Prins cyclization-mediated stereoselective synthesis of tetrahydropyrans and dihydropyrans: an inspection of twenty years

  1. ORCID Logo ,
  2. ,
  3. and
  4. ORCID Logo
Division of Molecular Synthesis & Drug Discovery, Centre of Biomedical Research (CBMR), SGPGIMS Campus Raebareli Road, Lucknow, 226014 Uttar Pradesh, India
  1. Corresponding author email
Associate Editor: S. Bräse
Beilstein J. Org. Chem. 2021, 17, 932–963. https://doi.org/10.3762/bjoc.17.77
Received 16 Jan 2021, Accepted 14 Apr 2021, Published 29 Apr 2021
Review
cc by logo
Album

Abstract

Functionalized tetrahydropyran (THP) rings are important building blocks and ubiquitous scaffolds in many natural products and active pharmaceutical ingredients (API). Among various established methods, the Prins reaction has emerged as a powerful technique in the stereoselective synthesis of the tetrahydropyran skeleton with various substituents, and the strategy has further been successfully applied in the total synthesis of bioactive macrocycles and related natural products. In this context, hundreds of valuable contributions have already been made in this area, and the present review is intended to provide the systematic assortment of diverse Prins cyclization strategies, covering the literature reports of the last twenty years (from 2000 to 2019), with an aim to give an overview on exciting advancements in this area and designing new strategies for the total synthesis of related natural products.

Introduction

6-Membered saturated oxygen heterocycles, known as tetrahydropyran (THP), are recognized as privileged scaffolds, present in a variety of biologically important natural products, such as polyether antibiotics, marine toxins, pheromones, and pharmaceutical agents. These structural motifs are frequently used as synthons and as key intermediates for natural product synthesis. Therefore, the development of stereoselective synthetic methods for the substituted THP subunit has long been the area of fundamental research in organic chemistry. Thus far, several methods have been devised for the construction of substituted tetrahydropyran rings. Since the year 2000, a number of conceptually different reactions have been developed for the efficient construction of THP rings and were eventually employed in the total synthesis of natural products [1-8]. Prins and related cyclization reactions [9,10], hetero-Diels–Alder cyclization [11], cyclization onto epoxides [12], Petasis–Ferrier rearrangement [13], intramolecular oxa-Michael reactions [14], cyclization through oxidative C–H bond functionalization [15], ring-closing metathesis (RCM) [16], halo etherification [17], reductive etherification [18,19], and metal-mediated cyclization [20,21], etc. are the most frequent strategies utilized for THP ring construction (Scheme 1). Amongst all, the Prins reaction has proven as a powerful technique in the stereoselective synthesis of the THP key scaffold and its application towards the total synthesis of natural products. Many advancements were also taking place in Prins cyclization methodologies over that period of time. This appraisal aims to bring together the work of many research groups in the area of the development of Prins and related cyclization strategies along with the discussion on the general mechanistic part. We sincerely hope that this review will deliver a snapshot of the up-to-date state of the inventiveness in this field, and most importantly, it will give an inspiration to the reader to take up the challenge and contribute greater advances in this area in the future. This review comprises the literature reports over the last twenty years and its advances. It is likely that some references may have escaped our attention unintentionally, for which we would greatly apologize to those whose contribution in this area has not been included.

[1860-5397-17-77-i1]

Scheme 1: General strategy for the synthesis of THPs.

Review

Prins cyclization: general

For the first time, in 1899, Kriewitz [22] reported the synthesis of unsaturated pinene alcohol through a thermal ene reaction using β-pinene and paraformaldehyde. After nearly twenty years, Prins explored this reaction further for the synthesis of diol by the condensation of styrene and paraformaldehyde in the presence of a Brønsted acid [23,24]. The major breakthrough for this reaction was reported by Hanschke in 1955, when the THP ring was selectively constructed through a Prins reaction involving 3-butene-1-ol and a variety of aldehydes or ketones in the presence of acid (Scheme 2) [25].

[1860-5397-17-77-i2]

Scheme 2: Developments towards the Prins cyclization.

Although the Kriewitz reaction was an ene reaction, the mechanism of the reaction was described to proceed via an oxocarbenium ion intermediate captured by a π-nucleophile, followed by the addition of an external nucleophile, leading to the formation of products. Since then, the Prins cyclization emerged as the most commonly used strategy for the stereoselective construction of the THP ring, and its application lead to some excellent reviews on the Prins reaction [26,27]. In general, an endo cyclization proceeds via an oxocarbenium ion intermediate in a stereoselective manner for THP ring formation as shown in Scheme 3.

[1860-5397-17-77-i3]

Scheme 3: General stereochemical outcome of the Prins cyclization.

The outcome of exclusive cis-stereoselectivity in the Prins cyclization might be attributed to the most favorable conformation adopted by 12 with equatorial orientation of the 2,6-substituents (R1 and R2). Alder and co-workers explained the formation of all-cis-2,4,6-trisubstituted THPs with the help of density functional theory (DFT) and stated that in the presence of an external nucleophile, the stabilization of the carbocation intermediate is favored through hyperconjugation [28]. The vacant p-orbital of C4 in TS 12a overlaps efficiently with the HOMO of the incoming nucleophile in an equatorial attack. Furthermore, this pseudoaxial C4 hydrogen atom in TS 12a leads to an optimal overlap between σ and σ* of C2–C3 and C5–C6 with the coplanar equatorial lone pair of the oxygen atom and the empty p-orbital at C4. These orbital stabilizations, along with the lack of 1,3-diaxial interaction experienced by the incoming nucleophile (mostly halide) leads to the preferential equatorial attack over an axial attack by the nucleophile (Scheme 3) to give all-cis-2,4,6-trisubstituted THPs. In the absence of an external nucleophile, the successive proton loss leads to the formation of the 2,6-disubstituted dihydropyran. The regioselectivity of the Prins reaction is explained through the intermediates formed during the course of the reaction (Scheme 4). The Z-homoallylic alcohol reacts with an activated aldehyde to give oxocarbenium ion 15, wherefrom two competing transition states, 15a and 15b, can possibly form. In the 6-membered chair-like transition state 15a, there is a 1,3-diaxial interaction between “H” and the substituent R2, while for the other five-membered transition state 15b, there is no such 1,3-diaxial interaction, which favors the formation of tetrahydrofuran product 17 instead of the tetrahydropyran 16 (Scheme 4).

[1860-5397-17-77-i4]

Scheme 4: Regioselectivity in the Prins cyclization.

Although the Prins cyclization is one of the powerful tools for the construction of 2,6-disubstituted THPs, there are some limitations that restrict a wide applicability. The major drawbacks identified with the Prins cyclization are the racemization due to competing oxonia-Cope rearrangement and side-chain exchange. Willis and co-workers studied the reactivity of the Prins reaction of different aryl group-substituted homoallylic alcohols 18 with propanal in the presence of a Lewis acid, which furnished the expected tetrahydropyran 23 as a single diastereomer via an oxocarbenium intermediate 21 (Scheme 5) [29,30].

[1860-5397-17-77-i5]

Scheme 5: Mechanism of the oxonia-Cope reaction in the Prins cyclization.

The reaction was dependent from the nature of the aromatic ring, which plays a crucial role in the product formation. Homoallylic alcohols with an electron-rich substituent at the arene ring produced predominantly symmetric THP product 26 over the desired trisubstituted heterocycle 23. The mechanism of the reaction was further investigated using enantioenriched homoallylic alcohol (S)-18 with 89% ee, which favored 2-oxonia-Cope rearrangement to give THP 23 only in 14% yield and <5% ee. The poor enantiomeric excess of the product 23 indicates that the racemization takes place during the course of the reaction. It was explained that the reason for the loss of optical purity was due to the formation of a benzylic cation, which is stabilized by the electron-rich aromatic substituent. In contrast, the reaction with aromatic aldehydes equipped with the electron-deficient substituent produced the desired trisubstituted THP along with recovered starting material. The enantioenriched homoallylic alcohol bearing an electron-deficient substituent, 27 (94% ee), was investigated with propanal, which proceeded with high selectivity to give the corresponding THP 28 (79% ee, 32% yield) along with some recovered starting material (47%), as shown in Scheme 6.

[1860-5397-17-77-i6]

Scheme 6: Cyclization of electron-deficient enantioenriched alcohol 27.

Partial racemization was also reported at the same time by reversible 2-oxonia-Cope rearrangement and via side-chain exchange [31-33]. The racemization occurs during allyl transfer as a result of 2-oxonia-Cope rearrangement through a 3,3-sigmatropic shift, which plays a crucial role during the reaction, as shown in Scheme 7.

[1860-5397-17-77-i7]

Scheme 7: Partial racemization through 2-oxonia-Cope allyl transfer.

The Prins cyclization between alcohol (R)-35 and aldehyde 36 was investigated under different Lewis acid conditions, as shown in Scheme 8 [33]. Cyclization promoted by BF3·OEt2/HOAc led to partial racemization of the desired product 37 (from 87% ee to 68% ee) and formation of side-chain exchange products 38 and 39 (symmetric tetrahydropyran). Presumably, this observation stands in support of the intervention of a 2-oxonia-Cope-mediated side-chain exchange reaction and is entirely consistent with Willis and co-workers’ result [29], which leads to the partial racemization observed in the desired product formation. Another Lewis acid, SnBr4, was found to be more efficient than BF3·OEt2/HOAc in terms of retention of enantiopurity in major product 37 during cyclization (from 87% ee to 85% ee, Scheme 8). This could probably be due to a faster rate of cyclization with SnBr4, which suppressed the competing 2-oxonia-Cope process.

