Open access peer-reviewed chapter

Tandem-, Domino- and One-Pot Reactions Involving Wittig- and Horner-Wadsworth-Emmons Olefination

Written By

Fatima Merza, Ahmed Taha and Thies Thiemann

Submitted: 21 October 2016 Reviewed: 11 July 2017 Published: 20 December 2017

DOI: 10.5772/intechopen.70364

From the Edited Volume

Alkenes

Edited by Reza Davarnejad and Baharak Sajjadi

Chapter metrics overview

2,343 Chapter Downloads

View Full Metrics

Abstract

The Wittig olefination utilizing phosphoranes and the related Horner-Wadsworth-Emmons (HWE) reaction using phosphonates transform aldehydes and ketones into substituted alkenes. Because of the versatility of the reactions and the compatibility of many functional groups towards the transformations, both Wittig olefination and HWE reactions are a mainstay in the arsenal of organic synthesis. Here, an overview is given on Wittig- and Horner-Wadsworth-Emmons (HWE) reactions run in combination with other transformations in one-pot procedures. The focus lies on one-pot oxidation Wittig/HWE protocols, Wittig/HWE olefinations run in concert with metal catalyzed cross-coupling reactions, Domino Wittig/HWE—cycloaddition and Wittig-Michael transformations.

Keywords

  • Wittig olefination
  • one-pot reactions
  • Domino reactions
  • tandem reactions
  • Horner-Wadsworth-Emmons olefination

1. Introduction

The Wittig olefination utilizing phosphoranes and the related Horner-Wadsworth-Emmons (HWE) reaction using phosphonates transform aldehydes and ketones into substituted alkenes. Because of the versatility of the reactions and the compatibility of many functional groups in the transformations, both Wittig olefination and HWE reactions are a mainstay in the arsenal of organic synthesis. The mechanism of the Wittig olefination has been the subject of intense debate [1]. While initially it was supposed that all Wittig olefination reactions lead via 1,2-addition to betaine structures 4 as zwitterionic intermediates that would form oxaphosphetane 3 with a final release of alkene and phosphine oxide by ring opening (syn-cycloreversion process), it has been seen more recently that especially under salt-free, aprotic conditions, many ylides undergo a π2s/π2a [2+2]-cycloaddition with the carbonyl component leading to the oxaphosphetane 3 directly [2], which in certain cases can be in equilibrium with betaine structures 4 (Scheme 1). In HWE reactions, the deprotonated phosphonate 6a undergoes a nucleophilic addition to the carbonyl compound (e.g., 7), which usually is the rate limiting step [3]. The elimination to the final products proceeds through oxaphosphetane 9 (Scheme 2). The Wittig olefination has been used industrially in the synthesis of terpenoids [4]. Recently, a one-pot synthesis of the vasodilator and anti-platelet agent Beraprost sodium, a prostacyclin analog, was communicated with the HWE reaction as the key transformation with the idea of using the approach in an industrial synthesis of the pharmaceutical [5].

Scheme 1.

Schematic presentation of the reaction mechanism of the Wittig olefination.

Scheme 2.

General reaction mechanism of the HWE reaction.

For years after the discovery of the Wittig olefination [6, 7], most Wittig transformations were carried out under inert atmosphere using dry solvents such as THF [8], DME [9], diethyl ether [10] and benzene [11]. Later it was realized that stabilized and semi-stabilized Wittig reagents can be reacted in non-de-aerated solvents, where the solvents need not be dried specifically. Most of these conjugated Wittig reagents are thermally stable and tolerate water, air and mild oxidants, while maintaining reactivity towards aldehydes and often also towards ketones. This allows for a plethora of reaction conditions for many Wittig olefination reactions such as obviating solvents altogether [12, 13] or running the reactions in aqueous solutions [14, 15] or in mixed solvents [16]. Also, it permits one-pot transformations of Wittig olefinations in combination with other reactions, also because the stabilized and to some extent the semi-stabilized phosphoranes are inert to mild oxidizing and reducing agents. However, also with non-stabilized phosphoranes, where reactions have to be performed under the exclusion of air and moisture, Wittig reactions can be performed in conjunction with further transformations [17].

This chapter is to give an insight into the types of transformations that can be combined with Wittig- and Horner-Wadsworth-Emmons olefinations in Domino-, tandem and one-pot reaction strategies. These include the preparation of phosphoranes and their reaction in situ, one-pot oxidation of alcohols to aldehydes and Wittig-olefination, in situ-recycling of phosphine oxides and catalytic Wittig reactions, one-pot Wittig-olefination metal catalyzed CC bond forming reactions such as Suzuki-Miyaura, Sonogashira- and Heck reactions, Wittig and Horner-Emmons reactions in combination with polar cyclizations, Wittig-reactions carried out in combination with electrocyclic reactions, one-pot Wittig and Horner Emmons-addition reactions; cascade reactions featuring (triphenylphosphoranylidene)-ethenone and similar phosphoranes.

Advertisement

2. Wittig and Horner-Wadsworth-Emmons (HWE) olefination reactions with phosphoranes and phosphonates prepared in situ

Primarily, phosphoranes as Wittig reagents are prepared by the reaction of a triarylphosphine, usually triphenylphosphine, or, more seldom, a trialkylphosphine, and an alkyl halide with subsequent dehydrohalogenation of the triaryl(alkyl)alkylphosphonium halide produced. Non-stabilized Wittig reagents are not stable enough to be stored over longer periods of time; therefore, it is the norm that the Wittig-ylide is formed in situ from the oftentimes stable phosphonium salt, usually with a strong base, and then reacted directly with the carbonyl compound. In the case of stabilized phosphoranes, they are often stable enough to store, and the dehydrohalogenation necessitates only a weak base such as sodium carbonate or even sodium bicarbonate [18]. Nevertheless, this likewise allows the preparation of the phosphorane and the subsequent Wittig olefination in one pot [19], where even protic solvents can be used, such as water. Similarly, semi-stabilized phosphoranes can be obtained in situ from their respective phosphonium salts, also even in aqueous medium, where LiCl promotes the Wittig olefination and suppresses the decomposition of the phosphoranes [14, 15]. Furthermore, all the catalytic Wittig reactions (see below) rely on the fact that the phosphorane is produced in situ.

Perhaps more interesting is the one-pot reaction of an alkyl halide, a phosphine and a carbonyl compound (Scheme 3). This can be achieved by consecutive addition of the components, when one or more of the components are sensitive, or by mixing of all components simultaneously. A consecutive addition of components in one pot was pursued by McNulty and Das who reacted air-sensitive triethylphosphine with benzyl bromides to the respective benzyltriethylphosphonium bromides, which were transformed to the phosphoranes with aq. NaOH, before being reacted with benzaldehydes to give (E)-stilbenes in an aqueous Wittig olefination [20]. Here, the triethylphosphine oxide by-product is water soluble. This reaction procedure has been diversified further by a one-pot preparation of benzyltriethylphosphonium bromides from the air-stable triethylphosphine hydrobromide and benzyl alcohols and subsequent Wittig olefination with aromatic aldehydes in aqueous medium [21]. Simultaneous mixing of alkyl halide such as α-haloesters (e.g., 13), α-halonitriles, α-halocarbonyl compounds and α-alkyl-α-halocarbonyl compounds, triphenylphosphine (12), and carbonyl compound (e.g., 11, 15, 18) in the presence either of a base [17, 22, 23, 24, 25, 26] or an alkene [27] was shown to give α,β-unsaturated esters [17, 22, 23, 24, 25, 26, 27] (e.g., 14, 17, 19), α,β-unsaturated nitriles [23, 26] and enones [27], respectively (Scheme 3). Epoxides are stable under these reaction conditions as can be seen in the transformation of 18 to 19 (Scheme 1). A one-pot, fluoride catalyzed Wittig-olefination has also been devised, where ethyl bromoacetate is reacted with carbaldehydes in the presence of tri-n-butylphosphine and tetrabutylammonium fluoride (Bu4NF) to give (E)-configured α,β-unsaturated esters in good yield [28]. The synthesis of α,β-unsaturated esters has also been achieved from their alkyl halide and aldehyde constituents using tributylarsine [29] or a substituted triarylarsine instead of triphenylphosphine [30]. The use of tributylarsine in the presence triphenyl phosphite [29] led to the creation of a catalytic system which was developed further with one-pot transformations that were managed with catalytic amounts (2 mol%) of poly(ethylene glycol) and (PEG)-supported tellurides in the presence of K2CO3 as base [31, 32, 33, 34]. Also, micellar reaction systems such as micellar solutions of sodium dodecyl sulfate (SDS) in water have been used, in which Wittig olefinations were carried out between aldehydes and phosphoranes, synthesized in situ [35, 36]. A. Galante has per Wittig reactions in the fluorous phase with in situ pre-formed perfluorinated ylides [37].

Scheme 3.

In situ preparation of phosphoranes and subsequent Wittig olefination.

Traditionally, stabilized halophosphoranes have been prepared by the halogenation of the nonhalogenated parent phosphoranes and a subsequent dehydrohalogenation of the halogenated phosphonium salt obtained. Karama et al. have combined this in situ halogenation: dehydrohalogenation step with the Wittig reaction itself. Additionally, an in situ alcohol oxidation to provide the aldehyde starting material was integrated into many of these reaction sequences (Scheme 4) [38, 39, 40, 41, 42].

Scheme 4.

One-pot oxidation, halogenation, and Wittig reaction to 2-haloacrylates.

