From Allylic Sulfoxides to Allylic Sulfenates - American Chemical Society

Jul 15, 2017 - 8.5. Synthesis of Natural Products by Sulfoxide−. Sulfenate Rearrangement. AD. 9. Conclusions. AF. Author Information. AG. Correspond...
21 downloads 11 Views 11MB Size
Review pubs.acs.org/CR

Cite This: Chem. Rev. 2017, 117, 14201−14243

From Allylic Sulfoxides to Allylic Sulfenates: Fifty Years of a NeverEnding [2,3]-Sigmatropic Rearrangement Ignacio Colomer, Marina Velado, Roberto Fernández de la Pradilla,* and Alma Viso* Instituto de Química Orgánica General, Consejo Superior de Investigaciones Científicas (IQOG-CSIC), Juan de la Cierva 3, 28006-Madrid, Spain ABSTRACT: The [2,3]-sigmatropic rearrangement of allylic sulfoxides to allylic sulfenates is a reversible process, generally shifted toward the sulfoxide. In the presence of thiophiles, the sulfenate is trapped, and allylic alcohols are obtained under mild conditions. In most cases, a good transfer of stereochemical information through an ordered transition state is obtained. Furthermore, the ease of coupling this process with other versatile, stereocontrolled reactions has enhanced the usefulness of this protocol. This review aims to provide a comprehensive survey of this rearrangement and its application in the synthesis of natural and bioactive products.

CONTENTS 1. Introduction 2. [2,3]-Sigmatropic Rearrangement of Allylic Sulfoxides: Early Results 3. Stereochemical and Mechanistic Aspects of the Mislow−Braverman−Evans Rearrangement 3.1. Stereochemical Outcome 3.1.1. Alkene Geometry 3.1.2. Transfer of Chirality 3.2. Mechanistic and Computational Studies 4. Synthesis of Allylic Sulfoxides 5. General Scope of [2,3]-Sigmatropic Sulfoxide− Sulfenate Rearrangements 5.1. From Allyl Sulfenates 5.2. From Allyl Sulfoxides 6. Sulfoxide−Sulfenate Rearrangements in Tandem Processes 6.1. [2,3]-Sigmatropic Rearrangement and [1,2]Elimination 6.2. Multiple [2,3]-Sigmatropic Rearrangements 6.3. Sulfoxide, Piperidine, and Carbonyl (SPAC) Reaction 6.3.1. SPAC-Like Reactions 6.3.2. Early SPAC Reaction and Selected Applications in Synthesis 6.3.3. Synthesis of E-γ-Hydroxy-α,β-unsaturated Sulfones by the SPAC Methodology 6.3.4. Enantiopure Sulfoxides in the SPAC Reaction 6.3.5. Applications of the SPAC Reaction in Total Synthesis 6.4. [4 + 2]-Cycloadditions and [2,3]-Sigmatropic Rearrangements 6.4.1. [4 + 2]-Cycloadditions with Sulfur-Containing Dienes or Dienophiles and [2,3]Sigmatropic Rearrangement © 2017 American Chemical Society

6.4.2. Other [4 + 2]-Cycloaddition/[2,3]-Sigmatropic Sequences 6.5. Tandem [3,3]-Sigmatropic and [2,3]-Sigmatropic Rearrangements 6.6. Michael-Type Addition and [2,3]-Sigmatropic Rearrangement 6.7. [2,3]-Sigmatropic Rearrangement and Radical Ring Expansion 7. [2,3]-Sigmatropic Sulfoxide−Sulfenate Rearrangements Involving Propargyl Moieties 7.1. [2,3]-Sigmatropic Rearrangement from Propargylic Sulfoxides 7.2. [2,3]-Sigmatropic Rearrangement from Propargylic Sulfenates 8. Mislow−Braverman−Evans Rearrangements in the Synthesis of Bioactive and Natural Products 8.1. Synthesis of Prostanoids 8.2. Synthesis of Carbohydrates 8.3. Synthesis of Steroids and Related Products 8.4. Synthesis of Natural Products by Sulfenate− Sulfoxide Rearrangement 8.5. Synthesis of Natural Products by Sulfoxide− Sulfenate Rearrangement 9. Conclusions Author Information Corresponding Authors ORCID Notes Biographies Acknowledgments References

14201 14202 14203 14203 14203 14203 14204 14205 14206 14206 14207 14211 14211 14211 14212 14212 14213

14213 14213

14219 14220 14221 14223 14224 14224 14224 14227 14228 14229 14229 14230 14230 14232 14233 14233 14233 14233 14233 14233 14233

14216 14217 Received: July 15, 2017 Published: November 29, 2017

14217 14201

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

1. INTRODUCTION Within the rich chemistry of the sulfinyl moiety in synthesis and catalysis, allylic sulfoxides have attracted considerable attention, especially because of their facile [2,3]-sigmatropic rearrangement. This transformation allows for the stereoselective preparation of substituted allylic alcohols with concomitant removal of the sulfur moiety by reaction with appropriate thiophiles. A particularly useful feature of this reversible rearrangement, also known as the Mislow−Braverman−Evans rearrangement, is that it can often participate in tandem processes, which broadens the usefulness of the method. The reversibility of the rearrangement is also advantageous for synthetic efforts, as either sulfoxides or allylic alcohols are viable starting materials. In most cases, the rearrangement is highly stereoselective, and the stereochemical aspects have been reviewed.1,2 The general issues and applications of this transformation have been discussed in several review articles.3−13 The purpose of this article is to compile a comprehensive survey of the rearrangement, briefly discussing the key features of the process with the aid of selected graphical schemes but focusing on newer methodological developments and on applications of the rearrangement in the field of the synthesis of bioactive and natural products. The closely related propargyl sulfenate-to-allenyl sulfoxide rearrangement is also included in the review, but the [2,3]-sigmatropic rearrangement of selenoxides is not included. Preliminary communications on the topic of sulfoxide−sulfenate rearrangements that are adequately covered and cited in subsequent full reports have not been included in this review.

Subsequent studies by Mislow and co-workers revealed that the rate of racemization of allyl p-tolyl sulfoxide is dependent on the polarity of the solvent, with polar solvents stabilizing the polar sulfoxide and thus increasing the energy of the transition state going from sulfoxide to apolar sulfenate.17 This and the regioselectivity found for deuterated lithium alkoxides 6 (Scheme 2) support the idea that the racemization takes place Scheme 2. Regiochemistry and Substituent Effects in Sulfoxide−Sulfenate Rearrangements

by a cyclic concerted reversible intramolecular rearrangement of the sulfoxide to the achiral sulfenate with the equilibrium displaced toward the sulfoxide 7. Electron-withdrawing groups and nonpolar solvents accelerate the racemization of sulfoxides 8 and increase the proportion of sulfenate 9 at equilibrium to produce equimolecular mixtures of sulfoxides and sulfenates in some cases.18 The possibility of transferring chirality from carbon to sulfur was investigated by Mislow and co-workers in an exploration of the rearrangement of (S)-sulfenate 10, formed in situ from (S)-3buten-2-ol, that led to (S)-trans-crotyl p-tolylsulfoxide (E)-(S)11 [37% enantiomeric excess (ee)], which, over a period of days, led to a 77:23 mixture of racemic E and Z sulfoxides (Scheme 3).17 Rautenstrauch investigated the stereoselectivity of the

2. [2,3]-SIGMATROPIC REARRANGEMENT OF ALLYLIC SULFOXIDES: EARLY RESULTS Most sulfoxides incapable of β-elimination are configurationally stable and racemize at about 200 °C by pyramidal inversion. In contrast, Mislow and co-workers found that allylic sulfoxides 1 racemize at 50−70 °C (ΔH‡ = 22 kcal mol−1 and ΔS‡ = −9 eu) without significant decomposition. These facts suggested a concerted mechanism involving a highly ordered transition state, with the sulfoxide in equilibrium with the achiral sulfenate 2 (Scheme 1).14 Treatment of suitably substituted lithium allyl

Scheme 3. Early Stereochemical Studies

Scheme 1. Early Results in Sulfoxide−Sulfenate Rearrangements

rearrangements of cis- and trans-butenyl p-toluenesulfenates 12, which afforded variable mixtures of racemic diastereomers 13 and 14 (76:24 from the trans-sulfenate and 22:78 from the cissulfenate) that equilibrated to ca. 52:48 mixtures with time. These results were justified in terms of preferred exo transition states 15 for both geometric isomers (Scheme 3).19

alkoxides 3 with p-toluenesulfenyl chloride afforded allyl p-tolyl sulfoxides 5 with a regiochemistry consistent with a concerted 1,3-allylic shift through sulfenate 4.15 Similarly, Braverman and Stabinsky found that the reaction between allyl alcohol and trichloromethanesulfenyl chloride directly gave rise to allyl trichloromethyl sulfoxide even at −70 °C.16 14202

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

introduction of 2-pyridyl and imidazolyl allylic sulfides that provide enhanced α selectivity upon alkylation and are then suitable for undergoing oxidation to the sulfoxide with smooth rearrangement and sulfenate trapping.28 Even at this early stage, it was noted that the rates of the [2,3]sigmatropic sulfoxide−sulfenate rearrangement in steroidal systems were dependent on the chirality at sulfur for a number of configurationally stable allylic methyl sulfoxides,29 and thermal diastereomerization at sulfur of 6β-allylsulfinyl-5α-cholestane was described, involving rearrangement of a steroidal sulfoxide moiety.30

Shortly afterward, Evans and co-workers demonstrated that several cyclic allylic alcohols afforded the expected allylic sulfoxides upon treatment of the lithium alkoxides with aryl sulfenyl chlorides in excellent yields. Upon heating of the allylic sulfoxide in the presence of a suitable thiophile (PhS−, piperidine) capable of trapping the sulfenate ester, the reverse reaction was also high yielding.20 Furthermore, α-deprotonation, α-alkylation, and [2,3]-sigmatropic rearrangement−trapping sequence of 17 efficiently afforded a single geometric isomer of the allylic alcohol 18 (Scheme 4),21 outlining the synthetic Scheme 4. Sequential α-Alkylation and [2,3]-Sigmatropic Rearrangements of Allylic Sulfoxides

3. STEREOCHEMICAL AND MECHANISTIC ASPECTS OF THE MISLOW−BRAVERMAN−EVANS REARRANGEMENT In this section, we summarize the more relevant stereochemical and mechanistic aspects of this useful rearrangement, building on the early findings that were detailed in the previous section. The stereochemical outcome of [2,3]-sigmatropic rearrangements was the subject of an insightful review by Hoffmann.1 3.1. Stereochemical Outcome

3.1.1. Alkene Geometry. The sigmatropic rearrangement of acyclic allylic sulfoxides using common thiophiles [P(OMe)3, PhSNa, etc.] leads to E-alkenes with good to excellent selectivity. In this manner, disubstituted alkenes 29 are obtained from 28 with very high E stereoselectivity by transoid arrangements (Scheme 6). In the case of trisubstituted alkenes 31 and 32, the E Scheme 6. Studies of Alkene Geometry equivalence of the sulfoxide-stabilized anion and a hypothetical vinyl anion derived from allyl alcohols. This protocol was later generalized, and the undesired competition of γ-alkylation of allylic sulfoxides (19, 20) was noted in some cases,22 particularly when the methodology was applied to precursors (21, 22) of trisubstituted olefins (23, 24).23−25 In contrast, the alkylation− rearrangement protocol was highly selective for a functionalized cycloalkenyl phenyl sulfinyl anion with perfect regio- and stereocontrol to produce a useful intermediate in prostaglandin synthesis.26 The Evans group also pursued mechanistic studies on the cleavage of allylic sulfenates with P(OMe)3 and on the rearrangement of allylic sulfoxide 25 to gain insight into the influence of sulfenate trapping with different thiophiles on the sulfoxide−sulfenate equilibrium in a sterically biased situation that ultimately results in varying ratios of alcohols 26 and 27 (Scheme 5). These results indicate a predominant rearrangement across the equatorial face of the ring, consistent with a reactant-like transition state.27 This group provided an excellent account of this research that includes, among other aspects, the

geometry is generally favored, although the precise selectivity depends on the relative size of R1 and R2 for 1,1-disubstituted allylic sulfoxides 30. For sulfoxides 33, with an additional substituent at C-2, despite the 1,2-allyl interaction, the E selectivity is maintained, to produce 34. 3.1.2. Transfer of Chirality. The absolute configuration at sulfur rarely influences the stereochemical outcome in acyclic systems with a stereogenic center at C-1 (Scheme 7). Thus, the suprafacial rearrangement of 35 brings about excellent “selfimmolative” 1,3-chirality transfer to allylic alcohol 36. Chirality transfer from sulfur is challenging and requires the use of an efficient thiophile to prevent extensive racemization because of the reversibility of the rearrangement.31 The case of the E and Z isomers of enantiopure acyclic allylic sulfoxides 37 was studied by Hoffman and co-workers (Scheme 8).32 The

Scheme 5. Influence of Thiophiles on the Ratio of Allylic Alcohols

14203

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 7. 1,3-Transfer of Chirality

Scheme 9. [2,3]-Sigmatropic Rearrangement of C-2Substituted Cyclic Allylic Sulfoxides

Scheme 8. Chirality Transfer from Sulfur in E- and Z-Allylic Sulfoxides

Scheme 10. Mechanistic Studies of Exo and Endo Transition States

rearrangement to the sulfenate can take place through two diastereomeric transition states, endo and exo, for each geometric isomer (38−41) to produce ultimately allylic alcohol 42. The asymmetric induction depends on the energy difference between these transition states, and this is related to interactions between the p-Tol group on sulfur and the substituent at C-3. The modest chirality transfer found for (E)-37 suggests a small energy difference between the two transition states, with endo 38 being slightly favored (70:30). In contrast, the exo-cis transition state 41 for (Z)-37 must be substantially destabilized, and this results in high enantiopurity of alcohol (S)-42, compatible with a 95:5 endo/exo ratio. Inspection of these transition states suggests that the presence of a sufficiently bulky substituent at C-2 of the allylic system could destabilize the endo arrangement and thus favor the exo transition state. This was also studied for aryl cycloalkenyl methyl sulfoxides that underwent preferential rearrangement through the exo transition state with chirality transfers that ranged from 44% ee for cyclopentenyl derivative 43 affording 44 to >90% ee for more hindered derivatives such as 45 (that contains an internal thiophile) affording 46 (Scheme 9).33,34 The diastereoselectivities of these rearrangements in cyclohexenyl methyl sulfoxides with an additional chiral center were also examined for diastereomers 47 and 48 (Scheme 9), which gave rise to allylic alcohols 49 and 50, respectively.35,36 The selectivities found suggested an important preference for the formation of an axial alcohol that, in the case of 48, overrides the exo effect noted for these sulfoxides.

terated compounds at position 3, and this allowed for the calculation of the rate constants and the construction of the freeenergy diagram. The difference in free energy between the exo 53 and endo 54 transition states was calculated as 2.7 kcal/mol, the sulfoxide was calculated to be 1.5 kcal/mol more stable than the sulfenate, the free energy difference between the sulfenate and the endo transition state was calculated to be 18.3 kcal/mol, and the ratio of rate constants kendo/kexo = 275 represents the maximum possible chirality transfer from chiral sulfur. However, these levels of chirality transfer have not been achieved, possibly because of inefficient cleavage of the sulfenates or peculiarities of this system.37 The [2,3]-sigmatropic rearrangement of allylic sulfoxide 55 to sulfenate 56 has been studied by ab initio methods, including an evaluation of the geometries and energies of the reactants and transition states. This study predicts a preference for the endo transition state (TS) over the exo by 1.76 kcal/mol (Scheme 11).38 Jones-Hertzog and Jorgensen performed ab initio and Monte Carlo calculations to clarify the reaction path and solvent effects of the rearrangement of allyl methyl sulfoxide 57 to sulfenate 58. They found that the endo TS was favored by 1.5− 2.2 kcal/mol and that the solvent effects and dipole moments indicated that the TS was more product-like but that the oxygen

3.2. Mechanistic and Computational Studies

The equilibrium between sulfenate 51 and sulfoxide 52 was studied (Scheme 10), including data for selectively monodeu14204

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 11. Calculations for Different Sulfoxide−Sulfenate Systems

Scheme 13. Effect of Pd(0) Catalysis or CF3 Substituents on [2,3]-Sigmatropic Rearrangements

Allylic trifluoromethyl sulfoxides 71 racemize spontaneously at room temperature through sulfenates 72. The process was studied by dynamic high-performance liquid chromatography (HPLC) and DFT calculations at different levels, which revealed that the endo and exo approaches were almost isoenergetic in these cases and that the calculated energy values agree well with experimental results.46

in the TS maintained sulfoxide-like character, with hydrogenbond-accepting ability.39 Jones-Hertzog and Jorgensen also employed ab initio calculations to investigate the four possible endo−exo TSs for diastereomeric sulfoxides 59a and 59b, which experimentally lead to a 97:3 mixture of (E)-60 and (Z)-61, and concluded that one diastereomer (R,S or S,R) rearranges preferentially through an endo TS and that the other diastereomer (R,R or S,S) rearranges preferentially through an exo TS. A detailed analysis of the steric interactions that justify this behavior was provided by the authors.40 The rearrangements of allyl sulfoxides 62 to the corresponding sulfenates 63 were examined computationally with several basis sets, leading to the conclusions that the rearrangements are concerted, the endo and exo transition states are very close in energy (ΔE ≤ 1 kcal/mol), and the activation barriers are in the 19.0−21.8 kcal/mol range.41 The somewhat related rearrangement of cinnamyl 4-nitrobenzenesulfenate 64 to produce cinnamyl p-nitrobenzenesulfoxide 67 was studied mechanistically, and a diradical pathway appeared to be the most likely mechanism,42 as was subsequently confirmed experimentally (Scheme 12).43 A density functional

4. SYNTHESIS OF ALLYLIC SULFOXIDES Oxidation of allylic sulfides is a general method for the synthesis of allylic sulfoxides that frequently employs H2O2 as a stoichiometric oxidant along with catalysts. Selected recent examples have used copper(II) complexes,47,48 Merrifield resinsupported peroxomolybdenum(VI) complexes,49 synthetic flavins,50 and 2,2,2-trifluoroacetophenone.51 Other stoichiometric oxidants such as calcium hypobromite, Ca(BrO)2,52 and mCPBA53,54 yield racemic allylic sulfoxides; however, biocatalytic oxidation of prochiral sulfides, using Corynebacterium equi IFO 373055 and Fe(II)/α-ketoglutarate-dependent dioxygenase from E. coli,56 have been reported to provide optically pure allylic sulfoxides. Interestingly, when allylic sulfides containing chiral substituents are oxidized, diastereomeric mixtures of allylic sulfoxides are obtained.57,58 An example is described in the synthesis of the camphor-derived allylic sulfoxide 74, unusually stable at room temperature and characterized by X-ray crystallography, from sulfide 73 with m-CPBA; hydroxy sulfoxide 74 underwent epimerization at the sulfoxide center upon heating for 3 h at 110 °C (74/75, 65:35). Furthermore, when compound 74 was heated to 145 °C, diastereomer 75 was obtained quantitatively (Scheme 14).59,60

Scheme 12. Alternative Mechanistic Proposals for [2,3]Sigmatropic Rearrangements

Scheme 14. Diastereoselective Oxidation of CamphorDerived Sulfide

theory (DFT) study on substrates 68 led to the conclusion that the rearrangement of alkyl and aromatic derivatives takes place by a radical breakdown pathway.44 The palladium-catalyzed allylic sulfinylation and the Mislow− Braverman−Evans rearrangement (from 69 to 70) of three model systems (Scheme 13) were investigated computationally and found to reveal moderate barriers (ΔG‡ = 82−104 kJ mol−1) for the free substrates, reduced to just 10−26 kJ mol−1 in the presence of the Pd(0) diphosphinoethane complex, thus suggesting a catalytic effect of palladium on the rearrangement.45

Aside from sulfides, other compounds have served as starting materials for the synthesis of allylic sulfoxides. Some examples are found in the reaction of alkenes with p-toluenesulfinamide catalyzed by Yb(III) and trimethylsilyl chloride (TMSCl)61 or with PhSOCl and ZnCl262 or EtAlCl2.63 Allenyl sulfoxides react with AlCl3 to provide 2-haloallyl sulfoxides,64 α-lithio sulfinyl 14205

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

carbanions give allylic sulfoxides by reaction with Cr and W Fischer carbenes,65 and β-sulfinylesters 76 generate sulfenate anions in basic media that are trapped with allylic acetates under Pd(0) catalysis to generate allylic sulfoxides 77−79 with different substitution (Scheme 15).66,67

Scheme 17. Synthesis of Enals and Epimerization of Norbornene Allylic Alcohols

Scheme 15. Synthesis of Allylic Sulfoxides from β-Sulfinyl Esters

unstable allylic sulfoxides 87, which rearrange to intermediates 88, which, in turn, are readily transformed into unsaturated aldehydes 89.73 This methodology was applied to prepare a fragment of papulacandin D.74 Starting from allylic alcohols, the sulfenate−sulfoxide rearrangement represents a valuable tool for the inversion of the configuration of hydroxyl groups that competes with Mitsunobu or oxidation−reduction protocols even in complex molecules. In this context, Brown and Fallis described the epimerization of an endo norbornane alcohol 90 by formation of a sulfenate that rearranged to sulfoxide 91, producing the target alcohol 92 upon a second rearrangement (Scheme 17). The unsaturated norbornene analogue of 91 was prepared by oxidation of the related sulfide and behaved similarly, producing a 9:1 mixture of norbornenols that underwent smooth oxy-Cope rearrangement.75 Whitham and colleagues reported that the rearrangement of sulfenate 93 afforded a single diastereomer of the sulfoxide, tentatively assigned as 94 (Scheme 18). In contrast, oxidation of

Finally, chiral sulfinate esters such as diacetone-D-glucose (DAG) or menthyl derivatives are among the more frequent precursors of all kinds of sulfoxides and are also useful for accessing enantiomerically enriched allylic sulfoxides by reaction with Grignard68 or lithium69 reagents. An interesting related approach developed by Vogel and Turks and co-workers70,71 is based on the sila- and bora-ene reactions of allylsilane and allylborane derivatives 80 and 81 with SO2, which allows for the synthesis of allylic sulfinates 82 and 83 (Scheme 16). The latter Scheme 16. Synthesis of Allylic Sulfoxides from Allylsilanes and Allylboranes

Scheme 18. Synthesis of Allylic Sulfoxides and Sulfinamides from Allylic Alcohols are subjected to chemoselective nucleophilic attack at sulfur with Grignard reagents, generating a series of allylic sulfoxides (84, 85) in fair yield. Further attempts to obtain enantiopure products using (R)-1,1′-bi-2-naphthol [(R)-BINOL] and (+)-α-pinenederived allylic boronates produced only moderate enantioselectivities.

5. GENERAL SCOPE OF [2,3]-SIGMATROPIC SULFOXIDE−SULFENATE REARRANGEMENTS Besides the examples of Mislow−Braverman−Evans rearrangements discussed in the tandem reactions or in the application in the synthesis of natural products (see below), there are numerous cases of interesting transformations that involve these rearrangements. In this section, we cover these examples that illustrate the scope of the process in different scenarios.

the sulfide afforded a 1:1 mixture of allylic sulfoxides 94 and 95. Treatment of this mixture with P(OMe)3 led to substantial amounts (ca. 20%) of the less reactive diastereomer 95.76 Stork and Kreft described several related rearrangements of cyclohexenyl propyl sulfenates;77 Hua et al. tested cyclopentenyl allylic sulfoxides, prepared from hydroxymethyl cyclopentenes, within synthetic efforts to access linear triquinanes;78 and the

5.1. From Allyl Sulfenates

Structural confirmation for a natural diterpene with interesting antipeptic ulcer activity was obtained by synthesis of an isomer involving sulfenate−sulfoxide rearrangement, alkylation, and sulfoxide-to-sulfenate rearrangement.72 Treatment of 3-trimethylsilyl allylic alcohols 86 (Scheme 17) with PhSCl gives 14206

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

In contrast to the related β-diastereomers of sulfoxides 105 (Scheme 20), which interconverted at 50 °C, α-isomers 105, upon thermolysis and hydrolytic workup, afforded alcohols 106, presumably through an intermediate ion pair 107.91 Treatment of chlorosulfoxide 108 with PhSK afforded α-phenylthio β,γunsaturated carbonyl compound 109 through cleavage of the unstable α,β-epoxy sulfoxide intermediate. Clean conversion to the rearranged allylic alcohol 110 required oxidation to the α,βunsaturated γ-phenylsulfinyloxy carbonyl compound precursor that was smoothly hydrolyzed with base. The use of enantiopure chlorosulfoxides allowed for the preparation of related cyclohexenols but in just 67−82% ee, indicating partial racemization at some point.92 Treatment of chloro alkenyl sulfoxides 111 with an excess of lithiated 2-piperidone afforded allylic alcohol 112 with the sulfur atom retained at the original carbon, presumably through sulfenate intermediate 113, cleaved by excess lithiated piperidone, followed by vinylic deprotonation and sulfenylation (Scheme 20).93 Oxidation of an allylic sulfide contained in a 14-membered ring and attempted isolation of the sulfoxides afforded a diene structurally compatible with [2,3]-rearrangement and elimination of sulfenic acid. Addition of P(OMe)3 to the oxidation mixture produced the expected allylic alcohol.94 The rearrangements of diastereomeric sulfoxides 114 (Scheme 21) and of the

rearrangement of a substituted cyclohexyl allylic alcohol was also used for the preparation of a semicyclic 1,3-diene.79 Treatment of a variety of allylic alcohols 96 with 4morpholinesulfenyl chloride at −78 °C afforded aminosulfenates 97, which, upon warming to room temperature, produced allylic sulfinamides 98 in good yields in a process that closely resembles the standard sulfenate−sulfoxide rearrangement (Scheme 18).80,81 A variety of difluoroallylic alcohols undergo smooth [2,3]-rearrangement to the allylic sulfoxides.82 The electrophilic trifluoromethylsulfenylation of allylic alcohols 99 with saccharin derivative 100 gives the sulfenates, which rearrange spontaneously into trifluoromethyl sulfoxides 101.83 These sulfoxides undergo spontaneous racemization at room temperature through the corresponding sulfenates.46 5.2. From Allyl Sulfoxides

In some cases, the rearrangement of allylic sulfoxides to E-allylic alcohols takes place under mild acidic conditions (0.1 equiv of pTsOH).84 This rearrangement also takes place by catalysis with antibodies followed by sulfenate cleavage with dithiothreitol with some substrate specificity as well as modest enantio- and diastereoselectivity.85 Sulfenate cleavage of base-sensitive allylic sulfoxides was successfully promoted by warming in pH 7.0 phosphate buffer.86 In some cases, the system has been designed to allow for intramolecular cleavage of the intermediate sulfenate and liberation of allylic alcohol.87 This strategy was used in an approach to N-alkylisothiazolidines.88 Treatment of a keto sulfoxide with 2 equiv of Ph3PCH2 in dimethyl sulfoxide (DMSO) generated an allylic sulfoxide that rearranged immediately to the allylic alcohol, presumably with trapping of the sulfenate by the ylide.89 Cycloadducts 102 derived from cyclopentadiene and alkanethial S-oxides rearrange smoothly in refluxing dichloromethane to afford sultenes 103 that are cleaved with PhLi to allylic alcohols 104 (Scheme 19).90

Scheme 21. Differential Reactivity of Sulfoxide Diastereomers in Sulfoxide−Sulfenate Rearrangements

Scheme 19. Sultenes from [2,3] Sulfoxide−Sulfenate Rearrangements

Scheme 20. Unexpected Results in the Synthesis of Allylic Alcohols by [2,3]-Sigmatropic Rearrangements