[1860-5397-17-77-i8]

Scheme 8: Partial racemization by reversible 2-oxonia-Cope rearrangement.

In order to stop racemization during the Prins cyclization, a substrate in which an oxocarbenium ion is generated from a masked aldehyde bearing a homoallylic alcohol moiety has been examined. In this context, the α-acetoxy ethers with different functionalities at C4 were examined in the presence of a variety of Lewis acids, and it was found that the α-acetoxy ether (R)-42 underwent Prins-type cyclization in the presence of BF3·OEt2 as well as SnBr4 to deliver the desired 37 and 40, respectively, without loss of optical purity (Scheme 9) [34,35].

[1860-5397-17-77-i9]

Scheme 9: Rychnovsky modification of the Prins cyclization.

This strategy was successfully utilized for the synthesis of the natural product (−)-centrolobine [33] and for the stereoselective synthesis of the C20–C27 tetrahydropyran segment of phorboxazole A (Scheme 10) [36].

[1860-5397-17-77-i10]

Scheme 10: Synthesis of (−)-centrolobine and the C22–C26 unit of phorboxazole A.

Axial selectivity in the Prins cyclization

To overcome the racemization process, the axially selective Prins cyclization was explored with a variety of substrates, which produced the corresponding THPs in excellent selectivity and good to excellent yield [37]. The experimental modification under segment coupling gave entirely the 4-axial product. For example, treatment of 47 with SnBr4 produced axial and equatorial products 48a and 48b in a 9:79 ratio under typical segment coupling. This selectivity was further improved for the formation of 48a by exclusively using TMSBr as a Lewis acid, as shown in Scheme 11.

[1860-5397-17-77-i11]

Scheme 11: Axially selective Prins cyclization by Rychnovsky et al.

The mechanistic rationale for an axially selective Prins cyclization is explained in Scheme 12 [38]. It is proposed that the reaction of 49 with TMSBr forms an intermediate 50, which, after solvolysis, affords an intimate ion pair 51. The proximal addition of a bromide ion to 51 produces axial adduct 56 exclusively. However, when SnBr4 is used as a Lewis acid, oxocarbenium ion 52 is formed via 50. The counterion SnBr4 being much less nucleophilic than the Br ion allows the formation of a solvent-separated ion pair 53, which results in the bromide addition to 53 preferentially from an equatorial position (Scheme 12).

[1860-5397-17-77-i12]

Scheme 12: Mechanism for the axially selectivity Prins cyclization.

Mukaiyama aldol–Prins cyclization

The Mukaiyama aldol–Prins (MAP) cyclization has also been explored for the synthesis of tetrahydropyran. In this approach, the side reaction is avoided by introducing a nucleophile into the enol ether, which traps the reactive oxocarbenium ion intermediate 60, leading to the formation of THP [39]. The first example of an MAP cascade reaction was reported by Rychnovsky and co-workers using allylsilane 62 as an internal nucleophile, as shown in Scheme 13 [40].

[1860-5397-17-77-i13]

Scheme 13: Mukaiyama aldol–Prins cyclization reaction.

This approach was further extended to the synthesis of the macrolide lecasacandrolide A [41]. BF3·OEt2 in combination with 2,6-di-tert-butylpyridine (DTBP) was a suitable combination for the synthesis of the THP unit of leucasacandrolide A, while TiBr4 [42] was found suitable in conjunction with DTBP for the synthesis of polyketide SCH 351448 [43], as shown in Scheme 14.

[1860-5397-17-77-i14]

Scheme 14: Application of the aldol–Prins reaction.

Hart and Bennett have also examined the trifluroacetic acid-catalyzed Prins cyclization of acetal 71 to afford 72 along with side-chain-exchanged product 73 (Scheme 15) [44].

[1860-5397-17-77-i15]

Scheme 15: Hart and Bennet's acid-promoted Prins cyclization.

This method was utilized for the synthesis of (−)-blepharocalyxin D29 [45] and the macrolide leucascandrolide A [46]. In another type, the triflic acid-catalyzed Prins cyclization was used for the synthesis of 2,4,5,6-tetrasubstituted tetrahydropyran with complete control of stereochemistry, which is an important core of a variety of natural products, such as polycarvernoside A [47], clavoslide A [48], and (−)-blepharocalyxin [49,50] and its analogs, as shown in Scheme 16.

[1860-5397-17-77-i16]

Scheme 16: Tetrahydropyran core of polycarvernoside A as well as (−)-clavoslide A and D.

Additionally, the reaction was used for the synthesis of rhoiptelol B, 7-desmethoxyfusarentin, and corresponding analogs [51]. Considering β-hydroxydioxinone as a better starting material for Prins cyclization, Scheidt and co-workers introduced a new method to access highly functionalized chiral THP efficiently (Scheme 17) [52].

[1860-5397-17-77-i17]

Scheme 17: Scheidt and co-workers’ route to tetrahydropyran-4-one.

Furthermore, the possible reaction pathway indicates the formation of oxocarbenium ion 82, followed by C–C bond formation via a chair-like transition state to afford 83 (Scheme 18). A sequence of reactions involving elimination of a proton from 83, treatment of 84 with an alkoxide, and protonation of the resulting enolate delivered thermodynamically favored equatorial ester 80 and 81.

[1860-5397-17-77-i18]

Scheme 18: Mechanism for the Lewis acid-catalyzed synthesis of tetrahydropyran-4-one.

The highly diastereoselective Brønsted superacid-catalyzed Prins cyclization of unsaturated enol ether 85 to cis-2,6-disubstituted 4-methylenetetrahydropyran 86 (55% yield) as shown in Scheme 19 was reported by Hoveyda and co-workers [53].

[1860-5397-17-77-i19]

Scheme 19: Hoveyda and co-workers’ strategy for 2,6-disubstituted 4-methylenetetrahydropyran.

Funk and Cossey demonstrated that ene-carbamate could be an excellent terminating group for Prins cyclization [54]. The reaction involved 87 in the presence of the mild Lewis acid InCl3 and benzaldehyde (88), which produced all-cis-tetrahydropyran-4-one 90 in excellent yield. The transformation proceeded through cyclization of a diequatorial chair-like conformation of the oxocarbenium ion 89 to provide an N-acyliminium ion, which upon hydrolysis produced 90. Similarly, the reaction of 91 produced all-cis-2,3,6-trisubstituted tetrahydropyran 93. The application of this reaction was further extended by an exceptionally concise formal total synthesis of the nuclear export inhibitor (+)-ratjadone A, as shown in Scheme 20.

[1860-5397-17-77-i20]

Scheme 20: Funk and Cossey’s ene-carbamates strategy.

Stereoselective Prins cyclization of substituted cyclopropylcarbinol 94 to 2,4,6-trisubstituted tetrahydropyran 97 was reported by Yadav and Kumar [55]. In this reaction, a homoallylic cation was generated from 94 by the opening of the cyclopropane ring in the presence of TFA, which upon reacting with an aldehyde delivered 2,4,6-trisubstituted tetrahydropyran 97 through Prins cyclization, as shown in Scheme 21.

[1860-5397-17-77-i21]

Scheme 21: Yadav and Kumar’s cyclopropane strategy for THP synthesis.

Similarly, an SnCl4-catalyzed Prins reaction was reported for the synthesis of 4-chlorotetrahydropyran 100. This intermediate was further utilized for the synthesis of the natural product centrolobine, as shown in Scheme 22 [56].

[1860-5397-17-77-i22]

Scheme 22: 2-Arylcylopropylmethanolin in centrolobine synthesis.

A strategy involving BiCl3-catalyzed microwave-assisted Prins cyclization of homoallylic alcohol 101 with an aldehyde 102 was successfully employed for the synthesis of 4-chloro-cis-2,6-disubstituted tetrahydropyran 103 as a single diastereomer [57], as shown in Scheme 23.

[1860-5397-17-77-i23]

Scheme 23: Yadav and co-workers’ strategy for the synthesis of THP.

In continuation, 4-amidotetrahydropyran derivative 106 was also synthesized from homoallylic alcohol 104 and an aldehyde 105 using a combination of cerium chloride and acetyl chloride following a Prins–Ritter reaction sequence (Scheme 24) [58]. 10 mol % cerium chloride was used as a reaction promotor, which dramatically improved the reaction rate and yield of the reaction.

[1860-5397-17-77-i24]

Scheme 24: Yadav and co-workers’ Prins–Ritter reaction sequence for 4-amidotetrahydropyran.

In a related study, the synthesis of polysubstituted tetrahydropyrans was described by Amberlyst® 15-catalyzed cyclization of homoallyl alcohol 107 and aldehydes 108. This method was further employed for the synthesis of highly substituted tetrahydropyrans with three contiguous stereocenters in one single operation [59]. The utility of this approach is showcased in the enantioselective total synthesis of (+)-prelactones B, C, and V, as shown in Scheme 25.

