27
2713 © 2021. The Author(s). Synthesis 2021, 53, 2713–2739 Georg Thieme Verlag KG, Rüdigerstraße 14, 70469 Stuttgart, Germany D. Roman et al. Review Syn thesis Applications of the Horner–Wadsworth–Emmons Olefination in Modern Natural Product Synthesis Dávid Roman 0000-0002-6223-7869 Maria Sauer Christine Beemelmanns* 0000-0002-9747-3423 Chemical Biology of Microbe-Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology – Hans-Knöll-Institute (HKI), Beutenbergstrasse 11a, 07745 Jena, Germany [email protected] Corresponding Author Christine Beemelmanns Chemical Biology of Microbe-Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology – Hans-Knöll-Institute (HKI), Beutenbergstrasse 11a, 07745 Jena, Germany eMail [email protected] O P O R 1 O R 1 O R 2 O P O R 1 O R 1 O S R 2 O P R 2 R 1 O R 1 O O P O R 1 O R 1 O N H R 2 O P S R 1 O R 1 O R 2 O NH O P O R 1 O R 1 O OR 2 X O P X O R 1 O R 1 O R 2 O P O O R 1 O R 1 O H R 2 R 3 H/R 4 R 2 H R 3 R 4 /H or H R 3 O R 4 R 3 O (Z) (E) R 2 Natural Product Synthesis Received: 20.01.2021 Accepted after revision: 28.04.2021 Published online: 28.04.2021 DOI: 10.1055/a-1493-6331; Art ID: ss-2021-z0027-r License terms: © 2021. The Author(s). This is an open access article published by Thieme under the terms of the Creative Commons Attribution-NonDerivative-NonCommercial-License, permitting copying and reproduction so long as the original work is given appropriate credit. Contents may not be used for commercial purposes or adapted, remixed, transformed or built upon. (https://creativecommons.org/licenses/by-nc-nd/4.0/) Abstract The Horner–Wadsworth–Emmons (HWE) reaction is one of the most reliable olefination reaction and can be broadly applied in or- ganic chemistry and natural product synthesis with excellent selectivity. Within the last few years HWE reaction conditions have been optimized and new reagents developed to overcome challenges in the total syn- theses of natural products. This review highlights the application of HWE olefinations in total syntheses of structurally different natural products covering 2015 to 2020. Applied HWE reagents and reactions conditions are highlighted to support future synthetic approaches and serve as guideline to find the best HWE conditions for the most compli- cated natural products. 1 Introduction and Historical Background 2 Applications of HWE 2.1 Cyclization by HWE Reactions 2.2.1 Formation of Medium- to Larger-Sized Rings 2.2.2 Formation of Small- to Medium-Sized Rings 2.3 Synthesis of ,-Unsaturated Carbonyl Groups 2.4 Synthesis of Substituted C=C Bonds 2.5 Late-Stage Modifications by HWE Reactions 2.6 HWE Reactions on Solid Supports 2.7 Synthesis of Poly-Conjugated C=C Bonds 2.8 HWE-Mediated Coupling of Larger Building Blocks 2.9 Miscellaneous 3 Summary and Outlook Key words HWE reaction, olefination, C=C bond formation, phospho- nate, aldehyde, alkenes, natural products 1 Introduction and Historical Background Carbon–carbon double bonds are ubiquitous in natural and biologically active products and common motifs in or- ganic building blocks. The formation of a C=C bond, often referred to as olefination, is a fundamental reaction in organic chemistry and allows further transformations, such as oxidation, metathesis, or polymerization. The challenge of selectively introducing double bond motifs into poly- functional molecules motivated organic chemists early on, and gained an enormous momentum when Wittig pub- lished his discovery on carbonyl olefination in 1954 (Figure 1), later called the Wittig reaction. He treated carbonyl compounds with substituted quaternary phosphonium salts in the presence of certain condensing agents and dis- covered that the intermediate phosphonium ylides reacted with the carbonyl compound to form an unstable interme- diate, which then broke down to phosphine oxide and an olefin. 1 This epochal discovery significantly changed the di- rection of olefination chemistry. Since then, the Wittig principle of double bond formation has been thoroughly studied and because of its effectiveness and applicability has led to the development of related transformations. 2 Most notably, in 1958 Horner showed that the acidity of a phosphonate correlates with its reactivity to form a stabi- lized phosphonate carbanion that reacts with carbonyl compounds to yield an olefin. 3 To prove his theory, he react- ed methyldiphenylphosphine oxide with benzophenone in the presence of sodium amide in benzene and obtained the predicted products diphenylphosphinic acid and 1,1-di- phenylpropene. Wadsworth and Emmons discussed the re- action reported by Horner and defined its limitations in 1961. 4 They found that both the Wittig and Horner reac- tions share same principles but with the major difference that phosphonate carbanions are stabilized by electron- withdrawing groups, which makes them more nucleophilic and less basic compared to phosphonium ylides (Figure 2). The systematic exploration of the now titled Horner– Wadsworth–Emmons (HWE) olefination opened a wide range of new possibilities and quickly became one of the most intensively studied and commonly applied reactions for selective double bond formation. SYNTHESIS0039-78811437-210X Georg Thieme Verlag KG Rüdigerstraße 14, 70469 Stuttgart 2021, 53, 2713–2739 review en

Applications of the Horner–Wadsworth–Emmons Olefination in

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Applications of the Horner–Wadsworth–Emmons Olefination in

2713

D. Roman et al. ReviewSynthesis

SYNTHESIS0 0 3 9 - 7 8 8 1 1 4 3 7 - 2 1 0 XGeorg Thieme Verlag KG Rüdigerstraße 14, 70469 Stuttgart2021, 53, 2713–2739

reviewen

Applications of the Horner–Wadsworth–Emmons Olefination in Modern Natural Product SynthesisDávid Roman 0000-0002-6223-7869

Maria Sauer

Christine Beemelmanns* 0000-0002-9747-3423

Chemical Biology of Microbe-Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology – Hans-Knöll-Institute (HKI), Beutenbergstrasse 11a, 07745 Jena, [email protected]

Corresponding AuthorChristine BeemelmannsChemical Biology of Microbe-Host Interactions, Leibniz Institute for Natural Product Research and Infection Biology – Hans-Knöll-Institute (HKI), Beutenbergstrasse 11a, 07745 Jena, GermanyeMail [email protected]

O

P

O

R1OR1O R2

O

P

O

R1OR1O S

R2

O

PR2

R1OR1O

O

P

O

R1OR1O N

H

R2

O

P SR1OR1O R2

ONH

O

P

O

R1OR1O OR2

X

O

PX

O

R1OR1O R2

O

PO

O

R1OR1O

H

R2

R3

H/R4

R2

H

R3

R4/H

or

H R3

O

R4 R3

O

(Z)(E)R2

Natural Product Synthesis

Received: 20.01.2021 referred to as olefination, is a fundamental reaction in

Accepted after revision: 28.04.2021Published online: 28.04.2021DOI: 10.1055/a-1493-6331; Art ID: ss-2021-z0027-r

License terms:

© 2021. The Author(s). This is an open access article published by Thieme under the terms of the Creative Commons Attribution-NonDerivative-NonCommercial-License, permitting copying and reproduction so long as the original work is given appropriate credit. Contents may not be used for commercial purposes or adapted, remixed, transformed or built upon. (https://creativecommons.org/licenses/by-nc-nd/4.0/)

Abstract The Horner–Wadsworth–Emmons (HWE) reaction is one ofthe most reliable olefination reaction and can be broadly applied in or-ganic chemistry and natural product synthesis with excellent selectivity.Within the last few years HWE reaction conditions have been optimizedand new reagents developed to overcome challenges in the total syn-theses of natural products. This review highlights the application ofHWE olefinations in total syntheses of structurally different naturalproducts covering 2015 to 2020. Applied HWE reagents and reactionsconditions are highlighted to support future synthetic approaches andserve as guideline to find the best HWE conditions for the most compli-cated natural products.1 Introduction and Historical Background2 Applications of HWE2.1 Cyclization by HWE Reactions2.2.1 Formation of Medium- to Larger-Sized Rings2.2.2 Formation of Small- to Medium-Sized Rings2.3 Synthesis of ,-Unsaturated Carbonyl Groups2.4 Synthesis of Substituted C=C Bonds2.5 Late-Stage Modifications by HWE Reactions2.6 HWE Reactions on Solid Supports2.7 Synthesis of Poly-Conjugated C=C Bonds2.8 HWE-Mediated Coupling of Larger Building Blocks2.9 Miscellaneous3 Summary and Outlook

Key words HWE reaction, olefination, C=C bond formation, phospho-nate, aldehyde, alkenes, natural products

1 Introduction and Historical Background

Carbon–carbon double bonds are ubiquitous in natural

and biologically active products and common motifs in or-

ganic building blocks. The formation of a C=C bond, often

organic chemistry and allows further transformations, such

as oxidation, metathesis, or polymerization. The challenge

of selectively introducing double bond motifs into poly-

functional molecules motivated organic chemists early on,

and gained an enormous momentum when Wittig pub-

lished his discovery on carbonyl olefination in 1954 (Figure

1), later called the Wittig reaction. He treated carbonyl

compounds with substituted quaternary phosphonium

salts in the presence of certain condensing agents and dis-

covered that the intermediate phosphonium ylides reacted

with the carbonyl compound to form an unstable interme-

diate, which then broke down to phosphine oxide and an

olefin.1 This epochal discovery significantly changed the di-

rection of olefination chemistry. Since then, the Wittig

principle of double bond formation has been thoroughly

studied and because of its effectiveness and applicability

has led to the development of related transformations.2

Most notably, in 1958 Horner showed that the acidity of a

phosphonate correlates with its reactivity to form a stabi-

lized phosphonate carbanion that reacts with carbonyl

compounds to yield an olefin.3 To prove his theory, he react-

ed methyldiphenylphosphine oxide with benzophenone in

the presence of sodium amide in benzene and obtained the

predicted products diphenylphosphinic acid and 1,1-di-

phenylpropene. Wadsworth and Emmons discussed the re-

action reported by Horner and defined its limitations in

1961.4 They found that both the Wittig and Horner reac-

tions share same principles but with the major difference

that phosphonate carbanions are stabilized by electron-

withdrawing groups, which makes them more nucleophilic

and less basic compared to phosphonium ylides (Figure 2).

The systematic exploration of the now titled Horner–

Wadsworth–Emmons (HWE) olefination opened a wide

range of new possibilities and quickly became one of the

most intensively studied and commonly applied reactions

for selective double bond formation.

© 2021. The Author(s). Synthesis 2021, 53, 2713–2739Georg Thieme Verlag KG, Rüdigerstraße 14, 70469 Stuttgart, Germany

Page 2: Applications of the Horner–Wadsworth–Emmons Olefination in

2714

D. Roman et al. ReviewSynthesis

Figure 1 Historical timeline capturing the most important discovery milestones in the development of the Horner–Wadsworth–Emmons reaction

In 1983 Still and Gennari modified the HWE reaction to

obtain almost exclusively Z-selective alkenes.5,6 This trans-

formation was achieved by using electrophilic bis(trifluoro-

ethyl)phosphonoacetate in combination with a strong dis-

sociating base, such as potassium hexamethyldisilazanide

(KHMDS) and 18-crown-6 in THF, which allowed highly ste-

reoselective formation of Z-unsaturated esters. In 1995,

Ando reported that the use of ethyl diphenylphosphono-

Georg WittigOlefination reaction with phosphonium ylides

Leopold HornerWittig reaction withstabilized phosphonatecarbanions

W. Clark Still &Cesare GennariDeveloped new reagents, conditions for Z-selective reaction

William S. Wadsworth &William D. EmmonsDefined scope of Wittig reaction and proposed mechanism

Michael W. Rathke &Michael NowakMild HWE conditions usingLiBr or MgBr2 and Et3N

Satoru Masamune & William R. Roush Mild HWE conditions usingweak bases DBU or DIPEA

Kaori AndoReported conditions for Z-selective alkene formation

Biographical Sketches

Christine Beemelmanns(1981) received her diploma inchemistry at the RWTH Aachen(Germany) in 2006. After a one-year research stay in Prof. MikikoSodeoka’s group at RIKEN(Wako-Shi, Japan) she then per-formed her Ph.D. under the su-pervision of Prof. Hans-UlrichReissig with a fellowship fromthe Fonds der Chemischen In-

dustrie and the Studienstiftungdes Deutschen Volkes. Afterpostdoctoral stays in Prof. Keis-uke Suzuki’s group at the TokyoInstitute of Technology (Tokyo,Japan) and Prof. Jon Clardy’sgroup at Harvard MedicalSchool (USA), supported by fel-lowships from the German Ex-change Service and GermanNational Academy of Sciences

Leopoldina, she joined the Leib-niz Institute for Natural ProductResearch and Infection Biology(Jena, Germany) as junior re-search group leader at the endof 2013. Her research focuseson the isolation, characteriza-tion, and total synthesis of mi-crobial signaling molecules.