14207

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

precursors that allowed for the transformation of allylic sulfide 128 (Scheme 23) into trisubstituted aldehyde 129.104 It was also

related 2-Z isomers (not shown) afforded allylic alcohol 115 in good yields. Interestingly, the rate of rearrangement of the major diastereomer at the sulfur of 114 was very slow, and this was attributed to reaction through transition state 116, with significant interaction between the Ph group and the bulky −CMe2OH group. In contrast, the rapidly reacting diastereomer at sulfur would place the Ph moiety far from the tertiary alcohol.95 Hauser et al. reported an even more remarkable case of differential reactivity of sulfoxide diastereomers in which, whereas 117 rearranged smoothly to allylic alcohol 118, the other sulfoxide diastereomer was completely unreactive, presumably because of the severe interactions between Ph and Me groups in the geometry necessary for a successful rearrangement.96 The Mislow−Braverman−Evans rearrangement of γ-chloroallyl sulfoxides takes place much faster than the rearrangement of the nonhalogenated counterparts and produces α,βunsaturated ketones in good yields if the process is carried out with cyclohexene as a PhSCl trap.97 This allows for a completely chemoselective rearrangement of bis-sulfoxide 119 to unsaturated keto sulfoxide 120 in excellent yield, along with 121 through intermediate 122 (Scheme 22).98 Oxidation of the

Scheme 23. Mislow−Braverman−Evans Rearrangement of γMethoxy Allylic Sulfoxides

Scheme 22. Mislow−Braverman−Evans Rearrangement of γChloroallyl Sulfoxides possible to prepare carbonyl adducts such as 130 with good α selectivity relative to sulfur, and the rearrangement of a variety of branched substrates at position β relative to the sulfoxide took place with excellent E selectivity to produce trisubstituted aldehydes 131 (Scheme 23). This preference appears to be general, and examples of all-carbon β-branching and terminal alkenes lacking the MeO substituent were reported.105 Similarly, the rearrangement of substrate 132, with a γ-hydroxyl substituent relative to sulfur, was highly selective to aldehyde 133 (E/Z, 90:10). Protection of the alcohol or removal of the MeO vinyl ether substituent resulted in selectivities in the 60:40 range. The unique influence of the free hydroxyl was rationalized in terms of an interaction between the hydroxyl and sulfinyl groups that favors a conformation of 134 that places the side chain pseudoequatorial.106 Zwanenburg and co-workers introduced the sulfoxide-induced sigmatropic rearrangement (SISR) of alkenyl sulfoxides with a sulfide substituent at the geminal position. Treatment of substrates 135 (Scheme 24) with NaSPh triggers isomerization to the allylic sulfoxide 136, followed by sigmatropic rearrangement and thiophilic cleavage to produce γ-hydroxy vinyl sulfides 137 in good yields and selectivities in a process reminiscent of the SPAC reaction.107 Warren explored phenylthio shifts in allylic systems and used the resulting allylic sulfides to prepare allylic

herbicide diallate gave unstable carbamoyl sulfoxide 123, which rearranged spontaneously to carbamoyl sulfenyl chloride 124 and 2-chloroacrolein 125.99 These processes were also examined in vitro and in vivo in an effort to understand the degradation pathways of diallate, a carcinogen in rats, and these studies revealed that this pathway was also operative to some extent with the production of 2-chloroacrolein as an ultimate mutagen.100 Similarly, the oxidation of cysteine S-conjugates of 1,3dichloropropene, a soil fumigant, was investigated, and a related rearrangement to finally release acrolein was proposed.101 This rearrangement was also facile for bicyclic fluorinated substrate 126, which produced unsaturated ketone 127 in excellent yield (Scheme 22).102 Following on earlier observations of selective α-alkylation of γmethoxy allylic sulfides and straightforward oxidation with concurrent facile [2,3]-rearrangement and hydrolysis to the unsaturated carbonyl compounds,103 Otera and colleagues reported an improved method for the preparation of the sulfide

Scheme 24. Sulfoxide−Sulfenate Rearrangements in SulfideSubstituted Allyl Sulfoxides

14208

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

alcohols by oxidation and [2,3]-rearrangement.108 In this manner, it was possible to prepare allylic sulfoxides such as 138 by selective oxidation of the disulfide precursor (Scheme 24) and to explore their rearrangements to afford allylic alcohols 139 with a vinyl sulfide functionality.109 Phenylthio migration of 140, prepared selectively by aldol chemistry and functional-group manipulation, gave allylic sulfide 141, which was oxidized to a 50:50 mixture of sulfoxide diastereomers that rearranged smoothly to syn derivative 142 (Scheme 25).110 Cyclic allyl silanes were also suitable precursors of allylic sulfides, which were oxidized to the sulfoxides and rearranged in the presence of P(OMe)3.111,112

Scheme 26. Mislow−Braverman−Evans Rearrangement in Acyclic Polyunsaturated Compounds

Scheme 25. Diastereoselective Synthesis of Allylic Alcohols

Scheme 27. Regioselective Synthesis of Cyclopentenyl Alcohols

dienes with a methyl sulfoxide at C-3 leads to [2,3]-sigmatropic rearrangement, and this approach has been applied in a synthesis of Yomogi alcohol.120 Walsh and co-workers investigated the use of sulfoxide 156 (Scheme 28) to induce the inactivation of two pyridoxal phosphate-dependent enzymes. The process is believed to start with the formation of pyridoxal enamine-iminium intermediate 157, which undergoes chloride elimination to generate an allyl sulfoxide−enzyme−pyridoxal adduct 158 and subsequent [2,3]sigmatropic rearrangement to sulfenate 159, which reacts with an enzymic nucleophile in a novel form of suicide inactivation affording 160.121 The use of [2,3]-sigmatropic rearrangement of allylic sulfoxides for inactivation of cyclohexanone oxygenase was explored with limited success.122 Within the context of studies directed toward clarifying the role of farnesylated proteins in ras oncogenes, Gordon and Pluscec prepared the sulfoxides of BocS-farnesylcysteinylvaline methyl ester 161 (Scheme 28) and explored the thermal interconversion of the separable sulfoxide diastereomers, as well as the reaction with different thiophiles of the intermediate sulfenates.123 The participation of [2,3]sigmatropic rearrangements of suitably functionalized nucleotides in interstrand cross-link formation was also examined.124 Lithiation of vinyl sulfoxide 162 (Scheme 29) and alkylation afforded allylic sulfoxide 163, which rearranged smoothly to allylic alcohol 164.125 The use of α-phosphoryl vinyl sulfoxide 165 allowed for a tandem Michael addition/Horner olefination and led to a number of heterocyclic vinyl sulfoxides. However, the reaction with o-aminoacetophenone 166 afforded hydrox-

In the course of a study on the asymmetric induction of the condensation between metalated allylic sulfoxides and benzaldehyde, a very facile rearrangement due to steric crowding was reported.113 Cinquini, Cozzi, and co-workers examined the condensation of lithiated allyl p-tolyl sulfoxide with chiral αmethyl aldehydes 143 (Scheme 25) seeking conditions to improve the α-/γ-regioselectivity (144/145, 5.6:1−9:1) and the diastereoselectivity in the α regioisomer (146, 2.1:1−28:1), which was determined by standard sigmatropic rearrangement of substrate 144 to produce diol 146.114 A similar study using 2pyridyl allyl sulfoxide and chiral alkoxy aldehydes was also undertaken by this group.115 1,3-Butadienes with an allylic sulfoxide such as 147 (Scheme 26) also participate in [2,3]-sigmatropic rearrangements to afford 148.116,117 It is possible to functionalize regioselectively the isopropylidene fragment of substrates such as 149 by reaction with PhSOCl and ZnCl2. The resulting allylic sulfoxide 150 undergoes smooth rearrangement to allylic alcohol 151.62,118 Trost and Matelich developed a phenylthio-substituted trimethylenemethane precursor that reacts with a variety of acceptors (alkenes, aldehydes, sulfonimines, etc.) to produce sulfides such as 152 (Scheme 27) that are readily isomerized to 153. Both regioisomeric allylic sulfides can be oxidized to the sulfoxides that lead to regioisomeric alcohols 154 and 155 upon sigmatropic rearrangement.119 The thermolysis of hexa-1,514209

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 28. Rearrangements of Allyl Sulfoxides in Biological Processes

Scheme 30. Diastereoselective Synthesis of 1,4-Diols and 1,4Aminoalcohols

Scheme 31. Metal-Mediated [2,3]-Sigmatropic Rearrangements

involve oxidative addition to the allyl sulfenate (in equilibrium with the starting allylic sulfoxide) to generate an η 3 allylpalladium(II) intermediate 175, followed by trapping with t-BuOK and release of Pd(0) and sulfenate anion. Oxidative addition to the aryl halide sets the stage for transmetalation with the sulfenate and reductive elimination.131 The dynamic kinetic resolution of allylic sulfoxides 176, combining a Rh-accelerated sigmatropic rearrangement in polar solvents by a π-allyl intermediate with olefin hydrogenation believed to involve Ocoordination of the sulfinyl moiety, results in good yields and enantiopurities of sulfoxides 177.132 Attempted Mislow−Braverman−Evans rearrangement of the sulfoxide derived from 178 (Scheme 32) failed, and a product derived from syn elimination and cyclization involving the

Scheme 29. Vinyl-to-Allyl Sulfoxide Isomerization Followed by [2,3]-Sigmatropic Rearrangement

ymethyl derivative 168 as the major product, presumably by vinyl-to-allyl isomerization of the reaction intermediate, and [2,3]-sigmatropic rearrangement followed by aromatization.126 This rearrangement was used for the epimerization of vinyl cyclobutanols,127 and it was mentioned in the optimization of a related rearrangement involving allyl sulfinyl carbanions.128 Raghavan et al. explored the reactivity of α-chloro sulfides bearing additional oxygen substituents with organozinc reagents to afford a variety of substrates related to 169 (Scheme 30) with good diastereoselectivities in most cases. Oxidation and sigmatropic rearrangement produced 1,4-diol derivatives 170 in excellent yields.129 Maruoka and co-workers reported the organocatalytic Mannich reaction of α-thio acetaldehydes with N-Boc imines followed by Takai olefination to afford amino sulfides 171 in excellent yields, as well as stereo- and enantioselectivities. Oxidation and [2,3]-sigmatropic rearrangement afforded unsaturated 1,4-aminoalcohols 172 in good yields.130 Allylic sulfoxides 173 are suitable precursors of sulfenate anions within a new Pd(0)-catalyzed domino process that results in aryl sulfoxides 174 (Scheme 31). The process is believed to

Scheme 32. Related Pummerer-like Transformations on Allylic Sulfoxides

14210

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

ketone was obtained. Alternatively, reduction of the carbonyl first, followed by oxidation and subsequent treatment with trifluoroacetic anhydride, afforded 179, resulting from a formal Mislow−Braverman−Evans rearrangement but likely to involve a different mechanism through activation of the sulfoxide and allylic displacement by the nucleophile.133 Raghavan and Ravi reported on related Pummerer-like transformations on allylic sulfoxide 180 to produce valuable synthetic intermediates 181 and 182 depending on the activation reagent.134

Scheme 34. Sulfoxide−Sulfenate Rearrangement Followed by [1,2]-Elimination

6. SULFOXIDE−SULFENATE REARRANGEMENTS IN TANDEM PROCESSES This section of the review aims to discuss several synthetically useful processes in which sulfoxide−sulfenate rearrangement (or the reverse rearrangement) is coupled with other important reactions. In some cases, the entire process takes place in a single step, but in other examples, there are some intermediate synthetic transformations between the key “tandem processes”.

sequence of sigmatropic rearrangements.141 The Reich protocol for the preparation of 1,3-dienes from allylic alcohols has been used in the synthesis of the plant growth regulator (−)-penienone,142 the natural product (5R)-thiolactomycin, an inhibitor of fatty acid synthesis,143 new related inhibitors,144 and a selectively deuterated cyclononadiene.145

6.1. [2,3]-Sigmatropic Rearrangement and [1,2]-Elimination

Reich et al. developed a general methodology for transforming a variety of cyclic and acyclic allylic alcohols into 1,3-dienes by a 1,4-dehydration shown to take place with cis stereochemistry in a cyclic system. The method entails the treatment of the alcohols (183−185) with 2,4-dinitrobenzenesulfenyl chloride and triethylamine to produce sulfenates such as 186 that undergo [2,3]-sigmatropic rearrangement and thermal syn elimination of the intermediate allylic sulfoxide to afford 1,3-dienes 188, 189, and 190, respectively (Scheme 33).135 The process was studied

6.2. Multiple [2,3]-Sigmatropic Rearrangements

Trost and Stanton reported that the sulfenylation of 3,4dehydro-β-ionol 195 under standard conditions gave rise to a good yield of allylic sulfoxide 196 as a complex isomeric mixture. Base-promoted sulfenylation triggered a second net [2,7]rearrangement to afford alcohol 197 in good yield (Scheme 35). These processes could take place by three sequential [2,3]-

Scheme 33. Sulfenate−Sulfoxide Rearrangement Followed by [1,2]-Elimination

Scheme 35. Sequential [2,3]-Sigmatropic Rearrangements

in depth, and its scope and limitations were defined,136 so that it was soon used by other groups in the synthesis of a pheromone137 and in the transformation of isocodeine into 6demethoxythebaine.138 The thermolysis of allylic sulfoxides that are unable to provide a typical syn elimination (191, 192) gives rise to 1,3-dienes (193, 194) presumably by [2,3]-rearrangement followed by thermal elimination of sulfenic acid, as has been suggested in these cases (Scheme 34).139,140 Within synthetic efforts to 9-deoxyprostanoids, the low-yielding preparation of 1,3-dienes by thermolysis of allylic sulfoxides lacking syn hydrogens has been attributed to a

sigmatropic rearrangements.146 Gaoni reported the thermal equilibration of pure (E)- or (Z)-sulfinyl allylic dienes 198 to produce the same 2:1 equilibrium mixture believed to occur through a symmetrical sulfenate ester 199.147 Kametani et al. reported the exceptionally facile anionaccelerated thermal ring opening of cyclobutenes with appropriately placed sulfoxide or sulfone substituents, such as 200 (Scheme 36). Whereas sulfones afforded dienes related to 201, in the case of sulfoxides, the sole products were dienes related to 202, presumably through a double [2,3]-sigmatropic rearrangement.148 Corey and Hoover described that lithiation of pentadienyl sulfoxide 203, followed by reaction with an aldehyde and benzoylation of the resulting alcohol to 204, gave, after standing at room temperature for 3 h, a 60% overall yield of doubly rearranged dienyl sulfoxide 205. After Pummerer rearrangement, 205 was finally transformed into 5-desoxyleukotriene D.149 In related studies, Corey et al. found a novel 1,7 sulfoxide−sulfoxide migration, and two mechanistic hypotheses were formulated to account for this result.150 The double [2,3]-sigmatropic process is believed to operate in the 1,5-sulfoxide migration of 2-sulfinyl-hexa-2,4-dienoate, which, upon treatment with 1,5-diazabicyclo[4.3.0]non-5-ene 14211

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 36. Double [2,3]-Sigmatropic Rearrangements of Dienyl Sulfoxides Generated in Situ

Scheme 38. General Scheme for the SPAC Reaction

reported that the treatment of cyclopentylidene sulfinyl acetate 212 with 2:1 Ac2O/pyridine led to an excellent yield of γ-acetoxy unsaturated ester 213 (Scheme 39).155 This group later Scheme 39. Stepwise SPAC-like Sequences

(DBN), affords 6-sulfinyl-hexa-2,4-dienoate,151 and additional examples of this chemistry were described by Florio and coworkers.152 Schreiber and Satake reported that the treatment of lithiated pentadienyl sulfoxide 203, prepared by sulfenate− sulfoxide rearrangement from divinyl carbinol, with a chiral bicyclic aldehyde and double sigmatropic rearrangement at room temperature gave an excellent yield of alcohol 206 as a 3:1 mixture of epimers at the new center (Scheme 37). This Scheme 37. Double Sigmatropic Rearrangements in Approaches to Natural Products

generalized these results to a number of acyclic and cyclic αphenylsulfinylacrylate derivatives.156 Trost and Rigby reported that the oxidation of sulfide 214 to the related sulfoxide and subsequent treatment with 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU)/MeOH afforded 215, a valuable intermediate for an approach to verrucarins. This transformation is rationalized by isomerization to allylic sulfoxide 217, rearrangement, sulfenate trapping, and spontaneous ring expansion to hemiketal 218 (Scheme 39).157 In 1981, Sammes and co-workers described the Knoevenagel condensation of phenylsulfinyl acetate 219 with aldehydes by means of a zinc enolate to afford 220. Subsequent treatment with aqueous pyridine promoted the isomerization, rearrangement and sulfenate trapping. Alternatively, the use of Mg(OMe)2 led directly to the rearranged alcohols 221 (Scheme 40).151,158

intermediate was eventually transformed into (±)-asteltoxin.153 Similar chemistry gave rise to sulfinyl dienol 207 that, after oxidation to the sulfone and α-lithiation, allowed for condensation with a functionalized aldehyde at the sulfone terminus and then led to a fragment of polycavernoside A.154 6.3. Sulfoxide, Piperidine, and Carbonyl (SPAC) Reaction

A particularly interesting variant of the preparation of allylic sulfoxides by isomerization of alkenyl sulfoxides is the Knoevenagel condensation between enolizable aldehydes, an excess of piperidine, and methylene-activated sulfoxides 208 (Scheme 38). In this case, the resulting vinyl sulfoxide 209 undergoes a C−C double-bond isomerization to 210, [2,3]sigmatropic rearrangement, and trapping of the sulfenate ester to produce γ-hydroxy-α,β-unsaturated derivatives 211. This tandem process is referred to as the SPAC (sulfoxide, piperidine, and carbonyl) condensation.11 6.3.1. SPAC-Like Reactions. Prior to the development of the tandem SPAC protocol, several groups outlined stepwise protocols that entailed the formation of the Knoevenagel condensation product and treatment with base to promote isomerization and sulfenate trapping. Uda and co-workers

Scheme 40. SPAC Reaction from Phenyl Sulfinyl Acetate

14212

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

the methodology was expanded to the preparation of unsaturated ketones171 and phosphonates.172 It was determined that the reverse rearrangement (from sulfenates of 4-hydroxy-2-alkenoates to allylic sulfoxides) does not take place,173 and the lipasemediated resolution of SPAC-derived unsaturated esters was reported.174 This methodology was also applied to the synthesis of ligands containing spirotetrahydrofuranyl motifs,175 and conditions to carry out the SPAC reaction between arylsulfinylacetonitrile and aldehydes in water under catalysis by N,Ndiethylaminopropylated silica gel were reported.176 6.3.3. Synthesis of E-γ-Hydroxy-α,β-unsaturated Sulfones by the SPAC Methodology. In 1990, Carretero and coworkers reported a facile SPAC protocol that entailed a piperidine-catalyzed condensation between p-tolylsulfinyl phenyl sulfone 234 and a variety of aldehydes 232 to produce hydroxy unsaturated sulfones 235 in just one synthetic operation, in good yields, and with complete E selectivity (Scheme 43).177,178 These unsaturated hydroxy sulfones were efficiently resolved by a lipase PS-catalyzed enantioselective acylation with vinyl acetate in i-Pr2O.179,180 Trost and Grese reported related studies using 4-chlorophenylsulfinyl phenylsulfonyl methane 236 and included aldehydes with a methyl alkyl chiral center next to the aldehyde that led to nonselective mixtures of isomers. In contrast, carbohydrate-derived aldehyde 237 produced sulfone 238 as a 7.6:1 mixture of diastereomers (Scheme 43).181 In subsequent years, Carretero’s group developed efficient methodologies to carry out stereoselective syn and anti additions of organolithium or organocopper reagents to the hydroxy vinyl sulfones exemplified by 239182,183 and further studied an iterative methodology to access polypropionate chains (Scheme 44).184,185 This protocol entailed the transformation of the sulfonyl groups of 240 and 241 into aldehydes 242 and 243, respectively, and subsequent SPAC condensation with enantiopure sulfonyl sulfinyl reagents. These γ-hydroxy vinyl sulfones have also been used for the synthesis of syn 2-aminoalcohol derivatives;186 for the preparation of carbocycles by free-radical cyclization;187 and for the preparation of polyhydroxylated quinolizidines,188 indolizidines,189 pyrrolizidines,190 and medium-sized cyclic amines.191 Additional applications include the Pd-catalyzed allylic substitution of the activated γ-hydroxyl,192 the Claisen rearrangement,193 and the endo selective intramolecular Pauson−Khand reaction.194 6.3.4. Enantiopure Sulfoxides in the SPAC Reaction. The possibility of achieving 1,4-chirality transfer from enantiopure sulfoxides to the γ-carbon atom was examined by the in situ SPAC methodology from enantiopure sulfinyl derivatives 244 and 245 or from enantiopure Knoevenagel adducts 246 and 247 (Scheme 45) in the presence of adequate bases and thiophiles to produce the desired γ-hydroxy unsaturated derivatives 248−250 in good yields but with moderate enantioselectivities.178,195−197 Burgess and colleagues explored the enzymatic resolution of a variety of sulfinyl acetates with Pseudomonas sp-K-10 with excellent results and studied the influence on the enantioselectivity of the SPAC reactions of different groups on sulfur. The best balance between yield and enantioselectivity was found for methyl p-chlorophenylsulfinyl acetate.198,199 The combined use of this protocol and a subsequent enzymatic resolution of the resulting hydroxy unsaturated esters provided valuable SPAC products of higher optical purity.200 A continuous flow variant of this resolution that solves some product inhibition issues was recently reported.201 A methodology to produce enantiopure γ-

Additional related observations are gathered in Scheme 41. Oxidation of phenylthiobutenolide 222 followed by reaction Scheme 41. Isomerization−Rearrangement Cascade Reactions

with Ac2O/pyr triggered the isomerization−rearrangement cascade producing acetoxy derivatives 223.159 Arylsulfinyl alkenoates 224, prepared by Knoevenagel condensation with catalytic piperidine,160 were refluxed in xylene with K2CO3 to produce mixtures of diene esters 225 and hydroxy alkenyl esters 226, presumably by isomerization and rearrangement to the intermediate sulfenate, which leads to alcohols 226 by sulfenate hydrolysis or to dienes 225 by elimination.139,161,162 6.3.2. Early SPAC Reaction and Selected Applications in Synthesis. In 1981, Nokami reported that the treatment of phenylsulfinylacetonitrile 227 with aldehydes or ketones 228 in the presence of piperidine gave γ-hydroxy-alkenenitriles 229 as single isomers and in good yields. Furthermore, in the case of methylketones, it was possible to control the regioselectivity of the introduction of the hydroxyl group (Scheme 42).163 This Scheme 42. Early Examples of the SPAC Reaction

report, commonly accepted as the first description of a SPAC protocol, was accompanied shortly afterward by the findings of Tanikaga et al. on the related chemistry found for methyl pchlorophenylsulfinyl acetate 231 and a variety of aldehydes 232, which led to the expected hydroxylated unsaturated esters 233 in good yields.164,165 Nokami’s group applied this chemistry to the efficient synthesis of several terpenoids,166 Blade and Robinson used the ester variant in the synthesis of piperovatine,167 Trost and Tometzki applied this chemistry to the synthesis of 1,4enediols,168 and other groups also used this straightforward methodology to access different substrates.169,170 The scope of 14213

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 43. E-γ-Hydroxy-α,β-unsaturated Sulfones by the SPAC Reaction

Scheme 44. Approach to Polypropionate Chains through an Iterative SPAC Methodology

Scheme 46. Double Diastereoselection in the SPAC Reaction

Scheme 45. Enantioselective Approaches to the SPAC Reaction

Trost and Mallart examined in detail the SPAC condensation between (S)-citronellal 256 and a variety of sulfinyl acetates with electron-withdrawing groups (Cl, NO2, PhSO2) at the para position of the aryl ring to determine the influence of the double diastereodifferentiation in the process. This study revealed that the stereochemical induction depends on the absolute stereochemistry of the sulfoxide, the substituent on the sulfoxide, and the amount of base.203 Scheme 47 gathers the results for enantiomeric sulfoxides (R)-257, which led to a 7.3:1 mixture of 258 and 259 in a matched scenario, and (S)-257, which afforded a 1:4.6 mixture of 258 and 259 in a mismatched scenario. These results led the authors to question the commonly accepted model for this process, which involved only the asymmetric protonation as the stereochemistry-determining step, because the chiral center from the aldehyde is remote from the center, being created α to the sulfoxide and ester groups, and the influence of excess piperidine on the dr of the process does not fit the former model. These authors rationalized the process in terms of reactive conformers 260 (matched) and 261 (mismatched), with excess piperidine facilitating interconversion of the diastereomers and trapping of the sulfenates in each case with a preferred C−O bond formation anti to the slightly more bulky R substituent and considering the energetics of the orientation of the Ar moiety as well. Carretero and Domı ́nguez carried out related studies on the preparation of γ-hydroxy alkenyl sulfones184,185 and found that the optimal conditions for the matched pairs [(R)-234/242, (S)234/243] involved CH2Cl2 at −20 °C to produce about 8:1 mixtures of diastereomers 262 and 263, respectively (Scheme 48); these results can be rationalized in terms of the stability of the envelope conformations of 264 and 265 with PhSO2 and Tol groups at equatorial positions to minimize steric strain. Nokami and co-workers recently introduced enantiopure 2(2-benzothiazolylsulfinyl)acetates 266, which gave the desired enoates 267 in higher yields and shorter times than the related p-

hydroxy alkenyl sulfones using biocatalytic resolution was also developed.179,180,185 In another strategy to enhance the optical purity of the resulting hydroxylated unsaturated esters, Burgess and Henderson examined the SPAC reactions of sulfinyl acetates 251 derived from several enantiopure alcohols with either configuration at sulfur to produce adducts 255 (Scheme 46). This study revealed that the asymmetry at sulfur was predominant and that a mismatched−matched scenario was operative, with the (−)-Cam auxiliary 252 being the most effective [diastereomeric ratio (dr) = 78:22/91:9 in the matched case].202 These results suggest that protonation α to the sulfoxide is the key step determining the enantioselectivity of the process. 14214

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 47. Double Diastereodifferentiation in the SPAC Reaction with (S)-Citronellal

Scheme 50. Enantioselective Synthesis of γ-Hydroxy αEnones through Vinyl-to-Allyl Isomerization

Scheme 48. Double Diastereodifferentiation in the SPAC Reaction with Enantiopure Sulfinyl Sulfones

The SPAC condensation between (S,S)-bis-p-tolylsulfinyl methane 270 and aldehydes 232 has been reported. The γhydroxy unsaturated sulfoxides 272 are obtained as inseparable 50:50 mixtures of diastereomers (Scheme 51),206 which were Scheme 51. SPAC Reaction Using (S,S)-Bis-p-tolylsulfinyl Methane and Aldehydes

chlorosulfinyl acetates (Scheme 49). These reagents were resolved by enzymatic kinetic resolution using porcine pancreatic Scheme 49. SPAC Reaction from 2-(2Benzothiazolylsulfinyl)acetates

readily separated in many cases by lipase PS-catalyzed transesterification in i-Pr2O, affording the (R)-alcohols and the (S)acetates.207 These diastereomerically pure allylic acetates underwent regio- and diastereoselective Pd-catalyzed allylic substitution using sodium dimethylmalonate; the reactivity of these substrates was controlled by the chiral sulfur and the size of the alkyl group at the γ-position, with the (S)-acetates being unreactive in most cases.208 In contrast to the poor diastereoselectivity found for the SPAC reaction of bis-sulfoxides and aldehydes, the reaction between lithiated bis-sulfoxides 270 and cyclic ketones 273 in the presence of an excess of PhSLi as thiophile affords good yields of γ-hydroxy alkenyl sulfoxides 277 and 278 as single diastereomers (Scheme 52). The reaction is believed to proceed by condensation and isomerization to allylic sulfoxide 274, which undergoes sigmatropic rearrangement, as shown in Scheme 52, to minimize nonbonding interactions (p-Tol/cycloalkenyl)

lipase (PPL).204 Addition of P(OMe)3 proved to increase the yield of the reaction slightly and was then tested with a variety of aldehydes to produce the desired products in good yields and with better optical purities than prior examples. This improvement was attributed to a fast elimination of the 2-benzothiazole moiety from the intermediate sulfenate. In related research, Miura et al. developed conditions to promote the vinyl-to-allyl isomerization of sulfinyl unsaturated ketones 268 with DBU in the presence of an excess of PPh3, followed by sigmatropic rearrangement and efficient sulfenate trapping with dilute H2O2. This protocol affords a variety of chiral γ-hydroxy α-enones 269 with good yields and generally high ee values (Scheme 50).205 14215

