [Google Scholar] [CrossRef]
  164. Semmelhack, M.F.; Bodurow, C.; Baum, M. Direct synthesis of pyran-lactones related to naphthoquinoneantibiotics. Tetrahedron Lett. 1984, 25, 3171–3174. [Google Scholar] [CrossRef]
  165. Tamaru, Y.; Kobayashi, T.; Kawamura, S.; Ochiai, H.; Hojo, M.; Yoshida, Z. Palladium catalyzed oxycarbonylation of 4-penten-1,3-diols: Efficient stereoselective synthesis of cis 3-hydroxytetrahydrofuran 2-acetic acid lactones. Tetrahedron Lett. 1985, 26, 3207–3210. [Google Scholar] [CrossRef]
  166. Gracza, T. Intramolecular oxycarbonylation in steroselective synthesis. In Stereoselective Synthesis of Drugs and Natural Products, 1st ed.; Andrushko, V., Andrushko, N., Eds.; Wiley: Oxford, UK, 2013; pp. 421–440. [Google Scholar]
  167. Semmelhack, M.F.; Shanmugam, P. Development of an approach to the synthesis of the plakortones. Tetrahedron Lett. 2000, 41, 3567–3571. [Google Scholar] [CrossRef]
  168. Patil, A.D.; Freyer, A.J.; Bean, M.F.; Carte, B.K.; Westley, J.W.; Johnson, R.K.; Lahouratate, P. The plakortones, novel bicyclic lactones from the sponge Plakortis halichondrioides: Activators of cardiac SR-Ca2+-pumping ATPase. Tetrahedron 1996, 52, 377–394. [Google Scholar] [CrossRef]
  169. Semmelhack, M.F.; Hooley, R.J.; Kraml, C.M. Synthesis of plakortone B and analogs. Org. Lett. 2006, 8, 5203–5206. [Google Scholar] [CrossRef]
  170. Hayes, P.Y.; Kitching, W. Total synthesis and absolute stereochemistry of plakortone D. J. Am. Chem. Soc. 2002, 124, 9718–9719. [Google Scholar] [CrossRef]
  171. Hayes, P.Y.; Chow, S.; Rahm, F.; Bernhardt, P.V.; De Voss, J.J.; Kitching, W. Synthesis of the sponge-derived plakortone series of bioactive compounds. J. Org. Chem. 2010, 75, 6489–6501. [Google Scholar] [CrossRef]
  172. Li, R.; Zhao, Q.; Li, S.; Han, Q.; Sun, H.; Lu, Y.; Zhang, L.; Zheng, Q. Micrandilactone A: A novel triterpene from Schisandra micrantha. Org. Lett. 2003, 5, 1023–1026, Erratum in Org. Lett. 2006, 8, 801. [Google Scholar] [CrossRef]
  173. Tang, Y.; Zhang, Y.; Dai, M.; Luo, T.; Deng, L.; Chen, J.; Yang, Z. A highly efficient synthesis of the FGH ring of micrandilactone A. Application of thioureas as ligands in the Co-catalyzed Pauson−Khand reaction and Pd-catalyzed carbonylative annulation. Org. Lett. 2005, 7, 885–888. [Google Scholar] [CrossRef]
  174. Dai, M.; Liang, B.; Wang, C.; Chen, J.; Yang, Z. Synthesis of a novel C2- symmetric thiourea and its application in the Pd-catalyzed cross-coupling reactions with arenediazonium salts under aerobic conditions. Org. Lett. 2004, 6, 221–224. [Google Scholar] [CrossRef]
  175. Dai, M.; Wang, C.; Dong, G.; Xiang, J.; Luo, Y.; Liang, B.; Chen, J.; Yang, Z. Development of thiourea-based ligands for the palladium-catalyzed bis(methoxycarbonylation) of terminal olefins. Eur. J. Org. Chem. 2003, 2003, 4346–4348. [Google Scholar] [CrossRef]
  176. Dai, M.; Liang, B.; Wang, C.; You, Z.; Xiang, J.; Dong, G.; Chen, J.; Yang, Z. A novel thiourea ligand applied in the Pd-catalyzed Heck, Suzuki and Suzuki carbonylative reactions. Adv. Synth. Catal. 2004, 346, 1669–1673. [Google Scholar]
  177. Nelson, R.A.; Pope, J.A., Jr.; Luedemann, G.M.; McDaniel, L.E.; Schaffner, C.P. Crisamicin A, a new antibiotic from Micromonospora. I. Taxonomy of the producing strain, fermentation, isolation, physico-chemical characterization and antimicrobial. J. Antibiot. 1986, 39, 334–344. [Google Scholar] [CrossRef] [PubMed]
  178. Li, Z.; Gao, Y.; Tang, Y.; Dai, M.; Wang, G.; Wang, Z.; Yang, Z. Total synthesis of crisamicin A. Org. Lett. 2008, 10, 3017–3020. [Google Scholar] [CrossRef] [PubMed]