[1860-5397-17-77-i25]

Scheme 25: Yadav and co-workers’ strategy to prelactones B, C, and V.

Yadav’s group reported the synthesis of 4-iodotetrahydropyrans (dr = 7.5:2.5) from aromatic aldehyde 111 and homoallylic alcohol 110 using TMSCl and NaI. Furthermore, the major diastereomer was utilized for the synthesis of centrolobine, as shown in Scheme 26 [60].

[1860-5397-17-77-i26]

Scheme 26: Yadav and co-workers’ strategy for the synthesis of (±)-centrolobine.

Loh and co-workers have shown the construction of cis-2,6-disubstituted tetrahydropyran 116 with an exocyclic double bond by reacting homoallylic alcohol 114 and aldehyde 115 in the presence of a catalytic amount of In(OTf)3 [61]. This approach was further used for the synthesis of a common intermediate 117 for (−)-zampanolide and (+)-dactyloide (Scheme 27) [62].

[1860-5397-17-77-i27]

Scheme 27: Loh and co-workers’ strategy for the synthesis of zampanolide and dactylolide.

Further improvement of this reaction was achieved by carrying out the Prins cyclization between homoallyl alcohol 118 (or using the corresponding aldehyde and allylsilyl chloride 119) and an aldehyde 120 in the presence of a catalytic amount of the mild Lewis acid In(OTf)3 and trimethylsilyl halide as an additive to produce cis-4-halo-2,6-disubstituted tetrahydropyran 121 (Scheme 28) [63,64]. It was noticed that the problem associated with epimerization of the substrate has been successfully overcome in this reaction, which was demonstrated in the enantioselective total synthesis of (−)-centrolobine using catalytic InBr3 as a mild Lewis acid.

[1860-5397-17-77-i28]

Scheme 28: Loh and Chan’s strategy for THP synthesis.

This strategy was further explored to construct tetrasubstituted cis-2,6-disubstituted 4,5-dibromotetrahydropyran 124 with high stereoselectivity using γ-brominated homoallylic alcohol (Z)-122 and aldehyde 123 in the presence of InBr3 and TMSBr in CH2Cl2 at 0 °C (Scheme 29) [65].

[1860-5397-17-77-i29]

Scheme 29: Prins cyclization of cyclohexanecarboxaldehyde.

Metzger and co-workers reported an AlCl3-catalyzed cyclization of methyl ricinoleate (127) with various aldehydes to produce 2,3,6-trialkyl-substituted 4-chlorotetrahydropyran 128 with excellent stereocontrol in all-cis-configuration (Scheme 30) [66].

[1860-5397-17-77-i30]

Scheme 30: Prins cyclization of methyl ricinoleate (127) and benzaldehyde (88).

The stereochemical outcome of this cyclization was rationalized by a chair-like transition state to produce predominantly the all-cis product. Similarly, cis- and trans-129 (1:4), generated in situ from methyl 10-undecenoate and an aldehyde 130 via ene reaction, undergo cyclization to form THPs 131 and 132 (Scheme 31).

[1860-5397-17-77-i31]

Scheme 31: AlCl3-catalyzed cyclization of homoallylic alcohol 129 and aldehyde 130.

Martín and co-workers reported a general strategy based on a reaction sequence of Evans aldol addition to construct a homoallylic alcohol, followed by Prins cyclization to furnish 2,3,4,5,6-pentasubstituted tetrahydropyrans 137 using β,γ-unsaturated N-acyloxazolidin-2-ones 134 as a key precursor [67]. In this Evans aldol−Prins (EAP) protocol, four new σ-bonds and five contiguous stereocenters were generated as shown in Scheme 32.

[1860-5397-17-77-i32]

Scheme 32: Martín and co-workers’ stereoselective approach for the synthesis of highly substituted tetrahydropyrans through an Evans aldol−Prins cyclization strategy.

Silyl-Prins cyclization

Extensive research efforts were made towards the synthesis of THP using the silyl-Prins cyclization reaction. In this reaction, an oxocarbenium ion is being trapped by allylsilanes, vinylsilanes, alkenyl methylsilanes, or propargylsilanes to produce a variety of the Prins-cyclized products. The allyl metalation, followed by intramolecular Sakurai cyclization (IMSC) provides an efficient route to a variety of tetrahydropyran derivatives, as described by Marko and Leroy [68,69]. In these approaches, an initial ene reaction between an aldehyde 139 and the allylsilane 138 was promoted by Et2AlCl to generate Z-configured homoallylic alcohol 140. Condensation of 140 with another aldehyde in the presence of BF3⋅OEt2 afforded the polysubstituted exo-methylene tetrahydropyran 142 in a completely stereocontrolled manner. The reaction proceeded via oxocarbenium 141, which upon intramolecular trapping by the allylsilane moiety through a chair-like transition state delivers the product (Scheme 33) [68].

[1860-5397-17-77-i33]

Scheme 33: Ene-IMSC strategy by Marko and Leroy for the synthesis of tetrahydropyran.

An analogous reaction was reported between (E)-enol carbamate 143 and an aldehyde 144 in the presence of BF3·OEt2 to provide THP 146 with exquisite diastereoselectivity. The carbamate substituent adopted the axial disposition in the proposed transition state 145, as shown in Scheme 34 [69].

[1860-5397-17-77-i34]

Scheme 34: Marko and Leroy’s strategy for the synthesis of tetrahydropyrans 146.

In another report by Rychnovsky and Gisinsky, two of the tetrahydropyran rings of the potent molluscicide cyanolide A were synthesized via a silyl-Prins cyclization and Sakurai macrocyclization/dimerization strategy to produce 150 in the presence of TMSOTf, as shown in Scheme 35 [70].

[1860-5397-17-77-i35]

Scheme 35: Sakurai dimerization/macrolactonization reaction for the synthesis of cyanolide A.

Hoye and Hu utilized camphor sulfonic acid (CSA) to construct a cis-2,6-disubstituted tetrahydropyran 153 via an intramolecular Sakurai cyclization reaction between the enal 151 and an allylisilane 152. Further manipulation of functional groups of 153 leads to the synthesis of (−)-dactyloide (Scheme 36) [71].

[1860-5397-17-77-i36]

Scheme 36: Hoye and Hu’s synthesis of (−)-dactyloide by intramolecular Sakurai cyclization.

The one-pot synthesis of a 2,6-disubstituted THP was reported by Minehan and co-workers and involved treating 3-iodo-2-[(trimethylsilyl)methyl]propene with an aldehyde in the presence of indium metal to produce homoallylic alcohol 156 (Scheme 37), which underwent a silyl-Prins cyclization to form polysubstituted exo-methylene THPs 157 [72].

[1860-5397-17-77-i37]

Scheme 37: Minehan and co-workers’ strategy for the synthesis of THPs 157.

Tandem allylation–silyl-Prins cyclization

Tetrahydropyran can also be synthesized stereoselectively by sequential allylation to an aldehyde, followed by silyl-Prins cyclization of the resulting homoallylic alcohol. For illustration, a facile enantioselective strategy for the synthesis of cis-2,6-disubstituted 4-methylenetetrahydropyran 161 (91% yield, dr = 5:1) was reported by Yu et al, utilizing, first, asymmetric allylation of an aldehyde by using [{(R)-BINOL}Ti(IV){OCH(CF3)2}2] as a chiral promotor in PhCF3, followed by cyclization using R2CHCl(OMe) in the presence of TMSNTf2, as shown in Scheme 38 [73]. The internal chirality transfer during cyclization probably took place due to the geometrical preference of 162 to minimize the allylic strain with the existing stereogenic center (pseudoaxial group), leading to the formation of cis-tetrahydropyran 161 rather than a trans-tetrahydropyran.

[1860-5397-17-77-i38]

Scheme 38: Yu and co-workers’ allylic transfer strategy for the construction of tetrahydropyran 161.

Floreancig and co-workers utilized a tandem allylation–silyl-Prins cyclization strategy to afford 2,6-disubstituted tetrahydropyran 167 by ionizing α,β-unsaturated acetals 164 in the presence of electron-rich olefins using Ce(NO3)3 and SDS in water [74]. The mechanism of the reaction is shown in Scheme 39, which plausibly proceeded through trapping of oxocarbenium ion 166 in a chair-like transition state.

[1860-5397-17-77-i39]

Scheme 39: Reactivity enhancement in intramolecular Prins cyclization.

The stability of the acetal under these reaction conditions reflected that the acid-sensitive functional groups are well tolerated in the cyclized product. Furthermore, a natural product, (+)-dactyloide, was synthesized by following the above strategy using an appropriate acetal (Scheme 40) [75].

[1860-5397-17-77-i40]

Scheme 40: Floreancig and co-workers’ Prins cyclization strategy to (+)-dactyloide.

The synthesis of enantiomerically enriched 172, cis-2,6-DHP and trans-2,6-DHP, respectively, was reported by a [4 + 2]-annulation strategy. The authors utilized crotylsilanes syn-170 and anti-170, respectively, with an aldehyde 171 in the presence of TMSOTf to deliver different DHPs 172 (Scheme 41) [76].

[1860-5397-17-77-i41]

Scheme 41: Panek and Huang’s DHP synthesis from crotylsilanes: a general strategy.

For syn-170, the reaction went via the favored boat-like transition state 173 instead of the disfavored chair-like transition state 174 to cis,trans-175 as a product (Scheme 42).