Dávid Roman received hisB.Sc. degree in chemistry(2015) and M.Sc. degree in or-ganic chemistry (2017) fromUniversity of Pavol Jozef Šafárikin Košice (Slovakia). After hisgraduation in 2017, he joinedthe lab of Dr Andreas Lerchner

in the DAx Global DiscoveryChemistry (GDC) department ofNovartis Institute for Biomedicalresearch (NIBR) in Basel (Swit-zerland) for a research stay.Since 2018, he has been per-forming his doctoral studies inthe group of Chemical Biology

of Microbe-Host Interactions ledby Dr Christine Beemelmanns atthe Hans-Knoll Institute in Jena(Germany). His main researchfocuses on total synthesis andderivatization of natural prod-ucts of microbial origin.

Maria Sauer received her B.Sc.and M.Sc. degrees in chemistry(2019) from Friedrich-Schiller-University in Jena (Germany). In

2019, she joined Dr. ChristineBeemelmanns’ group where sheis currently working on methoddevelopment and applications

in total synthesis of biologicallyactive natural products as partof her doctoral studies.

Synthesis 2021, 53, 2713–2739

Page 3: Applications of the Horner–Wadsworth–Emmons Olefination in

2715

D. Roman et al. ReviewSynthesis

acetate results in an almost exclusive formation of Z-

alkenes by using Still’s conditions [KHMDS/18-crown-6]

and benzyltrimethylammonium hydroxide (Triton B) as a

base.7

Shortly after the reports by Still and Gennari, Masamune

and Roush reported in 1984 that the presence of lithium

chloride leads to complexation of the phosphonate reagent,

which results in increased acidity and reactivity; weaker

bases like DBU or diisopropylethylamine (DIPEA) can be

used for deprotonation.8 Rathke and Nowak extended these

studies in 1985 and reported that the use of LiBr or MgBr2

and triethylamine leads to a conversion of aldehydes and

ketones with similar yields to those reactions reported with

strong bases.9 These studies triggered a cascade of explora-

tions to optimize the conditions for HWE reactions (phase

transfer catalyst,10 microwave irradiation, metal-promoted

HWEs),11 selectivity (enantioselective or diastereoselective

transformations),12 and substrate scope. While these semi-

nal findings unraveled the correlation of acidity, complex-

ation, and reactivity in phosphonate-based olefination re-

actions, they also stimulated the usage of other heteroatom-

stabilized carbanions. This mainly includes the Julia–

Kocienski olefination (methylated aryl alkyl sulfones)13 and

Peterson olefination (-silyl carbanion).14

The enormous numbers of HWE reaction possibilities in

addition to its functional group compatibility and selectivi-

ty resulted in its wide synthetic use, in particular in natural

product synthesis. Building up on preceding reviews that

discuss the mechanisms and reaction scopes of Wittig2 and

HWE reactions,15 this review aims to highlight the diversity

of employed HWE reagents and their applications in natural

product total syntheses covering 2015 to 2020 (Scheme 1).

Subsections are structured chronologically and according to

the formation of macrocycles, such as small to medium

rings, ,-unsaturated carbonyls, and conjugated double

bonds. The applied reagents and reaction conditions of the

herein described syntheses are highlighted in Table 1 and

allow the reader to navigate through the diversity of HWE

reactions and support identifying suitable reagents and

conditions for challenging natural product syntheses.

2 Applications of HWE

2.1 Cyclization by HWE Reactions

At the being of the 1970s the HWE reaction was exam-

ined a synthetic tool in the synthesis of lactones,39 lact-

ams,40 macrocycles,41 and hetero-polycycles.42 Since then,

intramolecular HWE-based cyclization reactions have be-

come very popular in natural product total synthesis due to

their functional group tolerance, high yields, and selectivi-

ty.43

2.2.1 Formation of Medium- to Larger-Sized Rings

Polyfunctionalized macrocyclic natural products are

ubiquitous in nature. Their stereochemical complexity and

ring size most often confers high bioactivities, thus making

them interesting targets for medicinal and organic chem-

ists.44 To showcase the application of HWE reaction, we se-

lected four recent literature examples.

Cytochalasans are a diverse group of polyketide-derived

macrocycles of fungal origin, which are known for their

broad range of biological activities.45 This includes cytotox-

ic, immunomodulatory, and nematicidal properties.46 The

cytochalasan family has been a popular synthetic target for

decades47 and numerous reports describe synthetic strate-

gies towards cytochalasin B,48 cytochalasin H,49 cytochala-

sin O,50 chaetochalasin A,51 aspochalasin B,52 and other re-

lated congeners of the cytochalasan family.

Figure 2 Reaction mechanism of the HWE reaction

Synthesis 2021, 53, 2713–2739

Page 4: Applications of the Horner–Wadsworth–Emmons Olefination in

2716

D. Roman et al. ReviewSynthesis

Scheme 1 Examples of some natural products synthesized using HWE reactions (highlighted red) presented herein incorporated in their total synthe-ses. Natural products are sorted according the outcome of the HWE reaction within their synthetic strategy for the formation of (A) macro- and medi-um-sized ,-unsaturated rings, (B) small rings, (C) polyconjugated C=C bonds, (D) ,-unsaturated esters and enones, and (E) coupling of larger building blocks.

O

OH

O

OH

OHOH

HO

OHO

O

1(–)-marinisporolide C

Dias and de Lucca (2017)63

O

COOH

H

O

O

HN CO2Me

OH

O

OH

11(–)-indoxamycin A

Liang and co-workers (2019)85

10tiancimycin B

Nicolaou and co-workers (2020)16

O

OMeO

O

HO

NH

Cl

OMe

O

14anaenamide A

Luesch and co-workers (2020)31

O

O

OH

O

O

O

OH O

OH6

(–)-lasonolide AHong and co-workers (2018)33

NO

HO

O

NH

OH

O

N

O

OH

HOO7

(+)-neooxazolomycinKim and co-workers (2019)25

N

O

OMe

HO

O

O O

O

16rhizoxin D

Fürstner and co-workers (2019)123

HN

OOHOO

2(–)-aspochalasin B

Trauner and co-workers (2018)53

Deng and co-workers (2018)34

NH

HOOH

O

O

8militarinone D

Sim and co-workers (2015)32

O

O

O

H

H

HO

H OOH

HO

O

4propindilactone G

Gui and co-workers (2020)70b

CHOH

H OCOAr

OH

13melleolide

Takasu and co-workers (2019)18

A)

N

O

O

H

N

O

O

H

H

5(–)-flueggenine C

Han and Jeon (2017)70a

HN

OO

H

OHOO

HH

OH

OMeHO3

asperchalasine DDeng and co-workers (2018)34

B)

OH

HN

OHO

O

9fumosorinone A

Schobert and co-workers (2018)28

C)

O

NHO

H OH

H

H

12pericoannosin A

Lindsley and co-workers (2019)80

O

AcO

O

O

NHSO3H

OH

O

O

OAc

OP

ONMe3

O O

11

15siladenoserinol A

Doi and co-workers (2018)119

Tong, Qian and co-workers (2019)17

D)

E)

Synthesis 2021, 53, 2713–2739

Page 5: Applications of the Horner–Wadsworth–Emmons Olefination in

2717

D. Roman et al. ReviewSynthesis

Table 1 Summary of Some HWE Reagents and Conditions Sorted According to the General Formulas Presented

HWE phosphonate reagents Conditions Products of HWE reaction Yield (%) Ratio (E/Z) Year (Section)a

17

NaH, THF–78 to 0 °C, ca 5 h

10

84 8:1 202016

(2.5)

18

DBU, LiCl, THFr.t., 2 h

19

77 only E 201917

(2.8)

20

NaH, toluene0 °C to r.t., 1.5 h

21

78 (2 steps) only E 201918

(–b)

22

n-BuLi, DMPU, THF0 °C, 4.5 h

23

98 20:1 201919

(–b)

24

NaH, t-BuOK, THF–78 °C to r.t., ca. 1 h

25

20 (2 steps) only E 202020

(–b)

26

LiCl, DIPEA, MeCN0 °C to r.t., 8 h

27

91 only E 202021

(2.8)

O

POR2

O

R1OR1O

O

POMe

O

EtOEtO

O

O

HN CO2Me

OH

O

OH

O

OTIPS

O

O

P

O

EtOEtO

PO

NMe3

OO

10

5

NBoc

O

OTBSO

H

H

O

OOTIPS

OO O

NMe3

OP

O

O

O

P

O

EtOEtO OEt

CO2Et

O

OTIPS

H

H

O

PMeOMeO

O

OMe TBSO

COOMe

O

P

OR1O

R1O R2

O

PMeOMeO

O

MeO

O

N

O

Me

O

O

PMeOMeO

O

O OTBDPS

OTBS

OPMB

O

O OTBDPS

Synthesis 2021, 53, 2713–2739

Page 6: Applications of the Horner–Wadsworth–Emmons Olefination in

2718

D. Roman et al. ReviewSynthesis

Table 1 (continued)

28

Ba(OH)2·8H2O, THFr.t., 1 h

29

85 only E 201822

(–b)

30

LiCl, Et3N, THF0 to 40 °C, 2.5 h

31

90 only E 201623

(2.3)

32

Ba(OH)2·8H2O, THFr.t., 12.5 h

33

73 (2 steps) only E 201924

(–b)

34

Ba(OH)2, THF/H2O–30 °C, 16 h

35

75 only E 201925

(2.7)

36

LiHMDS, HMPA, THF–78 °C to r.t., 2 h

37

69 only E 201926

(–b)

36

n-BuLi, THF–78 °C to r.t., ca 1 h

38

96 5.6:1 201727

(–b)

39

NaH, THF0 °C to r.t.

40

67 95% de 201828

(2.7)

41

LDA, THF–78 to 0 °C, 2 h

42

60 (2 steps) 5:1 201729

(2.7)

43

n-BuLi, THFr.t., 17 h

44

62 only E 201630

(2.7)

HWE phosphonate reagents Conditions Products of HWE reaction Yield (%) Ratio (E/Z) Year (Section)a

OBn

O

OTBSO

PEtOEtO

O O

O

TBSO

OTBS

OOBn

O

PMeOMeO

O

OTBS

SO2Ph

i-PrO2C OTBS

O

O

PMeOMeO

O

OTBS

OTBDPS

O

OTBS

i-Pr

i-Pr SiO

OO

O

N

O

O

PMeOMeO

O

O

O

N

O

OHO

O Si i-Pri-Pr

OHFmocHN

O

PR3

OR1O

R1O R2

O

PEtOEtO

O

OMe OMe

O

Bu3Sn

O

PEtOEtO

O

OMe

OTBS

OMe

O

O

PEtOEtO

O

OEt

O

OEt

O

PMeOMeO

O

OEt

O

O

O

OEt

O

PEtOEtO

O

OEtO

TBSO

NH

O

OEt

OH H

Synthesis 2021, 53, 2713–2739

Page 7: Applications of the Horner–Wadsworth–Emmons Olefination in

2719

D. Roman et al. ReviewSynthesis

Table 1 (continued)

45

n-BuLi, THF–78 °C, 3 h

46

50 6:1 202031

(2.4)

47

NaH, THF0 °C to r.t., 1 h

48

71 (2 steps) 1.2:3 201532

(2.7)

49

18-crown-6, KHMDS, THF –78 °C, 2 h

50

74 (2 steps) only Z 201833

(2.7)

49

18-crown-6, KHMDS, THF –78 °C, 1.5 h

51

90 1:20 201919

(–b)

52

KHMDS, THF0 °C, 10 min

53

82 7:1 201834

(2.2.1)

54

LDA, THFr.t., 12 h

55

86 only E 201623

(2.3)

56

n-BuLi, THF–78 to 0 °C, 1.5 h

57

55 only E 202035

(2.5)

HWE phosphonate reagents Conditions Products of HWE reaction Yield (%) Ratio (E/Z) Year (Section)a

O

P

O

R1OR1O OR2

R3

O

PMeOMeO

O

OMe

Cl

BocHN

Cl

OMe

O

O

PEtOEtO

O

OEt

Me

O

OEt

Me

O

PF3CH2CO

F3CH2CO

O

OMe

Me

O

OH

EtO2C

MeMeO2C

O

PF3CH2CO

F3CH2CO

O

OMe

Me OMeO

Me

OTBS

O

PR2

R1OR1O

O

PEtOEtO

TBSOOTBS

OTBS

O

PEtOEtO N

N

O

CHO

SMeO

PEtOEtO N

O

O

O

TESO

OTES O

N

OSMe

N

O

OTBS

Synthesis 2021, 53, 2713–2739

Page 8: Applications of the Horner–Wadsworth–Emmons Olefination in

2720

D. Roman et al. ReviewSynthesis

Table 1 (continued)

An excellent example of a ring-closing HWE reaction in

the total synthesis of (–)-aspochalasin D (72) was described

in 2018 by Trauner and co-workers (Scheme 2).53 Their

strategy started with epoxy alcohol 66 that was converted

into isoindolone moiety 67 in 7 steps. Addition of dimethyl

lithiomethylphosphonate (68) yielded N-debenzoylated

phosphonate 69. Aldehyde 70 was formed following selec-

tive TBS deprotection and oxidation with the Dess–Martin

periodinane (DMP) of the primary alcohol. The HWE reac-

tion using Masamune and Roush conditions (LiCl, DIPEA,

MeCN) afforded macrocycle 71 without epimerization in 3

steps and 46% yield.8 The synthesis of (–)-aspochalasin D

(72) was completed by TBS deprotection using tetra-

butylammonium fluoride, while acidic TBS deprotection

using HF in acetonitrile yielded aspergillin PZ (73) in 89%

yield. (–)-Aspochalasin D (72) was treated in 45% yield.