  179. Li, Z.; Gao, Y.; Jiao, Z.; Wu, N.; Wang, D.Z.; Yang, Z. Diversity-oriented synthesis of fused pyran γ-lactones via an efficient Pd-thiourea-catalyzed alkoxycarbonylative annulation. Org. Lett. 2008, 10, 5163–5166. [Google Scholar] [CrossRef] [PubMed]
  180. Nesbitt, C.L.; McErlean, C.S.P. Total synthesis of C19 lipid diols containing a 2,5-disubstituted-3-oxygenated tetrahydrofuran. Org. Biomol. Chem. 2011, 9, 2198–2208. [Google Scholar] [CrossRef] [PubMed]
  181. Nesbitt, C.L.; McErlean, C.S.P. An expedient synthesis of 2,5-disubstituted-3-oxygenated tetrahydrofurans. Tetrahedron Lett. 2009, 50, 6318–6320. [Google Scholar] [CrossRef]
  182. Lee, E.; Yoo, S.-K.; Choo, H.; Song, H.Y. Radical cyclization of β-alkoxyacrylates: A formal synthesis of (−)-kumausallene. Tetrahedron Lett. 1998, 39, 317–318. [Google Scholar] [CrossRef]
  183. Evans, P.A.; Murthy, V.S.; Roseman, J.D.; Rheingold, A.L. Enantioselective total synthesis of the nonisoprenoid sesquiterpene (−)-kumausallene. Angew. Chem. Int. Ed. 1999, 38, 3175–3177. [Google Scholar] [CrossRef]
  184. Bai, Y.; Davis, D.C.; Dai, M. Synthesis of tetrahydropyran/tetrahydrofuran-containing macrolides by palladium-catalyzed alkoxycarbonylative macrolactonizations. Angew. Chem. Int. Ed. 2014, 53, 6519–6522. [Google Scholar] [CrossRef]
  185. Gracza, T.; Hasenöhrl, T.; Stahl, U.; Jäger, V. Synthesis of 3,5-anhydro-2-deoxy-1,4-glyconolactones by palladium(II)-catalyzed, regioselective oxycarbonylation of C5- and C6-enitols. ω-Homologation of aldoses to produce intermediates for C-glycoside/C-nucleoside synthesis. Synthesis 1991, 1991, 1108–1118. [Google Scholar] [CrossRef]
  186. Babjak, M.; Kapitán, P.; Gracza, T. The first total synthesis of goniothalesdiol. Tetrahedron Lett. 2002, 43, 6983–6985. [Google Scholar] [CrossRef]
  187. Babjak, M.; Kapitán, P.; Gracza, T. Synthesis of (+)-goniothalesdiol and (+)-7-epi-goniothalesdiol. Tetrahedron 2005, 61, 2471–2479. [Google Scholar] [CrossRef]
  188. Bhaket, P.; Morris, K.; Stauffer, C.S.; Datta, A. Total synthesis of cytotoxic anhydrophytosphingosine pachastrissamine (jaspine B). Org. Lett. 2005, 7, 875–876. [Google Scholar] [CrossRef]
  189. Caletková, O.; Ďurišová, D.; Gracza, T.; Prónayová, N. Aminohydroxylation of divinylcarbinol and its application to the synthesis of bicyclic hydroxypyrrolidine and aminotetrahydrofuran building blocks. Chem. Pap. 2013, 67, 66–75. [Google Scholar] [CrossRef]
  190. Cintulová, D.; Slahúčková, M.; Paštrnák, J.; Prónayová, N.; Szolcsányi, P. Preparation of cis-fused tetrahydropyranyl lactones via palladium-catalysed cyclocarbonylation of enediols. Synthesis 2017, 49, 755–762. [Google Scholar]
  191. Kapitán, P.; Gracza, T. Stereocontrolled oxycarbonylation of 4-benzyloxy-hepta-1,6-diene-3,5-diols promoted by chiral palladium(II) complexes. Tetrahedron Asymmetry 2008, 19, 38–44. [Google Scholar] [CrossRef]
  192. Kapitán, P.; Gracza, T. Asymmetric intramolecular Pd(II)-catalysed oxycarbonylation of alkene-1,3-diols. Arkivoc 2008, 8, 8–17. [Google Scholar] [CrossRef]
  193. Doháňošová, J.; Lásiková, A.; Toffano, M.; Gracza, T.; Vo-Thanh, G. Kinetic resolution of pent-4-ene-1,3-diol by Pd(II)-catalysed oxycarbonylation in ionic liquids. New J. Chem. 2012, 36, 1744–1750. [Google Scholar] [CrossRef]
  194. Markovič, M.; Ďuranová, M.; Koóš, P.; Szolcsányi, P.; Gracza, T. Synthesis of bis-tetrahydrofuran subunit of (−)-neopallavicinin. Tetrahedron 2013, 69, 4185–4189. [Google Scholar] [CrossRef]
  195. Babjak, M.; Markovič, M.; Kandríková, B.; Gracza, T. Asymmetric formal synthesis of (+)-pyrenolide D. Synthesis 2014, 46, 817–821. [Google Scholar]
  196. Zhang, C.; Liu, J.; Du, Y. A concise total synthesis of (+)-pyrenolide D. Tetrahedron Lett. 2013, 54, 3278–3280. [Google Scholar] [CrossRef]