[1860-5397-17-77-i42]

Scheme 42: Panek and Huang’s DHP synthesis from syn-crotylsilanes.

In contrast, the reaction of anti-170 proceeded, however, through similar boat-like transition states 176 and 177 where the interaction between the methyl substituent and the alkyl group of aldehyde was less, leading to the formation of trans,trans-178 as a major product, as shown below in Scheme 43.

[1860-5397-17-77-i43]

Scheme 43: Panek and Huang’s DHP synthesis from anti-crotylsilanes.

A variety of natural products, such as (−)-apicularen A [77], the C1–C13 fragment of bistramide A71 [78], herboxideiene/GEX1A [79], (+)-kendomycin [80], and (+)-SCH351448 [81] were synthesized utilizing this [4 + 2]-annulation strategy. Following the above annulation route, later, Roush's group Introduced β-hydroxyallylsilanes for the synthesis of 2,6-disubstituted DHP (Scheme 44) [82].

[1860-5397-17-77-i44]

Scheme 44: Roush and co-workers’ [4 + 2]-annulation strategy for DHP synthesis [82].

This strategy was further utilized for the synthesis of the C29−C45 bispyran subunit (E−F) of spongistatin [82]. 2,6-Disubstituted 4-methylenetetrahydropyran was also synthesized from silylstannane and two units of aldehyde in a two-step protocol. The first step involves the addition between silylstannane 185 and an aldehyde in the presence of titanium BINOLate as a catalyst, which afforded hydroxyallylsilane 186 with excellent enantioselectivity (Scheme 45) [83-85]. This, upon further reaction with another aldehyde in the presence of TMSOTf, gave 2,6-disubstituted 4-methylenetetrahydropyran 187. This strategy was utilized for the synthesis of bryostatin and (+)-dactyloide analogs [86-88].

[1860-5397-17-77-i45]

Scheme 45: TMSOTf-promoted annulation reaction.

Similar to Prins cyclization of allylsilanes, Dobbs and co-workers recently utilized the corresponding vinylsilane as an alternative for the synthesis of cis-2,6-dihydropyran [89,90]. The synthesis involves tandem addition of vinylsilane, followed by silyl-Prins cyclization reaction. For example, 4-trimethylsilylpent-4-en-2-ol (188), upon reaction with phenylacetaldehyde (189) in the presence of InCl3, gave cis-2,6-dihydropyran 190 via chair-like transition state 191. This strategy was further elaborated for the synthesis of 5,6- and 6,6-ring-fused dihydropyrans 193 and 195, respectively, as shown in Scheme 46.

[1860-5397-17-77-i46]

Scheme 46: Dobb and co-workers’ synthesis of DHP.

A similar tandem strategy of an addition of vinylsilane 196, followed by silyl-Prins cyclization with an aldehyde 197 in the presence of 5 mol % BiBr3, was reported by Hinkle and co-workers to give the corresponding compound 198 (Scheme 47) [18].

[1860-5397-17-77-i47]

Scheme 47: BiBr3-promoted tandem silyl-Prins reaction by Hinkle et al.

The authors further investigated the Mukaiyama aldol reaction between the β,γ-unsaturated aldehyde 199 and acetal 200 in the presence of 10 mol % BiBr3 to obtain aldol product 201. However, the addition of 2 equiv of phenylacetaldehyde (189) and 10 mol % BiBr3 afforded dihydropyran 202 in 64% yield as a single isomer, as shown in Scheme 48 [91].

[1860-5397-17-77-i48]

Scheme 48: Substrate scope of Hinkle and co-workers’ strategy.

The cis-2,6-disubstituted tetrahydropyran 207 with two adjacent methylene groups at the C3 and C4 positions was synthesized via silyl-Prins cyclization of silane 205 with an aldehyde in the presence of Lewis acid TMSOTf [92]. The reaction proceeded through transition state 208 following silyl-Prins cyclization, as shown in Scheme 49.

[1860-5397-17-77-i49]

Scheme 49: Cho and co-workers’ strategy for 2,6 disubstituted 3,4-dimethylene-THP.

Unlike allyl- and vinylsilanes, as discussed earlier, Furman and co-workers introduced a new concept of synthesizing 211 utilizing silyl-Prins cyclization of propargylsilane 209 and aldehyde 210 in the presence of TMSOTf [93]. The oxocarbenium ion was intramolecularly trapped by the olefin, followed by removal of trimethylsilane (Scheme 50).

[1860-5397-17-77-i50]

Scheme 50: Furman and co-workers’ THP synthesis from propargylsilane.

The authors further explored this strategy for the asymmetric synthesis of 3-vinylidene-substituted tetrahydropyran by taking the chiral propargylsilane. A diastereoselective route to cis-2,6-disubstituted tetrahydropyran-4-one 215 was explored by introducing a silyl enol ether Prins cyclization concept in which oxocarbenium ion 214, generated by reacting hydroxy-substituted silyl enol ether 212 with aldehyde 213 (different types of aliphatic and aromatic as well as α,β-unsaturated aldehydes were used), was trapped by silyl enol ether [94]. A detailed mechanism similar to simple Prins cyclization, except trapping of oxocarbenium ion 214 with silyl enol ether instead of olefin, vinylsilane, or allylsilanes, was proposed as shown in Scheme 51.

[1860-5397-17-77-i51]

Scheme 51: THP synthesis from silyl enol ethers.

However, the reaction of silyl enol ether such as 216, upon reacting with an unsaturated aldehyde 217, produced a mixture of cis- and trans-220 (dr = 4.1:1.0). It was explained that the diastereoselectivity of the product depends on the size of the substituent. For example, when the substituent is sterically small, it occupies the pseudoaxial position in the reactive conformation 218 (Scheme 52).

[1860-5397-17-77-i52]

Scheme 52: Rychnovsky and co-workers’ strategy for THP synthesis from hydroxy-substituted silyl enol ethers.

Li et al. utilized allylic geminal bissilyl alcohol 221 for the construction of THP ring A of (−)-exiguolide via Prins cyclization with an aldehyde in the presence of Lewis acid as a promoter [95]. High yield and excellent diastereoselectivity were obtained under standard silyl-Prins cyclization conditions using TMSOTf as Lewis acid (Scheme 53).

[1860-5397-17-77-i53]

Scheme 53: Li and co-workers’ germinal bissilyl Prins cyclization strategy to (−)-exiguolide.

Recently Xu et al. reported the homoallylic silyl alcohol 224 containing a multisubstituted (Z)-alkene reacting with an aldehyde in the presence of TMSI and InCl3 to afford 226 in high diastereoselectivity [96]. The authors assumed that the Prins cyclization proceeded through Alder’s chair-like transition state 227 in which the (Z)-alkene accounts for the trans-stereocontrol at the C3 position and equatorial iodide addition accounts for the cis-stereocontrol at the C4 position, as shown below in Scheme 54.

[1860-5397-17-77-i54]

Scheme 54: Xu and co-workers’ hydroiodination strategy for THP.

The one-pot synthesis of tetrahydropyran by utilizing the Babier–Prins cyclization reaction of allyl bromide (228) with a carbonyl compound promoted by BBIMBr/SnBr2 complex under solvent-free conditions has been explored [97]. The mechanism of the reaction was shown to include a Barbier reaction of allyl bromide with an aldehyde in the presence of SnBr3 and a quaternary ammonium salt to produce allyltin compound 230, which subsequently reacts with an aldehyde to generate intermediate 231. This intermediate could be hydrolyzed by water during workup to afford 232, which does not give the required THP product. Desired product 235 was obtained only in the anhydrous conditions (Scheme 55).

[1860-5397-17-77-i55]

Scheme 55: Wang and co-workers’ strategy for tetrahydropyran synthesis.

The methodology of alkynylsilane Prins cyclization was explored for the synthesis of 2,6-dihydropyran 238 by reacting secondary homopropargyl alcohol 236, having a trimethylsilyl group at the triple bond, with an aldehyde (Scheme 56) [98-101]. The reaction follows alkyne Prins cyclization and minimizes the competitive 2-oxonia-[3,3]-sigmatropic rearrangement pathway. The reaction was highly stereoselective and afforded the cis-2,6-dihydropyran in the presence of Lewis acid FeCl3.

[1860-5397-17-77-i56]

Scheme 56: FeCl3-catalyzed synthesis of DHP from alkynylsilane alcohol.

From DFT calculations, the authors concluded that the Prins product is formed more rapidly than the α-trimethylsilylalkenyl cation 242 formed by the Grob-type fragmentation (Scheme 57), which was trapped by the subsequent attack of the halide anion, leading to the formation of Prins product 244. On the basis of theoretical calculations, the authors could conclude factors controlling the alkyne Prins cyclization over formal 2-oxonia-[3,3]-sigmatropic rearrangement.

[1860-5397-17-77-i57]

Scheme 57: Martín, Padrón, and co-workers’ proposed mechanism of alkynylsilane Prins cyclization for the synthesis of DHP.