Deng and co-workers accomplished the first total syn-

theses of asperchalasines A, D, E, and H in 2018, which in-

cluded a highly diastereoselective intermolecular Diels–

Alder reaction and an HWE macrocyclization step to yield

the monomer aspochalasin B (2) (Scheme 3).34 Their syn-

thesis started with the stereospecific synthesis of triene

fragment 53, starting from L-arabinose via a known 3-step

sequence to obtain hemiacetal 74.54 This was followed by a

Wittig reaction, TBS-protection, and the hydrogenation of

the double bond to yield methyl ketone 75 in 69%. Methyl

ketone 75 was directly subjected to an HWE reaction with

dienyl phosphonate 52 under strongly basic conditions to

give the conjugated triene 53 in 82% yield (E/Z ~7:1).55 The

synthesis of lactam 77 was achieved in 8 steps from N-Boc-

L-leucine 76 via C-acylation with methyl chloroformate,56

selenylation, and oxidative elimination. Heating a mixture

of 53 and 77 (100 °C, neat) resulted in the formation of the

Diels–Alder product 78 (E/Z 2:1). The phosphonate was in-

stalled by addition of dimethyl lithiomethylphosphonate to

the methyl ester 78. Deprotection and oxidation of the pri-

mary alcohol yielded aldehyde 79 as a substrate for HWE

macrocyclization. The 11-membered enone 72 was not

achieved after screening different reaction conditions

(KHMDS, NaH, LiBr/Et3N,9 K2CO3/18-crown-657 or NaOCH2CF358).

Only a mild Lewis acid, Zn(OTf)2, in combination with TMEDA

and triethylamine was found to promote intramolecular

HWE olefination. This was followed by TBS deprotection to

obtain enone (–)-aspochalasin D (72) in 79% with minimal

58

LiCl, DIPEA, MeCNr.t., 24 h

59

77 only E 202036

(2.4)

60

NaH, toluene0 °C, 11.5 h

61

78 (2 steps) only E 201918

(2.3)

62

t-BuOK, THFr.t., 5 h

63

85 9.2:0.8 201637

(–b)

64

n-BuLi, THF–78 °C, 2 h

65

98 only E 201638

(–b)

a Section in which relevant material can be found for this synthesis.b Not discussed in following sections.

HWE phosphonate reagents Conditions Products of HWE reaction Yield (%) Ratio (E/Z) Year (Section)a

O

P

O

R1OR1O N

H

R2

O

P

O

R1OR1O S

R2

O

P SR1OR1O R2

O NH

O

PEtO

O

NEtO O

O

Bn

OTBS O

N O

BnO

O

O

PEtO

O

SEtEtO

COSEt

O

OTIPS

H

H

O

PEtO

O

NHEtO

NH NH

NH

O

7

O

PEtO SEtO

O NH HN OS

Synthesis 2021, 53, 2713–2739

Page 9: Applications of the Horner–Wadsworth–Emmons Olefination in

2721

D. Roman et al. ReviewSynthesis

epimerization at C18.59 The synthesis of (–)-aspochalasin B

(2) was completed after oxidation of the allylic alcohol in

(–)-aspochalasin D (72) with TEMPO in 92% yield.

Naturally occurring 12, 14, or 16-membered macrolides

are most often of polyketide origin and have been shown to

exhibit diverse pharmacological activities.60 A series of new

14-membered macrolides, named pestalotioprolides B–H,

were discovered from the mangrove-derived endophytic

fungus Pestalotiopsis microspora by Liu, Proksch, and co-

workers in 2016; the derivatives showed significant cyto-

toxicity against the murine lymphoma cell lines.61 The first

total synthesis of derivatives pestalotioprolide E (83) and

pestalotioprolide F (84), with revision of its proposed struc-

ture, was accomplished by Goswami and co-workers

(Scheme 4).62 The key step of this synthesis was the HWE

reaction of 81a and 82b, which were obtained by partial hy-

drogenation of Sonogashira products 80a and 80b. Oxida-

tion of the primary alcohols yielded the corresponding

aldehydes 81a and 81b that were necessary for the intra-

molecular HWE reaction.

After testing different macrocyclization conditions, the

combination of LiCl/DIPEA or Ba(OH)2·8H2O were found to

provide macrocycles 82a and 82b in optimized 30% yield.

Subsequent desilylation of 82a and 82b resulted in pestalo-

tioprolide E (83) and the revised structure of pestalotiopro-

lide F (84), respectively.

Dias and de Lucca designed the first total synthesis of

the 30-membered (–)-marinisporolide C (1) in 2017, which

allowed the determination of the relative and absolute con-

figuration of the natural product (Scheme 5).63 Aldehyde 85

was prepared in 16 steps and elongated using Takai–Utimoto

olefination with CHI3 producing vinyl iodide 86 in 82%

yield.64 The free OH group in 86 was then subjected to

Yamaguchi esterification with 87 to afford 88 in 92%.65

Treatment with HF·pyridine led to global TBS deprotection

and spiroketal formation. The remaining hydroxyl groups

Scheme 2 Synthesis of (–)-aspochalasin D and aspergillin PZ by Traun-er and co-workers

OH

O

7 steps

OTBS

OTBS

OTBSBzN

O

CO2Me

H

66 67

OTBS

OTBS

OTBSHN

O

H

OP

O OMe

OMe

Li P OMe

O

OMe

75%

O

OTBS

OTBSHN

O

H

OP

O OMe

OMe

H

1. Et3N·3HF2. DMP

HNO

H

OTBSO

OTBS

LiCl, DIPEA, MeCN

46%(3 steps)

69

70

71

HNO

H

OOH

OH

HNO

H

72(–)-aspochalasin D

73aspergillin PZ

TBAF53%O

H

H

H O

OH

HF, MeCN89%

68

Scheme 3 Total synthesis of (–)-aspochalasin D and aspochalasin B by Deng and co-workers

O

OHHO

OH 3 steps

TBSO

OTBS

OTBS

O

P

52

74 75

52KHMDS

82%dr = 7:1 OTBSTBSO

TBSO

53

OH

O

NHBocNBz

O

CO2Me8 steps

76 77

neat, 100 °C

85%E/Z = 2:1

BzN

OCO2Me

H

TBSO

OTBS

OTBS78

1. BuLi, MePO(OMe)22. HF·pyridine3. DMAP

72% (over 3 steps)

79

HN

O

H

TBSO

OTBS

OP

OHN

OO

H

HOOH

1. Zn(OTf)2 Et3N, TMEDA2. TBAF

79%

72aspochalasin D

HN

OO

H

OH

2aspochalasin B

TsOH, H2OTEMPO

92%

O

18

O

OEtOEt

O

MeOMeO

Synthesis 2021, 53, 2713–2739

Page 10: Applications of the Horner–Wadsworth–Emmons Olefination in

2722

D. Roman et al. ReviewSynthesis

were again protected with tert-butyldimethylsilyl trifluoro-

methanesulfonate (TBSOTf) and afforded spiroketal 89.

Stille cross-coupling with 90 followed by oxidation of the

primary alcohol with TEMPO and bis(acetoxy)iodobenzene

(BAIB) provided the necessary aldehyde 91. HWE olefina-

tion to the macrocycle was accomplished using Masamune–

Roush conditions (LiCl, DBU) in acetonitrile. After global

deprotection with HF·pyridine and HCl, (–)-marinisporolide

C (1) was achieved in 25 steps (longest linear sequence)

with 1.15% overall yield.

2.2.2 Formation of Small- to Medium-Sized Rings

The selective synthesis of small (3–5) and medium-

sized (6–13) ring systems are still a challenge in total syn-

thesis.66 The HWE reaction has been applied in the forma-

tion of many five- and six-membered ring systems as

shown in the following three examples.

In 2015, Brimble and co-workers employed a one-pot

benzoyl transfer-intramolecular HWE reaction as a key step

in the construction of cis--hydroxycarvone-based building

blocks, which are important intermediates in the synthesis

of sesterterpenoid natural products (Scheme 6).67

The synthesis of the key starting material allylsilane 98

was accomplished in a 6-step synthesis from 92 and 93 via

a sequence of diastereoselective conjugate addition, nucleo-

philic phosphonate addition, and the key benzoyl transfer-

intramolecular HWE reaction to construct the cyclohexen-

one 98. Subsequent transformation into an allyl bromide al-

lowed Barbier coupling with farnesal and subsequent

deprotection of the benzoyl group yielded phorbin A (100)

and its C4′-epimer 99. This synthetic approach showed that

benzoyl enol ethers could serve as a masking group of

-ketophosphonates and allow a one-pot benzoyl transfer-

intramolecular HWE reaction. The same protocol was also

applied in the synthesis of sesterterpenoid natural prod-

ucts, such as phorbaketal A, which showed selective activi-

ty against a range of cancer cell lines.68

In the same year, Hiemstra and co-workers described

the total synthesis of sesquiterpene lactone aquatolide

(107), which was isolated from Asteriscus aquaticus

Scheme 4 Synthesis of pestalotioprolide E and pestalotioprolide F

OH

TBSO

O

O

P

R1

O

OEtOEt

80a R1 =80b R1 =

O

O

PR1

O

OEtOEt

O

OTBS1. Ni(OAc)2•4H2O NaBH4, H2, EDA EtOH, r.t., 1 h 90–91%

*

*

Ba(OH)2•8H2O THF/H2O 0 °C to r.t., 2 h 31%

OR1 O

OTBS

*OR1 O

OH

pestalotioprolide E 83 R1 = pestalotioprolide F 84 R1 =

OH

OH

CSA, DCM/MeOH0 °C to r.t., 6 h

95–96%*

OTBS

OTBS

81a R1 =81b R1 =

OTBS

OTBS

82a R1 =82b R1 =

OTBS

OTBS

2. DMP, NaHCO3

DCM 0 °C to r.t., 2 h quant.

Scheme 5 Final synthetic sequence within the total synthesis of (–)-marinisporolide C by Dias and de Lucca

OO O O O O O

OTBS

O

OTBSIPO(OEt)2

OO O O O O O

OTBS

O

OTBSP

4 OH

SnBu3

85

90

1. 90, PdCl2(MeCN)2 DMF, 98%2. TEMPO, BAIB, DCM

OH OH O

OH

O

OH

OH

HO

O

O

OH O O O O OTBS

OTBS

O

TBSO

OTBS

OHC

CHI3, CrCl2 dioxane, THF

82%dr = 89:11

OH O O O O OTBS

OTBS

O

TBSO

OTBS

I

87, TCBC, Et3N, DMAP, PhH

92%

O O O O O OTBS

OTBS

O

TBSO

OTBS

I

O

PO(OEt)2

86

88

1. HF·py, THF2. TBSOTf, 2,6-lutidine, DCM

87% (over 2 steps)

89

1. LiCl, DBU, MeCN2. HF·py, THF3. HCl, MeOH

15% (over 4 steps)

91 1(–)-marinisporolide C

O

P

O

OHEtOEtO

87

**

*

dr = 50:50

O

OEt

OEt

O

Synthesis 2021, 53, 2713–2739

Page 11: Applications of the Horner–Wadsworth–Emmons Olefination in

2723

D. Roman et al. ReviewSynthesis

(Scheme 7).69 The total synthesis started with a cross-aldol

reaction of isobutyraldehyde (101) and propynal (102) fol-

lowed by treatment with trimethyl orthoformate yielding

acetal 103. The intramolecular HWE-based cyclization was

introduced after esterification with phosphonate 104 by

treatment with an excess of NaH (1.5 equiv) under diluted

conditions (0.02 M) to form pentenolide 106 in 76% yield.

The synthesis required 11 more steps, including Crabbé ho-

mologation to an allene, photochemical [2+2]-cycloaddi-

tion, hydroboration, intramolecular Mukaiyama-type aldol

reaction to cyclize the eight-membered ring, Dess–Martin

oxidation, and Lewis acid catalyzed cyclization of silyl enol

ether to yield racemic aquatolide (107) (16 steps, longest

linear chain) with an overall yield of 2.2%.