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 52. SPAC Reaction Using (S,S)-Bis-p-tolylsulfinyl Methane and Cyclic Ketones

acid.215 The nitrile variant of this protocol has been applied in the total synthesis of (+)-allocyathin B2 (Scheme 55) from aldehyde

present in the alternative reaction conformer that involves rearrangement of the other sulfoxide. Cyclopentane-derived alkenyl sulfoxide 277 isomerized spontaneously to allylic sulfoxide 279, which underwent a second sulfoxide−sulfenate rearrangement to produce diacetate 280 with high diastereoselectivity.209 6.3.5. Applications of the SPAC Reaction in Total Synthesis. The SPAC reaction has been used in a number of syntheses of natural products. Scheme 53 summarizes the

Scheme 55. Approaches to (+)-Allocyathin B2 and to an Amphidinol 3 Fragment Using the SPAC Reaction

Scheme 53. SPAC Reaction in the Synthesis of (+)-Brefeldin A

hydroxylative Knoevenagel protocols used by Corey and Carpino210 and Nokami et al.211 in their syntheses of (+)-brefeldin A. Interestingly, seemingly minor changes in the reactants (281−284) and perhaps in the precise reaction conditions result in substantially different diastereoselectivities at the new chiral centers of products 285 and 286. Solladié and Gerber used a classical SPAC condensation to prepare the seco acids of pyrenophorin and patulolide A,212 and Nokami et al. carried out model studies of intramolecular SPAC processes on 287 to build macrolides 288 (Scheme 54). This protocol and a related dimerization approach from 289 resulted in the synthesis of (−)-pyrenophorin.213 The SPAC reaction with arylsulfinylalkanones has been applied to the synthesis of (+)-decarestrictine L214 and to the preparation of (±)-(11E)-13-hydroxy-10-oxooctadec-11-enoic

290, which underwent hydroxylative chain elongation and lactonization in the same step to produce 292 as a single isomer. Interestingly, a model aldehyde lacking the bulky CO2t-Bu fragment produced the (R)-allylic alcohol with dr > 20:1. These strikingly different stereochemical courses can be rationalized in terms of different orientations (perpendicular vs parallel) of the alkene of the allylic sulfoxide fragment relative to the plane of the

Scheme 54. Approaches to (−)-Pyrenophorin Using the SPAC Reaction

14216

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

bicycle and steric interactions modulating the ease of rearrangement of the sulfoxide moiety in 291.216 The ester SPAC protocol between aldehyde 293 and sulfoxide 257 has been used in the synthesis of 294 a precursor to an amphidinol 3 fragment.217

Indeed, the reaction between sulfenyl diene 313 and oxabicyclic dienophile 312 afforded a good yield of a moderately selective mixture of allylic sulfides 314 and 315 with good stereocontrol (Scheme 57). The oxidation of sulfide 314 followed by a facile sulfoxide−sulfenate rearrangement afforded allylic alcohol 316, which was further transformed into (+)-compactin. Interestingly, the sulfoxide derived from regioisomeric sulfide 315 does not rearrange under these conditions.222 Carreño, Garcı ́a-Ruano, and colleagues explored the Diels− Alder reactions of readily available enantiopure 1-sulfinyl-1,3butadienes 317 with N-methylmaleimide (NMM) that take place with complete endo and π-facial selectivities controlled by the sulfoxide both under thermal and catalytic conditions (Scheme 58). Catalysis by SnCl4 allowed for the isolation of allylic sulfoxides 318, and the slow thermal reactions with a large excess of NMM, presumably acting as a thiophile, led to excellent yields of mixtures of 318 and 319 (or just pure 319 in some cases), the expected product of sigmatropic rearrangement. Interestingly, standard thiophiles were not effective in producing the sigmatropic rearrangement, leading instead to either recovery of the starting material or decomposition.223 In contrast, diene 320 with an endocyclic double bond gave rise to cycloadduct 322 in good yield. This finding was rationalized in terms of Diels− Alder cycloaddition to the allylic sulfoxide followed by sigmatropic rearrangement, dehydration, and a further [4 + 2]cycloaddition of diene 321 with excess NMM.224−226 The reaction between sulfinyl diene 323 and maleic anhydride was examined by Carreño, Garcı ́a-Ruano, and colleagues with remarkably different results (Scheme 58). Indeed, under thermal conditions, a slow reaction at room temperature gave rise to lactones 325 and 326. The use of high pressure afforded a 70:30 mixture of the expected cycloadduct 324 and lactone 325. Upon standing in CH2Cl2 at room temperature, diastereomerization at sulfur of 324 was observed, along with the appearance of lactone 326. After 10 days, a 90:10 mixture of lactones was obtained. The enantiomeric purity of lactone 325 was substantially higher at short reaction times and higher temperatures. This observation was explained in terms of a competition between the intermolecular cycloaddition and sulfinyl oxygen acylation prior to cycloaddition to ultimately produce lactone 325 with opposite facial selectivity. 227,228 The behavior of a 3trimethylsilyloxy-1-sulfinyl butadiene has also been examined with mixed results.229 Carreño, Garcı ́a-Ruano, and colleagues examined the Diels− Alder reactions of enantiopure 1-sulfinyl-1,3-pentadiene with monosubstituted ethylenes, which either did not react (ethyl vinyl ether) or led to a complex mixture of regio- and stereoisomers (methyl acrylate).230 The cycloaddition between amino sulfinyl diene 327 and NMM gave rise to a separable mixture of cycloadducts 328 and 329 (Scheme 59) with moderate π-facial diastereoselectivity, presumably due to hindrance by the Boc or Ar groups on nitrogen. Interestingly, pure 329, upon standing in a freezer, spontaneously afforded tricyclic derivative 330 by sigmatropic rearrangement and intramolecular cyclization of the resulting alcohol.231 Sulfinyl-substituted 1,3-butadienes gave rise to interesting cycloadducts in hetero-Diels−Alder processes with a triazolinedione and with a nitrosoformate derivative (Scheme 60). In the case of the triazoline dienophile, at −10 °C, a quick reaction gave just sulfenate 332, and sulfoxide 331 was not detected. Sulfenate 332 was treated with P(OMe)3 to afford alcohol 333 in high yield, or the cycloaddition could be carried out in the presence of the thiophile to produce alcohol 333 (91%, >98% ee).

6.4. [4 + 2]-Cycloadditions and [2,3]-Sigmatropic Rearrangements

6.4.1. [4 + 2]-Cycloadditions with Sulfur-Containing Dienes or Dienophiles and [2,3]-Sigmatropic Rearrangement. The cycloaddition between 1,3-dienes appropriately substituted with sulfide or sulfoxide substituents leads to cycloadducts that are amenable to undergoing a sulfoxide− sulfenate rearrangement. Evans et al. pioneered this strategy by studying the reaction between 1-phenyl sulfinyl butadiene 295 and enamine 296, which gave rise to allylic sulfoxide 297 and ultimately to allylic alcohol 298 with the alkaloid hasubanan skeleton (Scheme 56).218 The Diels−Alder cycloaddition of Scheme 56. [4 + 2]-Cycloaddition and [2,3]-Sigmatropic Rearrangement from 1-Sulfinyl-Substituted 1,3-Dienes

pyrone sulfoxide 299 was examined by Posner et al., who obtained, under high pressure, a good yield of bicyclic sulfinyl lactone 300 as a 16:1 mixture of isomers at sulfur (Scheme 56). Methanolysis of the lactone triggered a sigmatropic rearrangement leading to diol 301 in excellent yield. Moderately selective protection and oxidation set the stage for a second sigmatropic rearrangement to afford triol derivative 303, provided that the pyrolysis was carried out in benzene; in contrast, the use of MeOH favored syn elimination of 302 to the expected cyclohexadiene derivative. This solvent effect is noteworthy. Unfortunately, it was not possible to obtain enantiopure pyrone 299 in quantities of more than a few milligrams.219 In the course of studies directed toward the synthesis of the cytochalasin D isoindolone fragment, Vedejs et al. reported the regio- and stereoselective Diels−Alder reaction between trienyl sulfide 304 and pyrrolinone 305 to produce cycloadduct 306, which was oxidized to the sulfoxide and subjected to sigmatropic rearrangement conditions to afford allylic alcohol 307 (Scheme 57).220 On the other hand, Overman et al. reported the reaction between sulfenyl dienyl carbamate 308 and acrolein, which gave cycloadduct 309, which, in turn, was epimerized with DBU, reduced, and subjected to the oxidation and rearrangement steps to produce diol carbamate 311.221 The Diels−Alder reaction/ sigmatropic rearrangement using sulfur-substituted butadienes was a pivotal sequence in Grieco et al.’s compactin synthesis.222 14217

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 57. [4 + 2]-Cycloaddition and [2,3]-Sigmatropic Rearrangement from Sulfenyl-Substituted 1,3-Dienes

Scheme 58. Diels−Alder Reaction of Enantiopure Sulfinyl Dienes with N-Methylmaleimide and Maleic Anhydride

Scheme 59. Diels−Alder Reaction and Sulfoxide−Sulfenate Rearrangement with Amino Sulfinyl Dienes

Interestingly, upon standing at room temperature, sulfenate 332 gave rise to a single isomer of a dimeric species, probably formed by SN2′ attack of alcohol 333 on sulfenate 332.232 In contrast to the extremely facile rearrangement observed for allylic sulfoxide

331, nitrosoformate adduct 334 was isolated uneventfully as a single isomer, with the enhanced stability of the sulfoxide being attributed to conformational issues that lock the sulfinyl group in an equatorial arrangement. Cycloadduct 334 was transformed smoothly into enantiopure pyrrolidines.233 Oxidation of a variety of sulfides functionalized with enol phosphates 335, obtained by Diels−Alder cycloaddition, delivered the expected sulfoxides 336, which rearranged smoothly to the desired alcohols 337 (Scheme 61).234 In some cases, the allylic sulfoxides were unexpectedly unreactive, the sigmatropic rearrangement failed under standard conditions, and it was advantageous to use the related selenides, which produced 14218

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

339) with high selectivity (Scheme 62). Treatment of the mixture with PhMgBr generated an allylic sulfoxide that underwent sigmatropic rearrangement and trapping with a thiophile to produce aminoalcohol 344 as a single isomer in excellent yield. The sequence of events that took place in this transformation was investigated by heating cis-allylic sulfoxide 340 in CDCl3 at 50 °C and monitoring by 1H NMR spectroscopy to produce a mixture of trans-allylic sulfoxides 343 epimeric at sulfur, presumably by sulfoxide−sulfenate rearrangement to 341 and conformational inversion to sulfenate 342, which rearranged to the mixture of allylic sulfoxides 343.235 It should be pointed out that the complete syn/anti selectivity found in this chemistry requires substitution at C-6 of the dihydrothiazine S-oxides; unsubstituted systems lead to nonselective mixtures of aminoalcohols.236 The methodology was also readily applied in an intramolecular manner, and the geometry of the diene allowed for control of the relative stereochemistry of the products. The intramolecular variant of the methodology was further applied to the synthesis of amino carbohydrates.237,238 Weinreb also developed a related methodology to prepare vicinal diamines with high selectivity.239 The possibility of achieving asymmetric induction in these cycloadditions by using N-sulfinylcarbamates derived from chiral alcohols such as 8-phenylmenthol or a camphor derivative has been addressed.240,241 Whitesell et al. examined the cycloaddition between carbamate 345 and (E,E)-hexa-2,4-diene under thermal conditions (mixture of four products) and in the presence of SnCl4, which led essentially to cycloadduct 346 as a single isomer (Scheme 63). Standard opening of the

Scheme 60. Hetero-Diels−Alder Reactions and Sulfoxide− Sulfenate Rearrangement

Scheme 61. Enol Phosphate Intermediates in Tandem [4 + 2]Cycloadditions and [2,3]-Sigmatropic Rearrangements

Scheme 63. Enantioselective Synthesis of 1,2-Aminoalcohols from Chiral N-Sulfinyl Carbamates

the desired allylic alcohols 337. In several examples, the use of benzene instead of MeOH for the sigmatropic rearrangement afforded an enhanced rearrangement as opposed to syn elimination for certain sulfoxides, as noted by Posner et al.219 6.4.2. Other [4 + 2]-Cycloaddition/[2,3]-Sigmatropic Sequences. Weinreb and co-workers explored the intermolecular hetero-Diels−Alder cycloaddition between 1,3-dienes of known geometry and readily available benzyl N-sulfinylcarbamate to produce the expected 3,6-dihydrothiazine 1-oxides (338,

dihydrothiazine S-oxide with PhMgBr and rearrangement− desulfurization gave rise to the expected aminoalcohol derivative 347. Interestingly, the use of MeMgBr afforded a methyl

Scheme 62. Diastereoselective Synthesis of 1,2-Aminoalcohols from N-Sulfinylcarbamates

14219

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

sulfoxide as a single isomer that did not rearrange under the conditions that promoted the rearrangement of the phenyl analogue.240 Weinreb and colleagues reported a good diastereoselectivity and yield for the reaction between 345 and cyclohexadiene and complete selectivity for a camphor-derived carbamate in the presence of TiCl4 at low temperature.241 The use of chiral Ti(IV) Lewis acids to promote the cycloaddition of cyclohexadiene and achiral N-sulfinyl dienophiles was examined with moderate success.242 In related research, Whitesell et al. examined the asymmetric induction in the ene reaction of N-sulfinylcarbamates derived from enantiopure trans-2-phenylcyclohexanol with a number of alkenes. In most cases, the reaction, in the presence of SnCl4, had the characteristics of a concerted process and proceeded with high levels of diastereoselectivity (95% de). Scheme 64 illustrates

Scheme 65. Hetero-Diels−Alder Reaction and [2,3]Sigmatropic Rearrangement in Syntheses of Natural Products

Scheme 64. Ene Reaction with Kinetic Resolution and [2,3]Sigmatropic Rearrangement

Scheme 66. Sulfinyl Dienes and Allenes in Cycloaddition/ [2,3]-Sigmatropic Rearrangement Tandem Processes a remarkable example of the methodology that entails kinetic resolution of an excess of racemic 349 to produce a 7.3:1 mixture of diastereomers of sulfinamide 350 that are separable by recrystallization. These sulfinamides had to be N-alkylated to allow for successful allyl sulfoxide formation followed by the rearrangement−desulfinylation protocol. In this case, allylic alcohol 351 was obtained in good yield and excellent optical purity.243 The hetero-Diels−Alder cycloaddition of cyclopentadiene with methyl N-sulfinylcarbamate afforded cycloadduct 352 (Scheme 65), which was transformed by reaction with PhMgBr into allylic sulfoxide 353, which, upon [2,3]-sigmatropic rearrangement, gave a mixture of bicyclic oxazolidinone 355 and uncyclized hydroxy carbamate 356 that could be cyclized. Oxazolidinone 355 was transformed into the alkaloid agelastatin A and into an oxazolidinone scaffold library.244,245 The intramolecular variant of this chemistry was also applied in a more complex setting to produce, from 357, cycloadduct 358, which underwent the allyl sulfoxide rearrangement sequence to afford allylic alcohol 359 in high yield and as a single isomer (Scheme 65). Alcohol 359 was transformed into a minor freshwater cyanobacteria hepatotoxin, 7-epicylindrospermopsin, where the total synthesis in this case allowed for a revision of the stereochemistry at one center of the more abundant cylindrospermopsin.246 This chemistry has also been applied to the synthesis of 1,4-benzodiazepin-5-one derivatives.247,248 The Diels−Alder reaction and sigmatropic rearrangement strategy has been used within other structural frameworks. Veselovskii et al. examined the Diels−Alder reactivity of racemic sulfinylmethyl-substituted butadienes 360 with a variety of dienophiles.249 The case of dimethylcyclopropene required high pressure to produce the expected cycloadduct 361 as a 1:1 mixture of diastereomers at sulfur that led to a 7.5:1 mixture of allylic alcohols 362 upon sigmatropic rearrangement under standard conditions (Scheme 66). On the other hand, Padwa et

al. explored the reactivity of sulfinyl allenes 364 and 368 in Diels−Alder and dipolar cycloadditions with nitrones. In both cases, the thermal conditions required for the cycloaddition also brought about the sigmatropic rearrangement to produce sulfenates of either meta aromatic derivatives 366 or isoxazolidines 370. Thus, the allenyl sulfoxides are formal equivalents of propargyl alcohol.250 6.5. Tandem [3,3]-Sigmatropic and [2,3]-Sigmatropic Rearrangements

The sequential use of Claisen and [2,3]-sigmatropic sulfoxide− sulfenate rearrangements constitutes a valuable strategy for acyclic stereocontrol. Aggarwal and Warren explored this protocol by transforming sulfenyl propionate 371 into carboxylic acid 372 and then to trisubstituted allylic alcohol 373 by sulfoxide−sulfenate rearrangement (Scheme 67).251 Heathcock and co-workers took advantage of the Evans aldol protocol followed by Claisen−Ireland rearrangement and a Mislow− Braverman−Evans rearrangement to complete the synthesis of ACRL toxin IIIb,252 as well as the synthesis of myxalamide A by transforming enantiopure propionate 374 into sulfide 375 and, 14220

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

obtained by [2,3]-rearrangement, to bis sulfinyl diene 385, has also been reported.256 Pelc and Zakarian reported that heating , vinyl sulfoxide 386, or its epimer at sulfur, at 150 °C triggered a sequential one-pot Claisen rearrangement to allylic sulfoxide 387 (not isolated) and [2,3]-sigmatropic rearrangement to afford densely functionalized allylic alcohol 388, which was transformed into the imine ring system fragment of pinnatoxins (Scheme 69).257 This one-pot

Scheme 67. Tandem Claisen−Ireland, Oxidation, and [2,3]Sigmatropic Rearrangement

Scheme 69. Tandem Claisen−Mislow−Evans Rearrangement in Syntheses of Natural Products

after chain elongation, carrying out the sulfoxide rearrangement on 376 and protection−deprotection scheme to afford 377.253 Posner et al. reported that the treatment of lithiated primary and secondary propargylic alcohols 378 with sulfinyl chloride 379 followed by heating at 100 °C afforded good yields of keto unsaturated esters 383 (Scheme 68). The process is believed to Scheme 68. Sigmatropic Rearrangements Followed by Allene Sulfoxide−Sulfenate Processes

Claisen−Mislow−Evans rearrangement can be carried out in just 20 min under microwave conditions.258 In contrast, the sulfoxides related to 389 were completely unreactive, but the [3,3]-rearrangement of vinyl sulfide 389 took place smoothly to produce allylic sulfide 390, which, after Wittig olefination and oxidation, gave sulfoxide 391, which rearranged to allylic alcohol 392 and was further transformed into mesembrine.259 Treatment of dienyl diol 393 (Scheme 70) with Cl3CCN afforded a bis-trichloroacetimidate that underwent a double Overman rearrangement to protected diamine 394. Oxidation gave rise to a 1:1 mixture of sulfoxides that rearranged smoothly to allylic alcohol 395 as a 1:1 mixture. Without separation, both isomers were transformed into oxazoline 396, which was subsequently elaborated into (−)-agelastatin A.260 6.6. Michael-Type Addition and [2,3]-Sigmatropic Rearrangement

involve chloride substitution, a [3,3]-sigmatropic rearrangement, an unusual [2,3]-rearrangement of allenic sulfoxide 381, and collapse of enol sulfenate 382.254 The methodology was also extended to lithiated allenic allylic alcohols affording the expected dienes.255 It should be noted that the related sulfenate-to-sulfoxide rearrangement involving an allene, 384,

The base-promoted (LDA, NaH, KH) cyclization of (Z,Z)hydroxy dienyl sulfoxides 397 took place in high yields and diastereoselectivities to produce allylic sulfoxides 398, which were completely stable toward epimerization at sulfur in most 14221

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 70. Approach to (−)-Agelastatin A Using Tandem Overman, Oxidation, and [2,3]-Sigmatropic Rearrangement

Scheme 72. Sulfoxide−Sulfenate Rearrangement/Trapping Sequence with Standard Thiophiles for the Synthesis of Dihydropyrans

cases (Scheme 71). On the other hand, the cyclization of selected examples with phosphazenes in CH2Cl2 gave variable mixtures of Scheme 71. Base-Promoted Cyclization of Hydroxy Dienyl Sulfoxides for the Synthesis of Dihydropyrans

to the syntheses of the cores of ent-dysiherbaine, deoxymalayamicin A,262 and enantiopure ethyl deoxymonate B.263 The marked difference in behavior of these substrates might be related to the reactive conformations of the diastereomeric allylic sulfoxides. Examination of the J2,3 coupling constants for 404 and 402 suggests conformer 407 [(H2/H3)gauche] as the preferred arrangement for 404, whereas 408 [(H2/H3)anti] would be preferred for 402. The latter can change to the alternate chair 409, which differs from 407 in that the bulky tolyl group is away from the pyran moiety in 409 whereas 407 places the tolyl group toward the dihydropyran moiety in an unfavorable scenario. The methodology has been extended to tetrahydropyridines 413 by two approaches (Scheme 73). First, enantiopure amino 1sulfinyl dienes 410, derived from lithiated sulfinyl dienes and sulfinimines, could cyclize to allylic sulfoxides 411, which could undergo the [2,3]-rearrangement. Second, the cyclization of amino 2-sulfinyl dienes 412, similarly to the alcohol analogues, could be a viable alternative.264 The feasibility of the first approach was established by treating sulfinyl diene 414 with K2CO3, which led in a single operation to piperidinol trans-ent415 in good yield. In contrast, the more readily available S diastereomer of 414 was very sluggish and provided mixtures of cis and trans products. Therefore, this approach was abandoned, and a Stille coupling to prepare a variety of amino 2-sulfinyl dienes 412 was developed. Cyclizations of N-Ts derivatives, such as 416, with DBU in toluene produced allylic sulfoxide 417 in good selectivities and yields. N-Boc and N-sulfinyl dienes such as 418 were found to undergo deprotection with trifluoroacetic acid (TFA) and, upon basification, led to N-unprotected allylic sulfoxides such as 419. These allylic sulfoxides were stable under standard conditions, and treatment with Et2NH in toluene proved to be effective in promoting the rearrangement and thiophilic cleavage to produce tetrahydropyridine-3-ol derivatives, such as 415 and 420. In fact, it was possible to carry out the cyclization and [2,3]-sigmatropic rearrangement in just one step (DBU followed by Et2NH) with identical selectivity and comparable overall yield.

trans-398, cis-399, and cis-allylic alcohol 400, presumably derived from 399 by sigmatropic rearrangement and thiophilic cleavage.261 The cyclization is also viable for (E,Z)-hydroxy dienyl sulfoxides 401, albeit with lower yields and different stereochemical outcome at the sulfur-bearing carbon for phenyl(402) and alkyl- (403) substituted substrates. Promoting the rearrangement/trapping sequence with P(OMe)3 on Ph-substituted 404 led to pure alcohol 405 (50%) along with recovered starting material 404 (50%) after 4 days. In contrast, diastereomer 402 afforded an excellent yield of ent-405 in just 12 h (Scheme 72).262 Focusing on 404, the use of Et2NH or Na2S led to variable amounts of cis-alcohol 406 and/or ent-405, suggesting alternative reaction pathways such as epimerization at C-3 of the starting material or ring-opening/ ring-closing sequence, supported by the 66:34 mixture of 406/ ent-405 found when 397 (R1 = R2 = H, R = Ph; Scheme 71) was treated with Na2S. Finally, 1,4-diazabicyclo[2.2.2]octane (DABCO) in toluene, a solvent reported to favor the sigmatropic process instead of syn elimination,219 led to trans-alcohol 405 as a single isomer. The synthetic potential of 3,6-dihydro-2H-pyran3-ol derivatives allowed for the application of this methodology 14222

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

high as 90:10 for piperidine. In contrast, diastereomers with an S allylic center (428) gave optimal results as silyl ethers and in EtOH to afford anti 1,4-diol derivatives (430) with high selectivities. In toluene, an intramolecular hydrogen bond might result in the major chairlike conformation being (1R,2S)-426, which lacks severe 1,3-diaxial interactions. In ethanol, gauche interactions would account for the relative stability of allylic sulfoxides, with (1S,2R)-429 being favored. An alternative to the route from 1-sulfinyl dienes 431 to allylic sulfoxides 432 and eventually 1,4-diol derivatives 434 relied on the addition of alcohols or amines to enantiopure 2-sulfinyl dienes 433 to produce essentially the same intermediate 432 (Scheme 75).266 A variety of sodium alkoxides of primary and

Scheme 73. Tandem Cyclization/[2,3]-Sigmatropic Rearrangement in the Diastereoselective Synthesis of Tetrahydropyridinols

Scheme 75. One-Pot Michael-Type Addition/Sulfoxide− Sulfenate Rearrangement of Amines and Alcohols to 2Sulfinyl Dienes

In an extension of the method to acyclic systems, 1-sulfinyl diene derivatives 421 undergo intermolecular addition of thiophilic amines to produce 422, which, by isomerization to allylic sulfoxides 423 and [2,3]-sigmatropic rearrangement and desulfurization with excess amine, produce unsaturated 1,4-diol derivatives 424 (Scheme 74).265 Optimal conditions for diastereomers with an R allylic center (425) involved the use of toluene, providing anti diols such as 427 with selectivities as Scheme 74. Acyclic 1,4-Diol Derivatives by Intermolecular Michael-Type Addition Followed by Sulfoxide−Sulfenate Rearrangement

secondary benzylic alcohols as well as primary and secondary amines afforded products such as 436 and 437 in good yields and diastereoselectivities. Extending the method to alkyl-substituted dienyl sulfoxides such as 438 gave moderate yields of 439 and 440. Diol 440, formed as a 92:8 mixture of enantiomers (αmethoxyphenyl acetates) when no BnOH was used, might derive from the regio- and stereocontrolled rearrangement of a bisallylic sulfoxide intermediate 441, with chirality transfer controlled solely by the chiral sulfoxide. 6.7. [2,3]-Sigmatropic Rearrangement and Radical Ring Expansion

Renaud and co-workers developed an efficient methodology to achieve a radical two-carbon ring expansion of cyclobutanones, 14223

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

(Scheme 78), which either evolves through head-to-tail dimerization into a bis-silylfuran 455 or is trapped with primary

which entails a cascade of Mislow−Braverman−Evans rearrangement of an allylic sulfoxide 442 and radical fragmentation through sulfenate 443, the minor component in the equilibrium, followed by a cyclization process to afford cyclohexanone 445 (Scheme 76).267 In fact, the allylic sulfoxide was prepared by

Scheme 78. Sulfoxide−Sulfenate Rearrangement in Alkynyl Propargyl Sulfoxides

Scheme 76. Mislow−Braverman−Evans Rearrangement Followed by Radical Ring Expansion

amines (BnNH2) to thiolactam 456. Both structures were determined by X-ray analysis.270 On the other hand, Julia and coworkers developed conditions that allow for the isolation of α,βunsaturated ketones from the transient allenyl sulfenates.271 This rearrangement has been used in approaches to natural products such as laulimalide272 and phoslactomycin B.273 The synthesis of a fragment of (−)-laulimalide, an antitumor marine macrolide, was carried out from propargylic sulfide 460, prepared by treatment of alkynyl zinc bromide 459 with chlorosulfide 458 (Scheme 79). Sulfide 460 was oxidized with m-CPBA and the

addition of vinyl lithium to the cyclobutanone, followed by trapping with PhSCl and sulfenate−sulfoxide rearrangement. This methodology was also extended to a radical oxy-Cope rearrangement that afforded ketone 447 from allylic sulfoxide 446.