Advertisement

3. In situ alcohol oxidation—Wittig/HWE reactions; other in situ aldehyde preparations run with subsequent Wittig/HWE sequences in one pot

The tolerance of stabilized phosphoranes towards mild oxidants allows for the oxidation of an alcohol to an aldehyde and its Wittig reaction in one-pot (Schemes 5 and 6). As oxidants, activated MnO2 [43, 44, 45, 46], barium permanganate [47, 48], tetra-n-propylammonium perruthenate (TPAP)/N-methylmorpholine N-oxide (NMO) [49, 50, 51, 52, 53, 54] and TPAP/N,N,N′,N′-tetramethylenediamine dioxide (TMEDAO2) [55], o-iodoxybenzoic acid (IBX) [56, 57, 58], Dess-Martin periodinane [59, 60, 61], DMSO-oxalyl chloride (Swern conditions) [62, 63, 64], DMSO-SO3-pyridine (Parikh-Doering oxidation) [38, 39] or DMSO-SO3-triethylamine [65], pyridinium chlorochromate (PCC) or PCC/celite [66, 67, 68, 69] as well as pyridinium dichromate (PDC) [70] such as PDC encapsulated in sol gel [71] have been used. In addition, metal catalyzed aerobic oxidation reactions of aldehydes with concomitant olefination reactions are known, where [(eta-p-cymene)RuCl2]2 (27) [72], nanoparticulate ruthenium supported on highly porous aluminum oxyhydroxide [73] or on silica gel [74], and nickel nanoparticles [75, 76] (Scheme 6) have been used as catalyst in the case of a concomitant Wittig reaction and gold/palladium bimetallic nanoparticles in the case of a concomitant Horner-Wadsworth-Emmons (HWE) reaction [77], Cu(I)-phenanthroline as a catalyst in an oxidation: HWE: sequential procedure [78].

Scheme 5.

One-pot MnO2-mediated oxidation—Wittig olefination.

Scheme 6.

One-pot metal catalyzed oxidation of alcohols utilizing oxygen—Wittig reaction.

Taylor et al. give a good overview of the tandem oxidation-Wittig processes developed until 2005, focusing especially on the tandem oxidation process (TOP) developed by his group [43, 44, 45, 46], using activated MnO2 [79]. Over the years, this process has been used more often [40, 80, 81, 82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97] than the other processes shown above. Recently, also MnO2 derived molecular sieve material such as OMS-2 [KMn4+Mn3+O16·nH2O] has been used with success in aerobic, catalytic one-pot oxidation Wittig reactions of benzylic and allylic alcohols to the respective cinnamates [98]. Overall, the Wittig transformations of the aldehydes produced in situ allows for the manipulation of aldehydes that are inherently instable such as of silyl substituted aldehydes, propargyl aldehyde [97], and chiral γ-aminoaldehydes, the latter without loss of stereochemical integrity (Scheme 5) [89]. In the case of Wittig transformations of chiral α-aminoaldehydes, β-aminoalcohols were oxidized to α-aminoaldehydes with NaOCl in the presence of AcNH-TEMPO, where the crude α-aminoaldehydes gained from the oxidation were subjected directly to olefination to give Wittig products without loss of stereochemical integrity [99, 100, 101].

Other preparation methods of aldehydes in conjunction with Wittig olefinations or HWE reactions have been reported. Thus, an oxidative cleavage of a glycol can be carried out in combination with a subsequent Wittig-olefination [102, 103, 104, 105] (Scheme 7). Also a one-pot carboxylic acid to aldehyde reduction and Wittig reaction is known [106]. Finally, a Domino hydroformylation/Wittig olefination procedure has been developed, starting from allylamines (Scheme 8). The aldehyde is not isolated [107]. Domino/hydroformylation/Wittig olefination protocols have been introduced with other olefinic starting materials, also [108, 109, 110].

Scheme 7.

Oxidative glycol cleavage—Wittig reaction.

Scheme 8.

Hydroformylation—Wittig reaction.

Advertisement

4. Wittig- and HWE reactions and CC-coupling reactions in one-pot procedures

Wittig- and Horner-Wadsworth-Emmons reactions can be combined with C─C-coupling reactions such as Suzuki cross-coupling [111, 112, 113], Mizoroki-Heck reaction [113, 114, 115, 116, 117, 118] and Sonogashira-reaction [119]. Initially, it was observed that conjugated phosphoranes were stable under reaction conditions used for Heck- or Suzuki reactions (Scheme 9). Thus, phosphoranes themselves could be functionalized by Suzuki- [120], Mizoroki-Heck- [121], or Sonogashira-type [119] cross-coupling reactions, either in solution or when polymer-bound [122]. These phosphoranes could then be subjected to normal Wittig-olefination reactions with ketones or aldehydes [120, 121, 122]. The one-pot Wittig-Heck-reaction strategy can be extended to include an O-alkylation, where the Wittig reaction of a p-hydroxybenzaldehyde (43) with methylenetriphenylphosphorane, obtained in situ from phosphonium salt 44 provides the p-hydroxystyrene as the olefin component in the Mizoroki-Heck reaction in the presence of an alkyl bromide (e.g., 45), which O-alkylates the phenoxy-function to give alkoxystilbenes 46 (Scheme 9) [123].

Scheme 9.

One-pot Heck cross-coupling/Wittig reaction.

Advertisement

5. One-pot Wittig- and HWE olefination/cycloaddition reaction

One can easily visualize that an alkene prepared by a Wittig olefination can easily be used as a 2-pi component in cycloaddition reactions, in one pot (Scheme 10). A typical such cycloaddition is [4+2]-cycloaddition, such as the classical Diels Alder reaction, which can be performed both inter-[69, 124] and intramolecularly [125, 126, 127, 128, 129, 130, 131, 132] in tandem with a Wittig-reaction.

Scheme 10.

Oxidation—Wittig-olefination—Diels-Alder reaction sequence.

Hilt and Hengst have published a cobalt(I)-catalyzed Diels Alder reaction of alkynyltriphenylphosphonium and 1,3-dienes with a consecutive Wittig reaction of the cycloadduct with various aldehydes in one pot that lead after a further dehydrogenative step to substituted stilbenes and styrenes (Scheme 11) [133].

Scheme 11.

Cobalt (I)-catalyzed Diels Alder reaction—Wittig reaction.

Interesting is the cycloaddition of in situ produced benzyne (55) to 1,4-diphenylbutadiene, prepared in situ by HWE reaction from cinnamaldehyde, (15) give 1,4-diphenylnaphthalene (56) (Scheme 12) [134].

Scheme 12.

One-pot HWE reactioncycloaddition of in situ produced benzyne.

The transformation sequence Diels-Alder/Wittig can be part of a more complex reaction chain. Thus, Ramachary and Barbas III [135] have forwarded a Domino Wittig/Knoevenagel/Diels-Alder sequence to spirotriones 58 (Scheme 13) and a Wittig/Knoevenagel/Diels-Alder/Huisgen cycloaddition sequence to polysubstituted triazoles 61 (Scheme 14).

Scheme 13.

Domino Wittig/Knoevenagel/Diels-Alder sequence.

Scheme 14.

Wittig/Knoevenagel/Diels-Alder/Huisgen cycloaddition sequence.

Oxidation of benzyl alcohols to benzaldehydes can be incorporated into a Wittig-Diels Alder sequence [69]. Also, hetero-Diels-Alder reactions can be run in tandem with a Wittig olefination as shown by Ramachary et al. in their synthesis of tetrahydropyrans 64 (Scheme 15) [136]. Here, diamine 63 is used as a catalyst. The reaction, however, gives the product only in low enantiomeric excess (Scheme 15).

Scheme 15.

Wittig-reaction/hetero-Diels Alder reaction.

Huisgen type [3+2]-cycloaddition reactions can be run also in a simple tandem process rather than incorporated in a more complex reaction chain (see above). A typical example is shown in Scheme 16, where azidoethyl-tetrahydro-hydroxyfuran 66 is treated with phosphorane 21 to give triazoline 68 alongside diazoamine 69 [137]. Further such approaches are known [138, 139].

Scheme 16.

Wittig reactionintramolecular Huisgen type [3+2]-cycloaddition.

Advertisement

6. One-pot Wittig- and HWE olefination/addition reaction

Electrophiles can be added to the alkene function obtained, in a one-pot reaction with the Wittig olefination. A typical example is the stereoselective bromination of the Wittig product with oxalyl bromide (71), where triphenylphosphine oxide (70) as side product of the olefination step acts as a catalyst in the bromination (Scheme 17) [140]. Hamza and Blum have developed a sol–gel entrapped tertiary phosphine by co-polycondensation of tetramethoxysilane, 2-diphenyl(phosphino)ethyltri(ethoxy)silane and N-2-(aminoethyl)-3-aminopropyltri(methoxy)silane. This could be reacted in a Wittig type olefination with benzyl chlorides (e.g., 76) and benzaldehydes, prepared in situ from benzyl alcohols (e.g., 75). The strategy allows for the combination of the process with a bromination step in one pot by addition of sol–gel-bound pyridinium hydrobromide perbromide after completion of the Wittig reaction (Scheme 18) [71].

Scheme 17.

Wittig olefination—Ph3PO-catalyzed addition of bromine.

Scheme 18.

Wittig olefination—addition of bromine.

Alternatively, the process can be combined with a hydrogenation step by the addition of hydrogen in the presence of an added heterogenized Wilkinson catalyst (Scheme 19) [141]. A further Wittig olefination—hydrogenation sequence was developed by Zhou et al. who obtained α-CF3-γ-ketoesters 82 by adding trichlorosilane to the reaction mixture where triphenylphosphine oxide (again as side product of the Wittig olefination) acts as a Lewis base and activates the silane as hydrogenating agent (Scheme 20) [142]. The routine was expanded to other aldehydes including alkanals as educts [143]. This reaction was also carried out with glyoxal derivatives 84 as starting materials, where after conjugate addition of trichlorosilane a few drops of methanol were added to the solution resulting in conversion of the trichlorosilylenol ether (86) to the keto compound 87 while at the same time generating HCl, which then promoted a Paal-Knorr reaction of 87 to the furan 88 (Scheme 21) [144].

Scheme 19.

Wittig olefination—hydrogenation.

Scheme 20.

Wittig reaction—triphenylphosphine oxide catalyzed hydrogenation.

Scheme 21.

Furan synthesis by one-pot Wittig olefination—hydrogenation—Paal-Knorr reaction.