Jeon and Han accomplished the first total synthesis of

(–)-flueggenine C (5) (Scheme 8) in 2017,70a which belongs

to the plant-derived securinega alkaloids family.71

Their synthesis started with the conversion of Boc-pro-

tected D-proline 108 in 8 steps. This included diastereo-

selective Rauhut–Currier (RC) reaction to acetylated dimer

109.72 The tertiary alcohols were both esterified by slow

addition of diethylphosphonoacetic acid to 109 in the pres-

ence of DCC affording phosphonate 110 in 51% yield. The

phosphonate 110 was then converted into olefin 111 via an

intramolecular HWE using NaH as base that was followed

by methanolic deacetylation. Subsequent treatment of alco-

hol 111 with methanesulfonyl chloride and triethylamine

followed by TFA-induced Boc-deprotection and final basic

treatment (aq K2CO3 in THF) resulted in (–)-flueggenine C

(5).

2.3 Synthesis of ,-Unsaturated Carbonyl Groups

,-Unsaturated esters and enones are common motifs

in chemical building blocks and natural products. Here, we

highlight five recent natural product syntheses that re-

quired the introduction of ,-unsaturated carbonyl groups

via HWE reactions.

Scheme 6 Synthesis of phorbin A and its C-4′ epimer by Brimble and co-workers

MeO

O

OO

Br

TMS

t-BuLi, CuCN, Li-thiopheneEt2O, –30 °C

MeO

O

OO

TMS

+76%

92 93 94

EtPO(OEt)2 n-BuLi, THF–78 °C

78%O

OO

TMS95

PO(OEt)2

1. BzCl, Et3N, DMAP CH2Cl2, r.t., 88%

BzO

HOHO

TMS96

PO(OEt)2

2. CSA MeCN/MeOH/H2O (2:2:1), r.t., 52% (88% brsm)

(COCl)2, DMSO Et3N, CH2Cl2 –78 °C to r.t.

NaH, THF0 °C to r.t.

53% (over 2 steps)

O

OBzTMS

98

O

OH

OH

O

OH

OH

100phorbin A

9937%:19%

2. Farnesal, Zn NH4Cl (aq) THF, 30 °C sonication3. K2CO3, THF MeOH, r.t.

+

BzO

OHO

TMS

P

1. NBS, THF –78 °C, 100%

1

44

1

O

OEt

OEt

97

Scheme 7 Total synthesis of aquatolide by Hiemstra and co-workers

PO(OEt)2

OBn

HO

O

104

OO

1. Ti(Oi-Pr)4, t-BuOK THF, –10 °C, 4 h

2. HC(OMe)3, p-TsOH MeOH, 16 h

62% (2 steps)

OH

+

1. 104, DCC DMAP, CH2Cl2 3 h

2. aq HCl acetone, 24 h

92% (2 steps)

O

O

O

PO(OEt)2

OBn

OMe

101 102

105

103

O O

OBn

106

11 stepsO

O

O

107aquatolide

OMe

NaH, THF16 h

76%

Scheme 8 Synthesis of (–)-flueggenine C by Jeon and Han

N

Boc O

OH8 steps O

OAc

NBoc

H

HO

AcO

O

OH

BocNH

O

OAc

NBoc

H

O

AcO

O

O

BocNH

PO

O

EtO

EtO

O

PO

EtO OEt

OH

NBoc

H

HO

BocNH

O

OO

O

DCC, THF51%

108 109

NaH, THFthen MeOH

52%

110

111

N

O

O

H

N

O

O

H

H1. MsCl, Et3N DCM, TFA

2. aq K2CO3

THF62% (2 steps) 5

(–)-flueggenine C

(EtO)2P(O)CH2CO2H

Synthesis 2021, 53, 2713–2739

Page 12: Applications of the Horner–Wadsworth–Emmons Olefination in

2724

D. Roman et al. ReviewSynthesis

The first example is pinolide (117), which is a member

of naturally occurring decanolides isolated in 201273 and

was synthesized in 2016 by Shelke and Suryavanshi

(Scheme 9).74 The synthesis started with asymmetric

-aminooxylation of pentanal (112) using L-proline and ni-

trosobenzene as an electrophilic oxygen source. -Anilino-

oxy aldehyde 113 was subjected to HWE olefination (LiCl

and DBU) in the second step with phosphonoacetate 20 to

yield -anilinooxy-,-unsaturated ester 114, which under-

went copper-mediated N–O bond cleavage to produce sec-

ondary alcohol 115 in 77% yield (99% ee). Total synthesis of

pinolide 117 was accomplished in 10 additional steps.

Scheme 9 Total synthesis of pinolide by Shelke and Suryavanshi

In the 2016 total synthesis of nakadomarin A by

Boeckman and co-workers, an HWE reaction was employed

to construct necessary building blocks.23 Nakadomarin A

belongs to the manzamine alkaloid family and was isolated

from marine sponges by Kobayashi and co-workers.75 Mem-

bers of this compound family exhibit a broad range of bio-

activities, such as cytotoxicity, antimicrobial, antileishma-

nial, antimalarial, and many more.76

As shown in Scheme 10, transformation of aldehyde 118

and phosphonate 30 under Masamune and Roush condi-

tions (LiCl/Et3N) resulted in the formation of the ,-unsat-

urated ketone 31 in 90% yield.25 A second reaction sequence

started with the dianion of phosphonate 24 being propar-

gylated and converted into enone 120 via an HWE olefina-

tion with 1,3-diacetoxyacetone (55% yield over 2 steps). Af-

ter deprotection and Swern oxidation the resulting furan-

carbaldehyde 121 was again subjected to an HWE reaction

with iminophosphonate 54 in the presence of LDA and sub-

sequent deprotection yielded 55. With both 119 and 55

building blocks in hand, the total synthesis of (–)-nakadom-

arin A (122) was accomplished in 17 steps (longest linear

sequence) with an overall yield of 3%.

Scheme 10 Synthesis of (–)-nakadomarin A by Boeckman and co-workers

Another example for the use of Masamune and Roush’s

HWE conditions for the synthesis of linear unsaturated car-

bonyl units in a natural product synthesis was described in

2017 by Scheidt and co-workers.77 They reported the total

synthesis of four sesquiterpenoids belonging to the protoil-

ludane, mellolide, and marasmane families (Scheme 11).

The synthesis of these bioactive natural products was based

on an HWE olefination of aldehyde 123 with phosphonate

124 in the presence of LiCl and DIPEA, which produced

enone 125 in good overall yield (70%). Intramolecular re-

ductive pinacol coupling was subsequently employed to

synthesize the key intermediate cis-cyclobutanediol 126.

The total synthesis of echinocidin D (127) was accom-

plished by desilylation of 126 with tris(dimethylamino)sul-

fonium difluorotrimethylsilicate (TASF). Intermediate 126

was also used to synthesize isovelleral (130) (3 additional

steps), amillaridin (128), and echinocidin B (129).

H

O

ONHPh

OEt

O

PhNO, L-proline

O

PEtO OEt

OEtO

112

20

CuSO4•5H2O (30 mol%)MeOH, 25 °C, 12 h

113

O

OO

O

OPMB

OHO

HO

OH

O

117pinolide116

H

O

ONHPh

MeCN, –20 °C, 24 h

then 20LiCl, DBU1 h

77%; 99% ee (over 2 steps)

OH

OEt

O

8 steps

114115

O

CO2i-Pr

SO2Ph

TBSO

O

P

30, LiCl, Et3N

THF0 °C to 40 °C2.5 h, 90%

TBSO

O

CO2i-Pr

SO2Ph

O

N

OPhO2S

HHN

O

118

30

31

119

9 steps

P

O O

1. NaH, n-BuLi THF, 0 °C, 30 min then 1-bromobut-2-yne 2 h, 72% O

OAc

OAc2. NaH, 1,3-diacetoxyacetone THF, 0 °C to r.t. 76%

O

O

1. 1 M HCl, H2O, EtOH 65 °C, 24 h, 80%2. (COCl)2, DMSO, Et3N –78 °C to r.t., 85%

P

O

NN

24 120

54

55

7 steps

NH

ON

H

H

122(–)-nakadomarin A

119 + 55

OMeOMe

OMeOMe

EtOEtO

O

121

O54, LDA, THFr.t., 12 h

then 1 M HClpetrol ether, 86%

Synthesis 2021, 53, 2713–2739

Page 13: Applications of the Horner–Wadsworth–Emmons Olefination in

2725

D. Roman et al. ReviewSynthesis

Takasu and co-workers also described the synthesis of

protoilludane sesquiterpenes melleolide (13), melleolide F

(132), echinocidins B (136) and D (134) (Scheme 12).18

Here, an HWE reaction of ketoaldehyde 131 with phospho-

nate 60 yielded ,-unsaturated thioester 61 (78% yield over

2 steps). Subsequent intramolecular Morita–Baylis–

Hillman reaction with trimethylphosphine yielded the cy-

clized precursor 133.

The synthesis of protoilludane by Takasu and co-work-

ers was accomplished by using a redox sequence including a

Corey–Kim oxidation to cyclobutanone, reduction with

NaBH(OAc)3 to 135, esterification of the alcohol with orsell-

inic acid, and Fukuyama reduction of the thioester to obtain

melleolide (13). Reduction of the aldehyde 13 with NaBH4

yielded melleolide F (132). The synthesis of other natural

products of the protoilludane family from intermediate 133

was also achieved. This includes echinocidin D (134) by re-

duction and desilylation with lithium aluminum hydride

and echinocidin B (136) from reduction of intermediate 135

with LiAlH4.

Kongkathip and co-workers described the asymmetric

total synthesis of (–)-epiquinamide (144) in 2019,78a which

is a quinolizidine alkaloid isolated from the skin of an Ecua-

dorian frog78b (Scheme 13). Quinolizidines are an important

group of natural products, which show a broad range of bi-

ological activity. For the total synthesis, furanoside 137,79

derived from D-glucose, was subjected to a Zn-mediated

Bernet–Vasella reaction. The resulting aldehyde 138 was

subjected to an HWE reaction with 20 in the presence of

NaH in toluene to obtain unsaturated ester 139 in 63% yield

(2 steps). After reduction of ester 139 to a primary alcohol

and silyl-protection, the secondary alcohol was trans-

formed into the respective azide 140 under Mitsunobu con-

ditions with inversion of the stereocenter. Mesylation of the

alcohol and intramolecular carbamate N-alkylation after

cleavage of silyl group with TBAF resulted in 2,3-cis-disub-

stituted piperidine 141. Boc-deprotection and coupling of

the amine with but-3-enoic acid provided diene 142, which

was subjected to ring-closing metathesis using the 2nd gen-

eration Hoveyda–Grubbs catalyst yielding an unsaturated

quinolizidinone. The total synthesis of (–)-epiquinamide

(144) was completed by hydrogenation of the saturated

amine 143, followed by N-acetylation and reduction of lact-

am with BH3·SMe2 complex.

2.4 Synthesis of Substituted C=C Bonds

The following two examples illustrate the formation of

substituted doubles bonds by HWE olefinations.

Lindsley and co-workers described the formal total syn-

thesis of pericoannosin A (12) in 2019 (Scheme 14),80 which

was isolated in 2015 from Periconia sp. F31 as a new PKS-

NRPS hybrid metabolite with a 3-(hexahydro-1H-iso-

chromen-3-yl)-5-isobutylpyrrolidin-2-one core structure.81

The new synthetic route included an HWE reaction to in-

stall the substituted olefin and required fewer steps with

Scheme 11 Synthesis of isovelleral, amillaridin, and echinocidins B and D and by Scheidt and co-workers

OEtO

O

O

EtO

O

O

124, LiClDIPEA

70%

4 steps

101 123 125

O

P

O

124

HO

Me

HO

OTBDPS

126

O

130isovelleral

O

TASF

HO

Me

OH

129echinocidin B

HOO

O

OHO

ClMeO

128amillaridin

87%

HO

127echinocidin D

HO

OH

HO

OEtOEt

Scheme 12 Total synthesis of melleolide, melleolide F, and echinoci-dins B and D by Takasu and co-workers

H

H

CHO

OH

O

OO

OTIPS

H

H

131

O

OTIPS

H

H

61

COSEt

60, NaH

PSEt

O

60

1. PMe3, 69%2. TBAF, NH4F, 91%3. NCS, SMe2, Et3N, 87%4. NaBH(OAc)3, 68%5. ArCOOH, EDC·HCl DMAP, 68%6. Pd/C, Et3SiH, MgSO4, 88%

13melleolide

H

H

OH

O

132melleolide F

OH

NaBH4, 90%

H

H

COSEt

OH

OTIPS

133

LiAlH4, 78%

H

H

OH

OH

echinocidin D 134

OH

HO

OH

Ar

O

Ar =

R

O

H

H

COSEt

OH

OH

135

LiAlH4, 78%

H

H

OH

OH

echinocidin B 136

OH

78% (2 steps)

O

EtOEtO

Synthesis 2021, 53, 2713–2739

Page 14: Applications of the Horner–Wadsworth–Emmons Olefination in

2726

D. Roman et al. ReviewSynthesis

improved overall yield in comparison to the total synthesis

of 12 described by Lücke, Kalesse, and co-workers in 2018.82

In short, an asymmetric Diels–Alder reaction with 145 and

isoprene (146) yielded an ,-unsaturated aldehyde, which

was subjected to a Wittig reaction to give terminal alkene

147. The corresponding -ketophosphonate 149 was ob-

tained by treatment of 147 with excess lithiated diethyl

ethylphosphonate 148 in 91% yield. An HWE reaction of ac-

etaldehyde with phosphonate 149 and Ba(OH)2 as base pro-

duced enone 150 in 81% yield. A Felkin–Anh stereoselective

reduction of enone 150 with DIBAL-H afforded the second-

ary alcohol. Protection of the alcohol followed by selective

hydroboration of the terminal alkene and basic oxidizing

work-up resulted in an alcohol, which underwent Dess–

Martin oxidation to aldehyde 151.