  197. Babjak, M.; Markovič, M.; Kandríková, B.; Gracza, T. Homogeneous cyclocarbonylation of alkenols with iron pentacarbonyl. Synthesis 2014, 46, 809–816. [Google Scholar] [CrossRef]
  198. Markoviča, M.; Koóša, P.; Gracza, T. A short asymmetric synthesis of sauropunols A–D. Synthesis 2017, 49, 2939–2942. [Google Scholar]
  199. Lopatka, P.; Gavenda, M.; Markovič, M.; Koóš, P.; Gracza, T. Flow Pd(II)-catalyzed carbonylative cyclisation in the total synthesis of jaspine B. Catalysts 2021, 11, 1513. [Google Scholar] [CrossRef]
  200. Lopatka, P.; Markovič, M.; Koóš, P.; Ley, S.V.; Gracza, T. Continuous Pd-catalyzed carbonylative cyclization using iron pentacarbonyl as a CO source. J. Org. Chem. 2019, 84, 14394–14406. [Google Scholar] [CrossRef]
  201. Markoviča, M.; Lopatka, P.; Gracza, T.; Koóš, P. Recent applications of continuous flow in homogeneous palladium catalysis. Synthesis 2020, 52, 3511–3529. [Google Scholar] [CrossRef]
Scheme 1. The Wacker reaction.
Scheme 1. The Wacker reaction.
Catalysts 15 00845 sch001
Scheme 2. PdII-catalysed oxidative cyclisation of alkenols and a related domino reaction.
Scheme 2. PdII-catalysed oxidative cyclisation of alkenols and a related domino reaction.
Catalysts 15 00845 sch002
Scheme 3. Five-membered rings from oxidative cyclisation of primary alkenols.
Scheme 3. Five-membered rings from oxidative cyclisation of primary alkenols.
Catalysts 15 00845 sch003
Scheme 4. Five-membered rings from oxidative cyclisation of a secondary or tertiary alkenol.
Scheme 4. Five-membered rings from oxidative cyclisation of a secondary or tertiary alkenol.
Catalysts 15 00845 sch004
Scheme 5. Oxidations leading to six-membered rings.
Scheme 5. Oxidations leading to six-membered rings.
Catalysts 15 00845 sch005
Scheme 6. Oxidative cyclisation with O2 as sole oxidant.
Scheme 6. Oxidative cyclisation with O2 as sole oxidant.
Catalysts 15 00845 sch006
Scheme 7. Oxidative cyclisation with H2O2: PdII or PdIV intermediate?
Scheme 7. Oxidative cyclisation with H2O2: PdII or PdIV intermediate?
Catalysts 15 00845 sch007
Scheme 8. Synthesis of dihydropyranones or furanones; selectivity.
Scheme 8. Synthesis of dihydropyranones or furanones; selectivity.
Catalysts 15 00845 sch008
Scheme 9. Access to highly substituted furans.
Scheme 9. Access to highly substituted furans.
Catalysts 15 00845 sch009
Scheme 10. Enantioselective difunctionalisation across alkenes.
Scheme 10. Enantioselective difunctionalisation across alkenes.
Catalysts 15 00845 sch010aCatalysts 15 00845 sch010b
Scheme 11. Bridged bicyclic ketals.
Scheme 11. Bridged bicyclic ketals.
Catalysts 15 00845 sch011
Scheme 12. A multistep domino reaction.
Scheme 12. A multistep domino reaction.
Catalysts 15 00845 sch012
Scheme 13. Lactones via dual Wacker-type reactions.
Scheme 13. Lactones via dual Wacker-type reactions.
Catalysts 15 00845 sch013
Scheme 14. Reactivity dependence on solvent, CuCl2 amount and additives.
Scheme 14. Reactivity dependence on solvent, CuCl2 amount and additives.
Catalysts 15 00845 sch014
Scheme 15. Functionalised 2,5-dioxabicyclo [2.2.1]heptanes from alkenyl polyols.
Scheme 15. Functionalised 2,5-dioxabicyclo [2.2.1]heptanes from alkenyl polyols.
Catalysts 15 00845 sch015
Scheme 16. Oxidative cyclisation and C=C migration; reactivity dependence on catalyst and oxidant.
Scheme 16. Oxidative cyclisation and C=C migration; reactivity dependence on catalyst and oxidant.
Catalysts 15 00845 sch016
Scheme 17. Uenishi’s cyclisations of ene-diols and the cis-oxypalladation/syn elimination mechanism.
Scheme 17. Uenishi’s cyclisations of ene-diols and the cis-oxypalladation/syn elimination mechanism.