Furthermore, Markó and co-workers successfully synthesized 2,6-anti-configured THP starting from allylsilane 245, following diethylaluminium chloride-promoted ene reaction and condensation with an aldehyde 246 [102]. Expected ene adduct 247 was obtained as a (Z)-olefin. The addition of ZnCl2·Et2O and (MeO)3CH to the resulting homoallylic alcohol 247 leads to the desired pyran derivative 248, having an acetal group at the C2 position (Scheme 58). By treatment of acetal 248 with allyltrimethylsilane gave 2,6-anti-configured THP 249 as a single diastereomer in the presence of TMSOTf.

[1860-5397-17-77-i58]

Scheme 58: Marko and co-workers’ synthesis of 2,6-anti-configured tetrahydropyran.

A new route to obtain 2,6-anti-configured THP ring 252 was reported using homoallylic α-hydroxy ester 250 in an In(OTf)3-catalyzed Prins cyclization with moderate selectivity. Although selectivity was not observed (almost 1:1), a variety of 2,6-syn and anti-4-chloro-trisubstituted THPs were prepared depending upon nature of R1 in 250. Whereas, particularly with benzoyl ester substituent (250), only syn product 251 was obtained in 69% yield (Scheme 59) [103].

[1860-5397-17-77-i59]

Scheme 59: Loh and co-workers’ strategy for 2,6-syn-tetrahydropyrans.

The possible mechanism for the formation of a variety of isomers was explained through transition state 254 and 255 (Scheme 60). Competition between electronically favored transition state 254 leads to the formation of anti-isomer 256, whereas the sterically preferred transition state 255 afforded syn-isomer 257.

[1860-5397-17-77-i60]

Scheme 60: Loh and co-workers’ strategy for anti-THP synthesis.

Unlike the well-explored selective synthesis of major cis-2,6-THP, a highly stereoselective route to the thermodynamically disfavored trans-2,6-tetrahydropyran 260 was reported by Cha and co-workers based on the coupling of hydroxyethyl-tethered cyclopropanol 258 and aliphatic aldehyde 259 using TiCl4 as a Lewis acid [104,105]. The reaction proceeded through the Prins cyclization (Scheme 61).

[1860-5397-17-77-i61]

Scheme 61: Cha and co-workers’ strategy for trans-2,6-tetrahydropyran.

The reaction proceeded via formation of a 7-membered cyclic acetal 263 as a single isomer in nearly quantitative yield, followed by Lewis acid-catalyzed rearrangement leading to the formation of tetrahydropyran. Under optimized reaction conditions, TMSOTf gave 7-membered cyclic acetal 263, which upon treatment with TiCl4 gave the desired THP as a 14:1 mixture of trans- and cis-265 in 80% yield. The trans-265 was obtained as a major isomer, where the reaction proceeded through the 6-membered chair-like transition state 264, and the electrophilic ring opening of cyclopropane by the oxocarbenium ion was believed to proceed via “corner attack” at the less substituted C–C bond. However, minor cis-265 was formed via the 6-membered boat like transition state 264’ (Scheme 62) [104].

[1860-5397-17-77-i62]

Scheme 62: Mechanism proposed by Cha et al.

Cha’s group also utilized the Rechnovsky convergent method where an α-acetoxy ether was used as a precursor for the oxocarbenium ion in the THP synthesis to complement the aforementioned 7-membered cyclic acetal strategy. The treatment of α-acetoxy ether 266 with Lewis acid produced the corresponding THP 267 with moderate diastereoselectivity in favor of the trans-2,6-stereoisomer, as shown in Scheme 63.

[1860-5397-17-77-i63]

Scheme 63: TiCl4-mediated cyclization to trans-THP.

A variety of 4-hydroxy-substituted THPs was exclusively generated via Prins reaction using FeCl3 as a Lewis acid catalyst. Excellent stereoselectivity was obtained for a remarkably broad range of substrates under mild reaction conditions (Scheme 64) [106].

[1860-5397-17-77-i64]

Scheme 64: Feng and co-workers’ FeCl3-catalyzed Prins cyclization strategy to 4-hydroxy-substituted THP.

The authors proposed fundamental insights into the mechanism of the reaction based on DFT calculations. A different [2 + 2]-cycloaddition process was suggested to rationalize the observed OH-selectivity.

In 2015, Padrón and co-workers also reported the Prins cyclization catalyzed by a Fe(III) and trimethylsilyl halide system for the synthesis of all-cis-2,4,6-trisubstituted THP [107]. As reported previously by Feng et al. [106], two mechanistic pathways via the classical oxocarbenium route and [2 + 2]-cycloaddition were considered for DFT calculations. Experimental and DFT studies suggested the preference of a classical oxocarbenium route over the [2 + 2]-pathway for those alcohols having unactivated and unsubstituted alkenes, whereas the substituent adjacent to the hydroxy group in the homoallylic alcohol controls the oxonia-Cope rearrangement (see 273ac). The alkyl substituent favored the exclusive formation of crossed THP derivatives, whereas 2-oxonia-Cope rearrangement was thermodynamically favored in the presence of a phenyl group (Scheme 65).

[1860-5397-17-77-i65]

Scheme 65: Selectivity profile of the Prins cyclization under participation of an iron ligand.

Matsumoto and co-workers reported a Lewis acid-mediated Prins cyclization between an alcohol 278 bearing a nonconjugated diene moiety and an aldehyde 277 with alkyl or aryl substituent in presence of BF3·Et2O and TMSCl at −40 °C to afford corresponding fluorinated bicyclic compound 284 [108]. A catalytic amount of TMSCl generates TMS-protected alcohol 279 and HCl. The activated aldehyde 280 reacts with 279 to form the intermediate 281. Then, the TMS group in 281 is attacked by F in the presence of HCl to give the alkoxycarbenium ion intermediate 282, which is followed by a sequential cyclization to form secondary carbocation 283, which in the presence of fluoride ions affords 284, as shown in Scheme 66.

[1860-5397-17-77-i66]

Scheme 66: Sequential reactions involving Prins cyclization.

Banerjee et al. explored the reactivity of cyclopropane carbaldehydes 285 with 3-butyn-1-ol in the presence of TiX4 for the stereoselective construction of the THF ring (Scheme 67) [109]. A series of geminal bishalogen-containing fused THPs was synthesized in high yield (up to 80%) and excellent diastereoselectivity. A Prins cyclization mechanism was proposed for the above transformation in the presence of TiCl4. Formation of the oxocarbenium ion 289, followed by an intramolecular nucleophilic attack by the alkynyl bond on the cyclopropane unit gave cyclic oxocarbenium intermediate 290. Further, the attack of a halide anion (from TiX4) leads to the Prins cyclization to give bishalogenated bicyclic THP with all-cis-stereochemistry in the major product.

[1860-5397-17-77-i67]

Scheme 67: Banerjee and co-workers’ strategy of Prins cyclization from cyclopropane carbaldehydes and propargyl alcohol.

Asymmetric Prins cyclization

Mullen and Gagné reported a first catalytic asymmetric Prins cyclization reaction between 2-allylphenol 292 and glyoxylate ester 293 using (R)-[(tolBINAP)Pt(NC6F5)2][SbF6]2 (294) as chiral catalyst [110]. An optimization study revealed that the enantioselectivity varied with the polarity of the solvent. The optimization study disclosed that the enantioselectivity increases with the decrease of the polarity of the solvent (Scheme 68).

[1860-5397-17-77-i68]

Scheme 68: Mullen and Gagné's (R)-[(tolBINAP)Pt(NC6F5)2][SbF6]2-catalyzed asymmetric Prins cyclization strategy to chromans.

Yu and co-workers reported a novel segment-coupling Prins cyclization involving sequential benzylic/allylic C–H bond activation via DDQ oxidation, followed by nucleophilic attack of an unactivated olefin to obtain all-cis-trisubstituted Prins products with high stereochemical precision [111]. A single-electron transfer (SET) mechanism was proposed for the above transformation (Scheme 69). A SET from an arene or alkene to DDQ and the subsequent abstraction of hydride from the benzylic or allylic position generated a charge-transfer complex 298. The complex 298 formed a tin-containing ate oxocarbenium ion complex 299 with SnBr4, and then rapid C–C bond formation took place to generate the cyclic intermediate 300. The subsequent trapping of the carbocation with the bromide ion led to all-cis-2,4,6-trisubstituted tetrahydropyran 297 (Scheme 69).

[1860-5397-17-77-i69]

Scheme 69: Yu and co-workers’ DDQ-catalyzed asymmetric Prins cyclization strategy to trisubstituted THPs.

Lalli and van de Weghe reported a chiral BINOL-derived bisphosphoric acid- and CuCl-catalyzed enantioselective tandem Prins–Friedel–Crafts cyclization between homoallylic alcohol 302 and substituted aromatic aldehydes 303 to form hexahydro-1H-benzo[f]isochromenes 305 with three new contiguous stereocenters in high enantio- and diastereoselectivity [112]. The three new contiguous stereogenic centers formed resulted from an attack of the alkene to the Si-face of the oxocarbenium ion, which was followed by a completely diastereoselective Friedel–Crafts reaction (Scheme 70).

[1860-5397-17-77-i70]

Scheme 70: Lalli and Weghe’s chiral-Brønsted-acid- and achiral-Lewis-acid-promoted asymmetric Prins cyclization strategy.