Anaenamides A (14) and B (155) are new anticancer cy-

anobacterial depsipeptides containing a chlorinated phar-

macophore and were isolated from a marine cyanobacteri-

um derived from the Anae Island reef system in Guam in

2020.31 Their scaffold has an unusual -chlorinated ,-un-

saturated ester and its double bond geometry impacts the

cytotoxicity. Luesch and co-workers achieved the synthesis

of anaenoic acid (154) in 9 steps that included an HWE re-

action to introduce the vinyl halide moiety (Scheme 15).31

Building blocks 153a and 153b were synthesized from com-

mercially available 152 and phosphonate 45 via HWE olefi-

nation, followed by Boc deprotection under acidic condi-

tions. Final coupling of anaenoic acid (154) with a crude

mixture of 153a and 153b by the use of EDC, HOBt, and

DIPEA led to the E- and Z-isomers (14 and 155, respectively)

of anaenamide.

Goswami and co-workers described the first total syn-

thesis and stereochemical assignment of penicitide A (163)

in 2020, which is a marine secondary metabolite (Scheme

16).36 Penicitide A (163) was isolated from a marine red al-

gal species by Wang and co-workers in 2011 and showed

moderate cytotoxicity against pathogen Alternaria brassi-

cae and human hepatocellular liver carcinoma cell lines.83

Scheme 13 Total synthesis of (–)-epiquinamide by Kongkathip and co-workers

O

BocHN O

OI

Zn, AcOH 50% O

NHBoc

OH

NHBoc

OH

OEt

O

137 138

20

139

OEt

O

P

20 NaH

63% (2 steps)

1. LiBH4, 81%2. TBDPSCl imidazole, 92%3. DIAD, PPh3 DPPA, 84%

NHBoc

N3

140

OTBDPSBocN

N3

1. TBAF, 70%

2. MsCl, DMAP3. NaH, 84% (2 steps)

141

N

N3

142O

1. Grubbs-II 86%

1. TFA2. but-3-enoic acid EDCI, DMAP

N

O

HNH2

143 (–)-epiquinamide

1. Ac2O, NaOH 99%

N

HNHAc

87%(2 steps)

1442. Pd/C, H2 95%

2. BH3·SMe2 68%

O

EtOEtO

Scheme 14 Formal synthesis of pericoannosin A by Lindsley and co-workers

EtO

O

O

O

OH

H

H

+2 Steps

14794% ee

O

H

P

H

O

H

H

H

O

NHO

H OH

H

H

12pericoannosin A

145 146

P

O

148

149150

148n-BuLi, 91%

Ba(OH)2·8H2Oacetaldehyde

O

OTBS

H

H

151

1. DIBAL-H, 80%2. TBSOTf 2,6-lutidine, 90%

literature known

81%3. 9-BBN then NaOH H2O2, 81%4. DMP, NaHCO3, 73%

O

OEtEtO

OEt

OEt

Scheme 15 Total synthesis of anaenamides A and B by Luesch and co-workers

OMe

O

O

O

OH

OH

O

154anaenoic acid

NHBoc

O

H n-BuLi

MeO

O

P

Cl

then TFAH3N

Cl

OMe

OH3N Cl

O OMe

15245153a 153b

OMe

O

O

O

OH

NH

O

14anaenamide A

Cl

OMe

O

153a and 153bEDC, HOBt, DIPEA

53%

OMe

O

O

O

OH

NH

O

155anaenamide B

Cl

OMeO

O

OMeOMe

+

++

Synthesis 2021, 53, 2713–2739

Page 15: Applications of the Horner–Wadsworth–Emmons Olefination in

2727

D. Roman et al. ReviewSynthesis

Scheme 16 Total synthesis of penicitide A by Goswami and co-workers

The syntheses commenced with alcohol 156, which was

first subjected to Swern oxidation and the resulting alde-

hyde 157 transformed by an HWE olefination with phos-

phonate 58 in the presence of LiCl and DIPEA to obtain pri-

marily E-isomer 59 in 77% yield. Alcohol 158 was obtained

after hydrogenation, Evans methylation, and reduction with

LiBH4. After another Swern oxidation aldehyde 159 was

treated again with phosphonate 58 under Masamune and

Roush conditions (LiCl, DIPEA) yielding the olefin in 73%. A

second sequence of Evans methylation and reduction fol-

lowing hydrogenation resulted in compound 160. A final

Dess–Martin oxidation and one carbon Wittig olefination

led to the first building block 161. Alkenes 161 and 162

were subjected to cross olefin metathesis, hydrogenation,

and deprotection to obtain penicitide A (163) in 15 linear

steps with an overall yield of 4.3%. Their synthesis allowed a

divergent synthetic approach with easy access to all possi-

ble stereoisomers, which led to confirmation of the stereo-

chemical assignments in penicitide A (163).

2.5 Late-Stage Modifications by HWE Reactions

Indoxamycins A–F were isolated in 2009 from a saline

culture of marine-derived actinomyces by Sato and co-

workers as new members of the polyketide structure

class.84 The structure elucidation of indoxamycins revealed

a highly congested [5.5.6] tricyclic carbon skeleton featuring

an ,-unsaturated carboxylic acid side chain and a trisub-

stituted alkene. The congested core structure consists of six

stereogenic centers with three quaternary carbons. This in-

triguing structure along with the remarkable biological ac-

tivity makes the indoxamycin family a captivating target for

total synthetic approaches.

Liang and co-workers successfully used the HWE reac-

tion in 2019 to install an ,-unsaturated ester at a late

stage in the total synthesis of (–)-indoxamycins A (11) and

B (169).85

The synthesis of indoxamycins required the generation

of 165 from (R)-carvone 164 (Scheme 17). This was done in

10 steps including acylation under KH/dimethyl carbonate

conditions, methylation of the generated potassium enolate

salt, addition of Comins’ reagent, and Pd-catalyzed reduc-

tive detriflation.

HWE olefination with 166, phosphonate 167, and NaH

followed by in situ removal of the tert-butyl group resulted

in the formation of (–)-indoxamycin A (11) in 92% yield. A

similar procedure was applied for the synthesis of indoxa-

mycin B (169). Reduction and oxidation were used to obtain

OH O

OH

O

PN

O

O

O

Bn58

(COCl)2, DMSO, Et3N

OH

OTBS

156

OTBS

N

O

O

O

Bn59

1. H2/Pd-C, 94%2. NaHMDS, MeI, 74%3. LiBH4, 81%

OTBS

158

OH

OTBS

160

OH

3. NaHMDS, MeI, 68%4. LiBH4, 82%

1. 58, LiCl, DIPEA, 73%2. Pd/C, H2, quant.

OTBS

161

1. DMP, NaHCO3 quant.2. Ph3PCH3Br t-BuOK, 68%

penicitide A

O

O

OH162

1. 161, Grubbs-II 60%

163

2. H2, Pd/C3. PTSA

58% (2 steps)

O

EtOEtO

O

OTBS 58, LiCl, DIPEA

quant. 77%

OTBS

O

(COCl)2, DMSOEt3N, quant.

157

159

Scheme 17 Total synthesis of (–)-indoxamycins A and B by Liang and co-workers

O

H

5 steps

H

O

H

H O

166

POt-Bu

O

167

167, NaH then TFA, 92%

H

O

H

H

(–)-indoxamycin A

COOH

H

O

H

H O

168

OH

POMe

O

17

1. Ac2O, py, DMAP, 84%2. 17, t-BuOK, 59%3. Me3SnOH, 80%

H

O

H

H

(–)-indoxamycin B

COOH

OH

6 steps164

H

O

H

H

165

O10 steps

11

169

O

EtOEtO

O

EtOEtO

Synthesis 2021, 53, 2713–2739

Page 16: Applications of the Horner–Wadsworth–Emmons Olefination in

2728

D. Roman et al. ReviewSynthesis

aldehyde 168. Next acetylation of the primary alcohol was

followed by HWE reaction with phosphonate 17 to give the

product with low conversion due to steric hindrance

around the aldehyde. Several bases were screened (NaH,

n-BuLi, and t-BuOK) and t-BuOK was relatively effective in

the HWE reaction giving the corresponding unsaturated

ester in 59% yield. Simultaneous hydrolysis of the methyl

ester and acetyl group removal with Me3SnOH led to

(–)-indoxamycin B (169).

Another example for late-stage introduction of ,-un-

saturated esters using HWE reactions was reported in 2020

by Nicolaou and co-workers in the total synthesis of tianci-

mycin B (10) (Scheme 18).16 TES-protected uncialamycin

170 was synthesized in 2016 from hydroxyisatin in 22

steps.86 It was transformed to uncialamycin analogue 171 in

four steps with a newly introduced ketone moiety. HWE

olefination of 171 with phosphonate 17 and NaH as a base

provided the E- and Z-isomers of tiancimycin B (10) in 84%

and 10% yield, respectively.

Late-stage HWE olefinations were also used in the mod-

ification of epothilone B analogues designed by Nicolaou

and co-workers to improve their potencies and pharmaco-

logical properties as anticancer drugs (Scheme 19).35 Intro-

duction of different heterocycle side chains to the macro-

cyclic core was achieved via HWE reactions in the presence

of NaHMDS, n-BuLi, or NaHMDS/t-BuOK, followed by desil-

ylation.

2.6 HWE Reactions on Solid Supports

Solid phase synthesis (SPS) is mainly applied in the syn-

thesis of longer and difficult to purify peptides and nucleo-

tides. However, this method has increasingly been used for

SP-supported HWE reactions in the synthesis of other small

molecules.87 Most notably, the synthesis of ,-unsaturated

amides,88 esters and nitriles,89 macrocycles,90 olefins,91 tri-

azolopyridazines, and peptides containing E-alkene amide

linkages have been reported.92 Furthermore, several reports

on the synthesis of Z-,-unsaturated esters,93 solvent-free

reactions,94 and crosslinking polysulfones under phase

transfer catalysis (PTC)95 have been published.

Beemelmanns and co-workers reported the application

of SP-based HWE reagents to prepare the vinylogous amino

acid containing natural product barnesin A (177) (Scheme

20),96 which was previously isolated from anaerobic bacte-

Scheme 18 Total synthesis of tiancimycin B by Nicolaou and co-work-ers

OH

HN OTESOOH

O

O

1. Ac2O, DIPEA DMAP, quant.2. 3HF·Et3N, 85%

3. DMP, 89%4. K2CO3, MeOH 96%

OH

HN OOOH

O

O

170 171

P COOMe

17

17, NaH, 94% E/Z = 8:1

OH

HNO

OH

O

O

tiancimycin B

COOMe

10

O

EtOEtO

Scheme 19 Synthesis of epothilone B analogues by Nicolaou and co-workers

O

RO

OR

O

O

X

H

O

O

HO

OH

O

O

X

H

1. NaHMDS or n-BuLi or NaHMDS, t-BuOK

2. HF·py or TFA

R2

EtO

P

O

EtO R2

N

SSMe

N

SN

Me

Boc N

SCF3

N

OSMe

S

N

R = TES, TBS, X = NR, CR2, O

NN

O

Scheme 20 Solid-phase based synthesis of cysteine proteases inhibitor barnesin A and derivatives by Beemelmanns and co-workers

2-chlorotrityl chloride resin

HO

O

P

O

OEtOEt

DIPEA, DCMthen MeOH

O

O

P

O

OEtOEt

172 173

H

O

R1

NHFmoc173 O

O

NHFmoc

R1

LiBr, DIPEADCM

HN

NHBocBocNNHBoc

R1 =

HO

OHN

R1

NH

O

O

R3

R2

R2 = Ph, 4-HOC6H4

R3 = CnHm

174175

176

HN

HO

O

HN

HN NH2

O

NH

O

177barnesin A

OH

Synthesis 2021, 53, 2713–2739

Page 17: Applications of the Horner–Wadsworth–Emmons Olefination in

2729

D. Roman et al. ReviewSynthesis

rium Sulfurospirrilum barnesii as the first secondary me-

tabolite from Epsilonproteobacteria and the first NRPS-PKS

hybrid molecule from this bacterial family.97

The synthesis began with loading a 2-chlorotrityl chlo-

ride resin (CTC resin) with diethylphosphonoacetic acid

(172),88 which served as a linker, and phosphonate for the

HWE reaction. Resin-bound phosphonate 173 was depro-

tonated with DIPEA and subjected to an HWE reaction with

Fmoc-protected aldehyde 174 (LiBr and DIPEA) yielding sol-

id-bound olefin 175, which allowed for further peptide cou-

pling steps. Subsequent global deprotection produced 14

lipodipeptides with micromolar to nanomolar inhibitory

activity against different cysteine proteases.