7. [2,3]-SIGMATROPIC SULFOXIDE−SULFENATE REARRANGEMENTS INVOLVING PROPARGYL MOIETIES

Scheme 79. Synthesis of the C-13−C-28 Fragment of (−)-Laulimalide by Sulfoxide−Sulfenate Rearrangement

7.1. [2,3]-Sigmatropic Rearrangement from Propargylic Sulfoxides

Interestingly, [2,3]-sigmatropic rearrangements also take place in sulfoxides with a higher degree of unsaturation such as alkynes and allenes.13 In early experiments,268,269 2-naphthyl propargyl sulfoxides 448 were heated at 80 °C to provide allenyl sulfenate 449 through a sulfoxide−sulfenate rearrangement followed by a [3,3]-sigmatropic reordering that involves the naphthyl ring to afford 450 (Scheme 77). Finally, an external nucleophile (solvent) reacts with enone 450 yielding naphthothiophenes 451 (X = SPh, OEt, or OAc) in high yield (87−93%). Similarly, refluxing alkynyl propargyl sulfoxides 452 in hexane triggers a cascade of sigmatropic rearrangements, [2,3] followed by [3,3], to yield (α-ketovinyl)thioketene intermediate 454 transient propargylic sulfoxide was subjected to a Mislow− Braverman−Evans rearrangement to provide α,β-unsaturated ketone 461. Five additional steps including an alkene transposition and Julia−Kocienski olefination gave the C-13−C-28 fragment of (−)-laulimalide.272

Scheme 77. Sulfoxide−Sulfenate Rearrangement in 2Naphthyl Propargyl Sulfoxides

7.2. [2,3]-Sigmatropic Rearrangement from Propargylic Sulfenates

Propargylic sulfenates are suitable starting materials for carrying out sulfenate−sulfoxide rearrangement providing allenyl sulfoxides, as initial experiments by Braverman and Stabinsky revealed.274 Afterward, using optically active starting materials, Stirling and co-workers showed that chiral allenyl sulfoxides, formed from menthyl (S)-p-toluenesulfinate, mutarotate upon standing in acetone solution at room temperature as a consequence of epimerization of the sulfoxide moiety through 14224

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

a sulfoxide−sulfenate rearrangement similar to that of the parent allylic sulfoxides, with the absolute configuration at the allene stereocenter maintained unaltered, as confirmed by oxidation to the allenyl sulfone.275,276 Within this context, Van Kruchten and Okamura designed an interesting study on the chemoselectivity of alkenynols 462, precursors of alkenyne sulfenates 463, which underwent sulfenate−sulfoxide rearrangement at −78 °C across the double bond to give 464 or across the triple bond to provide 465, depending on the substitution of the starting enynol (Scheme 80).277 Further heating in benzene of allylic sulfoxide 464 (R1 =

Scheme 81. Approach to Spiroketals and Indoles from Propargylic Alcohols

Scheme 80. Competitive Sulfenate−Sulfoxide Study Using Alkenynols

underwent ring cleavage by HBr elimination upon refluxing in dichloroethane to generate triene sulfoxide 480 in 61% yield (Scheme 82).288,289 Cyclization with PhSCl, SO2Cl2, or PhSeCl provides comparable yields. In contrast, when the ester group is attached to the allenyl sulfoxide (483), ring closure takes place by nucleophilic assistance of this group, allowing for the synthesis of 2,5-dihydrofuranones 484.290−292 In a number of examples, sulfenate−sulfoxide rearrangement provides allenyl sulfoxides that are transformed into sulfinyl or Scheme 82. Electrophile-Induced Cyclization of Allenyl Sulfoxides R2 = R3 = H) demonstrated the reversibility of the rearrangement, yielding a 1:20 mixture of allylic sulfoxides (E/Z) 464 and allene sulfoxide 465. Accordingly, 464 formed under kinetic control, and allenyl sulfoxide 465 formed under thermodynamic control. The versatile reactivity of allenyl sulfoxides and their easy access from propargylic sulfenates using [2,3]-sigmatropic rearrangement have made them ubiquitous synthetic intermediates that evolve through desulfinylation with n-BuLi,278 desulfinylation by electrophilic addition of iodine to the allene,279 α-deprotonation followed by alkylation,280 γ-deprotonation,271,281 chelotropic addition of SO2,282 and Michael nucleophilic additions to the allene central carbon of amines,283 alcohols,284 sulfides, and dimethyl cuprate.285 In particular, allenyl sulfoxides 468 and 474, prepared from propargylic alcohols 467 and 472, respectively, contain internal nucleophiles and provide simple access to heterocyclic structures such as spiroketals 470286 and indoles 475 (Scheme 81).287 Within the same context, allenyl sulfoxides can undergo intramolecular cyclization induced by electrophiles. In particular, Christov and Ivanov reported that enynol 476 is transformed into allenyl sulfoxide 478 through sulfenate−sulfoxide rearrangement of sulfenate 477 by treatment with PhSCl and Et3N. Further electrophilic addition of bromine to the allene double bond provided an 87% yield of oxathiolium bromide 479, which 14225

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

sulfonyl dienes255,256,293 that have eventually been used in [4 + 2]-cycloadditions.294 Also, allenyl sulfoxides formed by sulfenate−sulfoxide rearrangement have been used in Diels−Alder reactions with cyclopentadiene250 and maleic anhydride.295 Actually, an intramolecular cycloaddition involving allenyl sulfoxide 488, easily formed from enynol 485 and iodide 486 and not isolated, to afford cycloadduct 489 is the key step in the synthesis of sesquiterpene (+)-sterpurene, as nicely illustrated by Gibbs and Okamura (Scheme 83).296

sulfinyl radicals 492 that recombine through the oxygen atom (R1 = H) to give 493, which undergoes consecutive [3,3]sigmatropic rearrangement and intramolecular [2 + 2]-cycloaddition to yield 496, a dithiabicyclo [3.1.1] derivative containing the sulfur core of zwiebelane (>99%; R1 = H, R = Me). Also, for sterically demanding R groups (R = t-Bu), nucleophilic addition of sulfur onto the allene produces 495, which evolves through intramolecular 1,3-dipolar cycloaddition to oxadithiobicyclo [2.2.1] 497, obtained as a 71:29 mixture of diastereomers.297 Alternatively, recombination of 492 through the sulfur atom takes place mainly for R = H, generating a mixture of regioisomeric thiosulfonates 500 and 501 (R1 = TMS, 78:22; R1 = Et, 45:55).298,299 The sulfenate−sulfoxide rearrangement is also useful for transforming [2.2]paracyclophanes (Scheme 85).300 Consequently, bis-propargylic alcohols 502, prepared as single meso diastereomers from bis-aldehydes and lithium acetylides, were subjected to sulfenylation with R2SCl followed by double sigmatropic rearrangement to yield bis-allenyl sulfoxides 503 as a mixture of inseparable diastereomers that was oxidized to 504 with dimethyl dioxirane (DMDO, R2 = CCl3). Interestingly, biso-nitrophenylsulfinyl derivative 503 was more reactive under thermal conditions than the para and meta isomers and underwent a dyotropic rearrangement, induced by the intramolecular interaction of the two pseudogeminal allenyl sulfoxides, to generate bis α,β-unsaturated ketone 506.301,302 Groups other than sulfenates can participate in this type of sigmatropic rearrangement. For example, a series of propargylic aminosulfenates, prepared from propargylic alcohols 507, generated diastereomeric mixtures of racemic allenyl sulfinamides 508 (Scheme 86).303,304 The reaction is faster for dimethylamido sulfenyl chloride (R = Me, 1 h) than for the isopropyl analogue (R = i-Pr, 24−48 h). Marked differences in diastereoselectivity are probably due to the nature of the substituents, because allenyl sulfinamides 508 do not equilibrate upon standing at room temperature or after treatment with piperidine at room temperature (24 h). [2,3]-Sigmatropic rearrangements involving propargylic sulfenates have been used in synthetic approaches to natural products such as prostacyclins,305 steroids,306 and retinoids,307 among others. Within the context of the synthesis of steroids, mestranol has been employed as a starting material to provide a mixture of two diastereomeric allenyl sulfoxides 509 through sulfenate−sulfoxide rearrangement promoted by sulfenylation with PhSCl (Scheme 87). Subsequent transformations have generated valuable intermediates in the synthesis of steroids. Thus, treatment with NaOMe/MeOH and P(OMe)3 as a thiophile installs a hydroxymethyl ketone 510 as a single diastereoisomer at C-17,308,309 through Michael addition followed by a second sulfoxide−sulfenate rearrangement. Similarly, addition of sodium malonate to the allenyl sulfoxide affords a spiro γ-butyrolactone in this position (511).310 Another interesting application directed toward the synthesis of retinoids was developed by de Lera and co-workers.311,312 This approach starts from racemic α-ionone and involves a reversible [2,3]-sigmatropic rearrangement of the E-allylic sulfenate derived from alcohol 512, which promotes E/Z double-bond isomerization at C-9, followed by [2,3]-sigmatropic rearrangement of the propargylic sulfenate to an allenyl sulfoxide 513, which undergoes an irreversible and diastereoselective [1,5]sigmatropic hydrogen shift to finally generate Z vinyl sulfoxide intermediate 514, which is transformed into (9Z)-4,6-retroretinoic acid 515 in a three-step sequence (Scheme 88).

Scheme 83. Enantioselective Synthesis of (+)-Sterpurene through Sulfenate−Sulfoxide Rearrangement

Braverman et al. found that bis-propargylic dialkoxy disulfides 490 are stable in solution only at low temperature (−18 °C), whereas, upon reflux in CHCl3, they undergo a double sulfenate−sulfoxide rearrangement, providing bis-allenyl sulfoxide 491 (Scheme 84), which dissociates into two allenyl Scheme 84. Multiple Rearrangements of Bis-propargylic Sulfenates

14226

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 85. Sulfenate−Sulfoxide Rearrangement in [2.2]Paracyclophanes

Scheme 86. [2,3]-Sigmatropic Rearrangement in Propargylic Amino Sulfenates

Scheme 88. Sulfenate−Sulfoxide Rearrangements in an Approach to Retinoids

Scheme 87. Sulfenate−Sulfoxide Rearrangements in an Approach to Steroids

sulfoxide with t-BuLi/s-BuOH in each isolated compound (E518, Z-518, E-521, and Z-521) confirmed the stereochemical result. Interestingly, because the initial [2,3]-sigmatropic rearrangement is completely diastereospecific with respect to the allene carbon center while providing mixtures of diastereomers at sulfur (517; e.g., 60:40, R = H), it is the absolute configuration of the sulfoxide moiety that determines the stereochemical outcome for the electrocyclization to 518 and the [1,7]-hydrogen shift to 521, leading to Z or E diastereoisomers of the final compounds.

Finally, Okamura and co-workers have extensively studied the reactivity of allenyl sulfoxides, obtained through Mislow− Braverman−Evans rearrangement, in a search for approaches to the synthesis of drimanes, retinoids, and vitamin D derivatives.313−316 In a selected example of their investigations (Scheme 89),317,318 β-ionone was transformed into Z propargylic alcohol 516, which underwent sulfenate−sulfoxide rearrangement upon treatment with PhSCl and Et3N. Allenyl sulfoxide intermediate 517 was not observed because it evolved through two competitive processes. For compounds with small R groups (R = H, Me, Et), the preferred pathway is a stereospecific sixelectron electrocyclization leading to a Z/E mixture of drimatriene derivatives 518. In contrast, [1,7]-sigmatropic hydrogen shift takes place predominantly for substrates with sterically demanding R groups (R = i-Pr, t-Bu), providing a Z/E mixture of tetraenes 521. Further reductive cleavage of the

8. MISLOW−BRAVERMAN−EVANS REARRANGEMENTS IN THE SYNTHESIS OF BIOACTIVE AND NATURAL PRODUCTS This rearrangement has been used very often for the synthesis of prostaglandins, carbohydrates, and steroids, and therefore, these efforts are discussed as separate subsections of the review. Other synthetic applications of the rearrangement are grouped according to the nature of the starting material (sulfenates or sulfoxides). 14227

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Scheme 89. Drimatrienes and Tetraenes by [2,3]-Sigmatropic Rearrangement

Scheme 90. [2,3]-Sigmatropic Rearrangements for the Synthesis of PGE1 and Other Prostanoids

8.1. Synthesis of Prostanoids

The [2,3]-sigmatropic rearrangement involving allylic sulfoxides and allylic sulfenates has been applied extensively to the synthesis of prostaglandins and related products. The pioneering report by Untch and colleagues that described the straightforward double isomerization (geometry of alkene and configuration of C-15 alcohol) set the stage for the use of this rearrangement in the field.319 Treatment of Z allylic alcohol 524 with p-TolSCl gave sulfenate 525, which underwent [2,3]-rearrangement to produce allylic sulfoxide 526 (Scheme 90). Standard treatment with P(OMe)3 affords E allylic alcohol 527, which was transformed into racemic PGE1 methyl ester. The viability of the sequence for a very labile β-hydroxy ketone functionality is noteworthy. Taber reported the regio- and stereoselective cleavage of cyclopropane 528 with PhSK to afford allyl sulfide 529, which, upon oxidation and sigmatropic rearrangement, afforded hydroxy ester 530 (Scheme 90).320 In separate studies, Kondo and co-workers described similar chemistry that entailed the alkylation and decarboxylation of 529 to produce sulfide 531, which led to prostanoid 532.321 This chemistry was extended to the preparation of a 10-oxa analogue322 and to PGF2α.323 Nokami et al. explored the preparation of a sulfide related to 526 by conjugate addition of a lithiated allylic sulfide to the appropriate cyclopentenone, which gave rise to a nonselective 1:1 mixture of sulfides, transformed into a 1:1 mixture of C-15 alcohols related to 527.324 Another study by Taber et al. sought an enantioselective route to the key fused cyclopropanes (related to 528) but with the appropriate functionalization for the synthesis of (−)-PGE2 methyl ester using the sulfoxide rearrangement. This sequence proceeded through an unstable cis-8,12-dialkyl-11-silyloxy analogue of 529, which was retained as the rearrangement proceeded. The authors highlighted the retention of this stereochemical relationship as evidence of the neutrality of the conditions of the process. Epimerization at C-8 was subsequently carried out to access the correct trans stereochemistry of the target.325 In subsequent years, Taber’s group significantly

improved different aspects of the methodology and applied it to the synthesis of a variety of biologically important isoprostanes with cis stereochemistry at C-8 and C-12,326−333 as well as the preparation of the major urinary metabolite of (−)-PGE2, in enantiopure form.334 Grieco et al. reported the preparation of several 14fluoroprostaglandins utilizing the sulfenate−sulfoxide rearrangement of bicyclic alcohol 533 into isomeric 534 (Scheme 91).335 On the other hand, Fuchs and co-workers were able to transform the unwanted enantiomer ent-535, obtained by resolution of rac535, by a sulfoxide−sulfenate rearrangement sequence into monosilylated diol 540, which produced intermediate 536, identical to that obtained from 535, ultimately transformed into Scheme 91. Fluoro Prostaglandins and PGE2 by Mislow− Braverman−Evans Rearrangements

14228

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

(−)-PGE2 (Scheme 91).336 This rearrangement was also applied in the synthesis of carbacyclin analogues by transforming a Z allylic sulfide into an E allylic alcohol with correct stereochemistry,337 as well as in the preparation of prostacyclin analogues with enhanced chemical stability.338,339

isomer led to 546, as a single isomer, the minor isomer led to a 50:50 mixture of allylic sulfoxides 547, also obtained upon mild heating of 546.308 Desilylation of 549 followed by sulfide oxidation and rearrangement of the resulting sulfoxide gave rise to steroidal diol 550.346 The sigmatropic rearrangement of conjugated dienone 551, by standard procedures, gave rise to intermediate 552 if the process was performed in the absence of light and oxygen.347 The sigmatropic rearrangement was an important step in the synthesis of 19-norsteroids.348 Within the context of a study on the structure and stereochemistry of alkylation of metalated allylic sulfones, Trost and Schmuff examined the [2,3]-sigmatropic rearrangement of steroidal allylic alcohol 553, which gave a good yield of substituted allylic sulfoxide 554 (Scheme 94). In contrast, the

8.2. Synthesis of Carbohydrates

The Mislow−Braverman−Evans rearrangement has been fruitfully applied in the synthesis of carbohydrates. Dyong and Schulte reported the efficient transformation of pyranosyl sulfide 541 into allylic alcohol 542, which was transformed into a sibirosamine precursor.340 Danishefsky and colleagues developed the use of labile anomeric sulfoxides within this field. After some model studies,341 this chemistry was used in the synthesis of the esperamicin trisaccharide core.342 An important step was the preparation of differentially protected disulfide 543 (Scheme 92), which underwent smooth single oxidation at the anomeric

Scheme 94. Sulfenate−Sulfoxide Rearrangements in the Synthesis of Steroid Derivatives

Scheme 92. Sulfoxide−Sulfenate Rearrangements in the Synthesis of Carbohydrates

sulfide followed by sigmatropic rearrangement to afford glycal 544 in good yield.343 Related rearrangements of anomeric sulfoxides were also built into the syntheses of allosamidin344 and pyrenolide D.345 8.3. Synthesis of Steroids and Related Products

The propargylic sulfenate-to-allenyl sulfoxide rearrangement led to a separable 60:40 mixture of allenyl sulfoxides 545. The mixture was treated with NaOMe to afford allylic sulfoxides 546 and 547, which underwent smooth rearrangement to allylic alcohol 548 and then transformed into hydrocortisone acetate (Scheme 93). A detailed study of the reaction between the pure isomers of 545 and NaOMe revealed that, whereas the major

epimeric β alcohol gave a low yield of the sulfoxide, attributed to a 1,3-diaxial-like interaction between the angular methyl and the sulfenate intermediate.349 Somewhat related chemistry was also reported by Wang and Reusch350 and Morera and Ortar.351 A Zto-E isomerization of alkene (555, 556) by sulfenate−sulfoxide− sulfenate rearrangement, similar to the prostaglandin case, was an important step in the synthesis of the steroidal skeleton by a transannular Diels−Alder protocol.352 The sulfenate−sulfoxide rearrangement of allylic steroidal alcohol 557, followed by lithiation, alkylation, and sigmatropic rearrangement, gave trisubstituted allylic alcohol 558 as an 8.4:1 mixture of epimers that was ultimately transformed into brassinolide.353 Refluxing steroidal bis-sulfoxide 559 in toluene triggered a syn elimination, followed by subsequent isomerization to an allylic sulfoxide. After sigmatropic rearrangement and thiophilic cleavage, 16-methylene corticosteroid 560 was produced (Scheme 95).354 Treatment of allylic steroidal alcohol 561 with PhSCl and thiophilic cleavage gave an excellent yield of the carbinol epimer 562, which was transformed into an active potential oral contraceptive.355 Zard and co-workers reported an expedient synthesis of α-keto vinyl carbinols from a variety of cyclic and acyclic ketones, including several steroidal examples (563). The method entails enolate formation followed by Michael addition to alkynyl or allenyl sulfoxides and sigmatropic rearrangement with PPh3 as a thiophile to produce synthetically

Scheme 93. Mislow−Braverman−Evans Rearrangement in the Approach to Steroids and Related Products

14229

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Inversion of the allylic alcohol of advanced intermediate 570 using Mitsunobu conditions or oxidation−reduction failed. In contrast, sulfenate rearrangement afforded allylic sulfoxide 571, which underwent olefination and concurrent Mislow−Braverman−Evans rearrangement in a single step to produce 572, which was further transformed into the indoxamycins (Scheme 97).363 Sulfenate rearrangement of allylic alcohol 573 gave

Scheme 95. Syntheses of Steroidal Compounds Involving [2,3]-Sigmatropic Rearrangements

Scheme 97. Sulfenate−Sulfoxide Rearrangement Approach to Indoxamycins and Stemofoline Alkaloids

versatile α-ketovinyl carbinols 565.356 This approach nicely complements a previous method that relied on the radical addition of xanthates to ethyl vinyl sulfide, followed by elimination of the xanthate to afford mixtures of allyl and vinyl sulfides that rearranged to the desired keto vinyl carbinols upon oxidation and treatment with PPh3.357 8.4. Synthesis of Natural Products by Sulfenate−Sulfoxide Rearrangement

This section covers the syntheses of different products that commence with an allylic alcohol as the starting material of the [2,3]-rearrangement. Hua et al. studied the sulfenate-to-sulfoxide rearrangement of a variety of cyclohexenols, and the resulting substituted cyclohexenyl allylic sulfoxides were key starting materials in the synthesis of (+)- and (−)-12,13-epoxytrichothec-9-ene.358 Inversion of alcohol 566 by a double sigmatropic rearrangement was the last step of the synthesis of (±)-14deoxyisoamijiol 567 (Scheme 96).359 Inversion of a tricyclic vinyl

sulfoxide 574, which, upon lithiation and trapping with TMSCl, afforded a separable E/Z mixture of cage derivatives 575 that were further elaborated into the core of the stemofoline alkaloids.364

Scheme 96. Sulfenate−Sulfoxide Rearrangement Approach to (±)-14-Deoxyisoamijiol and Rhizoxins

8.5. Synthesis of Natural Products by Sulfoxide−Sulfenate Rearrangement

Masaki and colleagues developed a straightforward methodology to selectively functionalize the isopropylidene terminus of acyclic monoterpenes by addition of PhSCl, dehydrohalogenation, and sigmatropic rearrangement or an alternative protocol for allylic oxidation and applied the methodology to the synthesis of a variety of terpenes365,366 and isoprenoid quinones.367 A key step of the synthesis of vernolepin by Isobe and co-workers was the transformation of bicyclic allylic sulfide 576 by Mislow− Braverman−Evans rearrangement into allylic alcohol 577 (Scheme 98).368 The use of a p-MeOC6H4 residue on sulfur proved to be useful in model studies to minimize syn elimination of the sulfoxide intermediate and concurrent isomerization of the resulting diene.369 Sulfoxide−sulfenate rearrangement was also an important step in Vandewalle and co-workers’ synthesis of vernolepin370 as well as in Schultz and Godfrey’s model studies on this target.371 The alkylation of an allylic 2-pyridyl sulfide, followed by oxidation and rearrangement, was used for the synthesis of a pheromone.372 The 10-membered ring diterpene obscuronatin 579 was synthesized by oxidation and rearrangement of allylic sulfide 578 (Scheme 98).373 The synthesis of the

carbinol by a sulfenate sigmatropic rearrangement was an important step in Boeckman et al.’s synthesis of pleuromutilin.360 The rearrangement of alcohol 568 afforded trisubstituted allylic sulfoxide 569 with low E/Z selectivity that was significantly improved by heating the mixture at 65 °C. Oxidation with oxone afforded the related sulfone (60% overall), which was used in the synthesis of a rhizoxin fragment.361 A rearrangement and further oxidation to sulfone sequence was also used in a synthesis of (−)-erythrodiene.362 14230

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

oxone, which allowed for application of the allylic sulfoxide rearrangement in the presence of several alkenes.379 Treatment of selenide 584, an advanced intermediate in Vedejs and Wittenberger’s approach to cytochalasin, with m-CPBA resulted in oxidation at selenium and sulfur. After some warming, selenoxide elimination gave rise to Z enone 585 (66%), along with 20% of the E isomer, both as single isomers at sulfur. The sigmatropic rearrangement of 585 required rather harsh conditions (110 °C) to produce allylic alcohol 586 (Scheme 99).380 The last steps of the approach to bruceantin (Scheme 100) entailed an interesting preparation of allylic sulfoxide 588 by

Scheme 98. Mislow−Braverman−Evans Rearrangement in the Synthesis of Vernolepin and Obscuronatin

Scheme 100. Approach to Bruceantin by Sulfoxide−Sulfenate Rearrangement α-methylene-γ-lactones present in natural products by acidcatalyzed sigmatropic rearrangement, acetate migration, and lactonization was described.374 This rearrangement has also been utilized in the field of βlactam antibiotics. Thermolysis of allylic sulfoxide 580 (Scheme 99) triggered the [2,3]-sigmatropic rearrangement followed by Scheme 99. Penems, Ophiobolane, and Desmethylcytochalasin C Approaches Involving [2,3]Sigmatropic Rearrangements

dianion formation from 587 and acetylide capture to an alkynyl sulfoxide that undergoes cyclization to cyclohexenone sulfoxide 588 in excellent yield. Because the sigmatropic rearrangement of enone 588 failed to provide the desired alcohol cleanly, the enone was reduced to allylic alcohol 589 as a separable 3:2 mixture of sulfoxide epimers. Whereas the major isomer gave the desired trans diol (not shown) under mild conditions [P(OEt)3, EtOH, 40 °C], the minor diastereomer required heating to 80 °C to produce imino ether 590, which was also obtained under these conditions from the major diastereoisomer. The remarkable difference in reactivity observed at 40 °C between the sulfur diastereomers was rationalized by examining the preferred conformation for the rearrangement.381 The condensation between metalated allylic sulfides and suitable electrophiles, followed by oxidation to the sulfoxide and [2,3]-sigmatropic rearrangement, has been used in the synthesis of a fragment of rapamycin,382 and of 1-deoxycastanospermine and stereoisomers.383 Isomerization of a bicyclic vinyl sulfoxide to the related allylic isomer and concurrent rearrangement with aqueous pyridine was used in the preparation of intermediates for the synthesis of oxygenated elemanoids.384 Conjugate addition of lithiated allyl phenyl sulfide to 3,3-dimethylacrylonitrile afforded an allylic sulfide that, after oxidation and sigmatropic rearrangement, produced an allylic alcohol that was transformed into a chrysanthemic acid precursor.385 The oxidation and [2,3]sigmatropic rearrangement of imidazolyl allylic sulfides was used in several approaches to the synthesis of a fragment of the maytansines.386−389 Treatment of diene 591 with PhSCl or PhSeCl afforded densely functionalized intermediates 592, which, after oxidation, smoothly rearranged to allylic alcohol 593 under standard conditions (Scheme 101). The original approach followed by Crimmins et al. carried forward sulfide 592 to an advanced

intramolecular trapping to produce 581, which was transformed into penems.375 Related chemistry was reported by Ross and coworkers376 and by Yanagisawa and Ando.377 Oxidation of allylic sulfide 582 followed by rearrangement produced allylic alcohol 583, which was further elaborated into the ophiobolane ring system.378 In some cases, sulfide oxidation with m-CPBA led to undesired olefin epoxidation, as in the synthesis of petiodial; this issue could be solved by using tetrabutylammonium (TBA) 14231

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

The rearrangement of an allylic sulfoxide derived from quinic acid afforded a suitable precursor for the carbasugars gabosines.394 Similarly, rearrangement of a genipin derived sulfoxide was used to prepare the iridoid 5,6-dihydrovaltrate.395 Myers and co-workers developed a short and enantioselective approach to a variety of 6-deoxytetracycline antibiotics. From a common intermediate, it was possible to access precursor 600 (Scheme 103) and the related 5-deoxy analogue. The

Scheme 101. Synthetic Approach to (+)-Milbemycin Using [2,3]-Sigmatropic Rearrangement

Scheme 103. Mislow−Braverman−Evans Rearrangement in the Syntheses of Doxycycline, Conduritol B, Cyclophellitol, and Brefeldin A

intermediate that, upon oxidation and treatment with P(OMe)3, unfortunately afforded the diene of syn elimination, presumably by epimerization at either C-2 or C-3. The revised synthetic approach produced allylic alcohol 593 to eventually achieve the total synthesis of (+)-milbemycin D.390 Allylic dienyl sulfide 594 (Scheme 102), obtained from diene Fe(CO)3 complex chemistry, was smoothly oxidized to the Scheme 102. Synthetic Approaches to (−)-Halicholactone and Guanacastepene A Using [2,3]-Sigmatropic Rearrangements

preparation of 600 required a highly diastereoselective oxidation of 599 with a chiral reagent to produce the sulfoxide essentially as a single isomer at sulfur, which rearranged smoothly under standard conditions, unlike the other sulfoxide diastereomer. Intermediate 600 was further elaborated into (−)-doxycycline.396 Raghavan et al. explored the application of enantiopure allylic sulfides 601 (Scheme 103), prepared by ring-closing metathesis, to the synthesis of cyclitols such as conduritol B and cyclophellitol by a Mislow−Braverman−Evans rearrangement using 2-mercapto-1-methyl imidazole as a thiophile.397 The rearrangement of cyclopentenyl sulfide 603 under similar conditions produced allylic alcohol 604 in good yield, which was further elaborated to achieve a total synthesis of brefeldin A.398

9. CONCLUSIONS This review has attempted to provide a comprehensive summary of the [2,3]-sigmatropic rearrangement of allylic sulfoxides to allylic sulfenates with particular emphasis on the use of this transformation in synthetic organic chemistry. The process typically takes place under mild conditions and with high stereoselectivity to produce stereodefined allylic alcohols after treatment with a thiophile. This reversible rearrangement can be accessed either from sulfoxides or from allylic alcohols,

sulfoxide, and the ensuing sigmatropic rearrangement was effective in breaking the conjugation of the diene to afford bisallylic alcohol 595, which was eventually transformed into (−)-halicholactone.391 Oxidation of vinyl sulfide 596 produced sulfoxide 597, which was treated with DBU in EtCN at reflux to afford allylic alcohol 598, which was further elaborated to a fragment of guanacastepene A.392 A double Mislow−Braverman−Evans rearrangement of a bis-sulfide was one of the early steps of Wipf and Reeves’s approach to leucascandrolide A.393 14232

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

previously transformed into sulfenates, which is advantageous for synthetic purposes. Moreover, the synthetic versatility of sulfoxides and allylic alcohols has allowed for the design of numerous tandem or sequential processes in which this rearrangement participates. Although many aspects of this rearrangement are well established, it still attracts considerable interest particularly from a synthetic standpoint. In this regard, we believe that there is still considerable potential in the interaction between this rearrangement and organometallic chemistry as well as in additional developments of tandem methodologies involving acyclic scenarios.

formations, as well as pursuing straightforward applications of these methods to the synthesis of bioactive products. Alma Viso was born in Madrid, Spain, in 1964. She obtained a B.S. degree in Chemistry in 1987 and a Ph.D. degree in 1992 from Universidad Complutense de Madrid (UCM). In 1992, she moved to Massachusetts Institute of Technology (MIT) as a Fulbright fellow to work with Stephen L. Buchwald for 18 months. She joined the faculty at UCM (Madrid) in 1993 as an Assistant Professor and was promoted to Associate Professor in May 2002. In December 2002, she joined Instituto de Quı ́mica Orgánica General, CSIC, as a Staff Researcher (Cientı ́fico Titular), and she was promoted to Senior Staff Researcher (Investigador Cientı ́fico) in 2007. Her current research interests are focused on two main areas: the development of new methodologies in asymmetric synthesis using sulfoxides and sulfinamides and the application of these novel methods to efficient syntheses of therapeutically valuable products.