List and co-workers devised a strategy employing highly acidic confined iminoimidodiphosphate (iIDP) Brønsted acids 308 that catalyzed asymmetric Prins cyclizations of both aliphatic and aromatic aldehydes with alcohol 307 to obtain 309 (Scheme 71) [113]. The introduction of electron-withdrawing nitro groups on the BINOL backbone in the catalysts significantly enhanced the reactivity and enantioselectivity.

[1860-5397-17-77-i71]

Scheme 71: List and co-workers’ iIDP Brønsted acid-promoted asymmetric Prins cyclization strategy.

Zhou et al. reported an asymmetric Prins cyclization of in situ-generated quinone methides from phenol-tethered alkenyl alcohol 310 and o-aminobenzaldehyde 311 using chiral phosphoric acids (Scheme 72) [114]. Diverse functionalized trans-fused pyranotetrahydroquinoline derivatives 312 were synthesized in excellent yield and selectivity (up to 99% yield and 99% ee).

[1860-5397-17-77-i72]

Scheme 72: Zhou and co-workers’ strategy for chiral phosphoric acid (CPA)-catalyzed cascade Prins cyclization.

List et al. reported a chiral imidodiphosphoric acid-catalyzed asymmetric Prins cyclization with salicylaldehyde 316 and 3-methylbut-3-en-1-ol (317) to afford 4-methylenetetrahydropyrans 318 with high enantioselectivity (Scheme 73) [115]. A chiral bis-BINOL-based imidophosphoric acid 319 was efficient in this reaction, and the extreme bulkiness of this catalyst was the key to a successful transformation. This reaction proceeded via a Prins cyclization mechanism, activated by chiral acid 319.

[1860-5397-17-77-i73]

Scheme 73: List and co-workers’ approach for asymmetric Prins cyclization using chiral imidodiphosphoric acid 319.

Conclusion

Prins cyclization strategies have been proven as a reliable and robust method for the stereoselective construction of THP rings. Many of these strategies have been utilized for the elegant synthesis of natural products. In this review, we portrayed an inspection of twenty years in the arena of the development of Prins cyclizations and the further exploration of these strategies in the total synthesis of natural products. This up-to-date information showcases the knowledge gained in this area. In either case, it is hoped that the challenge of stereoselective construction of THP rings in the context of natural product synthesis will continue to inspire synthetic chemists to develop new methods in the coming years.

Acknowledgements

We sincerely acknowledge Dr. (Prof.) Ganesh Pandey for his critical suggestions in the compiling of this review.