Lu and Toy showed that the Wittig-olefination—trichloromethylsilane conjugate addition sequence can be coupled with the initial preparation of the phosphorane in one pot [145]. The conjugate addition to furnish the silyl enol ether can be combined with a reductive Aldol reaction, where for the Wittig reaction and for the reductive Aldol reaction two separate aldehydes can be used (Scheme 22) [145]. The reactions above can be run with a triarylphosphine-tertiary amine bifunctional polymeric reagent (Rasta-Resin-PPh3-NBniPr2), where the polymer bound triarylphosphine oxide also exerts a catalyzing effect on the addition of Cl3SiH while making it possible to recycle the polymer [146].

Scheme 22.

Wittig olefination—reductive Aldol reaction.

As many Wittig olefinations can be performed in aqueous medium, it is possible to combine the reaction with an enzymatic step. One such sequence is the enzymatic reduction of the olefinic moiety by a recombinant enoate reductase from Gluconobacter oxydans, carried out with an enzyme-coupled in situ cofactor regeneration with a glucose dehydrogenase as enzyme component and d-glucose as co-substrate (Scheme 23) [147].

Scheme 23.

Wittig-olefination—enzymatic ene-hydrogenation.

Interestingly, a Wittig reaction can also be run in combination with an enzymatic reduction, where the in situ prepared enone 93 is transformed to the alkenol 94 (Scheme 24) [148].

Scheme 24.

Wittig-olefination—enzymatic keto-reduction.

The possibility of a combination of a Wittig/HWE olefination and a Michael addition has been studied by a number of research groups. Thus, Piva and Comesse have added phosphonoesters to copper enolates derived from the 1,4 addition of cuprates 97 to enones 96with the idea that the enolate would deprotonate the phosphonoester 98 producing the reactive ketone and phosphonate, which undergo HWE reaction. Products 99 of the tandem Michael-HWE reaction are produced in acceptable yield (Scheme 25) [149, 150]. This strategy was used with p-methylcinnamaldehyde (100) as carbonyl component in the total synthesis of (±)-ar-turmerone (105), a bisabolane-type natural product found in Zingiber and Curcuma species (Scheme 26) [151].

Scheme 25.

One-pot Michael addition—HWE reaction.

Scheme 26.

Synthetic route to utilizing a one-pot Michael addition—HWE reaction.

Advertisement

7. One-pot Wittig-olefination/functional group interconversion

Wittig reactions can be performed with alkoxycarbonylmethylidenetriphenylphosphorane (21) in aq. NaOH, where the cinnamates formed are hydrolysed in situ to cinnamic acids 106 (Scheme 27) [152]. After completion of the reaction, triphenylphosphine oxide can be filtered off from the strongly basic, aqueous solution, and the cinnamic acids are isolated by simple filtration after acidification of the filtrate. Pinacol-acetal tripropylphosphonium salt 107 has been reacted in aq, 1 M NaOH with different benzaldehydes 37; the cinnamaldehyde O,O-pinacol acetal can be hydrolyzed directly to the cinnamaldehydes 108 with 25w% aq. H3PO4 (Scheme 28) [153].

Scheme 27.

One-pot Wittig reaction—ester hydrolysis.

Scheme 28.

One-pot Wittig reaction—acetal hydrolysis.

This procedure provides a nice alternative to the reaction of benzaldehydes with triphenylphosphoranylideneacetaldehyde, which often produces dienals and trienals as side-products. A tandem Wittig-cyanosilylation was developed by Zhou et al., where again Ph3PO as side product of the Wittig olefination acts as Lewis base to catalyze TMSCN in the cyanosilylation step. Chiral salen aluminum catalyst 109 was used as Lewis acid to activate the keto function in the cyanosilylation. Products were obtained with high enantioselectivity [68–93%ee]. TMSCN and chiral catalyst 109 were added after completion of the Wittig reaction, albeit in one pot (Scheme 29) [143].

Scheme 29.

Wittig reaction—cyanosilylation.

As Wittig reactions can be carried out in aqueous medium, enzymatic reactions can be integrated into the process (vide supra). In this regard, M. Krauβer et al. showed that 4-phenylbut-3-en-2-ones (93), obtained by Wittig olefination, are reduced to the corresponding 4-phenylbut-3-en-2-ols (94) in >99 ee(%) using (S)-alcohol dehydrogenase [(S)-ADH] from Rhodococcus sp. or (R)-ADH from Lactobacillus kefir [148].

Advertisement

8. One-pot Wittig- and HWE olefination/cyclization

Michael type cyclization—cyclic hemiacetals can be used efficiently as substrates in Wittig olefination reactions with stabilized Wittig reagents. After the Wittig reaction, the tethered alcohol function induces a cyclization through a Michael reaction. This reaction sequence has been used especially in the construction of functionalized C-glycosides such as in the stereospecific synthesis of ω-amino-β-d-furanoribosylacetic acid derivative 115 (Scheme 30) [154].

Scheme 30.

Synthesis of ω-amino-β-d-furanoribosylacetic acid derivative 115 utilizing a Wittig olefination-ring closure reaction en route.

In their synthesis to C-glycoside amphiphiles, Ranoux et al. followed a similar strategy, reacting non-protected sugars with HWE reagents in aqueous or solventless conditions, leading to C-glucosides 117 and 121 (Scheme 31) [155].

Scheme 31.

Synthesis of C-glucosides with a HWE—ring closure reaction.

A different mechanism to C-glucosides operates when 5,6-dideoxy-5,6-anhydro-6-nitro-d-glucofuranose 122 is reacted with an excess of phosphorane 21. Here, 21 acts as a base and 122 experiences an anion driven ring opening to 123, which undergoes an oxy-Michael addition to 124 with concomitant Wittig reaction, resulting in C-vinyl glycoside 125 (Scheme 32) [156].

Scheme 32.

Tandem oxy-Michael addition—Wittig reaction.

A highly stereoselective tandem Wittig-reaction-Michael addition has been developed by Liu et al. [157] when reacting 3-carboxy2-oxopropylidene)triphenylphosphorane 126 with enaldehydes (e.g., 15), using a chiral pyrrolidine-based catalyst such as 128 (Scheme 33). Most likely, the asymmetric Michael addition proceeds by the reaction of 15 with the iminium compound 129 (Scheme 33), formed from 15 with catalyst 128.

Scheme 33.

Asymmetric Michael-addition-Wittig-olefination.

Beltrán-Rodil et al. have elaborated a retro-aldol initiated Wittig-olefination-Michael addition sequence leading to an exchange of the hydroxyl function in 130 for a carbalkoxymethyl group in 134. The retro-aldol reaction is effected by the commercially available trimethylamine N-oxide (TMAO, 131) [158] (Scheme 34).

Scheme 34.

Retro-aldol-Wittig-olefination-Michael addition cascade.

Electrocyclizations, incl. photocyclizations, and pericyclic reactions: Electrocyclization can be run in concert with Wittig reactions. One such example is shown in Scheme 35, where allylic bromide 135, the product of a Morita-Baylis-Hillman transformation, is converted with triphenylphosphine to the corresponding phosphonium salt, which is reacted with benzaldehyde (11) to give triene 136. 136, heated under aeration, undergoes a 6π-electrocyclization—base catalyzed aerobic oxidation to o-terphenyl derivative 137 (Scheme 35) [159].

Scheme 35.

One-pot phosphorane synthesis—Wittig-reaction—6π-electrocyclization—oxidative dehydrogenation.

Similarly, Hamza and Blum [71], who developed a Wittig olefination with a sol-gel entrapped tertiary phosphine derived phosphorane (vide supra, Schemes 18 and 19) showed that the Wittig reaction can be run in concert with a photochemical cyclization under aerobic conditions to produce phenanthrene (138) (Scheme 36) [71].

Scheme 36.

Wittig olefination—photocyclization—oxidative dehydrogenation.

A number of tandem Wittig/HWE reaction—Claisen/Cope rearrangements have been reported [160, 161, 162, 163, 164, 165, 166, 167, 168, 169, 170, 171, 172, 173]. A typical example is shown in Scheme 37, where neat (4-fluorophenoxyacetyl)cyanomethylene)triphenylphosphorane 139 is subjected to microwave irradiation at 450 W in a sealed tube to undergo an intramolecular Wittig reaction—Claisen rearrangement to furnish benzofuran 134 (Scheme 37) [173].

Scheme 37.

Intramolecular Wittig reaction—Claisen rearrangement.

Mali et al. achieved the synthesis of seselin and angelicin derivatives (e.g., 148 and 150) by a tandem Wittig-olefination—Claisen rearrangement from propargyl and chloroalkyl ethers of 2,4-dihydroxybenzaldehyde and 2,4-dihydroxyacetophenone (e.g., 146 and 149) (Scheme 38) [164].

Scheme 38.

One-pot syntheses of seselin and angelicin derivatives.

Nevertheless, sometimes, these reactions are not easy to control. Thus, a cascade of Wittig reaction and double Claisen and Cope rearrangements starting from 2,4-prenyloxybenzaldehyde 151 leads to a plethora of products through the range of reactions that are possible with the intermediate 153, itself produced through the Wittig reaction and a first Claisen rearrangement. The final products found include gravelliferone (154, 10%), balsamiferone (155, 5%), and 6,8-diprenylumbelliferone (156, 15%) (Scheme 39) [169].

Scheme 39.

Wittig reaction—double Claisen and cope rearrangements.

Less common is the tandem Wittig and ene reaction. Tilve et al. have published such a combination of Wittig and ene reaction in their total synthesis of (±)-kainic acid (160), an amino acid found in different species of red algae [174]. Here, the product was formed in 65% yield as a mixture of diastereoisomers 1598a/159b in a ratio of 1:5. Previously, the authors had synthesized (±)-kainic acid (160) utilizing a Wittig—Michael reaction as the key step (Scheme 40) [175].

Scheme 40.

Wittig-ene cascade as a key step towards the synthesis of kainic acid (160).

Finally, the possibility of a tandem Wittig-olefination—aza-Wittig rearrangement should be mentioned—this combination was carried out on 2-benzoylaziridine 161 to give stereoisomeric dehydropiperidines 163/164 (Scheme 41) [176].