AUTHOR INFORMATION Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

ACKNOWLEDGMENTS This research was supported by MINECO (Spain) (CTQ201346459-C2-02-P, CTQ2016-77555-C2-2-R, CTQ2016-81797REDC). We are grateful to Dr. Carmen de la Torre (IQOGCSIC) and Dr. Miguel Á ngel Sierra (UCM) for encouragement and support.

Ignacio Colomer: 0000-0001-5542-7034 Marina Velado: 0000-0002-0683-591X Roberto Fernández de la Pradilla: 0000-0002-6633-8499 Alma Viso: 0000-0003-2622-4777 Notes

The authors declare no competing financial interest. Biographies

REFERENCES

Ignacio Colomer was born in Jerez de los Caballeros, Spain, and graduated with honors from University of Extremadura in 2007. He then pursued doctoral studies at the National Research Council (CSIC), mentored by Prof. Fernández de la Pradilla, working on acyclic stereocontrol on vinyl sulfoxides. In 2010, he moved to UC Berkeley, where he explored enantioselective gold catalysis under the guidance of Prof. Dean Toste. After earning his Ph.D. degree in 2012 from Universidad Complutense de Madrid (UCM), he was awarded a Marie Curie Fellowship, for work on developing new metal-free oxidation processes under the supervision of Professors Timothy Donohoe and Richard Compton at the University of Oxford from 2014 to 2016. He subsequently took a postdoctoral position with Prof. Stephen Fletcher at Oxford, where he is currently exploring new autocatalytic methods with potential interest for the origin of life.

(1) Hoffmann, R. W. Stereochemistry of [2,3]Sigmatropic Rearrangements. Angew. Chem., Int. Ed. Engl. 1979, 18, 563−572. (2) Hill, R. K. Chirality Transfer via Sigmatropic Rearrangements. In Asymmetric Synthesis; Morrison, J. D., Ed.; Academic Press: New York, 1984; Vol. 3, Chapter 8, pp 503−572. (3) Braverman, S. Rearrangements Involving Sulfoxides. In The Chemistry of Sulphones and Sulphoxides; Patai, S.; Rappoport, Z.; Stirling, C. J. M., Eds.; John Wiley & Sons: Chichester, U.K., 1988; Chapter 14, pp 717−757. (4) Braverman, S. Rearrangements. In The Chemistry of Sulphinic Acids, Esters and Their Derivatives; Patai, S., Ed.; John Wiley & Sons: Chichester, U.K., 1990; Chapter 11, pp 297−349. (5) Braverman, S. Rearrangements Involving Sulfenic Acids and Their Derivatives. In The Chemistry of Sulphenic Acids and Their Derivatives; Patai, S., Ed.; John Wiley & Sons: Chichester, U.K., 1990; Chapter 8, pp 311−359. (6) Brückner, R. 2,3-Sigmatropic Rearrangements. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon Press: New York, 1991; Vol. 6, Chapter 4.6, pp 873−908. (7) Ritter, K. In Methods of Organic Chemistry (Houben−Weyl), 4th ed.; Helmchen, G., Hoffmann, R. W., Mulzer, J., Schaumann, E., Eds.; Thieme: Stuttgart, Germany, 1996; Vol. E21, pp 4971−4995. (8) Ritter, K. In Methods of Organic Chemistry (Houben−Weyl), 4th ed.; Helmchen, G., Hoffmann, R. W., Mulzer, J., Schaumann, E., Eds.; Thieme: Stuttgart, Germany, 1996; Vol. E21, pp 5069−5072. (9) Mikołajczyk, M.; Drabowicz, J.; Kiełbasiński, P. Chiral Sulfur Reagents: Applications in Asymmetric and Stereoselective Synthesis; CRC Press: Boca Raton, FL, 1997. (10) Prilezhaeva, E. N. Rearrangements of Sulfoxides and Sulfones in the Total Synthesis of Natural Compounds. Russ. Chem. Rev. 2001, 70, 897−920. (11) Carretero, J. C.; Gómez Arrayás, R. In Science of Synthesis; Molander, G. A., Ed.; Thieme: Stuttgart, Germany, 2007; Vol. 33, pp 55−100. (12) Reggelin, M. [2,3]-Sigmatropic Rearrangements of Allylic Sulfur Compounds. Top. Curr. Chem. 2007, 275, 1−65. (13) Braverman, S.; Cherkinsky, M. [2,3]Sigmatropic Rearrangements of Propargylic and Allenic Systems. Top. Curr. Chem. 2007, 275, 67− 101.

Marina Velado was born in Madrid, Spain, in 1977. She obtained her Bachelor’s degree in Chemistry in 2001 from Universidad Autónoma de Madrid (UAM). In 2008, she joined Instituto de Quı ́mica Orgánica General (IQOG-CSIC) as a Research Assistant, within the department of Synthesis, Structure and Properties of Organic Compounds (SEPCO). She currently works with Dr. Roberto Fernández de la Pradilla and Dr. Alma Viso (Asymmetric Synthesis with Sulfoxides). Roberto Fernández de la Pradilla was born in Logroño, Spain, in 1958. He received his B.S. degree in Chemistry from Universidad Complutense de Madrid (UCM) in 1980 and a Ph.D. degree from The University of Michigan in 1985 under the supervision of Joseph P. Marino. After postdoctoral work at UCM (Madrid), he became Assistant Professor there in 1987. In 1988, he joined Instituto de Quı ́mica Orgánica General, CSIC, as a Staff Researcher (Cientı ́fico Titular), and he was promoted to Senior Staff Researcher (Investigador Cientı ́fico) in 2000 and to Research Professor in 2007. In 1992−1993, Dr. Fernández de la Pradilla was a Visiting Scientist with Rick L. Danheiser at Massachusetts Institute of Technology (MIT) for 18 months. His current scientific interests involve the development of novel synthetic strategies that rely on chiral sulfoxides with a particular emphasis on strategies that allow for multiple asymmetric trans14233

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

(14) Rayner, D. R.; Miller, E. G.; Bickart, P.; Gordon, A. J.; Mislow, K. Mechanisms of Thermal Racemization of Sulfoxides. J. Am. Chem. Soc. 1966, 88, 3138−3139. (15) Miller, E. G.; Rayner, D. R.; Mislow, K. Thermal Rearrangement of Sulfenates to Sulfoxides. J. Am. Chem. Soc. 1966, 88, 3139−3140. (16) Braverman, S.; Stabinsky, Y. The Rearrangement of Allylic Trichloromethanesulphenates. Chem. Commun. 1967, 270−271. (17) Bickart, P.; Carson, F. W.; Jacobus, J.; Miller, E. G.; Mislow, K. The Thermal Racemization of Allylic Sulfoxides and the Interconversion of Allylic Sulfoxides and Sulfenates. Mechanism and Stereochemistry. J. Am. Chem. Soc. 1968, 90, 4869−4876. (18) Tang, R.; Mislow, K. Rates and Equilibria in the Interconversion of Allylic Sulfoxides and Sulfenates. J. Am. Chem. Soc. 1970, 92, 2100− 2104. (19) Rautenstrauch, V. Stereoselectivity in the Ally1 SulphenateSulfoxide Rearrangement. J. Chem. Soc. D 1970, 526−527. (20) Abbott, D. J.; Stirling, C. J. M. Intramolecular Reactions. Part VIII. Interception of Sulphenate Esters formed in the Rearrangement of Sulfoxides. J. Chem. Soc. C 1969, 818−821. (21) Evans, D. A.; Andrews, G. C.; Sims, C. L. Reversible 1,3 Transposition of Sulfoxide and Alcohol Functions. Potential Synthetic Utility. J. Am. Chem. Soc. 1971, 93, 4956−4957. (22) Evans, D. A.; Andrews, G. C.; Fujimoto, T. T.; Wells, D. The Application of Allylic Sulfoxide Anions as Vinyl Anion Equivalents. A General Synthesis of Allylic Alcohols. Tetrahedron Lett. 1973, 14, 1385− 1388. (23) Evans, D. A.; Andrews, G. C.; Fujimoto, T. T.; Wells, D. Stereoselective Synthesis of Trisubstituted Olefins. Tetrahedron Lett. 1973, 14, 1389−1392. (24) Grieco, P. A. Stereospecific Trisubstituted Olefin Synthesis via Interconversion of Allylic Sulfoxides and Sulphenates. J. Chem. Soc., Chem. Commun. 1972, 702a. (25) Grieco, P. A.; Finkelhor, R. S. A Stereospecific Synthesis of (±)-(E)-Nuciferol via the [2,3]-Sigmatropic Rearrangement of Allylic Sulfoxides. J. Org. Chem. 1973, 38, 2245−2247. (26) Evans, D. A.; Crawford, T. C.; Fujimoto, T. T.; Thomas, R. C. A General Synthesis of 1-Alkyl-1-cyclopentene-cis-3,5-diols. Useful Intermediates in Prostaglandin Synthesis. J. Org. Chem. 1974, 39, 3176−3178. (27) Evans, D. A.; Andrews, G. C. Nucleophilic Cleavage of Allylic Sulfenate Esters. Mechanistic Observations. J. Am. Chem. Soc. 1972, 94, 3672−3674. (28) Evans, D. A.; Andrews, G. C. Allylic Sulfoxides: Useful Intermediates in Organic Synthesis. Acc. Chem. Res. 1974, 7, 147−155. (29) Jones, D. N.; Blenkinsopp, J.; Edmonds, A. C. F.; Helmy, E.; Taylor, R. J. K. Steroidal Sulphur Compounds. Part IX. 6α- and 6βMethylsulphinylcholest-4-ene and Related Compounds; Stereochemical Aspects of the Ally1 Sulfoxide−Sulphenate Rearrangement. J. Chem. Soc., Perkin Trans. 1 1973, 2602−2613. (30) Jones, D. N.; Blenkinsopp, J.; Edmonds, A. C. F.; Helmy, E.; Taylor, R. J. K. Steroidal Sulphur Compounds. Part X. Chiroptical Properties of Steroidal Ally1 Sulfoxides. J. Chem. Soc., Perkin Trans. 1 1974, 937−941. (31) Hoffmann, R. W.; Maak, N. Allylalkohole aus α,β-Ungesättigten Sulfoxiden. Tetrahedron Lett. 1976, 17, 2237−2240. (32) Goldmann, S.; Hoffmann, R. W.; Maak, N.; Geueke, K.-J. Stereochemie der [2,3]sigmatropen (Um)lagerung 3-Substituierter 2Alkenylsulfoxide. Chem. Ber. 1980, 113, 831−844. (33) Hoffmann, R. W.; Goldmann, S.; Maak, N.; Gerlach, R.; Frickel, F.; Steinbach, G. C-C-Verknüpfende Synthese von Allylalkoholen aus Aldehyden und Ketonen. Chem. Ber. 1980, 113, 819−830. (34) Hoffmann, R. W.; Goldmann, S.; Gerlach, R.; Maak, N. Stereochemie der [2,3]Sigmatropen Umlagerung von (1-Cycloalkenylmethyl)-sulfoxiden. Chem. Ber. 1980, 113, 845−855. (35) Hoffmann, R. W.; Gerlach, R.; Goldmann, S. Stereochemistry of a [2,3]Sigmatropic Allyl-Sulfoxide/Allyl Sulfenate Rearrangement. Tetrahedron Lett. 1978, 19, 2599−2602.

(36) Hoffmann, R. W.; Gerlach, R.; Goldmann, S. Konformationseinflüsse auf die Stereochemie der Allylsulfoxid-Sulfensäure-AllylesterUmlagerung. Chem. Ber. 1980, 113, 856−863. (37) Reich, H. J.; Yelm, K. E.; Wollowitz, S. Kinetics, Thermodynamics, and Stereochemistry of the Allyl Sulfoxide−Sulfenate and Selenoxide-Selenenate [2,3] Sigmatropic Rearrangements. J. Am. Chem. Soc. 1983, 105, 2503−2504. (38) Jursic, B. S. Ab Initio Study of [2,3] Sigmatropic Rearrangement of Allylic Nitroxides and Sulfoxides. J. Mol. Struct.: THEOCHEM 1995, 358, 131−138. (39) Jones-Hertzog, D. K.; Jorgensen, W. L. Elucidation of Transition Structures and Solvent Effects for the Mislow−Evans Rearrangement of Allylic Sulfoxides. J. Am. Chem. Soc. 1995, 117, 9077−9078. (40) Jones-Hertzog, D. K.; Jorgensen, W. L. Regioselective Synthesis of Allylic Alcohols Using the Mislow−Evans Rearrangement: A Theoretical Rationalization. J. Org. Chem. 1995, 60, 6682−6683. (41) Freeman, F.; Bathala, R. M.; Cavillo, J. E.; Huang, A. C.; Jackson, T. K.; Lopez-Mercado, A. Z.; Phung, S.; Suh, J.; Valencia, D. O., Jr. [2,3]Sigmatropic Rearrangements of Hydrogen and Alkyl 3-Propenyl Sulfoxides: A Computational Study. Int. J. Quantum Chem. 2006, 106, 2390−2397. (42) Amaudrut, J.; Pasto, D. J.; Wiest, O. Theoretical Studies of the Sulfenate−Sulfoxide Rearrangement. J. Org. Chem. 1998, 63, 6061− 6064. (43) Amaudrut, J.; Wiest, O. The Thermal Sulfenate−Sulfoxide Rearrangement: A Radical Pair Mechanism. J. Am. Chem. Soc. 2000, 122, 3367−3374. (44) Zborowski, K. K.; Alkorta, I.; Elguero, J. A Theoretical Study of the Sulfenate−Sulfoxide Rearrangement. Effect of the Hydrogen Bond Complexation. Int. J. Quantum Chem. 2010, 110, 2391−2397. (45) Kleimark, J.; Prestat, G.; Poli, G.; Norrby, P.-O. PalladiumCatalyzed Allylic Sulfinylation and the Mislow−Braverman−Evans Rearrangement. Chem. - Eur. J. 2011, 17, 13963−13965. (46) Bailly, L.; Petit, E.; Maeno, M.; Shibata, N.; Trapp, O.; Cardinael, P.; Chataigner, I.; Cahard, D. Enantiomerization of Allylic Trifluoromethyl Sulfoxides Studied by HPLC Analysis and DFT Calculations. Chirality 2016, 28, 136−142. (47) Gogoi, P.; Kalita, M.; Bhattacharjee, T.; Barman, P. Copper− Schiff Base Complex Catalyzed Oxidation of Sulfides with Hydrogen Peroxide. Tetrahedron Lett. 2014, 55, 1028−1030. (48) O’Mahony, G. E.; Eccles, K. S.; Morrison, R. E.; Ford, A.; Lawrence, S. E.; Maguire, A. R. Investigation of Steric and Electronic Effects in the Copper-Catalysed Asymmetric Oxidation of Sulfides. Tetrahedron 2013, 69, 10168−10184. (49) Boruah, J. J.; Das, S. P.; Ankireddy, S. R.; Gogoi, S. R.; Islam, N. S. Merrifield Resin Supported Peroxomolybdenum(VI) Compounds: Recoverable Heterogeneous Catalysts for the Efficient, Selective and Mild Oxidation of Organic Sulfides with H2O2. Green Chem. 2013, 15, 2944−2959. (50) Imada, Y.; Takagishi, M.; Komiya, N.; Naota, T. Oxidation of Sulfides with Hydrogen Peroxide Catalyzed by Vitamin B2 Derivatives. Synth. Commun. 2013, 43, 3064−3071. (51) Voutyritsa, E.; Triandafillidi, I.; Kokotos, C. G. Green Organocatalytic Oxidation of Sulfides to Sulfoxides and Sulfones. Synthesis 2017, 49, 917−924. (52) Pace, V.; Castoldi, L.; Holzer, W. Highly Chemoselective Synthesis of Aryl Allylic Sulfoxides through Calcium Hypobromite Oxidation of Aryl Allylic Sulfides. Tetrahedron Lett. 2012, 53, 967−972. (53) Fu, Z.; Wang, M.; Ma, Y.; Liu, Q.; Liu, J. Synthesis of Functionalized Allylic Sulfoxides and Their Use in the Construction of 2,3,4-Trisubstituted Furans via a [3 + 2] Annulation. J. Org. Chem. 2008, 73, 7625−7630. (54) Hunter, R.; Kaschula, C.; Stellenboom, N.; Cotton, J.; Parker, M. I. New Excursions into the Synthesis and Medicinal Chemistry of the Disulfide Bond. Phosphorus, Sulfur Silicon Relat. Elem. 2013, 188, 1497− 1507. (55) Ohta, H.; Okamoto, Y.; Tsuchihashi, G. Asymmetric Synthesis of Chiral Sulfoxides via Microbial Oxidation of Sulfides. Chem. Lett. 1984, 13, 205−208. 14234

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

(56) Hibi, M.; Kawashima, T.; Yajima, H.; Smirnov, S. V.; Kodera, T.; Sugiyama, M.; Shimizu, S.; Yokozeki, K.; Ogawa, J. Enzymatic Synthesis of Chiral Amino Acid Sulfoxides by Fe(II)/α-Ketoglutarate-Dependent Dioxygenase. Tetrahedron: Asymmetry 2013, 24, 990−994. (57) Swindell, C. S.; Blase, F. R.; Carroll, P. J. Evaluation of a Chiral Aryl Auxiliary Design for Sulfur: Construction of Auxiliary-Modified Reagents and Stereoselection in Sulfoxide Formation. Tetrahedron Lett. 1990, 31, 5405−5408. (58) Annunziata, R.; Cinquini, M.; Cozzi, F.; Farina, S.; Montanari, V. Synthesis of Configurationally Stable Allylic Sulfoxides via Diastereoselective Oxidation. Tetrahedron 1987, 43, 1013−1018. (59) Goodridge, R. J.; Hambley, T. W.; Haynes, R. K.; Ridley, D. D. Preparation of Stable, Camphor-Derived, Optically Active Allyl and Alkyl Sulfoxides and Thermal Epimerization of the Allyl Sulfoxides. J. Org. Chem. 1988, 53, 2881−2889. (60) Binns, M. R.; Goodridge, R. J.; Haynes, R. K.; Ridley, D. D. Preparation of Stable, Camphor-Derived, Optically Active Allylic Sulfoxides. Tetrahedron Lett. 1985, 26, 6381−6384. (61) Alajarín, M.; Pastor, A.; Cabrera, J. Arenesulfinamides as New Reagents for the Synthesis of Allylic Sulfoxides. Synlett 2004, 995−998. (62) Veselovskii, V. V.; Dragan, V. A.; Moiseenkov, A. M. Synthesis of E/Z-Sesquilavandulols from the Product of the Ene Reaction of Lavandulyl Acetate with Phenylsulfinyl Chloride. Izv. Akad. Nauk SSSR, Ser. Khim. 1990, 1906−1908. (63) Snider, B. B. Synthesis of Allylic Sulfoxides from Alkenes by EtAlCl2-Catalyzed Ene Reaction with p-Toluenesulfinyl Chloride. J. Org. Chem. 1981, 46, 3155−3156. (64) Ma, S.; Wei, Q. Hydrohalogenation Reaction of 1,2-Allenic Sulfoxides with AlX3 and H2O: Efficient Synthesis of 2-Haloallyl Sulfoxides. Eur. J. Org. Chem. 2000, 1939−1943. (65) Barluenga, J.; Fañanás-Mastral, M.; Aznar, F. A New Synthesis of Allyl Sulfoxides via Nucleophilic Addition of Sulfinyl Carbanions to Group 6 Fischer Carbene Complexes. Org. Lett. 2005, 7, 1235−1237. (66) Maitro, G.; Prestat, G.; Madec, D.; Poli, G. Preparation of Allyl Sulfoxides by Palladium-Catalyzed Allylic Alkylation of Sulfenate Anions. J. Org. Chem. 2006, 71, 7449−7454. (67) Maitro, G.; Prestat, G.; Madec, D.; Poli, G. An Escapade in the World of Sulfenate Anions: Generation, Reactivity and Applications in Domino Processes. Tetrahedron: Asymmetry 2010, 21, 1075−1084. (68) Daligault, F.; Arboré, A.; Nugier-Chauvin, C.; Patin, H. Chiral Homoallylic and Allylic Sulfoxides as Models for the Stereochemical Analysis of Sulfoxide Thiaoleates. Tetrahedron: Asymmetry 2004, 15, 917−924. (69) Zohar, E.; Stanger, A.; Marek, I. Synthesis of Chiral Methylenecyclopropane Derivatives. Synlett 2005, 2239−2241. (70) Turks, M.; Lawrence, A. K.; Vogel, P. The Bora-ene Reaction of Sulfur Dioxide and Prop-2-ene-1-boronic Esters. New Route to Sulfoxides. Tetrahedron Lett. 2006, 47, 2783−2786. (71) Stikute, A.; Peipiņs,̌ V.; Turks, M. Synthesis of Allyl Sulfoxides from Allylsilanes via Silyl Sulfinates. Tetrahedron Lett. 2015, 56, 4578− 4581. (72) Ogiso, A.; Kitazawa, E.; Kurabayashi, M.; Sato, A.; Takahashi, S.; Noguchi, H.; Kuwano, H.; Kobayashi, S.; Mishima, H. Isolation and Structure of Antipeptic Ulcer Diterpene from Thai Medicinal Plant. Chem. Pharm. Bull. 1978, 26, 3117−3123. (73) Cutting, I.; Parsons, P. J. A New Synthesis of Unsaturated Aldehydes. Tetrahedron Lett. 1981, 22, 2021−2022. (74) Barrett, A. G. M.; Peña, M.; Willardsen, J. A. Total Synthesis and Structural Elucidation of the Antifungal Agent Papulacandin D. J. Org. Chem. 1996, 61, 1082−1100. (75) Brown, W. L.; Fallis, A. G. Intramolecular Rearrangements: Epimerization of Bicyclic Vinyl Tertiary Alcohols via a [2,3] Sulfoxide Sigmatropic Rearrangement. Can. J. Chem. 1987, 65, 1828−1832. (76) Knight, D. J.; Lin, P.; Russell, S. T.; Whitham, G. H. Preparation of Some Allylic Sulphones; Base-catalysed Isomerisation and Deuteriation of Cyclohex-2-enyl p-Tolyl Sulphones. J. Chem. Soc., Perkin Trans. 1 1987, 2701−2705.

(77) Stork, G.; Kreft, A. F., III. Concerning the Stereochemistry of the SN2’ Reaction in Cyclohexenyl Systems. J. Am. Chem. Soc. 1977, 99, 3850−3851. (78) Hua, D. H.; Venkataraman, S.; Ostrander, R. A.; Sinai, G. Z.; McCann, P. J.; Coulter, M. J.; Xu, M. R. Asymmetric Synthesis of (+)-Hirsutene. J. Org. Chem. 1988, 53, 507−515. (79) Luparia, M.; Legnani, L.; Porta, A.; Zanoni, G.; Toma, L.; Vidari, G. Enantioselective Synthesis and Olfactory Evaluation of Bicyclic αand γ-Ionone Derivatives: The 3D Arrangement of Key Molecular Features Relevant to the Violet Odor of Ionones. J. Org. Chem. 2009, 74, 7100−7110. (80) Baudin, J.-B.; Julia, S. A. Unsaturated Sulfinamides II. An Efficient Preparation of 4-(2’-Alkenesulfinyl)-Morpholines by the Reaction of 4Morpholinesulfenyl Chloride with Allylic Alcohols. Tetrahedron Lett. 1988, 29, 3251−3254. (81) Baudin, J.-B.; Bkouche-Waksman, I.; Hareau, G.; Julia, S. A.; Lorne, R.; Pascard, C. Diastereoselectivity in the [2,3]-Sigmatropic Rearrangement of Substituted Allylic N,N-Dialkylamidosulfoxylates. XRay Molecular Structure of [(l’)S*,(S)S*]-(2’E)-4-[[3′-(4″-Bromophenyl)-1’-methyl-2’-propenyl]sulfinyl]-morpholine. Tetrahedron 1991, 47, 6655−6672. (82) Blades, K.; Patel, S. T.; Percy, J. M.; Wilkes, R. D. Facile [2,3]Rearrangements of Difluoroallylic Alcohols with C-P and C-S Bond Formation. Tetrahedron Lett. 1996, 37, 6403−6406. (83) Maeno, M.; Shibata, N.; Cahard, D. Trifluoromethyl Sulfoxides from Allylic Alcohols and Electrophilic SCF3 Donor by [2,3]Sigmatropic Rearrangement. Org. Lett. 2015, 17, 1990−1993. (84) Masaki, M.; Sakuma, K.; Kaji, K. Facile Synthesis of (E)-Allylic Alcohols by Acid-Catalyzed Modification of the Mislow−Evans Rearrangement of Allylic Sulfoxides. Chem. Pharm. Bull. 1985, 33, 2531−2534. (85) Zhou, Z. S.; Flohr, A.; Hilvert, D. An Antibody-Catalyzed Allylic Sulfoxide−Sulfenate Rearrangement. J. Org. Chem. 1999, 64, 8334− 8341. (86) Ortiz de Montellano, P. R.; Hsu, C. K. Regioselective Hydroxylation of α,β-Unsaturated Esters. Tetrahedron Lett. 1976, 17, 4215−4218. (87) Hoffmann, R. W.; Goldmann, S. Intramolekulare Substitution von Sulfenäure-allylestern. Chem. Ber. 1978, 111, 2716−2725. (88) Matsuyama, H.; Izuoka, A.; Kobayashi, M. Synthesis of NAlkylisothiazolidines. Heterocycles 1985, 23, 1897−1900. (89) Fujita, M.; Ishida, M.; Manako, K.; Sato, K.; Ogura, K. Significant Orbital Interactions in π-Facial Stereoselection of Electrophilic Addition Reactions to Vinylic Sulfoxides. Tetrahedron Lett. 1993, 34, 645−648. (90) Block, E.; Wall, A. 2-Thiabicyclo[2.2.1]hept-5-ene and Its SOxides and 3-Alkyl Derivatives: Sulfine and Sulfene Cyclopentadiene Diels-Alder Adducts. Conversion of the Cyclopentadiene-Sulfine Adducts into 2-Oxa-3-thiabicyclo[3.3.0]oct-7-enes, Novel Bicyclic Sultenes. J. Org. Chem. 1987, 52, 809−818. (91) Confalone, P. N.; Kulesha, I. D.; Uskoković, M. R. A New Synthesis of 7-Dehydrocholesterols. J. Org. Chem. 1981, 46, 1030−1032. (92) Satoh, T.; Motohashi, S.; Tokutake, N.; Yamakawa, K. A Novel Synthesis Including Asymmetric Synthesis of α,β-Unsaturated γHydroxy Carbonyl Compounds from Enones with Carbon Homologation. Bull. Chem. Soc. Jpn. 1992, 65, 2966−2973. (93) Satoh, T.; Miyagawa, T. First Example of the Mislow− Braverman−Evans Rearrangement Retaining the Sulfur Atom on the Original Carbon. Tetrahedron Lett. 2006, 47, 1981−1983. (94) Kodama, M.; Shimada, K.; Itô, S. Syntheses of Macrocyclic Terpenoids by Intramolecular Cyclization III. Model Reactions for the Synthesis of Cembrene and Thunbergol Derivatives. Tetrahedron Lett. 1977, 18, 2763−2764. (95) Kodama, M.; Shimada, K.; Takahashi, T.; Kabuto, C.; Itô, S. Biomimetic Transformations of Germacradienes. Stereospecific Conversion of Hedycaryol Phenyl Sulfides to Cadinanes. Tetrahedron Lett. 1981, 22, 4271−4274. (96) Hauser, F. M.; Ellenberger, W. P.; Adams, T. C., Jr. Total Synthesis of 2,6-Epoxy-1(2H)-benzoxocin Sugar Analogs. J. Org. Chem. 1984, 49, 1169−1174. 14235