Funding

A. B. thanks SERB-DST (YSS/2015/001085), India for financial assistance.

References

  1. Olier, C.; Kaafarani, M.; Gastaldi, S.; Bertrand, M. P. Tetrahedron 2010, 66, 413–445. doi:10.1016/j.tet.2009.10.069
    Return to citation in text: [1]
  2. Overman, L. E.; Pennington, L. D. J. Org. Chem. 2003, 68, 7143–7157. doi:10.1021/jo034982c
    Return to citation in text: [1]
  3. Boger, D. L.; Weinreb, S. M. Hetero Diels-Alder Methodology in Organic Synthesis; Academic Press: San Diego, CA, USA, 1987.
    Return to citation in text: [1]
  4. Nicolaou, K. C.; Prasad, C. V. C.; Somers, P. K.; Hwang, C. K. J. Am. Chem. Soc. 1989, 111, 5330–5334. doi:10.1021/ja00196a043
    Return to citation in text: [1]
  5. Kas’yan, L. I.; Pal’chikov, V. A.; Bondarenko, Ya. S. Russ. J. Org. Chem. 2011, 47, 1609–1652. doi:10.1134/s1070428011110017
    Return to citation in text: [1]
  6. Smith, A. B.; Fox, R. J.; Razler, T. M. Acc. Chem. Res. 2008, 41, 675–687. doi:10.1021/ar700234r
    Return to citation in text: [1]
  7. Padmaja, P.; Reddy, P. N.; Subba Reddy, B. V. Org. Biomol. Chem. 2020, 18, 7514–7532. doi:10.1039/d0ob00960a
    Return to citation in text: [1]
  8. Rapelli, C.; Sridhar, B.; Subba Reddy, B. V. Org. Biomol. Chem. 2020, 18, 7224. doi:10.1039/d0ob90122a
    Return to citation in text: [1]
  9. Pastor, I. M.; Yus, M. Curr. Org. Chem. 2007, 11, 925–957. doi:10.2174/138527207781024067
    Return to citation in text: [1]
  10. Pastor, I. M.; Yus, M. Curr. Org. Chem. 2012, 16, 1277–1312. doi:10.2174/138527212800564196
    Return to citation in text: [1]
  11. Heravi, M. M.; Ahmadi, T.; Ghavidel, M.; Heidari, B.; Hamidi, H. RSC Adv. 2015, 5, 101999–102075. doi:10.1039/c5ra17488k
    Return to citation in text: [1]
  12. Vilotijevic, I.; Jamison, T. F. Angew. Chem., Int. Ed. 2009, 48, 5250–5281. doi:10.1002/anie.200900600
    Return to citation in text: [1]
  13. Minbiole, E. C.; Minbiole, K. P. C. J. Antibiot. 2016, 69, 213–219. doi:10.1038/ja.2015.136
    Return to citation in text: [1]
  14. Nising, C. F.; Bräse, S. Chem. Soc. Rev. 2008, 37, 1218–1228. doi:10.1039/b718357g
    Return to citation in text: [1]
  15. Abrams, D. J.; Provencher, P. A.; Sorensen, E. J. Chem. Soc. Rev. 2018, 47, 8925–8967. doi:10.1039/c8cs00716k
    Return to citation in text: [1]
  16. Gradillas, A.; Pérez-Castells, J. Angew. Chem., Int. Ed. 2006, 45, 6086–6101. doi:10.1002/anie.200600641
    Return to citation in text: [1]
  17. Ko, C.; Hsung, R. P.; Al-Rashid, Z. F.; Feltenberger, J. B.; Lu, T.; Yang, J.-H.; Wei, Y.; Zificsak, C. A. Org. Lett. 2007, 9, 4459–4462. doi:10.1021/ol701768n
    Return to citation in text: [1]
  18. Evans, P. A.; Cui, J.; Gharpure, S. J.; Hinkle, R. J. J. Am. Chem. Soc. 2003, 125, 11456–11457. doi:10.1021/ja036439j
    Return to citation in text: [1] [2]
  19. Evans, P. A.; Cui, J.; Gharpure, S. J. Org. Lett. 2003, 5, 3883–3885. doi:10.1021/ol035438t
    Return to citation in text: [1]
  20. Miranda, P. O.; Díaz, D. D.; Padrón, J. I.; Bermejo, J.; Martín, V. S. Org. Lett. 2003, 5, 1979–1982. doi:10.1021/ol034568z
    Return to citation in text: [1]
  21. Marcos, R.; Rodríguez-Escrich, C.; Herrerías, C. I.; Pericàs, M. A. J. Am. Chem. Soc. 2008, 130, 16838–16839. doi:10.1021/ja8062887
    Return to citation in text: [1]
  22. Kriewitz, O. Ber. Dtsch. Chem. Ges. 1899, 32, 57–60. doi:10.1002/cber.18990320111
    Return to citation in text: [1]
  23. Bloys Van Treslong Prins, P. C. Chem. Weekbl. 1919, 1510.
    Return to citation in text: [1]
  24. Bloys Van Treslong Prins, P. C. Chem. Weekbl. 1919, 1072.
    Return to citation in text: [1]
  25. Hanschke, E. Chem. Ber. 1955, 88, 1053–1061. doi:10.1002/cber.19550880718
    Return to citation in text: [1]
  26. Perry, M. A.; Rychnovsky, S. D.; Sizemore, N. Synthesis of Saturated Tetrahydropyrans. In Synthesis of Saturated Oxygenated Heterocycles I; Cossy, J., Ed.; Topics in Heterocyclic Chemistry, Vol. 35; Springer: Berlin, Heidelberg, Germany, 2014; pp 978 ff. doi:10.1007/978-3-642-41473-2_2
    Return to citation in text: [1]
  27. Nasir, N. M.; Ermanis, K.; Clarke, P. A. Org. Biomol. Chem. 2014, 12, 3323–3335. doi:10.1039/c4ob00423j
    Return to citation in text: [1]
  28. Alder, R. W.; Harvey, J. N.; Oakley, M. T. J. Am. Chem. Soc. 2002, 124, 4960–4961. doi:10.1021/ja025902+
    Return to citation in text: [1]
  29. Crosby, S. R.; Harding, J. R.; King, C. D.; Parker, G. D.; Willis, C. L. Org. Lett. 2002, 4, 577–580. doi:10.1021/ol0102850
    Return to citation in text: [1] [2]
  30. Barry, C. S. J.; Crosby, S. R.; Harding, J. R.; Hughes, R. A.; King, C. D.; Parker, G. D.; Willis, C. L. Org. Lett. 2003, 5, 2429–2432. doi:10.1021/ol0346180
    Return to citation in text: [1]
  31. Al-Mutairi, E. H.; Crosby, S. R.; Darzi, J.; Hughes, R. A.; Simpson, T. J.; Smith, R. W.; Willis, C. L.; Harding, J. R.; King, C. D. Chem. Commun. 2001, 835–836. doi:10.1039/b101414p
    Return to citation in text: [1]
  32. Rychnovsky, S. D.; Marumoto, S.; Jaber, J. J. Org. Lett. 2001, 3, 3815–3818. doi:10.1021/ol0168465
    Return to citation in text: [1]
  33. Marumoto, S.; Jaber, J. J.; Vitale, J. P.; Rychnovsky, S. D. Org. Lett. 2002, 4, 3919–3922. doi:10.1021/ol026751i
    Return to citation in text: [1] [2] [3]
  34. Jaber, J. J.; Mitsui, K.; Rychnovsky, S. D. J. Org. Chem. 2001, 66, 4679–4686. doi:10.1021/jo010232w
    Return to citation in text: [1]
  35. Jasti, R.; Anderson, C. D.; Rychnovsky, S. D. J. Am. Chem. Soc. 2005, 127, 9939–9945. doi:10.1021/ja0518326
    Return to citation in text: [1]
  36. Rychnovsky, S. D.; Thomas, C. R. Org. Lett. 2000, 2, 1217–1219. doi:10.1021/ol005646a
    Return to citation in text: [1]
  37. Jasti, R.; Vitale, J.; Rychnovsky, S. D. J. Am. Chem. Soc. 2004, 126, 9904–9905. doi:10.1021/ja046972e
    Return to citation in text: [1]
  38. Jasti, R.; Rychnovsky, S. D. Org. Lett. 2006, 8, 2175–2178. doi:10.1021/ol0606738
    Return to citation in text: [1]
  39. Gesinski, M. R.; Van Orden, L. J.; Rychnovsky, S. D. Synlett 2008, 363–366. doi:10.1055/s-2008-1032053
    Return to citation in text: [1]
  40. Kopecky, D. J.; Rychnovsky, S. D. J. Am. Chem. Soc. 2001, 123, 8420–8421. doi:10.1021/ja011377n
    Return to citation in text: [1]
  41. Van Orden, L. J.; Patterson, B. D.; Rychnovsky, S. D. J. Org. Chem. 2007, 72, 5784–5793. doi:10.1021/jo070901r
    Return to citation in text: [1]
  42. Patterson, B.; Marumoto, S.; Rychnovsky, S. D. Org. Lett. 2003, 5, 3163–3166. doi:10.1021/ol035303n
    Return to citation in text: [1]
  43. Cheung, L. L.; Marumoto, S.; Anderson, C. D.; Rychnovsky, S. D. Org. Lett. 2008, 10, 3101–3104. doi:10.1021/ol8011474
    Return to citation in text: [1]
  44. Hart, D. J.; Bennett, C. E. Org. Lett. 2003, 5, 1499–1502. doi:10.1021/ol0342756
    Return to citation in text: [1]
  45. Ko, H. M.; Lee, D. G.; Kim, M. A.; Kim, H. J.; Park, J.; Lah, M. S.; Lee, E. Org. Lett. 2007, 9, 141–144. doi:10.1021/ol0627956
    Return to citation in text: [1]
  46. Barry, C. S.; Bushby, N.; Harding, J. R.; Willis, C. L. Org. Lett. 2005, 7, 2683–2686. doi:10.1021/ol050840o
    Return to citation in text: [1]
  47. Barry, C. S.; Elsworth, J. D.; Seden, P. T.; Bushby, N.; Harding, J. R.; Alder, R. W.; Willis, C. L. Org. Lett. 2006, 8, 3319–3322. doi:10.1021/ol0611705
    Return to citation in text: [1]
  48. Seden, P. T.; Charmant, J. P. H.; Willis, C. L. Org. Lett. 2008, 10, 1637–1640. doi:10.1021/ol800386d
    Return to citation in text: [1]
  49. Cons, B. D.; Bunt, A. J.; Bailey, C. D.; Willis, C. L. Org. Lett. 2013, 15, 2046–2049. doi:10.1021/ol400736w
    Return to citation in text: [1]
  50. Yadav, J. S.; Rahman, M. A.; Reddy, N. M.; Prasad, A. R.; Al Khazim Al Ghamdi, A. Synlett 2014, 25, 661–664. doi:10.1055/s-0033-1340181
    Return to citation in text: [1]
  51. Janardhan Reddy, P.; Srinivas Reddy, A.; Yadav, J. S.; Subba Reddy, B. V. Tetrahedron Lett. 2012, 53, 4051–4053. doi:10.1016/j.tetlet.2012.05.059
    Return to citation in text: [1]
  52. Morris, W. J.; Custar, D. W.; Scheidt, K. A. Org. Lett. 2005, 7, 1113–1116. doi:10.1021/ol050093v
    Return to citation in text: [1]
  53. Puglisi, A.; Lee, A.-L.; Schrock, R. R.; Hoveyda, A. H. Org. Lett. 2006, 8, 1871–1874. doi:10.1021/ol060430f
    Return to citation in text: [1]
  54. Cossey, K. N.; Funk, R. L. J. Am. Chem. Soc. 2004, 126, 12216–12217. doi:10.1021/ja046940r
    Return to citation in text: [1]
  55. Yadav, V. K.; Vijaya Kumar, N. J. Am. Chem. Soc. 2004, 126, 8652–8653. doi:10.1021/ja048000c
    Return to citation in text: [1]
  56. Yadav, V. K.; Verma, A. K.; Kumar, P.; Hulikal, V. Chem. Commun. 2014, 50, 15457–15460. doi:10.1039/c4cc07796b
    Return to citation in text: [1]
  57. Yadav, J. S.; Reddy, B. V. S.; Venugopal, C.; Srinivas, R.; Ramalingam, T. Synth. Commun. 2002, 32, 1803–1808. doi:10.1081/scc-120004059
    Return to citation in text: [1]