2.7 Synthesis of Poly-Conjugated C=C Bonds

Conjugated -bond systems allow delocalization of -

electrons through alternating single and double bonds. This

concept was introduced by Claisen in 1926 and later named

vinylogy.98 HWE reactions play a central role in the con-

struction of vinylogous systems and polyconjugated double

bonds, especially in natural product synthesis as shown in

the following eight examples.99,100

The polyunsaturated side chain of the alkaloids militari-

none D (8) (Scheme 21), isolated in 2003 from Paecilomyces

militaris,101 was accomplished by Sim and co-workers in

2015 using a sequence of HWE reactions.32 The first HWE

reaction (NaH in THF) of aldehyde 179 with phosphonate

47 yielded unsaturated ester 48 in 71% yield (2 steps, E/Z

1:2.3). A subsequent reduction/oxidation sequence pro-

duced the aldehyde as an E/Z mixture where the Z-isomer

was converted into the E-isomer 180 by treatment with cat-

alytic iodine. The second HWE olefination (NaH in THF) of

unsaturated aldehyde 180 and phosphonate 39 yielded

triene 40 in 80% yield, which was again transformed into al-

dehyde 181 in a reduction/oxidation sequence. Aldehyde

181 was added to lithiated pyridine 182 in the final se-

quence and the resulting allylic alcohol was oxidized to a

ketone. This was followed by a global MOM removal to

complete the synthesis of militarinone D (8) in 5% overall

yield.

A structurally related NRPS-PKS based fungal-derived

natural product, fumosorinone A (9),102 was isolated in

2017, and shown to contain a pentaene moiety attached to

a pyrrolidine-2,4-dione. Due to its potent protein tyrosine

phosphatase 1B inhibitor activity and intriguing structure,

it was synthesized in 2018 by Schobert and co-workers

(Scheme 22).28 The required aldehyde 183 was prepared

from commercially available (S)-citronellol in 10 steps to

synthesize the polyunsaturated PKS side-chain, which was

then subjected to the chain elongating HWE reaction with

phosphonate 39 and NaH as base to produce elongated ole-

fin 40 in 67% yield. Ester 40 was subjected to a reduc-

tion/oxidation sequence yielding aldehyde 181, which was

Scheme 21 Total synthesis of militarinone D by Sim and co-workers

1. DIBAL-H, DCM, 90%2. (COCl)2, DMSO, Et3N3. I2, Et2OO

OTBS

1. TBAF, THF quant.2. (COCl)2, DMSO Et3N

O

P

O

EtOEtO

OEt

O

47, NaH, THF

178 179

47

71% (over 2 steps)

O

O48

O

O

O

PEtOEtO

O

OEt

180

39

40

39, NaH, THF

80%

O

181

1. DIBAL-H, DCM, 78%2. MnO2, DCM

HO

NH

OH

O

O

R1

2. MnO2, DCM, quant.3. p-TsOH, THF/H2O, 56%

MOMO

N

Br

OMOM

OMOM1828

militarinone D

97% (over 2 steps)

1. 182, n-BuLi, THF 72% (over 2 steps)

R1

E/Z = 1:2.3

Scheme 22 Total synthesis of fumosorinone A by Schobert and co-workers

183

39

40

181

184

O

39, NaH, THF

67%(95% de)

P

EtO2C

R1

1. DIBAL-H, DCM 85%2. MnO2, DCM 92%

O

R1

O

t-BuS

OH

O

t-BuS

OH

P

184, BuLi, THF

95%

O

N

OH

R1

CO2Me

TBSO DMB

NHDMB

CO2Me

TBSO

185

186

186, AgOCOCF3

Et3N, THF89%

1. NaOMe, MeOH, quant.2. TFA, DCM, 30%3. KF, MeOH, 60%

OH

HN

OHO

O

187

9fumosorinone A

R1

O

OEtOEt

O

EtO

O

OEt

OEt

Synthesis 2021, 53, 2713–2739

Page 18: Applications of the Horner–Wadsworth–Emmons Olefination in

2730

D. Roman et al. ReviewSynthesis

subjected to a second HWE olefination following Loscher’s

protocol103 giving thioester 185 in a 95% yield as a 2:3 ke-

to/enol mixture. Silver-mediated acylation with amino acid

186 resulted in ketoamide 187. Subsequent cyclizations fol-

lowed by a partial TBS deprotection with TFA and global

desilylation with KF completed the first total synthesis of

fumosatinone A (9).

Another application of an HWE reaction in the synthesis

of conjugated C=C bonds was described by Gerstmann and

Kalesse in their total syntheses of nannocystin Ax (188) and

aetheramide A (189) (Scheme 23).104 Chiral aldehyde 191 [4

steps from benzaldehyde (190)] was subjected to an HWE

reaction using phosphonate 36 and n-BuLi to yield acrylic

ester 192 in 96% yield in favor of the desired E-isomer

(5.6:1). Ester 192 was reduced and subsequently oxidized to

aldehyde 193, which was used in a vinylogous Mukaiyama

aldol reaction with acetal 194 in the presence of catalyst

195 to give product 196. The alcohol 196 was then methyl-

ated and the ester was hydrolyzed to carboxylic acid 197,

which was used for completion of the total synthesis of

nannocystin Ax (188).

Similarly, the HWE reaction was used by Mori and co-

workers in their total synthesis of brevisamide (203)

(Scheme 24),30 a natural marine product isolated from di-

noflagellate containing a tetrahydropyran (THP) ring incor-

porated into a conjugated carbon chain.

The synthesis of the unsaturated side chain started from

commercial 4-(benzyloxy)butanol, which was converted

into diol 198 in 10 steps. A short reaction sequence afforded

mono-TBS deprotected alcohol 201. The primary alcohol

201 was oxidized with Dess–Martin periodinane to yield

Lindsley’s aldehyde 202 as a precursor for the HWE reac-

tion.100a,105 The olefination of 202 was accomplished using

the vinylogous phosphonate 43 and n-BuLi at room tem-

perature to provide dienoate 44 in 62% yield. Ethyl ester 44

was subsequently reduced to the corresponding alcohol us-

ing DIBAL-H. This was followed by desilylation of the sec-

ondary alcohol and final oxidation of the primary alcohol

with MnO2 gave brevisamide (203).

Hong and co-workers reported an extraordinary syn-

thetic strategy towards the total synthesis of lasonolide A

(6) in 2018 (Scheme 25).33 This is a natural product con-

tains two functionalized tetrahydropyran moieties and ex-

hibits high and selective cytotoxicity against cancer cells.

Scheme 23 Structure of nannocystin Ax and aetheramide A and the total synthesis of a building block by Gerstmann and Kalesse

O O

NHHO

O

Cl

HO

Cl

OMe

O

NHN

O

188

HO

MeO

N

NH

O

O O

OH

OMe

OO

189

O

4 steps

190

TBSO O

191

TBSO

OMe

O192

36

36, n-BuLi, THF–78 °C to r.t.

96%E/Z = 5.6:1

TBSO

O

1. DIBAL-H, DCM –78 °C, 92%2. MnO2, DCM, r.t.

193

POMe

OOEtO

EtO

TBSO

OH

OEt

O196

OEt

OTES

194

HNN

BO

Ts

O

Ph195

194, 195, EtCN–78 °C, acidic workup

1. MeO3+BF4

proton sponge DCM, r.t., 93%

2. LiOH, THF MeOH, H2O 0 °C to r.t., quant.

TBSO

OMe

OH

O197

77% (2 steps)dr 3.1:1

nannocystin Ax aetheramide A

Scheme 24 Synthesis of brevisamide by Mori and co-workers

OBnO OH

HH

OH

Tf2O, 2,6-lutidine–80 °C, 0.7 hthen TBSOTf0 °C, 0.8 h

88%

198

OBnO OTf

HH

OTBS

199

1.

KHMDS, 18-crown-6 THF, r.t., 0.5 h, 90%

Cbz

HN

O 200

2. Pd(OH)2, H2

EtOAc, r.t., 1 h, 96%

OHO

HN

HH

OTBS

O201

OO

HN

HH

OTBS

O

DMPDCM, r.t., 4 h

76%

P

O

OEtOEt

O

O

20243

43, n-BuLi, THF r.t., 17 h

62%

O

HN

HH

OTBS

O

EtO2C

1. DIBAL-H, DCM DCM, –78 °C, 15 min, 79%2. TBAF, THF, r.t., 2 h, 96%

3. MnO2, DCM, 88%

O

HN

HH

OH

O

44

203

O

Hbrevisamide

Synthesis 2021, 53, 2713–2739

Page 19: Applications of the Horner–Wadsworth–Emmons Olefination in

2731

D. Roman et al. ReviewSynthesis

Scheme 25 Total synthetic strategy lasonolide A by Hong and co-workers

A hydroboration-oxidation sequence starting from al-

lene 204 provided polyhydroxy alkane 206 in good overall

yields. The alcohol in 206 was first subjected to an oxida-

tion-Wittig elongation sequence to produce ester 207. Acid-

ic tetrahydropyran formation was followed by the oxidation

to aldehyde 209 that allowed the first HWE reaction with

Still–Gennari phosphonate 49 yielding trisubstituted cis-

conjugated methyl ester 50 in good yields (Z/E 10:1). Silyl-

ation of the secondary alcohol in 50 proceeded smoothly

and DIBAL-H mediated reduction of the ester delivered al-

dehyde 210, which was again subjected to HWE olefination

with a vinylogous phosphate 39 and LiHMDS as a base.

Subsequent Dess–Martin oxidation of the allylic alcohol

provided Z-enone 211 in 94% yield. An unusual 9-BBN com-

plexation of free carboxylic acid allowed successful Julia

olefination with sulfone 212 to give 213. Lactonization was

employed to complete the total synthesis of (–)-lasonolide

A (6).

The synthesis of polyconjugated mycolactone mimetics

by Altmann and co-workers in 2019 represents an addition

example for the efficient application of HWE reactions

(Scheme 26).106 Mycolactones A (227) and B (228) feature a

common 12-membered macrolactone core, which can in-

teract with several cellular targets, such as mTor,107 Sec61

translocon,108 Wiskott–Aldrich syndrome protein,109 and

the angiotensin II receptor.110 Their synthesis targeted my-

colactone analogues 220, 225, and 226 as configurationally

stable mimetics to investigate the importance of the double

bond for the activity and the mode of action in Buruli ulcer

pathogenesis (Scheme 26).

Their synthesis started from methyl 2-iodobenzoate

(214), which was converted into vinyl iodide 215 via Sono-

gashira coupling with TMS-acetylene, desilylation, reduc-

tion of the ester, and zirconium-catalyzed carboalumina-

tion/iodination of the terminal alkyne moiety. After oxida-

tion of the primary alcohol 215 with MnO2 the resulting

aldehyde was subjected to an HWE reaction with triethyl

phosphonoacetate (20) to yield ,-unsaturated ester 216

(99% over 2 steps). Stille coupling of vinyl iodine 216 with

stannane 217 followed by saponification gave acid 218. Fi-

nal esterification with macrolactone 219 and desilylation

yielded 220. Analogous syntheses for cyclopentene-derived

acids 224 and 225 starting from 2-oxocyclopentane-1-car-

boxylate (221) and 1,4-bis(hydroxymethyl)cyclohexane

were also reported.

The marine secondary metabolite nafuredin B (237) was

isolated in 2017 from deep-sea-derived Talaromyces aculea-

tus and Penicillium variabile together with penitalarins A–C

and known nafuredin A.111 Nafuredins contain a linear skel-

eton with an unsaturated chain linked to ,-unsaturated

lactone and showed cytotoxicity against several cell lines in

the micro-molar range thus making them attractive syn-

thetic targets.

TBDPSO

OAc

204

Sia2BH, THF–78 °C to –30 °Cthen TBAOH/H2O2

86%Z/E > 20:1

TBDPSO OHOH

205

9-BBN, THF–30 °C to 0 °Cthen (+)-IpcBH2, –20 °C

then NaOH, H2O2, r.t.73%

dr 6.4/1

TBDPSO OH OH

OH

206

1. CSA, Me2C(OMe)2 DCM, r.t. then TBAF, r.t. 80%

2. TEMPO, BAIB, THF, r.t. then Ph3P=CHCO2Et, r.t. 81% E/Z > 20:1

OHO

O

207

O

EtO2COH

O

Cl SO3H

1. 208, EtOH, reflux 82% dr 11:12. 4-PhCO2-TEMPO BAIB, DCM, r.t.