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

(97) Lansbury, P.; Rhodes, J. E. Synthesis of α,β-Unsaturated Ketones via [2,3] Sigmatropic Rearrangement of γ-Chloroallyl Sulfoxides and Amine Oxides. J. Chem. Soc., Chem. Commun. 1974, 21−22. (98) Lansbury, P. T.; Britt, R. W. γ-Chloroallyl Sulfoxides as Latent α,β-Unsaturated Carbonyl Compounds. A New Approach to Steroid Synthesis. J. Am. Chem. Soc. 1976, 98, 4577−4581. (99) Schuphan, I.; Casida, J. E. [2,3] Sigmatropic Rearrangement of S(3-Chloroallyl) Thiocarbamate Sulfoxides Followed by a 1,2-Elimination Reaction Yielding Unsaturated Aldehydes and Acid Chlorides. Tetrahedron Lett. 1979, 20, 841−844. (100) Schuphan, I.; Rosen, J. D.; Casida, J. E. Novel Activation Mechanism for the Promutagenic Herbicide Diallate. Science 1979, 205, 1013−1015. (101) Park, S. B.; Osterloh, J. D.; Vamvakas, S.; Hashmi, M.; Anders, M. W.; Cashman, J. R. Flavin-Containing Monooxygenase-Dependent Stereoselective S-Oxygenation and Cytotoxicity of Cysteine SConjugates and Mercapturates. Chem. Res. Toxicol. 1992, 5, 193−201. (102) Okano, T.; Ishihara, H.; Takakura, N.; Tsuge, H.; Eguchi, S.; Kimoto, H. Cyclization of Hindered Fluorinated Dichloro(5- and 6)alkenyl Radicals. Structural Effects of Olefins on the Selective Formation of 5-exo and 6-endo Cyclization Products. J. Org. Chem. 1997, 62, 7192−7200. (103) Mandai, T.; Moriyama, T.; Nakayama, Y.; Sugino, K.; Kawada, M.; Otera, J. α-Methoxyallyl Sulfides as a Novel Homoenolate Dianion Equivalent. Tetrahedron Lett. 1984, 25, 5913−5916. (104) Sato, T.; Okazaki, H.; Otera, J.; Nozaki, H. ThiostannaneMediated Preparation of γ-Alkoxyallyl Sulfides; Efficient β-Alkylation of α,β-Unsaturated Aldehydes. Tetrahedron Lett. 1988, 29, 2979−2982. (105) Sato, T.; Otera, J.; Nozaki, H. Efficient Stereocontrol in the [2,3] Sigmatropic Rearrangement of Allylic Sulfoxides. J. Org. Chem. 1989, 54, 2779−2780. (106) Sato, T.; Otera, J. Hydroxyl-Directed Stereoselective [2,3] Sigmatropic Rearrangement of γ-Hydroxyalkyl γ-Alkoxyallylic Sulfoxides. Tetrahedron Lett. 1994, 35, 6701−6704. (107) Bonini, B. F.; Franchini, M. C.; Mazzanti, G.; Slief, J.-W.; Wegman, M. A.; Zwanenburg, B. Sulfoxide Induced Sigmatropic Rearrangement (SISR) of Methyl 1-Methylsulfanylvinyl Sulfoxides. Chem. Commun. 1997, 1011−1012. (108) Brownbridge, P.; Warren, S. Organic Synthesis Using a Migrating Functional Group: [1,2] and [1,3] Phenylthio Shifts in Synthesis via Ally1 Sulphides. J. Chem. Soc., Perkin Trans. 1 1977, 1131− 1141. (109) Blatcher, P.; Warren, S. Phenylthio Migrations in Rearrangements of 2,2-Bisphenylthioethanols. J. Chem. Soc., Perkin Trans. 1 1985, 1055−1066. (110) Aggarwal, V. K.; Coldham, I.; McIntyre, S.; Warren, S. Transformation of Cyclic α-Phenylthio Aldehydes by Stereoselective Aldol Reactions and Phenylthio Migration into Spirocyclic Lactones and Ethers, and E-Allylic Alcohols with 1,4-Related Chiral Centres. J. Chem. Soc., Perkin Trans. 1 1991, 451−460. (111) Carter, M. J.; Fleming, I. Ally1 Silanes in Organic Synthesis: Some Reactions of 3-Trimethylsilylcyclohex-4-ene-1,2-dicarboxylic Acid and its Derivatives. J. Chem. Soc., Chem. Commun. 1976, 679−680. (112) Fleming, I.; Au-Yeung, B.-W. The Chemistry of an Allylsilane: The Synthesis of a Prostaglandin Intermediate and of Loganin. Tetrahedron 1981, 37 (Suppl. 1), 13−24. (113) Antonjuk, D. J.; Ridley, D. D.; Smal, M. A. Asymmetric Induction in the Additions of Anions of Allylic Sulfoxides to Benzaldehyde. Aust. J. Chem. 1980, 33, 2635−2651. (114) Annunziata, R.; Cinquini, M.; Cozzi, F.; Raimondi, L.; Stefanelli, S. Stereoselective Synthesis of (E)-2-Alkene-1,4-diols via Metallated Allylic Sulfoxides. Tetrahedron 1986, 42, 5443−5450. (115) Annunziata, R.; Cinquini, M.; Cozzi, F.; Raimondi, L.; Stefanelli, S. Stereoselective Synthesis of Polyols Precursors by Allyl Sulphinyl Anion Addition to Chiral Alkoxy Aldehydes. Tetrahedron 1986, 42, 5451−5456. (116) Veselovskii, V. V.; Shpiro, N. A.; Moiseenkov, M. A. Synthesis of Oxygen-Containing Derivatives of Ocimene Based on 4-Phenylsulfinylmyrcene. Izv. Akad. Nauk SSSR, Ser. Khim 1988, 1588−1591.

(117) Mandai, T.; Osaka, K.; Kawagishi, M.; Kawada, M.; Otera, J. Synthesis and Highly Regioselective Diels-Alder Reaction of Functionalized Isoprenes Involving a Terminal Alkoxy Group and Chemical Modification of the Resulting Adducts. J. Org. Chem. 1984, 49, 3595− 3600. (118) Veselovskii, V. V.; Dragan, V. A.; Moiseenkov, A. M. Regiospecific Ene Reaction of Benzenesulfinyl Chloride with Linear Isoprenoids. Izv. Akad. Nauk SSSR, Ser. Khim 1987, 2787−2990. (119) Trost, B. M.; Matelich, M. C. A Selectivity Control Element for Palladium-Catalyzed Trimethylenemethane Cycloaddition. J. Am. Chem. Soc. 1991, 113, 9007−9009. (120) Hwu, J. R.; Anderson, D. A. Zwitterion-Accelerated [3,3]Sigmatropic Rearrangements and [2,3]-Sigmatropic Rearrangements of Sulfoxides and Amine Oxides. J. Chem. Soc., Perkin Trans. 1 1991, 3199− 3206. (121) Johnston, M.; Raines, R.; Walsh, C.; Firestone, R. A. MechanismBased Enzyme Inactivation Using an Allyl Sulfoxide−Allyl Sulfenate Ester Rearrangement. J. Am. Chem. Soc. 1980, 102, 4241−4250. (122) Branchaud, B. P.; Walsh, C. T. Functional Group Diversity in Enzymatic Oxygenation Reactions Catalyzed by Bacterial FlavinContaining Cyclohexanone Oxygenase. J. Am. Chem. Soc. 1985, 107, 2153−2161. (123) Gordon, E. M.; Pluscec, J. Chemistry of Isoprenylated Cysteinyl Containing Peptides. [2,3] Sigmatropic Rearrangement of S-Farnesylcysteinyl Sulfoxides. Studies toward a Mild Method of Deprenylating Lipopeptides. J. Am. Chem. Soc. 1992, 114, 1521−1523. (124) Hong, I. S.; Ding, H.; Greenberg, M. M. Oxygen Independent DNA Interstrand Cross-Link Formation by a Nucleotide Radical. J. Am. Chem. Soc. 2006, 128, 485−491. (125) Ishibashi, H.; Nakatani, H.; Sakashita, H.; Ikeda, M. A New Synthesis of Vinyl Sulphides by Lewis Acid Promoted Reaction of [Chloro(arylthio)methyl]trimethylsilanes with Trisubstituted Alkenes. J. Chem. Soc., Chem. Commun. 1987, 338−340. (126) Mikołajczyk, M.; Krysiak, J. A.; Midura, W. H.; Wieczorek, M. W.; Róz ycka-Sokołowska, E. General Route to Racemic and Enantiomeric Carbo- and Heterocyclic Vinyl Sulfoxides via Tandem Michael Addition/Horner Olefination of α-Phosphorylvinyl Sulfoxides. J. Org. Chem. 2006, 71, 8818−8823. (127) Nemoto, H.; Miyata, J.; Yoshida, M.; Raku, N.; Fukumoto, K. A Novel Palladium-Mediated Cascade Reaction Triggered by Strain Release of the Cyclobutane System. A New General Route to Benzoand Naphthohydrindans. J. Org. Chem. 1997, 62, 7850−7857. (128) Fokin, A. A.; Kushko, A. O.; Kirij, A. V.; Yurchenko, A. G.; Schleyer, P. v. R. Direct Transformations of Ketones to α-Unsaturated Thiols via [2,3]-Sigmatropic Rearrangement of Allyl Sulfinyl Carbanions: A Combined Experimental and Computational Study. J. Org. Chem. 2000, 65, 2984−2995. (129) Raghavan, S.; Kumar, V. V.; Chowhan, L. R. Highly Stereoselective Preparation of Chiral α-Substituted Sulfides from αChloro Sulfides via 1,2-Asymmetric Induction. Synlett 2010, 1807− 1810. (130) Kano, T.; Sakamoto, R.; Maruoka, K. Remote Chirality Control Based on the Organocatalytic Asymmetric Mannich Reaction of α-Thio Acetaldehydes. Chem. Commun. 2014, 50, 942−944. (131) Bernoud, E.; Le Duc, G.; Bantreil, X.; Prestat, G.; Madec, D.; Poli, G. Aryl Sulfoxides from Allyl Sulfoxides via [2,3]-Sigmatropic Rearrangement and Domino Pd-Catalyzed Generation/Arylation of Sulfenate Anions. Org. Lett. 2010, 12, 320−323. (132) Dornan, P. K.; Kou, K. G. M.; Houk, K. N.; Dong, V. M. Dynamic Kinetic Resolution of Allylic Sulfoxides by Rh-Catalyzed Hydrogenation: A Combined Theoretical and Experimental Mechanistic Study. J. Am. Chem. Soc. 2014, 136, 291−298. (133) Johnston, J. N.; Tsui, H.-C.; Paquette, L. A. Application of an S→ O Allylic Transposition in the Context of a Bridgehead Olefinic System. New Opportunities for the Structural Modification of Bicyclo[6.2.1]undecanes via Transannular Ring Closure. J. Org. Chem. 1998, 63, 129− 136. 14236

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Northern Fragment of Polycavernoside A. Tetrahedron Lett. 2009, 50, 3366−3370. (155) Kosugi, H.; Uda, H.; Yamagiwa, S. Vinylogous Pummerer Rearrangement of Methyl Cyclopentylidene(phenylsulphinyl)acetate. J. Chem. Soc., Chem. Commun. 1975, 192−193. (156) Yamagiwa, S.; Sato, H.; Hoshi, N.; Kosugi, H.; Uda, H. New Rearrangement Reactions of α-Phenylsulphinylacrylate Derivatives. J. Chem. Soc., Perkin Trans. 1 1979, 570−583. (157) Trost, B. M.; Rigby, J. H. Synthetic Strategy toward Verrucarins. An Approach toward Verrucarol. J. Org. Chem. 1978, 43, 2938−2940. (158) Cass, Q. B.; Jaxa-Chamiec, A. A.; Sammes, P. G. Preparation and Rearrangement of Some Conjugated Phenylsulphinylacetate Derivatives. J. Chem. Soc., Chem. Commun. 1981, 1248−1250. (159) Kido, F.; Noda, Y.; Yoshikoshi, A. α-(Phenylthio)-β-vinylbutenolide, a Novel Annulating Reagent for the Synthesis of 4Oxygenated Perhydro-2-oxo-1-benzofuran Derivatives. J. Am. Chem. Soc. 1982, 104, 5509−5511. (160) Tanikaga, R.; Konya, N.; Tamura, T.; Kaji, A. Stereochemistry in the Knoevenagel Reaction of Methyl Arylsulphinylacetate and Aldehydes. J. Chem. Soc., Perkin Trans. 1 1987, 825−830. (161) Tanikaga, R.; Nozaki, Y.; Nishida, M.; Kaji, A. Conversion of 2Sulfinylated 2-Alkenoate Esters to (2E,4E)-2,4-Alkadienoate Esters by Pyrolysis. Bull. Chem. Soc. Jpn. 1984, 57, 729−733. (162) Nokami, J.; Nishiuchi, K.; Wakabayashi, S.; Okawara, R. Pyrolysis of Sulfoxide. III. Conversion of 2-(Phenylsulfinyl)enoate to 2,4-Dienoate and Synthesis of Pellitorine. Tetrahedron Lett. 1980, 21, 4455−4456. (163) Nokami, J.; Mandai, T.; Imakura, Y.; Nishiuchi, K.; Kawada, M.; Wakabayashi, S. Preparation of 4-Hydroky-2-alkenenitriles from Phenylsulfinylacetonitrile with Ketones or Aldehydes, and Its Synthetic Application to Substituted Furans. Tetrahedron Lett. 1981, 22, 4489− 4490. (164) Tanikaga, R.; Nozaki, Y.; Tanaka, K.; Kaji, A. Selective Reactions of Thiolate Anions with 4-Hydroxy-E-2-alkenoic Esters or 4Methanesulfonyloxy-E-2-alkenoic Esters. Synthesis of 2-Alken-4-olides (Δα, β-Butenolides) and E,E,-2,4-Alkadienoic Esters. Chem. Lett. 1982, 11, 1703−1706. (165) Tanikaga, R.; Nozaki, Y.; Tamura, T.; Kaji, A. Facile Synthesis of 4-Hydroxy-(E)-2-alkenoic Esters from Aldehydes. Synthesis 1983, 134− 135. (166) Ono, T.; Tamaoka, T.; Yuasa, Y.; Matsuda, T.; Nokami, J.; Wakabayashi, S. Reaction of α-(Phenylsulfinyl)acetonitrile with Aldehydes and Ketones to γ-Hydroxyalkenenitriles and Syntheses of Terpenoids. J. Am. Chem. Soc. 1984, 106, 7890−7893. (167) Blade, R. J.; Robinson, J. E. A Novel Application of Molybdenum Mediated Diene Synthesis in the Preparation of Piperovatine. Tetrahedron Lett. 1986, 27, 3209−3210. (168) Trost, B. M.; Tometzki, G. B. A Novel Palladium-Catalyzed Deoxygenation of Enediols to 1,3-Dienes. Synthesis 1991, 1235−1245. (169) Gung, B. W.; Francis, M. B. Diastereofacial Selectivity in Intramolecular l,4-Conjugate Additions: Conclusive Evidence for the Importance of Ground-State Conformations and the σ* Orbital Effect. J. Org. Chem. 1993, 58, 6177−6179. (170) Giardinà, A.; Marcantoni, E.; Mecozzi, T.; Petrini, M. Claisen− Johnson Orthoester Rearrangement of γ-Hydroxy α,β-Unsaturated Ketones and Nitriles. Eur. J. Org. Chem. 2001, 713−718. (171) Nokami, J.; Nishimura, A.; Sunami, M.; Wakabayashi, S. Facile Formation of 4-Hydroxyalk-2-enone Functionality. Tetrahedron Lett. 1987, 28, 649−650. (172) Nokami, J.; Taniguchi, A.; Honda, M.; Fukutake, S. Preparation of Diethyl 3-Hydroxy-1-alkenylphosphonates by the Reaction of Diethyl (Arylsulfinylmethyl)phosphonate with Aldehydes under the Knoevenagel Conditions. Chem. Lett. 1995, 24, 1025−1026. (173) Tanikaga, R.; Kaji, A. Allyl Benzenesulfenates in Isomeric Forms. Chem. Lett. 1988, 17, 677−680. (174) Burgess, K.; Henderson, I. Lipase-Mediated Resolutions of SPAC Reaction Products. Tetrahedron: Asymmetry 1990, 1, 57−60. (175) Paquette, L. A.; Tae, J.; Hickey, E. R.; Trego, W. E.; Rogers, R. D. Preorganized Ligand Arrays Based on Spirotetrahydrofuranyl Motifs.

(134) Raghavan, S.; Ravi, A. Synthesis of Crinane Utilizing an Allylic Sulfoxide for the Construction of a Hydroindole Ring via Vinylogous C−N Bond Formation. Org. Biomol. Chem. 2016, 14, 10222−10229. (135) Reich, H. J.; Reich, I. L.; Wollowitz, S. Conversion of Allyl Alcohols to Dienes by Sulfoxide and Selenoxide Syn Elimination. Synthesis of PCB Arene Oxides. J. Am. Chem. Soc. 1978, 100, 5981− 5982. (136) Reich, H. J.; Wollowitz, S. Conversion of Allyl Alcohols to 1,3Dienes by Sequential Sulfenate−Sulfoxide [2,3] Sigmatropic Rearrangement and Syn Elimination. J. Am. Chem. Soc. 1982, 104, 7051−7059. (137) Babler, J. H.; Invergo, B. J. A Convenient Stereoselective Route to the Sex Pheromone of the Red Bollworm Moth via an Allylic Sulfenate to Sulfoxide Rearrangement. J. Org. Chem. 1979, 44, 3723− 3724. (138) Hutchins, C. W.; Cooper, G. K.; Pürro, S.; Rapoport, H. 6Demethoxythebaine and Its Conversion to Analgesics of the 6,14Ethenomorphinan Type. J. Med. Chem. 1981, 24, 773−777. (139) Cookson, R. C.; Parsons, P. J. New Synthesis of Conjugated Ketones from 2-Phenylthio-allyl and -allenyl Alcohols. J. Chem. Soc., Chem. Commun. 1978, 821−822. (140) de Groot, A.; Jansen, B. J. M.; Reuvers, J. T. A.; Tedjo, E. M. Functionalized Dienes Via Thermolysis of Allylic Sulfoxides. Tetrahedron Lett. 1981, 22, 4137−4138. (141) Jones, D. N.; Meanwell, N. A.; Mirza, S. M. Convergent Syntheses of 9-Deoxy-12-phenylthioprostanoids and 9-Deoxy-Δ8(12)PGD1 Derivatives. J. Chem. Soc., Perkin Trans. 1 1985, 145−151. (142) Waterson, A. G.; Meyers, A. I. Bicyclic Lactams as Chiral Homoenolate Equivalents: Synthesis of (−)-Penienone. J. Org. Chem. 2000, 65, 7240−7243. (143) McFadden, J. M.; Frehywot, G. L.; Townsend, C. A. A Flexible Route to (5R)-Thiolactomycin, a Naturally Occurring Inhibitor of Fatty Acid Synthesis. Org. Lett. 2002, 4, 3859−3862. (144) McFadden, J. M.; Medghalchi, S. M.; Thupari, J. N.; Pinn, M. L.; Vadlamudi, A.; Miller, K. I.; Kuhajda, F. P.; Townsend, C. A. Application of a Flexible Synthesis of (5R)-Thiolactomycin To Develop New Inhibitors of Type I Fatty Acid Synthase. J. Med. Chem. 2005, 48, 946− 961. (145) Baldwin, J. E.; Raghavan, A. S.; Hess, B. A., Jr.; Smentek, L. Thermal [1,5] Hydrogen Sigmatropic Shifts in cis,cis-1,3-Cyclononadienes Probed by Gas-Phase Kinetic Studies and Density Functional Theory Calculations. J. Am. Chem. Soc. 2006, 128, 14854− 14862. (146) Trost, B. M.; Stanton, J. L. New Synthetic Methods. 1,3Alkylative Carbonyl Transposition. J. Am. Chem. Soc. 1975, 97, 4018− 4025. (147) Gaoni, Y. A 1,5-Shift of the Phenylsulfinyl Group in Phenyl Pentadienyl Sulfoxides. Phosphorus Sulfur Relat. Elem. 1979, 6, 101. (148) Kametani, T.; Tsubuki, M.; Nemoto, H.; Suzuki, K. Concerted Ring-Opening Reaction of Cyclobutenes. The Extraordinary Accelerating Effects of Arylsulfoxy and Sulfonyl Carbanion Substituents. J. Am. Chem. Soc. 1981, 103, 1256−1258. (149) Corey, E. J.; Hoover, D. J. Total Synthesis of 5-Desoxyleukotriene D. A New and Useful Equivalent of the 4-Formyl-trans, trans-1,3butadienyl Anion. Tetrahedron Lett. 1982, 23, 3463−3466. (150) Corey, E. J.; Oh, H.; Barton, A. E. Pathways for Migration and Cleavage of the S-Peptide Unit of the Leukotrienes. Tetrahedron Lett. 1982, 23, 3467−3470. (151) Cass, Q. B.; Jaxa-Chamiec, A. A.; Kunec, E. K.; Sammes, P. G. On the Preparation and Rearrangement of Some Vinylic Sulfoxides. J. Chem. Soc., Perkin Trans. 1 1991, 2683−2686. (152) Epifani, E.; Florio, S.; Ingrosso, G.; Ronzini, L.; Sgarra, R.; Troisi, L. Regio- and Stereo-chemistry of the Coupling Reaction of Phenylsulphenyl-, Phenylsulphinyl- and Phenylsulphonyl-pentadienyl Metals. Tetrahedron 1991, 47, 7489−7500. (153) Schreiber, S. L.; Satake, K. Total Synthesis of (±)-Asteltoxin. J. Am. Chem. Soc. 1984, 106, 4186−4188. (154) Jourdain, P.; Philippart, F.; Dumeunier, R.; Markó, I. E. Pentadienyl Sulfoxide in Triene Synthesis. Efficient Assembly of the 14237

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Synthesis of the Stereoisomeric 1,8,14-Trioxatrispiro[4.1.4.1.4.1]octadecanes and the Contrasting Conformational Features and Ionic Binding Capacities of These Belted Ionophores. J. Org. Chem. 2000, 65, 9160−9171. (176) Hagiwara, H.; Isobe, K.; Numamae, A.; Hoshi, T.; Suzuki, T. Sequential Knoevenagel Reaction/Mislow−Evans Rearrangement Catalyzed by Heterogeneous Amine Grafted on Silica in Water, Leading to γ-Hydroxy-α,β-unsaturated Nitrile. Synlett 2006, 1601−1603. (177) Domínguez, E.; Carretero, J. C. Facile Synthesis of E-γ-Hydroxyα,β-unsaturated Sulfones from Aldehydes. Tetrahedron Lett. 1990, 31, 2487−2490. (178) Domínguez, E.; Carretero, J. C. A Practical Route to (E)-γHydroxy-α,β-unsaturated Phenyl Sulfones. Tetrahedron 1990, 46, 7197−7206. (179) Domínguez, E.; Carretero, J. C.; Fernández-Mayoralas, A.; Conde, S. An Efficient Preparation of Optically Active (E)-γ-Hydroxyα,β]-Unsaturated Phenyl Sulfones Using Lipase-Mediated Acylations. Tetrahedron Lett. 1991, 32, 5159−5162. (180) Carretero, J. C.; Domínguez, E. Lipase-Catalyzed Kinetic Resolution of γ-Hydroxy Phenyl Sulfones. J. Org. Chem. 1992, 57, 3867− 3873. (181) Trost, B. M.; Grese, T. A. A Convenient Synthesis of γ-Hydroxy α,β-Unsaturated Sulfones. J. Org. Chem. 1991, 56, 3189−3192. (182) Alcaraz, C.; Carretero, J. C.; Domínguez, E. Synthesis and Conjugate Additions to (E)-γ-Alkoxy-α-substituted-α,β-unsaturated Sulfones. Tetrahedron Lett. 1991, 32, 1385−1388. (183) Domínguez, E.; Carretero, J. C. Diastereoselectivity of the Conjugate Addition of Organocopper Reagents to (E)-γ-Hydroxy-α-βunsaturated Sulfones. Tetrahedron Lett. 1993, 34, 5803−5806. (184) Carretero, J. C.; Domínguez, E. A New Method for the Iterative Construction of Enantiomerically Pure Polypropionate Chains. J. Org. Chem. 1993, 58, 1596−1600. (185) Domínguez, E.; Carretero, J. C. Flexible and Stereoselective Construction of Polypropionate Chains from Enantiomerically Pure γHydroxy-α,β-unsaturated Sulfones. Tetrahedron 1994, 50, 7557−7566. (186) de Blas, J.; Carretero, J. C.; Domínguez, E. Stereoselective Approach to Optically Pure syn 2-Amino Alcohol Derivatives. Tetrahedron Lett. 1994, 35, 4603−4606. (187) Adrio, J.; Carretero, J. C.; Gómez Arrayás, R. 5- and 6-Exo Radical Cyclizations of γ-Oxygenated-α,β-Unsaturated Sulfones. Synlett 1996, 640−642. (188) Carretero, J. C.; Gómez Arrayás, R.; Storch de Gracia, I. A Stereoselective Approach to Polyhydroxylated Quinolizidine Alkaloids. Tetrahedron Lett. 1997, 38, 8537−8540. (189) Carretero, J. C.; Gómez Arrayás, R. Stereoselective Synthesis of Polyhydroxylated Indolizidines from γ-Hydroxy α,β-Unsaturated Sulfones. J. Org. Chem. 1998, 63, 2993−3005. (190) de Vicente, J.; Gómez Arrayás, R.; Carretero, J. C. An Efficient and Stereoselective Synthesis of Enantiopure 1,2,7-Trihydroxylated Pyrrolizidines. Tetrahedron Lett. 1999, 40, 6083−6086. (191) Iradier, F.; Gómez Arrayás, R.; Carretero, J. C. Synthesis of Medium-Sized Cyclic Amines by Selective Ring Cleavage of Sulfonylated Bicyclic Amines. Org. Lett. 2001, 3, 2957−2960. (192) Alonso, I.; Carretero, J. C.; Garrido, J. L.; Magro, V.; Pedregal, C. Palladium-Catalyzed Allylic Substitution in γ-Oxygenated Vinyl Sulfones: One-Step Synthesis of Tetrasubstituted Dihydrofurans. J. Org. Chem. 1997, 62, 5682−5683. (193) Giovannini, R.; Marcantoni, E.; Petrini, M. Claisen Rearrangement of γ-Hydroxyvinyl Sulfones via Ketene Acetal Derivatives. A New Entry to Functionalized (2E,4E)-Alkadienoic Esters. Tetrahedron Lett. 1998, 39, 5827−5830. (194) Adrio, J.; Rodríguez Rivero, M.; Carretero, J. C. endo-Selective Intramolecular Pauson-Khand Reactions of γ-Oxygenated-α,β-unsaturated Phenylsulfones. Chem. - Eur. J. 2001, 7, 2435−2448. (195) Annunziata, R.; Cinquini, M.; Cozzi, F.; Raimondi, L.; Restelli, A. Synthesis of Optically Active γ-Hydroxy α,β-Unsaturated Acid Derivatives from Enantiomerically Pure Sulfoxides. Gazz. Chim. Ital. 1985, 115, 637−641.