Scheme 41.

Tandem Wittig-olefination—aza-Wittig-rearrangement.

Advertisement

9. Other transformations

A wealth of further transformations have been found to be possible in combination with Wittig/HWE reactions. Thus, cyclopropanation of alkenes using sulfur-ylide reagent 166 can be run in tandem with a Wittig reaction with a conjugated phosphorane such as 21. This combination of reactions can be performed with the preparation of the aldehyde as the Wittig substrate by oxidation of the corresponding alcohol 165 with MnO2 in one pot (Scheme 42) [177].

Scheme 42.

One-pot oxidation—Wittig-olefination—cyclopropanation.

Generally, non-stabilized phosphoranes are basic. This basicity has been used by Knüppel et al. in the transformation of α,α-dibromoenone 168 with excess methylenetriphenylphosphorane, where the phosphorane induces a Corey-Fuchs-reaction-type dehydrobromination/debromination to generate a terminal alkyne, which together with the concomitantly run Wittig-olefination delivers 169, an intermediate to the trisnorsesquiterpene (−)-clavukerin A (171) (Scheme 43) [178]. A metathesis reaction completes the sequence to 171. In this case, the metathesis reaction is not run in one pot with the previous transformations.

Scheme 43.

Wittig-olefination—Corey-Fuchs-reaction-type dehydrobromination/debromination.

Nevertheless, one-pot Wittig—metathesis reactions are well known from the literature [179, 180, 181]. A typical example is shown in Scheme 44, where catalyst 174 serves both as a catalyst for the metathesis as well as for the Wittig olefination, when the in situ produced aldehyde 175 is treated with triphenylphosphine and ethyl diazoacetate (176) in one pot (Scheme 44).

Scheme 44.

One-pot Wittig—Metathesis reaction.

Advertisement

10. Conclusion

Due to the fact that phosphoranes and phosphonates are stable under more diverse conditions than was initially realized, it has become possible to perform reaction cascades and one-pot reactions with Wittig- and HWE reactions as an integral part of the reaction sequence. Frequently, Wittig olefination reactions are carried out with in situ prepared phosphonium salts and phosphoranes [17, 22, 23, 24, 25, 26, 27]. One-pot oxidation—Wittig olefination reactions are also quite common [40, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 61, 62, 63, 64, 65, 66, 67, 68, 69, 70, 71, 72, 73, 74, 75, 76, 77, 78, 79, 80, 81, 82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97, 98, 99, 100, 101, 102, 103, 104, 105, 106, 107, 108, 109, 110], especially when the carbonyl component is labile [89, 97]. Often, the oxidant of choice is MnO2 [40, 43, 44, 45, 46, 79, 80, 81, 82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97], although a number of reactions are known where transformations were carried with air oxygen using metals and metal oxides as catalysts [72, 73, 74, 75, 76, 77, 78]. As many Wittig- and HWE reactions tolerate metal catalysts, this allows the running of Wittig/HWE reactions in combination with metal catalyzed cross coupling reactions and olefinations such as Heck [114, 115, 116, 117, 118, 119, 120, 121, 122, 123], Suzuki [111, 112, 113], Sonogashira [119, 120, 121], and metathesis reactions [179, 180, 181]. The alkenes gained in the olefination reactions can be submitted to cycloaddition reactions, including Diels Alder reactions [69]. Furthermore, the alkenes lend themselves to 1,2-addition reactions [71, 140] in one-pot procedures. In cases where enones or enaldehydes are produced in the olefination reaction, a 1,4-addition becomes a possibility; this includes the Michael addition [149, 150, 151]. Also, the combination of ring opening of cyclic hemiacetals or acetals, olefination reaction and a 1,4-addition leading to ring closure is quite common [154, 155, 156]. The outcome of one-pot sequences of olefination reaction—electrocyclic rearrangement can be predicted less easily. Nevertheless, one-pot Wittig olefination—Claisen- [173], Wittig olefination—Cope- [169], and Wittig olefination—aza Wittig [176] rearrangement reactions have been published. Lastly, Wittig olefination and HWE reactions have been combined with functional group transformations, including the hydrolysis of an ester function [152] and the reduction of a carbonyl group [148].

The prospects of multi-step, one-pot reactions and reaction cascades incorporating Wittig reagents can be seen in the rich chemistry of ketenylidenetriphenylphosphorane (178) (Scheme 45) [182, 183, 184, 185], which has been reviewed earlier [182, 186, 187]. Lastly, catalytic Wittig reactions can be seen as a subset of tandem reactions involving phosphoranes. Further research in specifically this area will help make the Wittig olefination more atom-economical and environmentally sustainable, so that this reliable alkene forming reaction will remain a competitive olefination strategy of choice.

Scheme 45.

Cascade reactions with ketenylidenetriphenylphosphorane (178).