208

209

74% Z/E = 10:1

(over 2 steps)

O

EtO2COH

MeO2C

50

1. 2,6-lutidine,TBSOTf DCM, r.t., 87%2. DIBAL-H, DCM, –78 °C

O

OTBS

HO

210

O

OTBS

O

HO2C

1. 39, LiHMDS, THF –78 °C to 0 °C then NaOH, 60 °C, 67% E/Z = 14:12. DMP, DCM, 0 °C, 94%

211

P

O

CO2MeO

OF3C

F3C

EtO2C P

O

OEtOEt

49

39

O

O

OTBS

O

OTMSTBSO

SBT

O O

O

O

OTBS

O

OTMSTBSO O

OHHO2C

1. 9-BBN, DME –20 to 50 °C2. 212, LiHMDS, –75 °C then HCl, r.t.

212

O

O

OH

O

OOH O

OHO

PhH, 90 °Cthen HF, r.t.

58%

213

6

EtO2C

65%E/Z = 12:1

(over 2 steps)

49, KHMDS THF, –78 °C

lasonolide A

O

Synthesis 2021, 53, 2713–2739

Page 20: Applications of the Horner–Wadsworth–Emmons Olefination in

2732

D. Roman et al. ReviewSynthesis

The first total synthesis of nafuredin B (237) was ac-

complished by Goswami and co-workers (Scheme 27)112

and was based on an elongation sequence of one HWE reac-

tion and two Julia–Kocienski olefinations. Commercially

available prenol was converted into unsaturated ester 229

using a known procedure. Ester 229 was then reduced with

DIBAL-H and oxidized with Dess–Martin periodinane and

the intermediate aldehyde was directly subjected to Julia–

Kocienski olefination with sulfone 230 to obtain the olefin

(E/Z 4:1). The isomers were separated and the E-isomer was

treated with TBAF to produce primary alcohol 231, which

was again oxidized and treated with Grignard reagent to

give the secondary alcohol 232. Subsequent oxidation to

ketone 233 and an HWE reaction (NaH) with triethyl phos-

phonoacetate (20) produced the elongated unsaturated side

chain in 234 (82% yield, E/Z 6:1). The final elongation via Ju-

lia–Kocienski olefination was pursued after another reduc-

tion/oxidation sequence and transformation with sulfone

235 to obtain 236.113,114 The last three steps involved treat-

ment with p-toluenesulfonic acid to afford a diol, subse-

quent acrylation, and then ring-closing metathesis provided

the desired nafuredin B (237).

Scheme 27 Total synthesis of nafuredin B by Goswami and co-workers

Separacenes A (242) and B (243) were isolated in 2013

from a marine actinomycete115 and contain an unusual tet-

raene moiety with a diols on both ends. Goswami and Das

reported the first convergent stereoselective total synthesis

(Scheme 28) shortly afterwards in 2017.29 The synthesis

started from D-tartaric acid which was transformed to un-

saturated ester 238 (see Table 1, Entry 14). A deprotec-

tion/protection and reduction/oxidation sequence yields

TBSOTf-protected aldehyde 239. They screened several con-

ditions for the HWE reaction of aldehyde 239 and keto-

phosphonate 240, such as LiCl, DIPEA, THF (0%, 24 h); LiCl,

DBU, THF (0%, 24 h); Cs2CO3, MeCN (60%, 5.5:1 dr, 12 h), and

were finally able to react 239 with 240 in the presence of

Scheme 26 Synthesis of mycolactone analogues by Altmann and co-workers

OTBS

OTBS OTBS

OHO

218

OMe

O

I

OH

I

I

COOEt

1. TMSCCH, Pd(PPh3)Cl2 CuI, Et3N2. K2CO3, 89% (2 steps)

3. DIBAL-H, 71%4. ZrCp2Cl2, AlMe3, I2, 84%

1. MnO22. 20, NaH

214 215

216

Bu3Sn

OTBS

OTBS OTBS217

220

1. 217, Pd(PPh3)4 CuI, CsF, 89%2. LiOH, 98%

POEt

O

20

O O

OO

OH

219

1. 219, 2,4,6-Cl3C6H2COCl DIPEA, DMAP, 97%

2. TBAF, then NH4F 37%

O

OOMe

OH

I

4 steps1. MnO22. 20, NaH

I

COOEt

223

221 222 1. 217, Pd(PPh3)4 CuI, CsF, 74%2. LiOH, 96%

OTBS

OTBS OTBS

OHO

224

225

1. 219, 2,4,6-Cl3C6H2COCl DIPEA, DMAP, 97%2. TBAF then NH4F 50%

OH OH

OO

O

OH

OH OH

O

Z: mycolactone A 227E: mycolactone B 228

220 225

cis-mimetics inactive

trans-mimetic active

226

A

B

O

EtOEtO

99% (2 steps)

93% (2 steps)

Si

t-Bu t-Bu

229

230231

O

CO2Et

O

OTBDPSS

OO

NN

N NPh

1. DIBAL-H, DCM, 97%2. DMP, NaHCO3, DCM3. 230, KHMDS, THF, 70%4. TBAF, THF, 98%

O

O

OH

232O

O

1. DMP, NaHCO3, DCM2. CH3MgCl, THF 84% (over 2 steps)

OH

233O

O ODMP, NaHCO3

DCM

88%

POEt

O

OEt

O

EtO

1. 20, NaH, THF, 82%2. DIBAL-H, DCM, 94%

20

234O

O

OH1. DMP, NaHCO3, DCM, quant.2. 235, KHMDS, THF, 62%

S

OO

NN

N NPh

235

236O

O

2. acryloyl chloride, DIPEA, 52%3. Grubbs-II, DCM, 70%

237O

O

OH

1. p-TsOH, MeOH, 78%

nafuredin B

Synthesis 2021, 53, 2713–2739

Page 21: Applications of the Horner–Wadsworth–Emmons Olefination in

2733

D. Roman et al. ReviewSynthesis

Ba(OH)2 to yield elongated ketone 241 (80%, 5.5:1 dr, 12 h).

They obtained 242 and 243 by stereoselective reduction of

ketone 241 with the Corey–Bakshi–Shibata reagent R-CBS

and S-CBS, respectively, and global deprotection with TBAF.

The asymmetric total synthesis of (–)-neooxazolomycin

(7) (Scheme 29) reported by Kim and co-workers is the fi-

nal example.25 The structurally complex natural product

consists of a unique lactam-lactone structure and an unsat-

urated polyketide-based sidechain that exhibits potent bio-

logical activities.116 The key substrate 244 was obtained

from D-serine in 16 steps and reacted with dimethyl

lithiomethylphosphonate to give phosphonate 246 in 82%

yield. Further modifications yielded dioxasilinane-protect-

ed phosphonate 34,117 which was then reacted with alde-

hyde 248 in a temperature-sensitive Ba(OH)2-mediated

HWE reaction to afford the partially deprotected and elon-

gated 35 in 75%. Ketone 35 was subsequently reduced to the

corresponding alcohol and acetylated to give 249. The sec-

ond fragment, 250, was synthesized in 13 steps and cou-

pled with 249. This was then treated with K2CO3 and MeOH

to produce globally deprotected (+)-neooxazolomycin (7).

2.8 HWE-Mediated Coupling of Larger Building Blocks

The intramolecular HWE reaction, where one building

block contains a phosphonate moiety and the other has an

aldehyde or ketone, has found immediate application in

synthetic chemistry and natural product synthesis. Here,

we give four recent examples.

Siladenoserinol A (15) was isolated in 2013 from a tuni-

cate collected in Indonesia,118 and found to contain a 6,8-di-

oxabicyclo[3.2.1]octane skeleton with two long chains that

each contain a sulfamated serinol or glycerophosphocho-

line moiety. Doi and co-workers reported the first total syn-

thesis of siladenoserinol A (15) in 2018 (Scheme 30).119

Their synthetic strategy started with commercially avail-

able D-malic acid, which was converted into aldehyde 251

in 21 steps. The glycerophosphocholine moiety was intro-

Scheme 28 Total synthesis of separacenes A and B by Goswami and Das

238

O

EtO2C

O

239

OTBS

OTBS

1. CSA, EtOH, 87%2. TBSOTf, 2,6-lutidine, DCM, 95%3. DIBAL-H, DCM, 95%4. IBX, EtOAc, 90%

OTBS

OTBSO

OTBS

OTBS

O

P

O

OMeOMe

240

241

240, Ba(OH)2·H2O, THF

67% (5.5:1 dr)

OH

OHOH

OH

1. (R)-CBS, BH3·SMe2

THF, 60%2. TBAF, THF, 80%

242separacene A

OH

OHOH

OH

separacene B243

1. (S)-CBS, BH3·SMe2

THF, 60%2. TBAF, THF, 80%

O

Scheme 29 Total synthesis of (+)-neooxazolomycin accomplished by Kim and co-workers

244

NMeO

O

O

N

TMS

HO

O

OCO2t-Bu

(MeO)2P–Me

245

245, n-BuLi, THF

82%

246

P

O

O

N

TMS

HO

O

OCO2t-Bu

O

MeO

MeO

1.TBAF, THF, 80%2. FeCl3, Et3SiH, DCM 51% P

O

NHO

O

OHO

MeO

MeO

O

O

P

O

NO

O

O

MeO

MeO

O

O

OSii-Pri-Pr

Si(i-Pr)2(OTf)2, 2,6-lutidine99%

FmocHNO

247

34

248

Ba(OH)2, 248 THF/H2O

75%O

NHO

O

O

O

O Si i-Pri-Pr

OH

35OAc

NHO

O

O

O

R1O Si i-Pr

i-Pr

OH1. L-Selectride, THF2. Ac2O, pyridine

75% (over 2 steps)

249

BzO

O

OH

NO

250

R1

R1

OH

NHO

O

O

O

NH

OH

HO

O

NO

7

1. BOPCl, Et3N, DCM DBU, DCM 51%

2. K2CO3, MeOH then Amberlyst IRC-86 90%

(+)-neooxazolomycin

O

Synthesis 2021, 53, 2713–2739

Page 22: Applications of the Horner–Wadsworth–Emmons Olefination in

2734

D. Roman et al. ReviewSynthesis

duced by an HWE reaction using the previously prepared

phosphonate 252 yielding the unsaturated ester 253 in 64%

yield. The Boc group was removed with HCl and sulfamate

formation finished the synthesis of siladenoserinol A (15).

Scheme 30 Total synthesis of siladenoserinol A 15 by Doi and co-workers

Tong, Qian, and co-workers described a second asym-

metric synthesis of 15 in 2019 (see Table 1, formation of

19).17 They also applied an HWE reaction for the construc-

tion of the ,-unsaturated ester moiety and included the

synthesis of siladenoserinol H, which is a structural ana-

logue, to test for antibacterial activity of this compound

class.120

Rhizoxin is a natural product produced by species of the

bacterial genus Burkholderia.121 It was isolated in 1984 and

showed high in vitro cytotoxicity and in vivo antitumor ac-

tivity.122 Fürstner and co-workers presented their own syn-

thetic effort in 2019 that extended earlier studies and re-

sulted in a different modern strategy (Scheme 31).123 The

synthesis was based on intermolecular HWE olefination for

coupling 254 and 255 followed by macrocyclization to form

a 16-membered ring through ring-closing alkyne meta-

thesis.

Their synthesis started from commercially available (R)-

oct-1-en-3-ol in its optically pure form, which was convert-

ed into substrate 254 in 14 steps. The second segment 255

was prepared from crotyl alcohol in 11 steps. A straightfor-

ward HWE reaction between phosphonate 254 and alde-

hyde 255 in the presence of LiCl and DBU was performed to

couple both segments and afforded cyclization precursor

256 in 86% yield. Ring-closing alkyne metathesis was fol-

lowed by TBAF deprotection of silyl groups and yielded de-

sired macrolactone 257. Next steps such as trans-hydro-

stannation, C-methylation, and other known procedures led

to rhizoxin D (16).

Formosalides A (261) and B (262) were isolated in 2009

by Lu and co-workers from the dinoflagellate Prorocentrum

sp. from southern Taiwan.124 These macrolides consist of

substituted tetrahydropyran (THP), a tetrahydrofuran ring,

and a C-14 linear unsaturated chain containing four cis-

double bonds.125 Both compounds exhibited good cytotoxic

activity against human T-cell, acute leukemia cells, and hu-

man colon adenocarcinoma cells and were thus selected for

total synthesis.

Mohapatra and co-workers reported a simple and effi-

cient synthesis of the C1–C16 fragment of the proposed

structure of formosalide B (262) (Scheme 32).21 The synthe-

sis started from commercially available materials and was

completed in 9 steps. Aldehyde 258 was prepared from pen-

tane-1,5-diol in 5 steps. Both building blocks were then

coupled using an HWE reaction in the presence of LiCl and

DIPEA in acetonitrile to furnish the E-isomer enone 27 in

91% yield. Olefin 27 underwent dihydroxylation using a

Sharpless ligand to give diol 259 (dr 9:1). Finally, 259 was

subjected to acid-mediated ketalization to form the THP

ring 260 and complete the fully functional C1–C16 frag-

ment of formosalide B (262).