(196) Nokami, J.; Mandai, T.; Nishimura, A.; Takeda, T.; Wakabayashi, S.; Kunieda, N. Synthesis of Optically Active 4Hydroxyalk-2-enenitriles; Reaction of Optically Active 2-(pChlorophenylsulfinyl)acetonitrile with Aldehydes in the Presence of Piperidine in Acetonitrile. Tetrahedron Lett. 1986, 27, 5109−5112. (197) Kosugi, H.; Kitaoka, M.; Takahashi, A.; Uda, H. Highly Enantioselective Rearrangement of (S)-E-α-p-Tolylsulphinyl-α,βenoates to (R)-E-γ-Hydroxy-α,β-enoates. J. Chem. Soc., Chem. Commun. 1986, 1268−1270. (198) Burgess, K.; Henderson, I. A Facile Route to Homochiral Sulfoxides. Tetrahedron Lett. 1989, 30, 3633−3636. (199) Burgess, K.; Henderson, I.; Ho, K.-K. Biocatalytic Resolutions of Sulfinylalkanoates: A Facile Route to Optically Active Sulfoxides. J. Org. Chem. 1992, 57, 1290−1295. (200) Burgess, K.; Cassidy, J.; Henderson, I. Optically Active Building Blocks from the SPAC Reaction: A Completely Asymmetric Synthesis of (4S-cis)-5-(Cyclohexylmethyl)-4-hydroxy-2-pyrrolidinone, a Statine Analogue. J. Org. Chem. 1991, 56, 2050−2058. (201) Liu, Z.; Burgess, K. Continuous Flow Biocatalytic Resolutions of Methyl Sulfinylacetates. Tetrahedron Lett. 2011, 52, 6325−6327. (202) Burgess, K.; Henderson, I. Stereochemically Matched Sulfinylacetates for Double Diastereoselection in the SPAC Reaction. Tetrahedron 1991, 47, 6601−6616. (203) Trost, B. M.; Mallart, S. Double Diastereodifferentiation in the Hydroxylative Knoevenagel Condensation. Tetrahedron Lett. 1993, 34, 8025−8028. (204) Du, Z.; Kawatani, T.; Kataoka, K.; Omatsu, R.; Nokami, J. Alkyl 2-(2-benzothiazolylsulfinyl)acetates as Useful Synthetic Reagents for Alkyl 4-Hydroxyalk-2-enoates by Sulfinyl-Knoevenagel Reaction. Tetrahedron 2012, 68, 2471−2480. (205) Miura, M.; Toriyama, M.; Kawakubo, T.; Yasukawa, K.; Takido, T.; Motohashi, S. Asymmetric Synthesis of γ-Hydroxy α-Enones by 1,8Diazabicyclo[5.4.0]undec-7-ene-Catalyzed Stereoselective Rearrangement of Chiral α-Sulfinyl Enones. Org. Lett. 2010, 12, 3882−3885. (206) Guerrero de la Rosa, V.; Ordóñez, M.; Alcudia, F.; Llera, J. M. A Convenient Synthesis of (E)-γ-Hydroxy- and (E)-γ-Keto-α,β-unsaturated Sulfoxides. Tetrahedron Lett. 1995, 36, 4889−4892. (207) Guerrero de la Rosa, V.; Ordóñez, M.; Llera, J. M. Efficient Preparation of Enantiomerically Pure (E)-γ-Hydroxy-α,β-unsaturated pTolylsulfoxides Using Lipase-Mediated Acylations. Tetrahedron: Asymmetry 2000, 11, 2991−3001. (208) Guerrero de la Rosa, V.; Ordóñez, M.; Llera, J. M. PalladiumCatalyzed Allylic Nucleophilic Substitution Reactions of (E)-γ-Acetoxyα,β-unsaturated p-Tolylsulfoxides. Tetrahedron: Asymmetry 2001, 12, 1089−1094. (209) Brebion, F.; Nàjera, F.; Delouvrié, B.; Lacôte, E.; Fensterbank, L.; Malacria, M. Addition of Bis-sulfinyl Anions to Ketones: Stereoselective Synthesis of Allylic Alcohols through Evans−Mislow Rearrangement Based Domino Reactions. Synthesis 2007, 2273−2278. (210) Corey, E. J.; Carpino, P. A Simplified Synthesis of (+)-Brefeldin A. Tetrahedron Lett. 1990, 31, 7555−7558. (211) Nokami, J.; Ohkura, M.; Dan-Oh, Y.; Sakamoto, Y. Facile Synthesis of (+)-Brefeldin A. Tetrahedron Lett. 1991, 32, 2409−2412. (212) Solladié, G.; Gerber, C. Asymmetric Synthesis of the Seco-Acids of (R,R)-Pyrenophorin and (R)-Patulolide A Monitored by a Chiral Sulfoxide Group. Synlett 1992, 449−450. (213) Nokami, J.; Taniguchi, T.; Gomyô, S.; Kakihara, T. A New Approach to the Synthesis of γ-Hydroxy-α,β-unsaturated Macrolides and (−)-Pyrenophorin by Intramolecular C = C Bond Formation with Oxidative Functionalization from ω-[α-(p-Chlorophenylsulfinyl)acetoxy]alkanal. Chem. Lett. 1994, 23, 1103−1106. (214) Nokami, J.; Taniguchi, T.; Ogawa, Y. Convenient Synthesis of (+)-Decarestrictine L. Chem. Lett. 1995, 24, 43−44. (215) Nokami, J.; Kataoka, K.; Shiraishi, K.; Osafune, M.; Hussain, I.; Sumida, S. Convenient Formation of 4-Hydroxyalk-2-en-1-one Functionality via a Knoevenagel-type Carbon Chain Elongation Reaction of Aldehyde with 1-Arylsulfinylalkan-2-one. J. Org. Chem. 2001, 66, 1228−1232. 14238

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

the Synthesis of Optically Pure 1,4-Imino-L-ribitol Derivatives. Org. Lett. 2000, 2, 3165−3168. (234) Koprowski, M.; Krawczyk, E.; Skowroñska, A.; McPartlin, M.; Choi, N.; Radojevic, S. Stereospecific [2,3] Sigmatropic Rearrangement of Allylic Sulfoxides and Selenoxides. Synthesis of Novel Polycyclic Allylic Alcohols and α-Hydroxy Ketones. Tetrahedron 2001, 57, 1105− 1118. (235) Garigipati, R. S.; Freyer, A. J.; Whittle, R. R.; Weinreb, S. M. Diastereoselective Synthesis of Unsaturated Vicinal Amino Alcohols via Diels-Alder Reactions of N-Sulfinyl Dienophiles. J. Am. Chem. Soc. 1984, 106, 7861−7867. (236) Natsugari, H.; Turos, E.; Weinreb, S. M.; Cvetovich, R. J. Further Studies on Stereoselective Synthesis of Vicinal Diamines from 3,6Dihydrothiazine-1-imines. Heterocycles 1987, 25, 19−24. (237) Remiszewski, S. W.; Whittle, R. R.; Weinreb, S. M. Synthesis of 5-epi-Desosamine via a Stereoselective Intramolecular N-Sulfinyl DielsAlder Cycloaddition. J. Org. Chem. 1984, 49, 3243−3244. (238) Remiszewski, S. W.; Stouch, T. R.; Weinreb, S. M. Synthesis of Methyl 3-Deoxy-3-methylaminoarabinopyranoside, a Component of Some Aminoglycoside Antibiotics. Tetrahedron 1985, 41, 1173−1178. (239) Weinreb, S. M. Synthetic Methodology Based upon N-Sulfinyl Dienophile [4 + 2]-Cycloaddition Reactions. Acc. Chem. Res. 1988, 21, 313−318. (240) Whitesell, J. K.; James, D.; Carpenter, J. F. Asymmetric Induction in the Cycloaddition Reactions of the N-Sulphinylcarbamate of 8-Phenylmenthol. J. Chem. Soc., Chem. Commun. 1985, 1449−1450. (241) Remiszewski, S. W.; Yang, J.; Weinreb, S. M. Enantioselective NSulfinyl Dienophile Diels-Alder Cycloadditions. Tetrahedron Lett. 1986, 27, 1853−1856. (242) Bayer, A.; Gautun, O. R. Asymmetric N-Sulfinyl Dienophile Diels−Alder Cycloadditions Using Chiral Ti(IV)-based Lewis Acids. Tetrahedron Lett. 2000, 41, 3743−3746. (243) Whitesell, J. K.; Carpenter, J. F.; Yaser, H. K.; Machajewski, T. Asymmetric Induction in the Ene Reactions of N-Sulfinylcarbamates. J. Am. Chem. Soc. 1990, 112, 7653−7659. (244) Stien, D.; Anderson, G. T.; Chase, C. E.; Koh, Y.; Weinreb, S. M. Total Synthesis of the Antitumor Marine Sponge Alkaloid Agelastatin A. J. Am. Chem. Soc. 1999, 121, 9574−9579. (245) Cully, S. J.; Storr, T. E.; Rawling, M. J.; Abeysena, I. R.; Hamza, D.; Jones, G.; Pearce, C. A.; Quddus, A.; Lewis, W.; Stockman, R. A. Expedient Synthesis of an Atypical Oxazolidinone Compound Library. Bioorg. Med. Chem. 2016, 24, 5249−5257. (246) Heintzelman, G. R.; Fang, W.-K.; Keen, S. P.; Wallace, G. A.; Weinreb, S. M. Stereoselective Total Syntheses and Reassignment of Stereochemistry of the Freshwater Cyanobacterial Hepatotoxins Cylindrospermopsin and 7-Epicylindrospermopsin. J. Am. Chem. Soc. 2002, 124, 3939−3945. (247) Anwar, B.; Grimsey, P.; Hemming, K.; Krajniewski, M.; Loukou, C. A Thiazine-S-oxide, Staudinger/aza-Wittig Based Synthesis of Benzodiazepines and Benzothiadiazepines. Tetrahedron Lett. 2000, 41, 10107−10110. (248) Hemming, K.; Loukou, C. The Synthesis of Bicyclic 1,2,3,4Tetrahydro-1,4-benzodiazepin-5-ones from 2-(o-Nitrobenzoyl)-1,2thiazine-1-oxide Precursors. Tetrahedron 2004, 60, 3349−3357. (249) Veselovskii, V. V.; Makarova, Z. G.; Lutsenko, A. I.; Shpiro, N. A.; Zhulin, V. M.; Moiseenkov, A. M. 2-Phenylsulfinylmethyl-1,3butadiene in the Thermal Diels-Alder Reaction. Izv. Akad. Nauk SSSR, Ser. Khim. 1988, 810−815. (250) Padwa, A.; Bullock, W. H.; Norman, B. H.; Perumattam, J. (Nitroaryl)sulfinyl-Substituted Allenes. Novel and Convenient Propargyl Alcohol Synthons in 4 + 2 Cycloaddition Chemistry. J. Org. Chem. 1991, 56, 4252−4259. (251) Aggarwal, V. K.; Warren, S. Synthesis of Compounds with 1,4Related Chiral Centres by Phenylthio Migration: Syn and Anti-E-2,4Dimethylhex-3-ene-1,5-diol. Tetrahedron Lett. 1987, 28, 1925−1928. (252) Munchhof, M. J.; Heathcock, C. H. Total Synthesis of ACRL Toxin IIIb: A Protocol for Parlaying Aldols into Synthons Containing Three Stereocenters Having an [n, n+1, n+4] Relationship. J. Org. Chem. 1994, 59, 7566−7567.

(216) Trost, B. M.; Dong, L.; Schroeder, G. M. Exploiting the Pd- and Ru-Catalyzed Cycloisomerizations: Enantioselective Total Synthesis of (+)-Allocyathin B2. J. Am. Chem. Soc. 2005, 127, 10259−10268. (217) de Vicente, J.; Huckins, J. R.; Rychnovsky, S. D. Synthesis of the C31−C67 Fragment of Amphidinol 3. Angew. Chem., Int. Ed. 2006, 45, 7258−7262. (218) Evans, D. A.; Bryan, C. A.; Sims, C. L. The Complementarity of (4 + 2) Cycloaddition Reactions and [2,3] Sigmatropic Rearrangements in Synthesis. A New Synthesis of Functionalized Hasubanan Derivatives. J. Am. Chem. Soc. 1972, 94, 2891−2892. (219) Posner, G. H.; Haces, A.; Harrison, W.; Kinter, C. M. Highly Stereocontrolled Synthesis of Some Polyfunctionalized Cyclohexenes. A Short Formal Total Synthesis of (±)-Chorismic Acid. J. Org. Chem. 1987, 52, 4836−4841. (220) Vedejs, E.; Campbell, J. B., Jr.; Gadwood, R. C.; Rodgers, J. D.; Spear, K. L.; Watanabe, Y. Synthesis of the Cytochalasin D Isoindolone Unit: Solutions to the Problem of Regiochemistry in N-Benzoylpyrrolinone Diels-Alder Reactions. J. Org. Chem. 1982, 47, 1534−1546. (221) Overman, L. E.; Petty, C. B.; Ban, T.; Huang, G. T. Preparation and Diels-Alder Reactions of 1,3-Dienes Containing both Sulfur and Nitrogen Substituents. Complete Orientational Control by the Acylamino Group. J. Am. Chem. Soc. 1983, 105, 6335−6337. (222) Grieco, P. A.; Lis, R.; Zelle, R. E.; Finn, J. Convergent, Enantiospecific Total Synthesis of the Hypocholesterolemic Agent (+)-Compactin. J. Am. Chem. Soc. 1986, 108, 5908−5919. (223) Arce, E.; Carreño, M. C.; Cid, M. B.; García-Ruano, J. L. First Diels-Alder Reactions of Enantiomerically Pure 1-p-Tolylsulfiny1 Dienes: Straightforward Access to Cyclohexenols through Tandem Cycloaddition/[2,3]-Sigmatropic Rearrangement. J. Org. Chem. 1994, 59, 3421−3426. (224) Carreño, M. C.; Cid, M. B.; Colobert, F.; García-Ruano, J. L.; Solladié, G. Diels-Alder Reactions of 1-Sulfinyldienes with an Endocyclic Double Bond: The Unexpected Evolution of the NMethylmaleimide Adducts. Tetrahedron: Asymmetry 1994, 5, 1439− 1442. (225) Pindur, U.; Lutz, G.; Rogge, M. First Synthesis of Chiral 3Vinylindoles as 4π-Components for Diels-Alder Reactions. J. Heterocycl. Chem. 1995, 32, 201−206. (226) Aversa, M. C.; Barattucci, A.; Bonaccorsi, P.; Giannetto, P.; Nicolò, F.; Rizzo, S. Diels-Alder Reactions of Enantiopure [(1S)Isoborneol-10-sulfinyl]- and [(1S-exo)-2-Bornylsulfinyl] vinylcyclohexenes with Maleimides. Tetrahedron: Asymmetry 1999, 10, 3907−3917. (227) Carreño, M. C.; Cid, M. B.; García-Ruano, J. L. Diels-Alder Reactions of [(S)R]-(1E,3E)-l-p-Tolylsulfinyl-1,3-pentadiene: The Unexpected Evolution of Maleic Anhydride Adducts. Tetrahedron: Asymmetry 1996, 7, 2151−2158. (228) Aversa, M. C.; Barattucci, A.; Bonaccorsi, P.; Giannetto, P. Chiral Sulfinyl-l,3-dienes. Synthesis and Use in Asymmetric Reactions. Tetrahedron: Asymmetry 1997, 8, 1339−1367. (229) Aranda, M. T.; Aversa, M. C.; Barattucci, A.; Bonaccorsi, P.; Carreño, M. C.; Cid, M. B.; García-Ruano, J. L. Synthesis of Chiral 1and 2-(p-Tolylsufinyl)-3-trimethylsilyloxybuta-1,3-dienes and their Behaviour in Diels-Alder Cycloadditions. Tetrahedron: Asymmetry 2000, 11, 1217−1225. (230) Carreño, M. C.; Cid, M. B.; García-Ruano, J. L.; Santos, M. DielsAlder Reactions of [(S)R]-(1E,3E)-l-p-Tolylsulfinyl-l,3-pentadiene with Monosubstituted Ethylenes. Tetrahedron: Asymmetry 1997, 8, 2093− 2097. (231) Blasco, S.; Carreño, M. C.; Cid, M. B.; García-Ruano, J. L.; Martín, M. R. Direct Enantioselective Approach to Oxazolo[4,5e]isoindoles from [(S)R]-1-Aminosubstituted-4-(p-tolylsulfinyl)-1,3butadienes. Tetrahedron: Asymmetry 1999, 10, 3473−3477. (232) Carreño, M. C.; Cid, M. B.; García-Ruano, J. L.; Santos, M. Short Enantioselective Approach to Substituted Triazolopyridazines from [(S)R]-1-(1E,3E)-1-p-Tolylsulfinyl-1,3-pentadiene. Tetrahedron Lett. 1998, 39, 1405−1408. (233) Arribas, C.; Carreño, M. C.; García-Ruano, J. L.; Rodríguez, J. F.; Santos, M.; Sanz-Tejedor, M. A. First Asymmetric Hetero Diels−Alder Reaction of 1-Sulfinyl Dienes with Nitroso Derivatives. A New Entry to 14239

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

(253) Mapp, A. K.; Heathcock, C. H. Total Synthesis of Myxalamide A. J. Org. Chem. 1999, 64, 23−27. (254) Posner, G. H.; Carry, J.-C.; Crouch, R. D.; Johnson, N. OneFlask, Consecutive [3,3] and [2,3] Sigmatropic Rearrangements for Conversions of Propargylic Alcohols into Two-Carbon-Extended 4Oxo-2-alkenoate Esters. Use of a New 1-Chloro-1-ethoxy-2-sulfinylethylene. J. Org. Chem. 1991, 56, 6987−6993. (255) Wang, X.; Ni, Z.; Lu, X.; Hollis, A.; Banks, H.; Rodriguez, A.; Padwa, A. [2,3]-Sigmatropic Rearrangement of β-Phenylsulfonyl Propargylic Sulfenates as a Method for Preparing 1,4-Bis(phenylsulfonyl)-1,3-butadienes. J. Org. Chem. 1993, 58, 5377−5385. (256) Jeganathan, S.; Okamura, W. H. 2,3-Di(phenylsulfinyl)-1,3butadiene via a Double [2,3]-Sigmatropic Rearrangement of the BisSulfenate Ester of 2-Butyne-1,4-diol. Tetrahedron Lett. 1982, 23, 4763− 4764. (257) Pelc, M. J.; Zakarian, A. An Approach to the Imine Ring System of Pinnatoxins. Org. Lett. 2005, 7, 1629−1631. (258) Pelc, M. J.; Zakarian, A. Synthesis of the A,G-Spiroimine of Pinnatoxins by a Microwave-Assisted Tandem Claisen−Mislow−Evans Rearrangement. Tetrahedron Lett. 2006, 47, 7519−7523. (259) Ilardi, E. A.; Isaacman, M. J.; Qin, Y.-C.; Shelly, S. A.; Zakarian, A. Consecutive Sigmatropic Rearrangements in the Enantioselective Total Synthesis of (−)-Joubertinamine and (−)-Mesembrine. Tetrahedron 2009, 65, 3261−3269. (260) Hama, N.; Matsuda, T.; Sato, T.; Chida, N. Total Synthesis of (−)-Agelastatin A: The Application of a Sequential Sigmatropic Rearrangement. Org. Lett. 2009, 11, 2687−2690. (261) Fernández de la Pradilla, R.; Tortosa, M.; Lwoff, N.; del Á guila, M. A.; Viso, A. Sulfur-Directed Enantioselective Synthesis of Functionalized Dihydropyrans. J. Org. Chem. 2008, 73, 6716−6727. (262) Fernández de la Pradilla, R.; Lwoff, N.; del Á guila, M. A.; Tortosa, M.; Viso, A. [2,3]-Sigmatropic Rearrangements of 3-Sulfinyl Dihydropyrans: Application to the Syntheses of the Cores of entDysiherbaine and Deoxymalayamicin A. J. Org. Chem. 2008, 73, 8929− 8941. (263) Fernández de la Pradilla, R.; Lwoff, N.; Viso, A. Stereoselective Functionalization of Dihydropyran-3-ols: Application to the Synthesis of Enantiopure Ethyl Deoxymonate B. Eur. J. Org. Chem. 2009, 2312− 2322. (264) Simal, C.; Bates, R. H.; Ureña, M.; Giménez, I.; Koutsou, C.; Infantes, L.; Fernández de la Pradilla, R.; Viso, A. Synthesis of Enantiopure 3-Hydroxypiperidines from Sulfinyl Dienyl Amines by Diastereoselective Intramolecular Cyclization and [2,3]-Sigmatropic Rearrangement. J. Org. Chem. 2015, 80, 7674−7692. (265) Fernández de la Pradilla, R.; Colomer, I.; Ureña, M.; Viso, A. Enantiopure 1,4-Diols and 1,4-Aminoalcohols via Stereoselective Acyclic Sulfoxide−Sulfenate Rearrangement. Org. Lett. 2011, 13, 2468−2471. (266) Fernández de la Pradilla, R.; Velado, M.; Colomer, I.; Simal, C.; Viso, A.; Gornitzka, H.; Hemmert, C. Remote Stereocontrol in the Synthesis of Acyclic 1,4-Diols and 1,4-Aminoalcohols from 2-Sulfinyl Dienes. Org. Lett. 2014, 16, 5200−5203. (267) Chuard, R.; Giraud, A.; Renaud, P. Allyl Sulfoxides as Precursors for Radical Two-Carbon Ring Expansion of Cyclobutanones. Angew. Chem., Int. Ed. 2002, 41, 4323−4325. (268) Makisumi, Y.; Takada, S. Synthesis of Condensed Thiophens via [2,3] and [3,3] Sigmatropic Rearrangements of Aryl Prop-2-ynyl Sulfoxides. J. Chem. Soc., Chem. Commun. 1974, 848−849. (269) Majumdar, K. S.; Thyagarajan, B. S. Unusual ortho-Claisen Rearrangements of Aryl Prop-2-ynyl Sulfoxides. J. Chem. Soc., Chem. Commun. 1972, 83−84. (270) Aoyagi, S.; Makabe, M.; Shimada, K.; Takikawa, Y.; Kabuto, C. Novel Generation of (α-Ketovinyl)thioketenes as Intermediates through Tandem [2,3]/[3,3] Sigmatropic Rearrangement of Alkynyl Propargyl Sulfoxides. Tetrahedron Lett. 2007, 48, 4639−4642. (271) Baudin, J. B.; Julia, S. A.; Lorne, R. Sulfoxydes Propargyliques Substitués: Préparation et Étude de leur Transposition Sigmatropique [2,3] en Présence d’Agents Thiophiles. Bull. Soc. Chim. Fr. 1992, 129, 440−456.

(272) Raghavan, S.; Samanta, P. K. Stereoselective Synthesis of the C13−C28 Subunit of (−)-Laulimalide Utilizing an α-Chlorosulfide Intermediate. Synlett 2013, 24, 1983−1987. (273) Raghavan, S.; Patel, J. S. A Stereoselective Synthesis of the Carbon Backbone of Phoslactomycin B. Eur. J. Org. Chem. 2017, 2981− 2997. (274) Braverman, S.; Stabinsky, Y. The Rearrangement of Propargylic Trichloromethanesulfenates. Isr. J. Chem. 1967, 5, 125−126. (275) Cinquini, M.; Colonna, S.; Stirling, C. J. M. Asymmetric Induction and Racemisation in Allenic Sulfoxides. J. Chem. Soc., Chem. Commun. 1975, 256−257. (276) Cinquini, M.; Colonna, S.; Cozzi, F.; Stirling, C. J. M. Asymmetric Induction and Racemisation in Allenic Sulfoxides. J. Chem. Soc., Perkin Trans. 1 1976, 2061−2067. (277) Van Kruchten, E. M. G. A.; Okamura, W. H. On the [2,3]Sigmatropic Rearrangements of Sulfenate Esters Derived from Alkenynols: Synthesis of Vinylallene and Vinylacetylene Sulfoxides. Tetrahedron Lett. 1982, 23, 1019−1022. (278) Cutting, I.; Parsons, P. J. New Transformations of Substituted Allene Sulfoxides. Tetrahedron Lett. 1983, 24, 4463−4464. (279) Wang, M.; Fu, C.; Ma, S. Cleavage of Oxygen-sulfur Double Bonds and Carbon-sulfur Bonds: Unusual Highly Reactive Electrophilic Addition of Allenylic Sulfoxides. Chem. Sci. 2013, 4, 1016−1022. (280) Cookson, R. C.; Parsons, P. J. Convenient Synthesis of Allenyl Sulphides; Application to the Synthesis of α,β-Unsaturated Ketones. J. Chem. Soc., Chem. Commun. 1978, 822−824. (281) Bridges, A. J.; Fischer, J. W. Highly Substituted Dienes from Butyne-1,4-diol. A One-flask Synthesis and the Mechanism of Formation of 3-(N,N-Diethylamino)-2-phenylsulphinylbutadiene. J. Chem. Soc., Chem. Commun. 1982, 665−666. (282) Christov, V. Ch.; Ivanov, I. K. Alkatrienyl Sulfoxides and Sulfones. Part VI. Cheletropic Addition of Sulfur Dioxide to 1- and 3Vinylallenyl Sulfoxides and Sulfones. Heterocycles 2004, 63, 2203−2206. (283) Altenbach, H.-J.; Soicke, H. Phosphor- und Schwefelsubstituierte Allene in der Synthese, III. β-Ketosulfoxide aus 3-Alkin1-olen über Allensulfoxide. Liebigs Ann. Chem. 1982, 1096−1104. (284) Cookson, R. C.; Gopalan, R. A New Synthesis of Conjugated Dienones. J. Chem. Soc., Chem. Commun. 1978, 608. (285) Cutting, I.; Parsons, P. J. Chemistry of Allene Sulfoxides: Novel Transformations of Acetylenic Alcohols. J. Chem. Soc., Chem. Commun. 1983, 1209−1210. (286) Pairaudeau, G.; Parsons, P. J.; Underwood, J. M. Chemistry of Allene Sulfoxides: a New Methodology for the Synthesis of Spiroketals and 1,5-Dioxadecalin Ring Systems. J. Chem. Soc., Chem. Commun. 1987, 1718−1720. (287) Gray, M.; Parsons, P. J.; Neary, A. P. A New Synthesis of Indoles Involving a Tandem Sigmatropic Rearrangement/Michael Addition Sequence. Synlett 1993, 281−282. (288) Christov, V. C.; Ivanov, I. K. Alkatrienyl Sulfoxides and Sulfones. Part I. 3-Methyl-1,2,4-pentatrienyl Phenyl Sulfoxide-Synthesis and Electrophile-Induced Cyclization Reactions. Phosphorus, Sulfur Silicon Relat. Elem. 2002, 177, 2445−2454. (289) Christov, V. C.; Ivanov, I. K. Alkatrienyl Sulfoxides and Sulfones. Part II. (5-Methyl-1,3,4-hexatriene-3-yl) Phenyl Sulfoxide-Synthesis and Electrophile-Promoted Cyclization Reactions. Sulfur Lett. 2002, 25, 191−198. (290) Ivanov, I. K.; Parushev, I. D.; Christov, V. C. Bifunctionalized Allenes, Part IX: An Efficient Method for Regioselective Synthesis of 4Heteroatom-Functionalized Allenecarboxylates. Heteroat. Chem. 2013, 24, 322−331. (291) Ivanov, I. K.; Parushev, I. D.; Christov, V. C. Bifunctionalized Allenes. Part XII. Electrophilic Cyclization and Addition Reactions of 4Sulfinylated or 4-Sulfonylated Allenecarboxylates. Phosphorus, Sulfur Silicon Relat. Elem. 2014, 189, 1503−1513. (292) Christov, V. C.; Prodanov, B. Bifunctionalized Allenes. VI. Synthesis of 3-Sulfinylfuran-2(5H)-ones and 5H-1,2-Oxathiole-3carboxylate 2-Oxides by Bromination of 2-Sulfinyl-2,3-alkadienoates. Synth. Commun. 2004, 34, 1577−1588. 14240