References

  1. 1. Maryanoff BE, Reitz AB. The Wittig olefination reaction and modifications involving phosphoryl stabilized carbanions, stereochemistry, mechanism, and selected synthetic aspects. Chemical Reviews. 1989;89:863-927
  2. 2. Vedejs E, Marth CF. Mechanism of the Wittig reaction: The role of substituents at phosphorus. Journal of the American Chemical Society. 1988;110:3948-3958
  3. 3. Larsen EO, Aksnes G. Kinetic studies of the Horner reaction I. Phosphorus Silicon. 1983;15:219-228
  4. 4. Pommer H. The Wittig reaction in industrial practice. Angewandte Chemie (International Ed. in English). 1977;16:423-429
  5. 5. Wang Y, Hou DL, Hu HZ, Zhang BZ, Wang ES. Efficient synthesis of beraprost sodium. Chemical Research in Chinese Universities. 2016;32:581-585
  6. 6. Wittig G, Schöllkopf U. Über Triphenylphosphinmethylene als olefinbildende Reagenzien (1. Mitteil.). Chemische Berichte. 1954;87:1318-1330
  7. 7. Wittig G, Haag W. Ueber Triphenylphosphinmethylene als olefinbildende Reagenzien (2. Mitteil.). Chemische Berichte. 1955;88:1654-1666
  8. 8. Taber C, Nelson CG. Potassium hydride in paraffin: A useful base for organic synthesis. The Journal of Organic Chemistry. 2006;71:8973-8974
  9. 9. Corey EJ, Clark DA, Goto G, Marfat A, Mioskowski C, Samuelsson B, Hammerström B. Stereospecific total synthesis of a “slow reacting substance” of anaphylaxis, leukotriene C-1. Journal of the American Chemical Society. 1980;192:1436-1439
  10. 10. Wittig G, Schöllkopf U. Methylenecyclohexane. Organic Syntheses. 1960;40:66-66
  11. 11. Aksnes G, Berg TJ, Gramstad T. Temperature and solvent effects in Wittig reactions. Phosphorus, Sulfur, and Silicon. 1995;106:79-84
  12. 12. Thiemann T, Watanabe M, Tanaka Y, Mataka S. Solvent-free Wittig olefination with stabilized phosphoranes—scope and limitations. New Journal of Chemistry. 2004;28:578-584
  13. 13. Thiemann T. Solventless Wittig reactions with fluorinated benzaldehydes. Journal of Chemical Research. 2007;31:336-341
  14. 14. Wu J, Zhang D, Wei S. Wittig reactions of stabilized phosphorus ylides with aldehydes in water. Synthetic Communications. 2005;35:1213-1222
  15. 15. Wu J, Zhang D. Aqueous Wittig reactions of aldehydes with in situ formed semistabilized phosphorus ylides. Synthetic Communications. 2005;35:2543-2551
  16. 16. Watanabe M, Ribeiro Morais G, Mataka S, Ideta K, Thiemann T. Two variations of solvent-reduced Wittig olefination reactions—comparison of solventless Wittig reactions to Wittig reactions under ultrasonication with minimal work-up. Zeitschrift für Naturforschung B. 2005;60:909-915
  17. 17. Parvartur PT, Torney PS, Tilve SG. Recent developments of Wittig reaction in organic synthesis through tandem or sequential processes. Current Organic Synthesis. 2013;10:288-317
  18. 18. Kelly MJB, Fallot LB, Gustaffson JL, Bergdahl BM. Water mediated Wittig reactions of aldehydes in the teaching laboratory: Using sodium bicarbonate for the in situ formation of stabilized ylides. Journal of Chemical Education. 2016;93:1631-1636
  19. 19. Al Jasem Y, El-Esawi R, Thiemann T. Wittig- and Horner-Wadsworth-Emmons (HWE) olefination reactions with stabilized and semi-stabilized phosphoranes and phosphonates under non-classical conditions. Journal of Chemical Research. 2014;38:453-463
  20. 20. McNulty J, Das P. Highly stereoselective and general synthesis of (E)-stilbenes and alkenes by means of an aqueous Wittig reaction. European Journal of Organic Chemistry. 2009;4031-4035
  21. 21. McNulty J, Das P, McLeod D. Microwave-assisted, aqueous Wittig reactions: Organic-solvent- and protecting-group-free chemoselective synthesis of funcationalized alkenes. European Journal of Chemistry. 2010;16:6756-6760
  22. 22. El-Batta A, Jiang C, Zhao W, Anness R, Cooksy AL, Bergdahl M. Wittig reaxctions in water media emplying stabilized ylides with aldehydes. Synthesis of α,β-unsaturated esters from mixing aldehydes, α-bromoesters, and Ph3P in aqueous NaHCO3. The Journal of Organic Chemistry. 2007;72:5244-5259
  23. 23. Wu J, Yue C. One-pot Wittig reactions in aqueous media: A rapid and environmentally benign synthesis of α,β-unsaturated carboxylic esters and nitriles. Synthetic Communications. 2006;36:2939-2947
  24. 24. Westman J. An efficient combination of microwave dielectric heating and the use if solid-supported triphenylphosphine for Wittig reactions. Organic Letters. 2001;3:3745-3747
  25. 25. Choudhary BM, Mahendar K, Kantam ML, Ranganath KVS, Athar T. The one-pot Wittig reaction: A facile synthesis of α,β-unsaturated esters and nitriles by using nanocrystalline magnesim oxide. Advanced Synthesis and Catalysis. 2006;348:1977-1985
  26. 26. Kantam ML, Kumar KBS, Balasubramanyam V, Venkanna GT, Figueras F. One-pot Wittig reaction for the synthesis of α,β-unsaturated esters using highly basic magnesium/lanthanum mixed oxide. Journal of Molecular Catalysis A: Chemical. 2010;321:10-14
  27. 27. Liu D-N, Tian S-K. Stereoselective synthesis of polysubstituted alkenes through a phosphine-mediated three-component system of aldehydes, α-halo carbonyl compounds, and terminal alkenes. Chemistry—A European Journal. 2009;15:4538-4542
  28. 28. Fumagalli T, Sello G, Orsini F. One-pot, fluoride promoted Wittig reaction. Synthetic Communications. 2009;39:2178-2195
  29. 29. Shi L, Wang W, Wang Y, Huang Y. The first example of a catalytic Wittig-type reaction. Tri-n-butylarsine catalysed olefination in the presence of triphenyl phosphite. The Journal of Organic Chemistry. 1989;54:2027-2028
  30. 30. Wang C, Yang L, Zhou R, Mei G. Solvent-free, one-pot synthesis of α,β-unsaturated esters in the presence of a C3-symmetric arsine. Synthetic Communications. 2016;46:1074-1079
  31. 31. Huang Z-Z, Ye S, Xia W, Yu Y-H, Tang Y. Wittig-type olefination catalysed by PEG-telluride. The Journal of Organic Chemistry. 2002;67:3096-3103
  32. 32. Huang Z-Z, Ye S, Xia W, Tang Y. A practical catalytic Wittig-type reaction. Journal of the Chemical Society, Chemical Communications. 2001;1384-1385
  33. 33. Huang Z-Z, Tang Y. Unexpected catalyst for Wittig-type and dehalogenation reactions. The Journal of Organic Chemistry. 2002;67:5320-5326
  34. 34. Tang Y, Ye S, Huang Z-Z, Huang Y-Z. Telluronium ylides in cyclopropanation and catalytic olefination. Heteroatom Chemistry. 2002;13:463-466
  35. 35. Orsini F, Sello G, Fumigalli T. One-pot Wittig reactions in water and in the presence of a surfactant. Synlett. 2006;1717-1718
  36. 36. Agarkar SV, Pathan RU. Green synthesis of morpholino cinnamides using sodium lauryl sulphate and water. Chemical Science Review and Letters. 2014;4:1-3
  37. 37. Galante A, Llhoste P, Sinou D. Wittig reaction using perfluorinated ylides. Tetrahedron Letters. 2001;42:5425-5427
  38. 38. Karama U, Mahfouz R, Al-Othman Z, Warad I, Almansoor A. One-pot combination of the Wittig olefination with bromination and oxidation reactions. Synthetic Communications. 2013;43:893-898
  39. 39. Karama U, Alshamari H, Abdelall H, Sultan MA. A facile one-pot synthesis of (Z)-α-chloro-α,β-unsaturated esters from alcohols. Arabian Journal of Chemistry. in press
  40. 40. Karama U, Al-Othman Z, Al-Majid A, Almansour A. A facile one-pot synthesis of α-bromo-α,β-unsaturated esters from alcohols. Molecules. 2010;15:3276-3280
  41. 41. Karama U. One-pot synthesis of (E)-α-chloro-α,β-unsaturated esters. Journal of Chemical Research. 2009;405-406
  42. 42. Karama U. One-pot approach to the conversion of alcohols into α-iodo-α,β-unsaturated esters. Synthetic Communications. 2010;40:3447-3451
  43. 43. Wei X, Taylor RJK. In situ oxidation – Wittig reactions. Tetrahedron Letters. 1998;39:1815-3818
  44. 44. Blackburn L, Wei X, RJK T. Manganese oxide can oxidize unactivated alcohols under in situ oxidation-Wittig conditions. Journal of the Chemical Society, Chemical Communications. 1999;1337-1338
  45. 45. Wei X, Taylor RJK. In situ manganese dioxide alcohol oxidation – Wittig reactions: Preparation of bifunctional dienyl building blocks. The Journal of Organic Chemistry. 2000;65:616-620
  46. 46. Taylor RJK, Campbell L, McAllister GD. (±)-trans-3,3’-(1,2-cyclopropanediyl)bis-2-(E)-propenoic acid, diethyl ester: Tandem oxidation procedure (TOP) using MnO2 oxidation – Stabilized phosphorane trapping. Organic Syntheses. 2008;85:15-26
  47. 47. Shuto S, Niizuma S, Matsuda S. One-pot conversion of α,β-unsaturated alcohols into the corresponding carbon-elongated dienes with a stable phosphorus-ylide – BaMnO4 system. Synthesis of 6’-methylene derivatives of neplanocin A as potential antiviral nucleosides. New neplanocin analogues. 11. The Journal of Organic Chemistry. 1998;63:4489-4493
  48. 48. Gholinejad M, Firouzabadi H, Bahrami M, Nájera C. Tandem oxidation – Wittig reaction using nanocrystalline barium manganate (BaMnO4): An improved one-pot protocol. Tetrahedron Letters. 2016;57:3773-3775
  49. 49. MacCoss RN, Balkus EP, Ley SV. A sequential tetra-n-propylammonium perrhuthenate (TPAP)-Wittig oxidation olefination protocol. Tetrahedron Letters. 2003;44:7779-7781
  50. 50. Hoshi M, Kaneko O, Nakajima M, Arai S, Nishida A. Total synthesis of ±−lundurine B. Organic Letters. 2014;16:768-771