Catalysts 15 00845 sch017aCatalysts 15 00845 sch017b
Scheme 18. The competing anti-addition/anti-elimination proposal.
Scheme 18. The competing anti-addition/anti-elimination proposal.
Catalysts 15 00845 sch018
Scheme 19. Diastereoselectivity dependences.
Scheme 19. Diastereoselectivity dependences.
Catalysts 15 00845 sch019
Scheme 20. Access to a vinylmorpholine leading to chelonin A.
Scheme 20. Access to a vinylmorpholine leading to chelonin A.
Catalysts 15 00845 sch020
Scheme 21. Absence of 1,5-chirality transfer.
Scheme 21. Absence of 1,5-chirality transfer.
Catalysts 15 00845 sch021
Scheme 22. Towards the total synthesis of yessotoxin.
Scheme 22. Towards the total synthesis of yessotoxin.
Catalysts 15 00845 sch022
Scheme 23. Towards the synthesis of natural products.
Scheme 23. Towards the synthesis of natural products.
Catalysts 15 00845 sch023
Scheme 24. Highly functionalised tetrahydrofuran cores.
Scheme 24. Highly functionalised tetrahydrofuran cores.
Catalysts 15 00845 sch024
Scheme 25. Deprotection/heterocyclisation.
Scheme 25. Deprotection/heterocyclisation.
Catalysts 15 00845 sch025
Scheme 26. Intramolecular domino Wacker/Heck reactions.
Scheme 26. Intramolecular domino Wacker/Heck reactions.
Catalysts 15 00845 sch026
Scheme 27. Intermolecular domino Wacker/Heck reactions.
Scheme 27. Intermolecular domino Wacker/Heck reactions.
Catalysts 15 00845 sch027
Scheme 28. Intermolecular domino Wacker/dehydrogenative Heck reactions.
Scheme 28. Intermolecular domino Wacker/dehydrogenative Heck reactions.
Catalysts 15 00845 sch028
Scheme 29. Alkoxy- and hydroxycarbonylation of Wacker intermediates.
Scheme 29. Alkoxy- and hydroxycarbonylation of Wacker intermediates.
Catalysts 15 00845 sch029aCatalysts 15 00845 sch029b
Scheme 30. Semmelhack and Kitching syntheses of plakortones.
Scheme 30. Semmelhack and Kitching syntheses of plakortones.
Catalysts 15 00845 sch030
Scheme 31. Yang approach and synthesis of “micrandilactone A” and crisamicin A.
Scheme 31. Yang approach and synthesis of “micrandilactone A” and crisamicin A.
Catalysts 15 00845 sch031aCatalysts 15 00845 sch031b
Scheme 32. McErlean’s approach to natural products.
Scheme 32. McErlean’s approach to natural products.
Catalysts 15 00845 sch032
Scheme 33. Dai’s macrolactonisations.
Scheme 33. Dai’s macrolactonisations.
Catalysts 15 00845 sch033
Scheme 34. The last 20 years of the domino cyclisation/alkoxycarbonylative lactonisation journey of the Slovak University of Technology.
Scheme 34. The last 20 years of the domino cyclisation/alkoxycarbonylative lactonisation journey of the Slovak University of Technology.
Catalysts 15 00845 sch034aCatalysts 15 00845 sch034bCatalysts 15 00845 sch034c
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Muzart, J. PdII Catalysis: Recent Advances in the Intramolecular Wacker-Type Reaction of Alkenols and Related Domino Reactions. Catalysts 2025, 15, 845. https://doi.org/10.3390/catal15090845

AMA Style

Muzart J. PdII Catalysis: Recent Advances in the Intramolecular Wacker-Type Reaction of Alkenols and Related Domino Reactions. Catalysts. 2025; 15(9):845. https://doi.org/10.3390/catal15090845

Chicago/Turabian Style

Muzart, Jacques. 2025. "PdII Catalysis: Recent Advances in the Intramolecular Wacker-Type Reaction of Alkenols and Related Domino Reactions" Catalysts 15, no. 9: 845. https://doi.org/10.3390/catal15090845

APA Style

Muzart, J. (2025). PdII Catalysis: Recent Advances in the Intramolecular Wacker-Type Reaction of Alkenols and Related Domino Reactions. Catalysts, 15(9), 845. https://doi.org/10.3390/catal15090845

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