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

Chirality Transfer: Synthesis of Drimatrienes and Related transDecalins. J. Am. Chem. Soc. 1985, 107, 1034−1041. (314) Okamura, W. H.; Shen, G.-Y.; Tapia, R. Sulfoxide Induced Acceleration and Enhancement of Geometric Selectivity of a [1,5]Sigmatropic Hydrogen Shift. J. Am. Chem. Soc. 1986, 108, 5018−5019. (315) Shen, G.-Y.; Tapia, R.; Okamura, W. H. Substituent Effect Studies on the Thermal [1,5]-Sigmatropic Hydrogen Shifts of Vinylallenes. J. Am. Chem. Soc. 1987, 109, 7499−7506. (316) Haces, A.; van Kruchten, E. M. G. A.; Okamura, W. H. On the Lithium Dimethylcuprate Induced Conversion of Propargylic Esters into Allenes Using a C/D Fragment Derived from Vitamin D3 as a Stereochemical Probe. Tetrahedron Lett. 1982, 23, 2707−2710. (317) Elnagar, H. Y.; Okamura, W. H. Thermal [1,7]-Sigmatropic Hydrogen Shift versus Electrocyclization of Dienylallene Sulfoxides. J. Org. Chem. 1988, 53, 3060−3066. (318) Reischl, W.; Okamura, W. H. Electrocyclization of 1Allenyldienes: Novel Synthesis of Drimatrienes and Related trans− Decalins. J. Am. Chem. Soc. 1982, 104, 6115−6117. (319) Miller, J. G.; Kurz, W.; Untch, K. G.; Stork, G. Highly Stereoselective Total Syntheses of Prostaglandins via Stereospecific Sulfenate−Sulfoxide Transformations. 13-cis-l5β-Prostaglandins E1 to Prostaglandins El. J. Am. Chem. Soc. 1974, 96, 6774−6775. (320) Taber, D. F. Cyclopentanone Ring Formation with Control of Side Chain Stereochemistry. A Simple Stereoselective Route to the Prostaglandins. J. Am. Chem. Soc. 1977, 99, 3513−3514. (321) Kondo, K.; Umemoto, T.; Takahatake, Y.; Tunemoto, D. Stereoselective Synthesis of 11-Deoxyprostaglandin El. Tetrahedron Lett. 1977, 18, 113−116. (322) Tunemoto, D.; Takahatake, Y.; Kondo, K. Stereoselective Synthesis of 10-Oxa-11-deoxy-prostaglandin E1. Chem. Lett. 1978, 7, 189−190. (323) Kondo, K.; Umemoto, T.; Yako, K.; Tunemoto, D. A New Stereoselective Synthesis of dl-Prostaglandin F2α. Tetrahedron Lett. 1978, 19, 3927−3930. (324) Nokami, J.; Ono, T.; Iwao, A.; Wakabayashi, S. Regioselective Reaction of Allylic Carbanion Derived from 1-Phenylthio-2-octene with Conjugated Cyclopentenone Derivative and Synthesis of Prostaglandin E1. Bull. Chem. Soc. Jpn. 1982, 55, 3043−3044. (325) Taber, D. F.; Hoerrner, R. S. Enantioselective Rh-Mediated Synthesis of (−)-PGE2 Methyl Ester. J. Org. Chem. 1992, 57, 441−447. (326) Taber, D. F.; Herr, R. J.; Gleave, D. M. Diastereoselective Synthesis of an Isoprostane: (±)-8-epi-PGF2α Ethyl Ester. J. Org. Chem. 1997, 62, 194−198. (327) Taber, D. F.; Kanai, K. A Synthetic Approach to 15-D2cIsoprostane Ethyl Ester. J. Org. Chem. 1998, 63, 6607−6609. (328) Taber, D. F.; Kanai, K. Synthesis of the Four EnantiomericallyPure Isomers of 15-F2t-Isoprostane. Tetrahedron 1998, 54, 11767− 11782. (329) Taber, D. F.; Kanai, K.; Pina, R. 5-F2t-Isoprostane, A Human Hormone? J. Am. Chem. Soc. 1999, 121, 7773−7777. (330) Taber, D. F.; Kanai, K. Synthesis of 2,3-Dinor-5,6-dihydro-15F2tisoprostane. J. Org. Chem. 1999, 64, 7983−7987. (331) Taber, D. F.; Jiang, Q. Total Synthesis of the Four Enantiomerically Pure Diastereomers of 8-F2t-Isoprostane. J. Org. Chem. 2001, 66, 1876−1884. (332) Taber, D. F.; Xu, M.; Hartnett, J. C. Synthesis of the Eight Enantiomerically Pure Diastereomers of the 12-F2-Isoprostanes. J. Am. Chem. Soc. 2002, 124, 13121−13126. (333) Taber, D. F.; Hoerrner, R. S.; Herr, R. J.; Gleave, D. M.; Kanai, K.; Pina, R.; Jiang, Q.; Xu, M. The Diazo Ketone Approach to the Isoprostanes. Chem. Phys. Lipids 2004, 128, 57−67. (334) Taber, D. F.; Gu, P. Preparation of the Major Urinary Metabolite of (−)-Prostaglandin E2. Tetrahedron 2009, 65, 5904−5907. (335) Grieco, P. A.; Schillinger, W. J.; Yokoyama, Y. C(14)-Fluorinated Prostaglandins: Synthesis and Biological Evaluation of the Methyl Esters of (+)-14-Fluoro-, (+)-15-epi-14-Fluoro-, (+)-13(E)-14-Fluoro-, and (+)-13(E)-15-epi-14-Fluoroprostaglandin F2α. J. Med. Chem. 1980, 23, 1077−1083.

(293) Smith, G.; Stirling, C. J. M. Intramolecular Reactions. Part IX. Stereochemistry and Mechanism of the Rearrangement of Acetylenic Esters of Sulphinic and Sulphenic Acids to Allenes. J. Chem. Soc. C 1971, 1530−1535. (294) Bull, J. R.; Gordon, R.; Hunter, R. Intramolecular Diels−Alder Reactions of 1-Phenylsulfonylalka-1,2,(ω−3),(ω−1)-tetraenes. J. Chem. Soc., Perkin Trans. 1 2000, 3129−3139. (295) Christov, V. C.; Ivanov, I. K. Alkatrienyl Sulfoxides and Sulfones, VIII: Diels−Alder Reaction of 3-Methylpenta-1,2,4-trienyl Phenyl Sulfoxide with Maleic Anhydride. Synth. Synth. Commun. 2007, 37, 3201−3209. (296) Gibbs, R. A.; Okamura, W. H. A Short Enantioselective Synthesis of (+)-Sterpurene: Complete Intramolecular Transfer of Central to Axial to Central Chiral Elements. J. Am. Chem. Soc. 1988, 110, 4062− 4063. (297) Braverman, S.; Pechenick, T.; Gottlieb, H. E.; Sprecher, M. A Convenient Preparation of Mixed Allylic and Allenic Thiosulfinates. Tetrahedron Lett. 2004, 45, 8235−8238. (298) Braverman, S.; Pechenick, T.; Gottlieb, H. E.; Sprecher, M. New Structures from Multiple Rearrangements of Propargylic Dialkoxy Disulfides. J. Am. Chem. Soc. 2003, 125, 14290−14291. (299) Braverman, S.; Pechenick, T.; Gottlieb, H. E. Tandem Sigmatropic Rearrangements and Cyclizations of Propargylic Dialkoxy Disulfides. Tetrahedron Lett. 2003, 44, 777−780. (300) Birsa, M. L.; Hopf, H. Synthesis of α,β-Unsaturated Pseudogeminal [2.2]Paracyclophane Bisketones. Synlett 2007, 2753− 2755. (301) Birsa, M. L.; Jones, P. G.; Braverman, S.; Hopf, H. PseudoGeminally-Substituted [2.2]Paracyclophanes as Spacers for Bisallenyl Sulfoxides and Sulfones. Synlett 2005, 640−642. (302) Birsa, M. L.; Hopf, H. Pseudo-Geminal [2.2]-Paracyclophane as Spacer for Bisallenyl Sulfoxides and Sulfones. Phosphorus, Sulfur Silicon Relat. Elem. 2005, 180, 1453−1454. (303) Baudin, J.-B.; Bkouche-Waksman, I.; Julia, S. A.; Pascard, C.; Wang, Y. Stereochemical Course of the [2,3]-Sigmatropic Rearrangement of Substituted Propargyl N,N-Dialkylamidosulfoxylates. X-ray Molecular Structure of [S*,(S)R*]-4-[1,4,4-Trimethyl-1,2pentadienyl)sulfinyl]-morpholine. Tetrahedron 1991, 47, 3353−3364. (304) Baudin, J.-B.; Julia, S. A.; Wang, Y. 4-(Alca-1,2-diénylsulfinyl)morpholines substituées: Préparations, Passage aux Dérivés Sulfiniques Correspondants: Esters ou Acides Précurseurs d’alcynes. Bull. Soc. Chim. Fr. 1995, 132, 739−753. (305) Djuric, S. W.; Miyano, M.; Clare, M.; Rydzewski, R. M. A Stereocontrolled Synthesis of a Novel Prostacyclin Analog “Allenecarbacyclin”. Application of Molecular Mechanics Calculations to Organic Synthesis. Tetrahedron Lett. 1987, 28, 299−302. (306) Neef, G.; Eder, U.; Seeger, A. Reaction of Unsaturated Sulfoxides with Alkyllithiums. Tetrahedron Lett. 1980, 21, 903−906. (307) Okamura, W. H. Pericyclic Reactions of Vinylallenes: From Calciferols to Retinoids and Drimanes. Acc. Chem. Res. 1983, 16, 81−88. (308) VanRheenen, V.; Shephard, K. P. New Synthesis of Cortico Steroids from 17-Keto Steroids: Application and Stereochemical Study of the Unsaturated Sulfoxide−Sulfenate Rearrangement. J. Org. Chem. 1979, 44, 1582−1584. (309) Lu, H.-L.; Wu, Z.-W.; Song, S.-Y.; Liao, X.-D.; Zhu, Y.; Huang, Y.-S. An Improved Synthesis of Nomegestrol Acetate. Org. Process Res. Dev. 2014, 18, 431−436. (310) Altenbach, H.-J.; Soicke, H. Phosphor- und Schwefelsubstituierte Allene-IV. Eine Einfache Synthese von 2[5H]Furanonen und βMethylen-γ-butyrolactonen durch Addition von Malonsäureester an Allensulfoxide. Tetrahedron Lett. 1986, 27, 1561−1562. (311) Iglesias, B.; Torrado, A.; de Lera, A. R.; López, S. A Pericyclic Cascade to the Stereocontrolled Synthesis of 9-cis-Retinoids. J. Org. Chem. 2000, 65, 2696−2705. (312) de Lera, A. R.; Castro, A.; Torrado, A.; López, S. Stereoselective Isomerization of 10-Arylsulfenate-11,12-Dehydroretinoids to 9-cisRetinoids. Tetrahedron Lett. 1998, 39, 4575−4578. (313) Okamura, W. H.; Peter, R.; Reischl, W. Allenyldiene Electrocyclization. A Stereospecific Tandem Center-Axis-Center 14241

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

(336) Donaldson, R. E.; Saddler, J. C.; Byrn, S.; McKenzie, A. T.; Fuchs, P. L. A Triply Convergent Total Synthesis of L-(−)-Prostaglandin E2. J. Org. Chem. 1983, 48, 2167−2188. (337) Kojima, K.; Koyama, K.; Amemiya, S. A Novel Synthesis of Carbacyclin Analogs via a Stereoselective Introduction of 15α-Hydroxy Group. Tetrahedron 1985, 41, 4449−4462. (338) Bannai, K.; Toru, Y.; Oba, T.; Tanaka, T.; Okamura, N.; Watanabe, K.; Hazato, A.; Kurozumi, S. Synthesis of Chemically Stable Prostacyclin Analogs. Tetrahedron 1983, 39, 3807−3819. (339) Mori, S.; Takechi, S. Synthesis of Benzodioxane Prostacyclin Analogue. Heterocycles 1990, 31, 1189−1193. (340) Dyong, I.; Schulte, G. Derivate des Sibirosamins (4,6-Didesoxy3-C-methyl-4-methyl-amino-D-altropyranose) durch 4 → 2-Chiralitätstransfer und vic. cis-Oxyaminierung eines 3-C-Methyl-hex-3-enopyranosids. Chem. Ber. 1981, 114, 1484−1502. (341) Wittman, M. D.; Halcomb, R. L.; Danishefsky, S. J.; Golik, J.; Vyas, D. A Route to Glycals in the Allal and Gulal Series: Synthesis of the Thiosugar of Esperamicin A1. J. Org. Chem. 1990, 55, 1979−1981. (342) Halcomb, R. L.; Wittman, M. D.; Olson, S. H.; Danishefsky, S. J.; Golik, J.; Wong, H.; Vyas, D. Synthesis of the Core Trisaccharide of Esperamicin: Corroboration of the Proposed Structure for Its Rearrangement Product and Stabilization of the Core Trisaccharide Domain. J. Am. Chem. Soc. 1991, 113, 5080−5082. (343) Halcomb, R. L.; Boyer, S. H.; Wittman, M. D.; Olson, S. H.; Denhart, D. J.; Liu, K. K. C.; Danishefsky, S. J. Studies Related to the Carbohydrate Sectors of Esperamicin and Calicheamicin: Definition of the Stability Limits of the Esperamicin Domain and Fashioning of a Glycosyl Donor from the Calicheamicin Domain. J. Am. Chem. Soc. 1995, 117, 5720−5749. (344) Griffith, D. A.; Danishefsky, S. J. The Total Synthesis of Allosamidin. Expansions of the Methodology of Azaglycosylation Pursuant to the Total Synthesis of Allosamidin. A Surprising Enantiotopic Sense for a Lipase-Induced Deacetylation. J. Am. Chem. Soc. 1996, 118, 9526−9538. (345) Engstrom, K. M.; Mendoza, M. R.; Navarro-Villalobos, M.; Gin, D. Y. Total Synthesis of (+)-Pyrenolide D. Angew. Chem., Int. Ed. 2001, 40, 1128−1130. (346) Kyler, K. S.; Watt, D. S. Partial Synthesis of (20R,22R)-20,22Dihydroxycholesterol. J. Org. Chem. 1981, 46, 5182−5188. (347) Hirayama, M.; Gamoh, K.; Ikekawa, N. Stereoselective Synthesis of Withaferin A and 27-Deoxywithaferin A. Tetrahedron Lett. 1982, 23, 4725−4728. (348) Bull, J. R.; Bischofberger, K. 14-Methyl Steroids. Part 2. Total Synthesis of (±)-14α-Methyl-19-norsteroids. J. Chem. Soc., Perkin Trans. 1 1983, 2723−2727. (349) Trost, B. M.; Schmuff, N. R. Stereochemistry of Allyl Sulfones. On the Structure of Metalated Allyl Sulfones and Their Stereochemistry of Alkylation. J. Am. Chem. Soc. 1985, 107, 396−405. (350) Wang, W.-Y.; Reusch, W. A Model for Triterpene Side-Chain Synthesis. Tetrahedron 1988, 44, 1007−1014. (351) Morera, E.; Ortar, G. Preparative and Stereochemical Features of the Sulfoxide−Sulfenate [2,3] Sigmatropic Rearrangement in 17-Vinyl17-hydroxy Steroids. J. Org. Chem. 1983, 48, 119−121. (352) Takahashi, T.; Shimizu, K.; Doi, T.; Tsuji, J.; Fukazawa, Y. Macroring Contraction Methodology. 4. A Novel Route to Steroid A, B, C Rings by the Transannular Diels-Alder Reaction of the 14-Membered (E,E,E)-Macrocyclic Triene. J. Am. Chem. Soc. 1988, 110, 2674−2676. (353) Zhou, W.-S.; Shen, Z.-W. Study on the Synthesis of Brassinolide and Related Compounds. Part 15. Formal Synthesis of Brassinolide via Stereoselective Sulphenate-Sulfoxide Transformation. J. Chem. Soc., Perkin Trans. 1 1991, 2827−2830. (354) Boivin, J.; Chauvet, C.; Zard, S. Z. A Convenient Access to 16Methylene and 16β-Methyl Corticosteroids by a Novel Transformation of Geminal Bis(Phenylsulphinyl) Intermediates. Tetrahedron Lett. 1992, 33, 4913−4916. (355) Tuba, Z.; Bardin, C. W.; Dancsi, A.; Francsics-Czinege, E.; Molnár, C.; Csörgei, J.; Falkay, G.; Koide, S. S.; Kumar, N.; Sundaram, K.; Dukát−Abrók, V.; Balogh, G. Synthesis and Biological Activity of a

New Progestogen, 16-Methylene-17α-hydroxy-18-methyl-19-norpregn-4-ene-3,20-dione Acetate. Steroids 2000, 65, 266−274. (356) Debien, L.; Zard, S. Z. A Two-Step Synthesis of α-Keto Vinyl Carbinols from Ketones. Org. Lett. 2013, 15, 6066−6069. (357) Braun, M.-G.; Zard, S. Z. An Approach to α-Keto Vinyl Carbinols. Org. Lett. 2011, 13, 776−779. (358) Hua, D. H.; Venkataraman, S.; Chan-Yu-King, R.; Paukstelis, J. V. Enantioselective Total Synthesis of (+)-12,13-Epoxytrichothec-9-ene and Its Antipode. J. Am. Chem. Soc. 1988, 110, 4741−4748. (359) Majetich, G.; Song, J.-S.; Ringold, C.; Nemeth, G. A.; Newton, M. G. Intramolecular Additions of Allylsilanes to Conjugated Dienones. A Direct Stereoselective Synthesis of (±)-14-Deoxyisoamijiol. J. Org. Chem. 1991, 56, 3973−3988. (360) Boeckman, R. K., Jr.; Springer, D. M.; Alessi, T. R. Synthetic Studies Directed toward Naturally Occurring Cyclooctanoids. 2. A Stereocontrolled Assembly of (±)-Pleuromutilin via a Remarkable Sterically Demanding Oxy-Cope Rearrangement. J. Am. Chem. Soc. 1989, 111, 8284−8286. (361) Keck, G. E.; Savin, K. A.; Weglarz, M. A.; Cressman, E. N. K. Synthetic Studies on the Rhizoxins. II. An Approach to the C10-C26 Subunit Using ″Substrate Directed″ Allylstannane Additions to Aldehydes. Tetrahedron Lett. 1996, 37, 3291−3294. (362) Deng, K.; Chalker, J.; Yang, A.; Cohen, T. Allyl Sulfones as Precursors to Allylzincs in the Palladium-Catalyzed Zinc-ene Cyclization: Highly Efficient Synthesis of Enantiopure (−)-Erythrodiene. Org. Lett. 2005, 7, 3637−3640. (363) He, C.; Zhu, C.; Dai, Z.; Tseng, C.-C.; Ding, H. Divergent Total Synthesis of Indoxamycins A, C, and F. Angew. Chem., Int. Ed. 2013, 52, 13256−13260. (364) Ideue, E.; Shimokawa, J.; Fukuyama, T. Synthesis of the Common Core Structure of the Stemofoline Alkaloids. Org. Lett. 2015, 17, 4964−4967. (365) Masaki, Y.; Hashimoto, K.; Sakuma, K.; Kaji, K. Facile Regio- and Stereo-specific Allylic Oxidation of gem-Dimethyl Olefins via Addition of Benzenesulphenyl Chloride. Synthesis of Allylic Oxygenated Terpenes. J. Chem. Soc., Perkin Trans. 1 1984, 1289−1295. (366) Masaki, Y.; Hashimoto, K.; Kaji, K. Facile Functionalization of the Isopropylidene Terminus of Acyclic Monoterpenes by Way of Benzenesulfenyl Chloride Addition. Tetrahedron 1984, 40, 3481−3490. (367) Masaki, Y.; Hashimoto, K.; Sakuma, K.; Kaji, K. Synthetic Studies on Isoprenoidquinones. I. A Facile Regio- and Stereocontrolled Synthesis of Protected Hydroquinones with an Omega-Hydroxyprenyl or Omega-Hydroxygeranyl Side Chain. Chem. Pharm. Bull. 1984, 32, 3952−3958. (368) Iio, H.; Isobe, M.; Kawai, T.; Goto, T. Total Synthesis of Vernolepin. 2. Stereocontrolled Synthesis of (±)-Vernolepin. J. Am. Chem. Soc. 1979, 101, 6076−6081. (369) Isobe, M.; Iio, H.; Kitamura, M.; Goto, T. Substituent Effects on the Ratio of [2,3]-Sigmatropic Rearrangement vs syn-Elimination of Allylic Aryl Sulfoxides. Chem. Lett. 1978, 7, 541−544. (370) Zutterman, F.; De Wilde, H.; Mijngheer, R.; De Clercq, P.; Vandewalle, M. Sesquiterpene Lactones. A Total Synthesis of (±)-Vernolepin. Tetrahedron 1979, 35, 2389−2396. (371) Schultz, A. G.; Godfrey, J. D. An Annelation Approach to the Synthesis of Eudesmane and Elemane Sesquiterpene Lactones. Total Synthesis of dl-Dihydrocallitrisin, dl-7,8-Epialantolactone, dl-7,8Epiisoalantolactone and dl-Atractylon. J. Am. Chem. Soc. 1980, 102, 2414−2428. (372) Mori, K.; Ueda, H. Synthesis of the Optically Active Forms of 2,6-Dimethyl-1,5-heptadien-3-ol Acetate, the Pheromone of Comstock Mealybug. Tetrahedron 1981, 37, 2581−2583. (373) Kodama, M.; Okumura, K.; Kobayashi, Y.; Tsunoda, T.; Itô, S. Synthesis of Macrocyclic Terpenoids by Intramolecular Cyclization IX. Total Synthesis of (±)-Obscuronatin and (±)-Biflora-4,10(19),15Triene. Tetrahedron Lett. 1984, 25, 5781−5784. (374) Corbet, J.-P.; Benezra, C. Allergenic α-Methylene-γ-lactones. General Method for the Preparation of β-Acetoxy- and β-Hydroxy-αmethylene-γ-butyrolactones from Sulfoxides. Application to the Synthesis of a Tuliposide B Derivative. J. Org. Chem. 1981, 46, 1141−1147. 14242

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243

Chemical Reviews

Review

(375) Wasserman, H. H.; Han, W. T. Penem Synthesis through C3-N Ring Closure of a β-Lactam Precursor. J. Am. Chem. Soc. 1985, 107, 1444−1446. (376) Daniels, N. J.; Johnson, G.; Ross, B. C.; Yeomans, M. A. Use of the Allylic Sulfoxide−Sulphenate Ester Rearrangement for the Synthesis of a 2-Thiacephem and a Penem. J. Chem. Soc., Chem. Commun. 1982, 1119−1120. (377) Yanagisawa, H.; Ando, A. Conversion of Cepham-1-oxide into 1Oxadethiacepham. Tetrahedron Lett. 1982, 23, 3379−3382. (378) Rigby, J. H.; Senanayake, C. A Synthetic Entry into the Ophiobolane Ring System. J. Org. Chem. 1987, 52, 4634−4635. (379) Trost, B. M.; Matsuda, K. A Biomimetic Synthesis of (±)-Petiodial. A Novel Palladium-Catalyzed Enallene Cyclization. J. Am. Chem. Soc. 1988, 110, 5233−5235. (380) Vedejs, E.; Wittenberger, S. J. The Total Syntheses of dlZygosporin E and dl-C18-Desmethylcytochalasin D. J. Am. Chem. Soc. 1990, 112, 4357−4364. (381) Darvesh, S.; Grant, A. S.; Magee, D. I.; Valenta, Z. Synthetic Studies towards Bruceantin. Part 2. The Synthesis of a Pentacyclic Intermediate. Can. J. Chem. 1991, 69, 723−731. (382) Romo, D.; Johnson, D. D.; Plamondon, L.; Miwa, T.; Schreiber, S. L. Synthetic Investigations of Rapamycin. 2. Synthesis of a C22-C42 Fragment. J. Org. Chem. 1992, 57, 5060−5063. (383) St-Denis, Y.; Chan, T.-H. Synthesis of 1-Deoxycastanospermine and Stereoisomers. J. Org. Chem. 1992, 57, 3078−3085. (384) Kato, M.; Kido, F.; Watanabe, M.; Masuda, Y.; Awen, B. Z. Preparation of Key Intermediates for the Asymmetric Synthesis of Oxygenated Elemanoids. J. Chem. Soc., Perkin Trans. 1 1993, 2831− 2836. (385) Lambs, L.; Singh, N. P.; Biellmann, J.-F. Enantiospecific Synthesis of a Chrysanthemic Acid Precursor. J. Org. Chem. 1992, 57, 6301−6304. (386) Ho, P.-T. Synthesis of Maytansinoids: the Aromatic Moiety of Maytansine. Can. J. Chem. 1980, 58, 861−863. (387) Goodwin, T. E.; Ratcliff, D. G.; Crowder, C. M.; Seitzinger, N. K. Synthesis of (E)-1-Aryl-2-methyl-3-alkyl-2-propen-l-ones via Allylic Sulfoxide−Sulfenate Ester Rearrangements. J. Org. Chem. 1982, 47, 815−820. (388) Goodwin, T. E.; Orlicek, S. G.; Adams, N. R.; Covey-Morrison, L. A.; Jenkins, J. S.; Templeton, G. L. Preparation of an Aromatic Synthon for Maytansinoid Synthesis. J. Org. Chem. 1985, 50, 5889− 5892. (389) Goodwin, T. E.; Cousins, K. R.; Crane, H. M.; Eason, P. O.; Freyaldenhoven, T. E.; Harmon, C. C.; King, B. K.; LaRocca, C. D.; Lile, R. L.; Orlicek, S. G.; et al. Synthesis of Two New Maytansinoid Model Compounds from Carbohydrate Precursors. J. Carbohydr. Chem. 1998, 17, 323−339. (390) Crimmins, M. T.; Al-awar, R. S.; Vallin, I. M.; Hollis, W. G., Jr.; O’Mahony, R.; Lever, J. G.; Bankaitis-Davis, D. M. Asymmetric Total Synthesis of (+)-Milbemycin D. J. Am. Chem. Soc. 1996, 118, 7513− 7528. (391) Baba, Y.; Saha, G.; Nakao, S.; Iwata, C.; Tanaka, T.; Ibuka, T.; Ohishi, H.; Takemoto, Y. Asymmetric Total Synthesis of Halicholactone. J. Org. Chem. 2001, 66, 81−88. (392) Magnus, P.; Ollivier, C. Synthesis of the Hydroazulene Portion of Guanacastepene A Using a [2.3]Sigmatropic Sulfoxide Rearrangement: Observations on Silyl Enol Ether Electrophilic Chemistry for the Introduction of the C-13 Hydroxyl Group. Tetrahedron Lett. 2002, 43, 9605−9609. (393) Wipf, P.; Reeves, J. T. A Formal Total Synthesis of Leucascandrolide A. Chem. Commun. 2002, 2066−2067. (394) Shinada, T.; Fuji, T.; Ohtani, Y.; Yoshida, Y.; Ohfune, Y. Syntheses of Gabosine A, B, D, and E from Allyl Sulfide Derived from (−)-Quinic Acid. Synlett 2002, 1341−1343. (395) Tamura, S.; Fujiwara, K.; Shimizu, N.; Todo, S.; Murakami, N. Concise Synthesis of 5,6-Dihydrovaltrate Leading to Enhanced Revexport Inhibitory Congener. Bioorg. Med. Chem. 2010, 18, 5975−5980.

(396) Charest, M. G.; Lerner, C. D.; Brubaker, J. D.; Siegel, D. R.; Myers, A. G. A Convergent Enantioselective Route to Structurally Diverse 6-Deoxytetracycline Antibiotics. Science 2005, 308, 395−398. (397) Raghavan, S.; Chiluveru, R. K.; Subramanian, S. G. Stereoselective Formal Synthesis of (+)- and (−)-Cyclophellitol and (−)-Conduritol-B and Synthesis of (−)-Conduramine-B Derivative Using a Sulfinyl Moiety for C−O Bond Formation and α-Chloro Sulfide for C−C Bond Formation. J. Org. Chem. 2016, 81, 4252−4261. (398) Raghavan, S.; Yelleni, M. K. R. Application of Ru(II)-Catalyzed Enyne Cyclization in the Synthesis of Brefeldin A. J. Org. Chem. 2016, 81, 10912−10921.

14243

DOI: 10.1021/acs.chemrev.7b00428 Chem. Rev. 2017, 117, 14201−14243