  51. 51. Lagouette R, Šebesta P, Jiroš P, Kalinová B, Jirošová A, Straka J, Černá K, Šobotník J, Cvačka J, Jahn U. Total synthesis, proof of absolute configuration, and biosynthetic origin of stylopsal, the first isolated sex pheromone of Strepsiptera. Chemistry—A European Journal. 2013;19:8515-8524
  52. 52. Candy M, Tomas L, Parat S, Hernan V, Bienaymé H, Pons J-M, Bressy C. A convergent approach to (−)-callystatin A based on local symmetry. Chemistry—A European Journal. 2012;18:14267-14271
  53. 53. Matovic NJ, Hayes PY, Penman K, Lehmann RP, De Voss JJ. Polyunsaturated alkyl amides from Echinacea: Synthesis of diynes, enynes, and dienes. The Journal of Organic Chemistry. 2011;76:4467-4481
  54. 54. Boyer A, Isono N, Lackner S, Lautens M. Domino rhodium(I)-catalysed reactions for the efficient synthesis of substituted benzofurans and indoles. Tetrahedron. 2010;66:6468-6482
  55. 55. Read CDG, Moore PW, Williams CM. N,N,N’,N’-Tetramethylenediamine dioxide (TMEDAO2) facilitates atom economical/open atmosphere Ley-Griffith (TPAP) tandem oxidation – Wittig reactions. Green Chemistry. 2015;17:4537-4540
  56. 56. Crich D, Mo X-S. One pot selective 5’-oxidation/olefination of 2’-deoxynucleosides. Synlett. 1999;67-68
  57. 57. Maiti A, Yadav JS. One-pot oxidation and Wittig olefination of alcohols using o-iodoxybenzoic acid and stable Wittig ylide. Synthetic Communications. 2001;31:1499-1506
  58. 58. Chen J, Fu X-G, Zhou L, Zhang J-T, Qi X-L, Cao X-P. A convergent route for the total synthesis of malynamides O, P, Q, and R. The Journal of Organic Chemistry. 2009;74:4149-4157
  59. 59. Huang CC. Synthesis of C-14 labelled 1-(4-methoxybenzoyl)-5-oxo-2-pyrrolidinepropanoic acid (CI-933). Journal of Labelled Compounds and Radiopharmaceuticals. 1987;24:676-681
  60. 60. Barrett AGM, Hamprecht D, Ohkubo M. Dess-Martin periodinane oxidation of alcohols in the presence of stabilized phosphorus ylides: A convenient method for the homologation of alcohols via unstabe aldehydes. The Journal of Organic Chemistry. 1997;62:9376-9378
  61. 61. Mori K. Pheromone synthesis. Part 257: Synthesis of methyl (2E,4Z,7Z)-2,4,7-decatrienoate and methyl (E)-2,4,5-tetradecatrienoate, the pheromone components of the male dried bean beetle, Acanthoscelides obtectus (Say). Tetrahedron. 2015;71:5589-5596
  62. 62. Ireland RE, Norbeck DW. Application of the Swern oxidation to the manipulation of highly reactive carbonyl compounds. The Journal of Organic Chemistry. 1985;50:2198-2200
  63. 63. Salgado A, Mann E, Sanchez-Sancho F, Herradon B. Synthesis of heterocyclic γ-amino-α,β-unsaturated acid derivatives and peptide-heterocycle hybrids. Heterocycles. 2003;60:57-71
  64. 64. Lin S, Deiana L, Tseggai A, Córdova A. Concise total synthesis of dihydrocorynanthenol, Protoemetinol, protoememtine, 3-epi-protoemetinol and emetine. European Journal of Organic Chemistry. 2012;398-408
  65. 65. Crisóstomo FRP, Carrillo R, Martín T, García-Tellado F, Martín VS. A convenient one-pot oxidation/Wittig reaction for the C2-homologation of carbohydrate-derived glycols. The Journal of Organic Chemistry. 2005;70:10099-10101
  66. 66. Bressette AR, Glover IV LC. A convenient one-pot PCC oxidation Wittig reaction of alcohols. Synlett 2004:738-740.
  67. 67. Shet J, Desai V, Tilve S. Domino primary alcohol oxidation-Wittig reaction: Total synthesis of ABT-418 and (E)-4-oxonon-2-enoic acid. Synthesis. 2004;1859-1863
  68. 68. Majik MS, Parameswaran PS, Tilve SG. Total synthesis of (−)- and (+)-tedanalactam. The Journal of Organic Chemistry. 2009;74:6378-6381
  69. 69. Harris GH, Graham AE. Efficient oxidation-Wittig olefination-Diels-Alder multicomponent reactions of α-hydroxyketones. Tetrahedron Letters. 2010;51:6890-6892
  70. 70. Dhumaskar KL, Bhat C, Tilve SG. PDC-mediated tandem oxidative Wittig olefination. Synthetic Communications. 2014;44:1501-1506
  71. 71. Hamza K, Blum J. One-pot combination of the Wittig olefination with oxidation, hydrogenation, bromination and photocyclization reactions. Tetrahedron Letters. 2007;48:293-295
  72. 72. Kim G, Lee DG, Chang S. In situ aerobic alcohol oxidation-Wittig reactions. Bulletin of the Korean Chemical Society. 2001;22:943-944
  73. 73. Lee EY, Kim Y, Lee JS, Park J. Ruthenium-catalyzed, one-pot alcohol oxidation-Wittig reaction producing α,β-unsaturated esters. European Journal of Organic Chemistry. 2009;2943-2946
  74. 74. Carillo AI, Schmidt LC, Marín ML, Scaiano JC. Mild synthesis of mesoporous silica supported ruthenium nanoparticles as heterogeneous catalysts in oxidative Wittig coupling reactions. Catalysis Science & Technology. 2014;4:435-440
  75. 75. Alonso F, Riente P, Yus M. Wittig-type olefination of alcohols promoted by nickel nanoparticles: Synthesis of polymethoxylated and polyhydroxylated stilbenes. European Journal of Organic Chemistry. 2009;6034-6042
  76. 76. Alonso F, Riente P, Yus M. One-pot synthesis of stilbenes from alcohols through a Wittig-type olefination reaction promoted by nickel nanoparticles. Synlett. 2009;1579-1582
  77. 77. Miyamura H, Suzuki A, Yasukawa T, Kobayashi S. Integrated process of aerobic oxidation-olefination-asymmetric C-C bond formation catalysed by robust heterogeneous gold/palladium and chirally modified rhodium nanoparticles. Advanced Synthesis and Catalysis. 2015;357:3815-3819
  78. 78. Reddy AGK, Mahendar L, Satyanarayana G. Simple copper(I)-catalysed oxidation of benzylic/allylic alcohols to carbonyl compounds: Synthesis of functionalized cinnamates in one pot. Synthetic Communications. 2014;44:2076-2087
  79. 79. Taylor RJK, Reid M, Foot J, Raw SA. Tandem oxidation processes using manganese dioxide. Accounts of Chemical Research. 2005;38:851-869
  80. 80. MacDonald G, Alcaraz L, Wei X, Lewis NJ, Taylor RJK. Unsaturated amides derived from 2-amino-3-hydroxycyclopentenone: A Stille approach to Asuka mABA, 2880-II, and limocrocin. Tetrahedron. 1998;54:9823-9836
  81. 81. Alcaraz L, Macdonald G, Ragot J, Leiws NJ, Taylor RJK, Manumycin A. Revision of structure and synthesis of the (+)-enantiomer. The Journal of Organic Chemistry. 1998;63:3526-3527
  82. 82. Alcaraz L, Macdonald G, Ragot J, Leiws NJ, Taylor RJK. Synthetic approaches to the manumycin A, B, and C antibiotics: The first total synthesis of (+)-manumycin-A. Tetrahedron. 1999;55:3707-3716
  83. 83. Zeng F, Negishi E. A highly efficient, selective, and general method for the synthesis of conjugated (all-E)-oligoenes of the (CH=CH)n type via iterative hydrozirconation-palladium-catalysed cross-coupling. Organic Letters. 2002;4:703-706
  84. 84. Petrowski RJ. Straightforward preparation of (2E,4Z)-2,4-heptadien-1-ol and (2E,4Z)-2,4-heptadienal. Synthetic Communications. 2003;33:3233-3241
  85. 85. Nicolaou KC, Li Y, Fylaktakidou KC, Monenschein H, Li Y, Weyershausen B, Mitchell HJ, Wei H-X, Guntupalli P, Hepworth D, Sugita K. Total synthesis of apopotolidin. Journal of the American Chemical Society. 2003;125:15433-15442
  86. 86. Aitken DJ, Faure S, Roche S. Synthetic approaches to the southern part of cyclotheonamide C. Tetrahedron Letters. 2003;44:8827-8830
  87. 87. Mladenova M, Tavlinova M, Momchev M, Alami M, Ourévitch M, Brion JD. An iterative procedure for the synthesis of conjugated omega-chlorotrienoic and -tetraenoic esters and related derivatives. European Journal of Organic Chemistry. 2003;2713-2718
  88. 88. Phillips DJ, Pillinger KS, Li W, Taylor AE, Graham AE. Desymmetrization of diols by a tandem oxidation/Wittig olefination reaction. Journal of the Chemical Society, Chemical Communications. 2006;2280-2282
  89. 89. Davies SB, McKervey MA. Convenient in-situ synthesis of nonracemic N-protected β-amino aldehydes from β-amino acids. Applications in Wittig reaction and heterocycle synthesis. Tetrahedron Letters. 1999;40:1229-1232
  90. 90. Lang S, Taylor RJK. Tandem oxidation-Wittig-Wittig sequences for the preparation of functionalized dienoates. Tetrahedron Letters. 2006;47:5489-5492
  91. 91. Blackburn L, Pei C, Taylor RJK. In situ alcohol oxidation-Wittig reactions using non-stabilized phosphoranes. Synlett. 2002;215-218
  92. 92. Blackburn L, Kanno H, Taylor RJK. In situ alcohol oxidation-Wittig reactions using N-methoxy-N-methyl-2-(triphenylphosphoranylidine)acetamide: Application to the synthesis of a novel analogue of 5-oxo-eicosatetraenoic acid. Tetrahedron Letters. 2003;44:115-118
  93. 93. Runcie KA, Taylor RJK. The in situ oxidation-Wittig reaction of α-hydroxyketones. Journal of the Chemical Society, Chemical Communications. 2002;974-975
  94. 94. Quesada E, Raw SA, Reid M, Roman E, Taylor RJK. One-pot conversion of activated alcohols into 1,1-dibromoalkenes and terminal alkynes using tandem oxidation processes with manganese dioxide. Tetrahedron. 2006;62:6673-6680
  95. 95. McAllister GA, Oswald MF, Paxton RJ, Raw SA, Taylor RJK. The direct preparation of functionalized cyclopropanes from allylic alcohols or α-hydroxyketones using tandem oxidation processes. Tetrahedron. 2006;62:6681-6694
  96. 96. Raw SA, Reid M, Roman E, Taylor RJK. A tandem oxidation procedure for the conversion of alcohols into 1,1-dibromoalkenes. Synlett. 2004;819-822
  97. 97. Pavlakos E, Georgiou T, Tofi M, Montagnon T, Vassilikogiannakis G. γ-Spiroketal γ-lactones from 2-(γ-hydroxyalkyl)furans: Syntheses of epi-pyrenolides D and crassalactone D. Organic Letters. 2009;11:4556-4559