  58. Yadav, J. S.; Reddy, B. V. S.; Kumar, G. G. K. S. N.; Reddy, G. M. Tetrahedron Lett. 2007, 48, 4903–4906. doi:10.1016/j.tetlet.2007.05.056
    Return to citation in text: [1]
  59. Yadav, J. S.; Reddy, M. S.; Prasad, A. R. Tetrahedron Lett. 2005, 46, 2133–2136. doi:10.1016/j.tetlet.2005.01.121
    Return to citation in text: [1]
  60. Sabitha, G.; Reddy, K. B.; Reddy, G. S. K. K.; Fatima, N.; Yadav, J. S. Synlett 2005, 2347–2351. doi:10.1055/s-2005-872668
    Return to citation in text: [1]
  61. Loh, T.-P.; Hu, Q.-Y.; Tan, K.-T.; Cheng, H.-S. Org. Lett. 2001, 3, 2669–2672. doi:10.1021/ol016228o
    Return to citation in text: [1]
  62. Loh, T.-P.; Yang, J.-Y.; Feng, L.-C.; Zhou, Y. Tetrahedron Lett. 2002, 43, 7193–7196. doi:10.1016/s0040-4039(02)01666-0
    Return to citation in text: [1]
  63. Chan, K.-P.; Loh, T.-P. Tetrahedron Lett. 2004, 45, 8387–8390. doi:10.1016/j.tetlet.2004.09.049
    Return to citation in text: [1]
  64. Chan, K.-P.; Loh, T.-P. Org. Lett. 2005, 7, 4491–4494. doi:10.1021/ol051951q
    Return to citation in text: [1]
  65. Liu, F.; Loh, T.-P. Org. Lett. 2007, 9, 2063–2066. doi:10.1021/ol070506n
    Return to citation in text: [1]
  66. Biermann, U.; Lützen, A.; Metzger, J. O. Eur. J. Org. Chem. 2006, 2631–2637. doi:10.1002/ejoc.200500701
    Return to citation in text: [1]
  67. Álvarez-Méndez, S. J.; Fariña-Ramos, M.; Villalba, M. L.; Perretti, M. D.; García, C.; Moujir, L. M.; Ramírez, M. A.; Martín, V. S. J. Org. Chem. 2018, 83, 9039–9066. doi:10.1021/acs.joc.8b01182
    Return to citation in text: [1]
  68. Leroy, B.; Markó, I. E. J. Org. Chem. 2002, 67, 8744–8752. doi:10.1021/jo025899c
    Return to citation in text: [1] [2]
  69. Markó, I. E.; Leroy, B. Tetrahedron Lett. 2000, 41, 7225–7230. doi:10.1016/s0040-4039(00)01179-5
    Return to citation in text: [1] [2]
  70. Gesinski, M. R.; Rychnovsky, S. D. J. Am. Chem. Soc. 2011, 133, 9727–9729. doi:10.1021/ja204228q
    Return to citation in text: [1]
  71. Hoye, T. R.; Hu, M. J. Am. Chem. Soc. 2003, 125, 9576–9577. doi:10.1021/ja035579q
    Return to citation in text: [1]
  72. Pham, M.; Allatabakhsh, A.; Minehan, T. G. J. Org. Chem. 2008, 73, 741–744. doi:10.1021/jo7016857
    Return to citation in text: [1]
  73. Yu, C.-M.; Lee, J.-Y.; So, B.; Hong, J. Angew. Chem., Int. Ed. 2002, 41, 161–163. doi:10.1002/1521-3773(20020104)41:1<161::aid-anie161>3.0.co;2-n
    Return to citation in text: [1]
  74. Aubele, D. L.; Lee, C. A.; Floreancig, P. E. Org. Lett. 2003, 5, 4521–4523. doi:10.1021/ol0359259
    Return to citation in text: [1]
  75. Aubele, D. L.; Wan, S.; Floreancig, P. E. Angew. Chem., Int. Ed. 2005, 44, 3485–3488. doi:10.1002/anie.200500564
    Return to citation in text: [1]
  76. Huang, H.; Panek, J. S. J. Am. Chem. Soc. 2000, 122, 9836–9837. doi:10.1021/ja002087u
    Return to citation in text: [1]
  77. Su, Q.; Panek, J. S. J. Am. Chem. Soc. 2004, 126, 2425–2430. doi:10.1021/ja037957x
    Return to citation in text: [1]
  78. Lowe, J. T.; Panek, J. S. Org. Lett. 2005, 7, 3231–3234. doi:10.1021/ol050982i
    Return to citation in text: [1]
  79. Zhang, Y.; Panek, J. S. Org. Lett. 2007, 9, 3141–3143. doi:10.1021/ol701427k
    Return to citation in text: [1]
  80. Lowe, J. T.; Panek, J. S. Org. Lett. 2008, 10, 3813–3816. doi:10.1021/ol801499s
    Return to citation in text: [1]
  81. Zhu, K.; Panek, J. S. Org. Lett. 2011, 13, 4652–4655. doi:10.1021/ol201863b
    Return to citation in text: [1]
  82. Roush, W. R.; Dilley, G. J. Synlett 2001, 955–959. doi:10.1055/s-2001-14632
    Return to citation in text: [1] [2] [3]
  83. Micalizio, G. C.; Pinchuk, A. N.; Roush, W. R. J. Org. Chem. 2000, 65, 8730–8736. doi:10.1021/jo001236o
    Return to citation in text: [1]
  84. Sanchez, C. C.; Keck, G. E. Org. Lett. 2005, 7, 3053–3056. doi:10.1021/ol051040g
    Return to citation in text: [1]
  85. Keck, G. E.; Welch, D. S.; Vivian, P. K. Org. Lett. 2006, 8, 3667–3670. doi:10.1021/ol061173h
    Return to citation in text: [1]
  86. Keck, G. E.; Kraft, M. B.; Truong, A. P.; Li, W.; Sanchez, C. C.; Kedei, N.; Lewin, N. E.; Blumberg, P. M. J. Am. Chem. Soc. 2008, 130, 6660–6661. doi:10.1021/ja8022169
    Return to citation in text: [1]
  87. Keck, G. E.; Covel, J. A.; Schiff, T.; Yu, T. Org. Lett. 2002, 4, 1189–1192. doi:10.1021/ol025645d
    Return to citation in text: [1]
  88. Keck, G. E.; Welch, D. S.; Poudel, Y. B. Tetrahedron Lett. 2006, 47, 8267–8270. doi:10.1016/j.tetlet.2006.09.094
    Return to citation in text: [1]
  89. Dobbs, A. P.; Guesné, S. J. J.; Martinović, S.; Coles, S. J.; Hursthouse, M. B. J. Org. Chem. 2003, 68, 7880–7883. doi:10.1021/jo034981k
    Return to citation in text: [1]
  90. Dobbs, A. P.; Martinović, S. Tetrahedron Lett. 2002, 43, 7055–7057. doi:10.1016/s0040-4039(02)01558-7
    Return to citation in text: [1]
  91. Lian, Y.; Hinkle, R. J. J. Org. Chem. 2006, 71, 7071–7074. doi:10.1021/jo060738k
    Return to citation in text: [1]
  92. Cho, Y. S.; Karupaiyan, K.; Kang, H. J.; Pae, A. N.; Cha, J. H.; Koh, H. Y.; Chang, M. H. Chem. Commun. 2003, 2346–2347. doi:10.1039/b307320c
    Return to citation in text: [1]
  93. Dziedzic, M.; Lipner, G.; Furman, B. Tetrahedron Lett. 2005, 46, 6861–6863. doi:10.1016/j.tetlet.2005.08.014
    Return to citation in text: [1]
  94. Tay, G. C.; Huang, C. Y.; Rychnovsky, S. D. J. Org. Chem. 2014, 79, 8733–8749. doi:10.1021/jo501580p
    Return to citation in text: [1]
  95. Lu, J.; Song, Z.; Zhang, Y.; Gan, Z.; Li, H. Angew. Chem., Int. Ed. 2012, 51, 5367–5370. doi:10.1002/anie.201201323
    Return to citation in text: [1]
  96. Xu, Y.; Yin, Z.; Lin, X.; Gan, Z.; He, Y.; Gao, L.; Song, Z. Org. Lett. 2015, 17, 1846–1849. doi:10.1021/acs.orglett.5b00485
    Return to citation in text: [1]
  97. Zhao, X.-L.; Liu, L.; Chen, Y.-J.; Wang, D. Tetrahedron 2006, 62, 7113–7120. doi:10.1016/j.tet.2006.04.075
    Return to citation in text: [1]
  98. Miranda, P. O.; Ramírez, M. A.; Padrón, J. I.; Martín, V. S. Tetrahedron Lett. 2006, 47, 283–286. doi:10.1016/j.tetlet.2005.11.032
    Return to citation in text: [1]
  99. Miranda, P. O.; Ramírez, M. A.; Martín, V. S.; Padrón, J. I. Org. Lett. 2006, 8, 1633–1636. doi:10.1021/ol060247m
    Return to citation in text: [1]
  100. Scoccia, J.; Pérez, S. J.; Sinka, V.; Cruz, D. A.; López-Soria, J. M.; Fernández, I.; Martín, V. S.; Miranda, P. O.; Padrón, J. I. Org. Lett. 2017, 19, 4834–4837. doi:10.1021/acs.orglett.7b02270
    Return to citation in text: [1]
  101. Miranda, P. O.; Ramírez, M. A.; Martín, V. S.; Padrón, J. I. Chem. – Eur. J. 2008, 14, 6260–6268. doi:10.1002/chem.200800281
    Return to citation in text: [1]
  102. Dubost, C.; Markó, I. E.; Bryans, J. Tetrahedron Lett. 2005, 46, 4005–4009. doi:10.1016/j.tetlet.2005.04.041
    Return to citation in text: [1]
  103. Chan, K.-P.; Seow, A.-H.; Loh, T.-P. Tetrahedron Lett. 2007, 48, 37–41. doi:10.1016/j.tetlet.2006.11.023
    Return to citation in text: [1]
  104. Parida, B. B.; Lysenko, I. L.; Cha, J. K. Org. Lett. 2012, 14, 6258–6261. doi:10.1021/ol3030204
    Return to citation in text: [1] [2]
  105. Lee, H. G.; Lysenko, I. L.; Cha, J. K. Angew. Chem., Int. Ed. 2007, 46, 3326–3328. doi:10.1002/anie.200700172
    Return to citation in text: [1]
  106. Zheng, K.; Liu, X.; Qin, S.; Xie, M.; Lin, L.; Hu, C.; Feng, X. J. Am. Chem. Soc. 2012, 134, 17564–17573. doi:10.1021/ja3062002
    Return to citation in text: [1] [2]
  107. Pérez, S. J.; Purino, M.; Miranda, P. O.; Martín, V. S.; Fernández, I.; Padrón, J. I. Chem. – Eur. J. 2015, 21, 15211–15217. doi:10.1002/chem.201502488
    Return to citation in text: [1]
  108. Matsumoto, K.; Yanagi, R.; Yamaguchi, K.; Hayashi, E.; Yasuda, E.; Nokami, T.; Nishiwaki, K.; Kashimura, S.; Kuriyama, K. Heterocycles 2018, 96, 1363–1372. doi:10.3987/com-18-13940
    Return to citation in text: [1]
  109. Kumar, P.; Dey, R.; Banerjee, P. Org. Lett. 2018, 20, 5163–5166. doi:10.1021/acs.orglett.8b02094
    Return to citation in text: [1]
  110. Mullen, C. A.; Gagné, M. R. Org. Lett. 2006, 8, 665–668. doi:10.1021/ol052845r
    Return to citation in text: [1]
  111. Yu, B.; Jiang, T.; Li, J.; Su, Y.; Pan, X.; She, X. Org. Lett. 2009, 11, 3442–3445. doi:10.1021/ol901291w
    Return to citation in text: [1]
  112. Lalli, C.; van de Weghe, P. Chem. Commun. 2014, 50, 7495–7498. doi:10.1039/c4cc02826k
    Return to citation in text: [1]
  113. Liu, L.; Kaib, P. S. J.; Tap, A.; List, B. J. Am. Chem. Soc. 2016, 138, 10822–10825. doi:10.1021/jacs.6b07240
    Return to citation in text: [1]
  114. Sun, H.-R.; Zhao, Q.; Yang, H.; Yang, S.; Gou, B.-B.; Chen, J.; Zhou, L. Org. Lett. 2019, 21, 7143–7148. doi:10.1021/acs.orglett.9b02714
    Return to citation in text: [1]
  115. Tsui, G. C.; Liu, L.; List, B. Angew. Chem., Int. Ed. 2015, 54, 7703–7706. doi:10.1002/anie.201500219
    Return to citation in text: [1]
Other Beilstein-Institut Open Science Activities