  98. 98. Kona J, King’ondu CK, Howell AR, Suib SL. OMS-2 for aerobic, catalytic, one- pot alcohol oxidation-Wittig reactions: Efficient access to α,β-unsaturated esters. ChemCatChem. 2014;6:749-752
  99. 99. Loukas V, Noula C, Kokotos G. Efficient protocol for the synthesis of enantiopure gamma-amino acids with proteinogenic side chains. Journal of Peptide Science. 2003;9:312-319
  100. 100. Loukas V, Markidis T, Kokotos G. Amino acid based synthesis of chiral long chain diamines and tetramines. Molecules. 2002;7:767-776
  101. 101. Magrioti V, Antonopoulou G, Pantoleon E, Kokotos G. Synthesis of 2-amino alcohols and unnatural amino acids from serine. ARKIVOC. 2002;Xiii:51-55
  102. 102. Dunlap NK, Mergo W, Jones JM, Carrick JD. A general procedure for a one-pot oxidative cleavage/Wittig reaction of glycols. Tetrahedron Letters. 2002;43:3923-3925
  103. 103. Terrell LR, Ward JS II, Maleczka RE Jr. Synthetic studies toward amphidinolide A: A synthesis of fully functionalized subunits. Tetrahedron Letters. 1999;40:3097-3100
  104. 104. Ray PC, Roberts SM, Juliá-Colonna SM. Stereoselective epoxidation of some α,β-unsaturated enones possessing a stereogenic centre at the γ-position: Synthesis of a protected galactonic acid derivative. Journal of the Chemical Society, Perkin Transactions 1. 2001;149-153
  105. 105. Outram HS, Raw SA, Taylor RJK. In situ oxidative diol cleavage-Wittig process. Tetrahedron Letters. 2002;43:6185-6187
  106. 106. Duan Y, Yao P, Du Y, Feng J, Wu Q, Zhu D. Synthesis of α,β-unsaturated esters via a chemo-enzymatic chain elongation approach by combining carboxylic acid reduction and Wittig reaction. Beilstein Journal of Organic Chemistry. 2015;11:2245-2251
  107. 107. Farwick A, Helmchen G. Stereoselective synthesis of β-proline derivatives from allylamines via Domino hydroformylation/Wittig olefination and Aza-Michael addition. Advanced Synthesis and Catalysis. 2010;352:1023-1032
  108. 108. Ruan Q, Zhou L, Breit B. New phosphorus self-assembling ligands for the tandem hydroformylation-Wittig olefination reaction of homoallylic alcohols—A key step for stereoselective pyran synthesis. Catalysis Communications. 2014;53:87-90
  109. 109. Breit B, Zahn SK. Domino hydroformylation-Wittig reactions. Angewandte Chemie, International Edition. 1999;38:969-971
  110. 110. Wong GW, Landis CR. Iterative asymmetric hydroformylation/Wittig olefination sequence. Angewandte Chemie, International Edition. 2013;52:1564-1567
  111. 111. Thiemann T, Watanabe M, Tanaka Y, Mataka S. One-pot Wittig-olefination/Suzuki-reaction—The compatibility of conjugated phosphoranes in Pd(0) catalysed C-C-bond forming reactions. New Journal of Chemistry. 2006;30:359-369
  112. 112. Thiemann T, Watanabe M, Tanaka Y, Mataka S. One pot Suzuki coupling—Wittig olefination reactions. Journal of Chemical Research. 2004;28:723-727
  113. 113. Chaudhary AR, Bedekar AV. 1-(α-aminobenzyl)-2-naphthol as phosphine-free ligand Pd-catalysed Suzuki and one-pot Wittig-Suzuki reaction. Applied Organometallic Chemistry. 2012;26:430-437
  114. 114. Burmester C, Mataka S, Thiemann T. Synthesis of non-symmetric Divinylarenes by a Heck/Wittig reaction combination. Synthetic Communications. 2010;40:3196-3208
  115. 115. Saiyed AS, Bedekar AV. One-pot synthesis of stilbenes by dehydrohalogenation–Heck olefination and multicomponent Wittig–Heck reaction. Tetrahedron Letters. 2010;51:6227-6231
  116. 116. Saiyed AS, Patel KN, Kamath BV, Bedekar AV. Synthesis of stilbene analogues by one-pot oxidation-Wittig and oxidation-Wittig–Heck reaction. Tetrahedron Letters. 2012;53:4692-4696
  117. 117. Patel HA, Patel AL, Bedekar AV. Polyaniline-anchored palladium catalyst-mediated Mizoroki-Heck and Suzuki-Miyaura reactions and one-pot Wittig-Heck and Wittig-Suzuki reactions. Applied Organometallic Chemistry. 2015;29:1-6
  118. 118. Patel KN, Bedekar AV. One-pot synthesis and study of spectroscopic properties of oligo(phenylenevinylene)s. Tetrahedron Letters. 2015;56:6617-6621
  119. 119. Watanabe M, Mataka S, Thiemann T. One pot Sonogashira-coupling/Wittig olefination procedures. Journal of Chemical Research. 2005;29:636-639
  120. 120. Thiemann T, Umeno K, Ohira D, Inohae E, Sawada T, Mataka S. Suzuki-Kumada coupling with Bromoaroylmethylidenephosphoranes. New Journal of Chemistry. 1999;23:1067-1070
  121. 121. Thiemann T, Umeno K, Inohae E, Mataka S. Novel elongated Phosphoranes by Heck-reaction and Pd(0)-catalysed Alkynylation and their use in C-7 group functionalisation in Estrones. The Reports of Institute of Advanced Material Study, Kyushu University. 2000;14(1):17-29 Chem. Abstr. 2000;133:335380w
  122. 122. Thiemann T, Umeno K, Wang J, Arima K, Watanabe M, Tabuchi Y, Tanaka Y, Gorohmaru H, Mataka S. Elongated phosphoranes by C-C coupling of haloaroylmethylidenephosphoranes—Synthesis and applications. Journal of the Chemical Society, Perkin Transactions 1. 2002;2090-2110
  123. 123. Patel KN, Kamath BV, Bedekar AV. Synthesis of alkoxy stilbenes by one-pot O-alkylation-Wittig and O-alkylation Wittig-Heck reaction sequence. Tetrahedron Letters. 2013;54:80-84
  124. 124. Thiemann T, Ohira D, Li YQ, Sawada T, Mataka S, Rauch K, Noltemeyer M, de Meijere A. [4+2]-Cycloaddition of Thiophene-S-monoxides onto activated Methylenecyclopropanes. Journal of the Chemical Society, Perkin Transactions 1. 2000;17:2968-2976
  125. 125. Jarosz S, Szewczyk K. Stability of regioisomeric sugar allyltins. Cleavage of the carbon-oxygen bond under radical conditions. Tetrahedron Letters. 2001;42:3021-3024
  126. 126. Jarosz S, Kozlowska E, Jeżewski A. Intramolecular Diels-Alder reaction of chiral, highly oxygenated trienoates derived from sugar allyltins. Tetrahedron. 1997;53:10775-10782
  127. 127. Jarosz S, Skóra S. A convenient route to enantiomerically pure highly oxygenated decalins from sugar allyltin derivatives. Tetrahedron: Asymmetry. 2000;11:1433-1448
  128. 128. Jarosz S. Tandem Wittig-Diels-Alder reaction between sugar-derived phosphoranes and sugar aldehydes. An easy route to optically pure highly oxygenated decalins. Journal of the Chemical Society, Perkin Transactions 1. 1997;3579-3580
  129. 129. Tilve SG, Torney PS, Patre RE, Kamat DP, Srinivasan BR, Zubkov FI. Domino Wittig-Diels Alder reaction: Synthesis of carbazole lignans. Tetrahedron Letters. 2016;57:2266-2268
  130. 130. Torney P, Patre R, Tilve S. A rapid assembly of furo[3,4-b]- and pyrrolo[3,4-b]carbazolones by Domino Wittig Diels Alder reaction. Synlett. 2011;639-642
  131. 131. Wu J, Sun L, Dai W-M. Microwave-assisted tandem Wittig-intramolecular Diels-Alder cycloaddition. Product distribution and stereochemical assignment. Tetrahedron. 2006;62:8360-8372
  132. 132. Wu J, Jiang X, Xu J, Dai W-M. Tandem Wittig-intramolecular Diels-Alder cycloaddition of ether-tethered 1,3,9-decatrienes under microwave heating. Tetrahedron. 2011;67:179-192
  133. 133. Hilt G, Hengst C. A concise synthesis of substituted stilbenes and styrenes from propargylic phosphonium salts by a cobalt-catalysed Diels-Alder/Wittig olefination reaction sequence. The Journal of Organic Chemistry. 2007;72:7337-7342
  134. 134. Chen Z, Shou W, Wang Y. One-pot synthesis of 1,4-diarylnaphthalenes via a Wittig-Horner reaction/[4+2]-cycloaddition/dehydrogenation sequence. Synthesis. 2009;1075-1080
  135. 135. Ramachary DB, Barbas CF. Towards organo-click chemistry: Development of organocatalytic multicomponent reactions through combinations of Aldol, Wittig, Knoevenagel, Michael, Diels-Alder and Huisgen cycloaddition reactions. Chemistry—A European Journal. 2004;10:5323-5331
  136. 136. D. B. Ramachary, R. Mondal, S. Jain, Direct organocatalytic Wittig/hetero-Diels-Alder reactions in one-pot: Synthesis of highly-substituted tetrahydropyranes. ARKIVOC 2016;ii:98-115.
  137. 137. Herdeis C, Schiffer T. Synthesis of nonracemic 2,3,6-trisubstituted piperidine derivatives from sugar lactones via tandem Wittig [2+3] cycloaddition reaction. A novel entry to prosopis and cassia alkaloids. Tetrahedron. 1999;55:1043-1056
  138. 138. Herdeis C, Telser J. A stereoselective synthesis of nonracemic (+)-desoxoprosophylline by a tandem Wittig [2+3]-cycloaddition reaction. European Journal of Organic Chemistry. 1999;1407-1414
  139. 139. Herdeis C, Schiffer T. Synthesis of chiral nonracemic homo-1-deoxyazasugars with D-talo- and L-allo-configuration via tandem Wittig [2+3] cycloaddition reaction. Tetrahedron. 1999;52:14745-14756
  140. 140. Yu T-Y, Wei H, Luo Y-C, Wang Y, Wang Z-Y, Xu P-F. PPh3O as an activating reagent for one-pot stereoselective syntheses of di- and polybrominated esters from simple aldehydes. The Journal of Organic Chemistry. 2016;81:2730-2736
  141. 141. Gelman F, Blum J, Avnir D. One-pot reactions with opposing reagents: Sol-gel entrapped catalyst and base. Journal of the American Chemical Society. 2000;122:11999-12000
  142. 142. Chen L, Shi T-D, Zhou J. Waste as catalyst: Tandem Wittig/conjugate addition reduction sequence to α-CF3γ-keto esters that uses Ph3PO as catalyst for the chemoselective conjugate reduction. Chemistry, An Asian Journal. 2013;556-559
  143. 143. Cao J-J, Zhou F, Zhou J. Improving the atom efficieny of the Wittig reaction by a “waste as catalyst/co-catalyst” strategy. Angewandte Chemie, International Edition. 2010;49:4976-4980
  144. 144. Chen L, Du Y, Zeng X-P, Shi T-D, Zhou F, Zhou J. Successively recycle waste as catalyst: A one-pot Wittig/1,4-reduction/Paal-Knorr sequence for modular synthesis of substituted furans. Organic Letters. 2015;17:1557-1560
  145. 145. Lu J, Toy PH. Tandem one-pot Wittig/reductive aldol reactions in which the waste from one catalyzes a subsequent reaction. Chemistry, An Asian Journal. 2011;6:2251-2254
  146. 146. Teng Y, Lu J, Toy PH. Rasta resin-PPh3-NBniPr2 and its use in one-pot Wittig cascades. Chemistry, An Asian Journal. 2012;7:351-359
  147. 147. Krauβer M, Winkler T, Richter N, Dommer S, Fingerhut A, Hummel W, Gröger H. Combination of C=C bond formation by Wittig reaction and enzymatic C=C bond reduction in a one-pot process in water. ChemCatChem. 2011;3:293-296
  148. 148. Krauβer M, Hummel W, Gröger H. Enantioselective one-pot two-step synthesis of hydrophobic allylic alcohols in aqueous medium through the combination of a Wittig reaction and an enzymatic ketone reduction. European Journal of Organic Chemistry. 2007;5175-5179
  149. 149. Piva O, Comesse S. Tandem Michael-Wittig-Horner reaction: One-pot synthesis of δ-substituted α,β-unsaturated esters. Tetrahedron Letters. 1997;38:7191-7194
  150. 150. Piva O, Comesse S. Tandem Michael-Wittig-Horner reaction: One-pot synthesis of δ-substituted α,β-unsaturated carboxylic acid derivatives – Application to a concise synthesis of (Z)- and (E)-ochtoden-1-al. European Journal of Organic Chemistry. 2000;2417-2424
  151. 151. Chuzel O, Piva O. Tandem Michael-Wittig-Horner reaction: Application to the synthesis of bisabolenes. Synthetic Communications. 2003;33:393-402
  152. 152. Thiemann T, Elshorbagy MW, Salem M, Ahmadani SAN, Al-Jasem Y, Al Azani M, Al-Sulaibi M, Al-Hindawi B. Facile, direct reaction of benzaldehydes to 3-arylprop-2-enoic acids and 3-arylprop-2-ynoic acids in aqueous medium. International Journal of Organic Chemistry. 2016;6:26-141
  153. 153. McNulty J, Zepeda-Velazquez C, McLeod D. Development of a robust reagent for the two-carbon homologation of aldehydes to (E)-α,β-unsaturated aldehydes. Green Chemistry. 2013;15:3146-3149
  154. 154. Ma D, Li X, Hao L, Zhang P, Chen H, Zhao Y. Stereospecific synthesis of ω-amino-β-D-furanoribosyl acetic acid based on microwave assisted Wittig-Michael tandem reaction. Chemical Journal of Chinese Universities. 2014;35:959-964
  155. 155. Ranoux A, Lemiègre L, Benoit M, Guégan J-P, Benvegu T. Horner-Wadsworth-Emmons reaction of unprotected sugars in water or in the absence of any solvent: One-step access to C-glycoside amphiphiles. European Journal of Organic Chemistry. 2010;32:1314-1323
  156. 156. Kumar RS, Karthikeyan K, Kumar BVNP, Muralidharan D, Perumal PT. Synthesis of densely functionalized C-glycosides by a tandem oxy Michael addition-Wittig olefination pathway. Carbohydrate Research. 2010;345:457-461
  157. 157. Liu Y-K, Ma C, Jiang K, Liu T-Y, Chen Y-C. Asymmetric tandem Michael addition – Wittig reaction to cyclohexenone annulation. Organic Letters. 2009;11:2848-2851
  158. 158. Beltrán-Rodil S, Donald JR, Edwards MG, Raw SA, Taylor RJK. Tandem retro-aldol/Wittig/Michael and related cascade processes. Tetrahedron Letters. 2009;50:3378-3380
  159. 159. Lim CH, Kim SH, Kim KH, Kim JN. An efficient synthesis of o-terphenyls from Morita-Baylis-Hillman adducts of cinnamaldehydes: A consecutive bromination, Wittig reaction, 6π-electrocyclization, and an aerobic oxidation process. Tetrahedron Letters. 2013;54:2476-2479
  160. 160. Quesada E, Taylor RJK. Tandem Horner-Wadsworth-Emmons olefination/Claisen rearrangement/hydrolysis sequence: Remarkable acceleration in water with microwave irradiation. Synthesis. 2005;3193-3195
  161. 161. Ajmeri AA, Rao SSM. Studies in synthesis and characterization of tandem Wittig and Claisen reactions of cinnamylated and prenylated 2, 4-dihydroxybenzaldehyde. Indian. Journal of Heterocyclic Chemistry. 2014;24:175-180
  162. 162. Mali RS, Sandhu PK. One-pot Wittig reaction and Claisen rearrangement of 2-hydroxy-4-prenyloxybenzaldehyde and 2-hydroxy-4-prenyloxyacetophenone: Synthesis of 2’,3’,3’-trimethyl-2’,3’-dihydroangelicins. Journal of Chemical Research (S). 1996;148-149
  163. 163. Mali RS, Joshi P. Useful syntheses of prenylated and pyrano-3-arylcoumarins. Synthetic Communications. 2001;31:2883-2891
  164. 164. Mali RS, Pandhare NA, Sindkhedkar MD. Convenient two-step syntheses of seselin and angelicin derivatives. Tetrahedron Letters. 1995;36:7109-7110
  165. 165. Mali RS, Walture AS, Babu KN. Synthesis of 7-desoxy-8-allylcoumarins. Organic Preparations and Procedures International. 1996;28:217-221
  166. 166. Mali RS, Joshi PP, Sandhu PK, Manekar-Tilve A. Efficient syntheses of 6-prenylcoumarins and linear pyranocoumarins: Total synthesis of suberosin, toddaculin, O-methylapigravin (O-methylbrosiperin), O-methylbalsamiferone, dihydroxanthyletin, xanthyletin, and luvangetin. Journal of the Chemical Society, Perkin Transactions 1. 2002;371-376
  167. 167. Rehman H, Rao JM. Tandem intramolecular Wittig and Claisen rearrangement reactions in the thermolysis of 2-methyl-2-phenoxypropionylcyanomethylenetriphenylphosphoranes: Synthesis of substituted 2H-1-benzopyrans and benzofurans. Tetrahedron. 1987;43:5335-5340
  168. 168. Yadla R, Rao JM. Thermolysis of phenoxyacetylcyanomethylenetriphenylphosphoranes – Tandem intramolecular Wittig and Claisen rearrangement reactions. Heterocycles. 1987;26:329-331
  169. 169. Patre RE, Shet JB, Parameswaran PS, Tilve SG. Cascade Wittig reaction-double Claisen and Cope rearrangements: One-pot synthesis of dipenylated coumarins gravelliferone, balsamiferone, and 6,8-diprenylumbelliferone. Tetrahedron Letters. 2009;50:6488-6490
  170. 170. Schmidt B, Riemer M. Synthesis of allyl- and prenylcoumarins via microwave-promoted tandem Claisen rearrangement/Wittig olefination. Synthesis. 2016;48:141-149
  171. 171. Kawasaki T, Terashima R, Sakaguchi K-e, Sekiguchi H, Sakamoto M. A short route to “reverse-prenylated” pyrrolo[2,3-b]indoles via tandem olefination and Claisen rearrangement of 2-(3,3-dimethylallyloxy)indol-3-ones: First total synthesis of flustramine C. Tetrahedron Letters. 1996;37:7525-7528
  172. 172. Rama Rao VVVNS, Reddy GV, Yadla R, Narsaiah B, Rao PS. Synthesis of fluorine containing 3-cyano/ethoxycarbonyl-2-ethyl-benzo[b]furans via microwave assisted tandem intramolecular Wittig and Claisen rearrangement. ARKIVOC. 2005;3:211-220
  173. 173. Rama Rao VVVNS, Reddy GV, Maitraie D, Ravikanth S, Yadla R, Narsaiah B, Rao PS. One-pot synthesis of fluorine containing 3-cyano/ethoxycarbonyl-2-methyl-benzo[b]furans. Tetrahedron. 2004;60:12231-12237
  174. 174. Majik MS, Parameswaran PS, Tilve SG. Tandem Wittig-Ene reaction approach to Kainic acid. The Journal of Organic Chemistry. 2009;74:3591-3594
  175. 175. Bhat C, Tilve SG. Tandem approach for the synthesis of functionalized pyrrolidones: Efficient routes toward allokainic acid and kainic acid. Tetrahedron Letters. 2013;54:245-248
  176. 176. Coldham I, Collis AI, Mould RJ, Rathmell RE. Synthesis of 4-phenylpiperidines by tandem Wittig Olefination – aza- Wittig rearrangement of 2-benzoylaziridines. Journal of the Chemical Society, Perkin Transactions 1. 1995;2739-2745
  177. 177. Huang WH, Wang LL. One-pot synthesis of cyclopropane derivatives with a cis/trans stereoselectivity by Wittig olefination - sulfur ylide cyclopropanation sequence. Journal of Chemical Research. 2013;380-384
  178. 178. Knȕppel S, Rogachev VO, Metz P. A concise catalytic route to the marine sesquiterpenoids (−)-clavukerin A and (−)-isoclavukerin A. European Journal of Organic Chemistry. 2010;(32):6145-6148
  179. 179. Sirasani G, Paul T, Andrade RB. Sequencing cross-metathesis and non-metathesis reactions to rapidly access building blocks for synthesis. Tetrahedron. 2011;67:2197-2205
  180. 180. Paul T, Andrade RB. Sequential cross-metathesis/phosphorus-based olefination: Stereoselective synthesis of 2,4-dienoates. Tetrahedron Letters. 2007;48:5367-5370
  181. 181. Murelli RP, Snapper ML. Ruthenium-catalyzed tandem cross-metathesis/Wittig olefination: Generation of conjugated dienoic esters from terminal olefins. Organic Letters. 2007;9:1749-1752
  182. 182. Schobert R, Gordon GJ. Bioactive heterocycles from Domino Wittig-pericyclic reactions. Current Organic Chemistry. 2002;6:1181-1196
  183. 183. Schobert R, Löffler J, Siegfried S. Phosphacumulene ylides as versatile C2 building blocks: New pathways to tetronates, coumarins and azacycles. Targets in Heterocyclic Systems. 1999;3:245
  184. 184. Schobert R, Dietrich M, Mullen G, Urbina-Gonzalez J-M. Phosphorus ylide based functionalizations of tetronic and tetramic acids. Synthesis. 2006;3902-3914
  185. 185. Marcos IS, Pedrero AB, Sexmero MJ, Diez D, Basabe P, Hernandez FA, Urones JG. Synthesis and absolute configuration of three natural ent-halimanoilides with biological activity. Tetrahedron Letters. 2003;44:369-372
  186. 186. Schobert R. Domino syntheses of bioactive tetronic and tetramic acids. Naturwissenschaften. 2007;94:1-11
  187. 187. Schobert R, Hölzl C. Heterocycles from unsaturated phosphorus ylides. Topics in Heterocyclic Chemistry. 2008;12:193-218

Written By

Fatima Merza, Ahmed Taha and Thies Thiemann

Submitted: 21 October 2016 Reviewed: 11 July 2017 Published: 20 December 2017