O

AcO O

O

O11

BocN

O

PO

OEtOEt

O

O

OAc

OP

ONMe3

O O

O

AcO

O

O

BocN

O

O

O

OAc

OP

ONMe3

O O

11

DBU, LiBr, THF

O

AcO

O

O

NHSO3H

OH

O

O

OAc

OP

ONMe3

O O

11

64%

1. HCl, 83%2. SO3·py, Et3N, 58%

252

15

253

251

siladenoserinol A

+

Scheme 31 Total synthesis of rhizoxin D by Fürstner and co-workers

TBSO

OMe

O

OTBS

O

P

O

OEt

OEt

O

O

O

TBSO

OMe

O

TBSO

O

O O

LiCl, DBU86%

254 255

256

HO

OMe

O

HO

O

O O

OMe

O

HO

O

O O

NO

1. [Mo(N(t-Bu)Ar)3], 67%2. TBAF, THF, 90%

257

16rhizoxin D

Synthesis 2021, 53, 2713–2739

Page 23: Applications of the Horner–Wadsworth–Emmons Olefination in

2735

D. Roman et al. ReviewSynthesis

Scheme 32 Total synthesis of the C1–C16 fragment of formosalides by Mohapatra and co-workers

Exiguolide (271) was isolated in 2006 from a marine

sponge Geodia exigua Thiele by Ohta, Ikegami, and co-work-

ers.126 It contains a unique 20-membered macrolide core

and two tetrahydropyran rings and was shown to have in-

hibitory activity against human lung cell carcinoma, human

lung adenocarcinoma cells, pancreatic cancer, and breast

cancer cells.127 Exiguolide (271) has been a captivating tar-

get for synthesis due to its interesting structure and re-

markable biological properties. The first total synthesis of it

was accomplished in 2010 by Sasaki and Fuwa.128 Several

synthetic strategies have since been developed and de-

scribed.129 In 2020, Ishihara and co-workers reported an

asymmetric total synthesis of 271 (Scheme 33).130 The first

building block, 263, was prepared in 8 steps and the THP-

based fragment, 264, was prepared from propane-1,3-diol

in 7 steps. Phosphonate 263 and aldehyde 264 were then

subjected to an HWE reaction using NaH to give enone 265

in a 68% yield. This was then reacted with Stryke’s reagent

to provide a saturated compound.131 Cleavage of MOM and

Bn groups by treatment with trimethylsilyl iodide was ac-

complished, which was followed by silane reduction that

yielded bis(tetrahydropyran) 266 as single isomer.132 Sa-

ponification, macrolactonization, removal of the TBDPS

group, and subsequent oxidation of the secondary alcohol

to the corresponding aldehyde yielded compound 267. An

asymmetric HWE reaction based on Fuji’s method was em-

ployed to convert aldehyde 267 with chiral phosphonate

268 in the presence of NaHMDS to obtain 269 in a quantita-

tive yield (Z/E 3.6:1). Hydrostannylation and Stille coupling

produced pure (–)-exiguolide 271 (Z-isomer).

Scheme 33 Total synthesis of (–)-exiguolide by Ishihara and co-work-ers

2.9 Miscellaneous

Xu and co-workers described the first enantioselective

total synthesis of dapholdhamine B (276) in 2019, an alka-

loid with a rare aza-adamantane core skeleton (Scheme

34A).133 The critical homologation of the key building block

272 was achieved via an HWE reaction with 273 in the

presence of NaH to give intermediate 274, which then un-

derwent a one-pot acidic hydrolysis to obtain thioester 275.

This was followed by basic hydrolysis to obtain daphold-

hamine B (276) (21 steps).

P

OOMe

OMe

O

OO

OTBS

PMBO

O

O

OTBS

PMBO

LiCl, DIPEA91%

O

O

OTBS

OH

OH OPMBH

H

(DHQD)2PHAL, K3Fe(CN)6

K2CO3, MeSO2NH2

80%

OOH

OH

PMBO

OTBDPSO

HMeOH

H

PPTS, MeOH

82%

26 258

27

259

260OH

OROH

OO

O

HO

HO

O

261 formosalide A, R = H262 formosalide B, R = Me

TBDPSO

TBDPSO

TBDPSO

+

O

MeO2C

TBDPSO

OBn

OMOM O

PO

OMeMeO

O

MeO2C

TBDPSO

OBn

O

MOMO

264

263

NaH, THF

265

O

MeO2C

TBDPSO

O

HO

O

O

O

OO

1. LiOH, MeOH, H2O2. 2,4,6-Cl3C6H2COCl 85% (over 2 steps)

1. Cu(OAc)2·H2O, BSPB PMHS, 93%2. TMSI, DCM, 67%3. BF3·OEt2, Et3SiH 78% (dr > 20:1)

266

268

O

O

PO

CO2Me

267

O O

OO

MeO2C

O O

OO

MeO2C

MeO2C

NaHMDS, 268quant.Z/E = 3.6:1

1. Bu3SnH, Pd2(dba)3

Cy3P·HBF4, 54%2. CuTC, NMP, 270, 75%

I

MeOOC

269

270

271

3. Bu4NF, THF, 87%4. DMP, DCM, quant.

68%

(–)-exiguolide

O

Synthesis 2021, 53, 2713–2739

Page 24: Applications of the Horner–Wadsworth–Emmons Olefination in

2736

D. Roman et al. ReviewSynthesis

Scheme 34 Total synthesis of dapholdhamine B and caldaphnidine O by Xu and co-workers

Xu and co-workers also reported the total synthesis of

alkaloid (–)-caldaphnidine O (279) in 22 steps from cyclo-

hexane-1,3-dione (Scheme 34B).134 After the synthesis of

aldehyde 277, they employed an HWE reaction with thio-

ketal phosphonate 273 and n-BuLi followed by acidic and

basic workup to produce methyl ester 278. After diastereo-

selective hydrogenation of the alkene motif they obtained

(–)-caldaphnidine O (279).

In a second example, Lim and Choi employed an HWE

reaction for synthesis of indoles and azaindoles (Scheme

35A).135 In detail, reaction of commercially available N-Cbz-

protected phosphonoglycine trimethyl ester 280 and sub-

stituted 2-bromobenzaldehydes afforded enamines 281,

which were converted into indole 283 via Cu-catalyzed cy-

clization. This method was also envisaged for the synthesis

of thiophene- and benzothiophene-linked pyrroles, which

makes it a versatile tool in drug discovery.

Scheme 35 Synthesis of substituted indoles (A) and bicyclic imines and amines (B)

Another application is the synthesis of bicyclic imines

286 and amines 287 of the hexahydroquinoxaline-2(1H)-

one class by Wojaczyńska, Olszewski, and co-workers

(Scheme 35B). These compounds show broad pharmacolog-

ical properties.136 Subsequent HWE reactions of chiral

phosphonate 285 with different aldehydes in the presence

of NaH resulted in a tautomeric mixture of imine 286 and

enamine 287, in favor of the imine form. Additional reduc-

tion of the imine or enamine with NaBH4 led to bicyclic

amines.

3 Summary and Outlook

In recent decades, the HWE reaction has been an active

area for applications to natural product synthesis. To solve

increasingly challenging synthetic problems the required

phosphonate reagents have been diversified from simple

phosphonates to highly activated, branched, and modified

phosphonate reagents. Here, we have summarized repre-

sentative examples of functionalized reagents used in natu-

ral product synthesis reported in 2015–2020. Our compre-

hensive review will support future synthetic approaches

and serve as guideline to find the best HWE conditions for

the most complicated natural products.

Conflict of Interest

The authors declare no conflict of interest.

S S

P

273

N

H

H

H

H

O

N

H

H

H

H

OH273, NaH

SS

N

H

H

H

H

OH

SO

HS

2 M HClaq. NaOH

N

H

OO

H

H

H

H

OH

276

A

B

N

On-BuLi, 273

p-TsOH then NaOMe

N

CO2Me

N

H

CO2Me

H2, Pt/C

272 274

275

277 278

279caldaphnidine O

dapholdhamine B

OH

OEtO

EtO

OR

X

POMe

O

HNPG

R

XHN

PG

OMe

O

+

R

CuN

PG

OMe

O

L

X = B(OR)3, Br, IPG = Cbz, BocL = bpy

Z-enamine

RN

O

OMe

PG– CuL

21 examples47–98%

NH

HN O

P

1. NaH, THF

2. R-CHON

HN O

R

A

B

N

HN O PO(OMe)2, Et3N

PhMe

NH

HN O

R+

O

MeOMeO

O

OMeMeO284 285

286 287

280 281

282 283

Synthesis 2021, 53, 2713–2739

Page 25: Applications of the Horner–Wadsworth–Emmons Olefination in

2737

D. Roman et al. ReviewSynthesis

Funding Information

We greatly acknowledge funding from the European Research Council

(ERC) under the European Union’s Horizon 2020 research and innova-

tion program (Project: 802736 MORPHEUS) and Boehringer Ingel-

heim Stiftung (Exploration Grant Engina).H2020 European Research Council (802736 MORPHEUS)Boehringer Ingelheim Stiftung (Exploration Grant Engina)

References

(1) Wittig, G.; Haag, W. Chem. Ber. 1955, 88, 1654.

(2) (a) Maryanoff, B. E.; Reitz, A. B. Chem. Rev. 1989, 89, 863.

(b) Byrne, P. A.; Gilheany, D. G. Chem. Soc. Rev. 2013, 42, 6670.

(c) Rocha, D. H.; Pinto, D. C.; Silva, A. M. Eur. J. Org. Chem. 2018,

2018, 2443. (d) Heravi, M. M.; Ghanbarian, M.; Zadsirjan, V.;

Jani, A. B. Monatsh. Chem. 2019, 150, 1365.

(3) Horner, L.; Hoffmann, H.; Wippel, H. G. Chem. Ber. 1958, 91, 61.

(4) Wadsworth, W. S.; Emmons, W. D. J. Am. Chem. Soc. 1961, 83,

1733.

(5) Still, W. C.; Gennari, C. Tetrahedron Lett. 1983, 24, 4405.

(6) Janicki, I.; Kiełbasiński, P. Adv. Synth. Catal. 2020, 362, 2552.

(7) (a) Ando, K. Tetrahedron Lett. 1995, 36, 4105. (b) Ando, K. J. Org.

Chem. 1997, 62, 1934.

(8) Blanchette, M. A.; Choy, W.; Davis, J. T.; Essenfeld, A. P.;

Masamune, S.; Roush, W. R.; Sakai, T. Tetrahedron Lett. 1984, 25,

2183.

(9) Rathke, M. W.; Nowak, M. J. Org. Chem. 1985, 50, 2624.

(10) Pascariu, A.; Ilia, G.; Bora, A.; Bora, A.; Iliescu, S.; Popa, A.;

Dehelean, G.; Pacureanu, L. Cent. Eur. J. Chem. 2003, 1, 491.

(11) (a) Teichert, A.; Jantos, K.; Harms, K.; Studer, A. Org. Lett. 2004,

6, 3477. (b) Schauer, D. J.; Helquist, P. Synthesis 2006, 3654.

(12) (a) Sano, S.; Yokoyama, K.; Teranishi, R.; Shiro, M.; Nagao, Y. Tet-

rahedron Lett. 2002, 43, 281. (b) Sano, S.; Abe, S.; Azetsu, T.;

Nakao, M.; Shiro, M.; Nagao, Y. Lett. Org. Chem. 2006, 3, 798.

(13) (a) Julia, M.; Paris, J.-M. Tetrahedron Lett. 1973, 14, 4833.

(b) Blakemore, P. R.; Sephton, S. M.; Ciganek, E. Org. React. (N. Y.)

2018, 95, 1–422.

(14) (a) Peterson, D. J. J. Org. Chem. 1968, 33, 780. (b) van Staden, L.

F.; Gravestock, D.; Ager, D. J. Chem. Soc. Rev. 2002, 31, 195.

(15) (a) Mikołajczyk, M.; Midura, W. H.; Mohamed Ewas, A. M.;

Perlikowska, W.; Mikina, M.; Jankowiak, A. Phosphorus, Sulfur

Silicon Relat. Elem. 2008, 183, 313. (b) Bisceglia, J. A.; Orelli, L. R.

Curr. Org. Chem. 2012, 16, 2206. (c) Bisceglia, J. A.; Orelli, L. R.

Curr. Org. Chem. 2015, 19, 744. (d) Kobayashi, K.; Tanaka, K.;

Kogen, H. Tetrahedron Lett. 2018, 59, 568.

(16) Nicolaou, K.; Das, D.; Lu, Y.; Rout, S.; Pitsinos, E. N.; Lyssikatos, J.;

Schammel, A.; Sandoval, J.; Hammond, M.; Aujay, M. J. Am.

Chem. Soc. 2020, 142, 2549.

(17) Márquez-Cadena, M. A.; Ren, J.; Ye, W.; Qian, P.; Tong, R. Org.

Lett. 2019, 21, 9704.

(18) Shimoda, K.; Yamaoka, Y.; Yoo, D.; Yamada, K.-i.; Takikawa, H.;

Takasu, K. J. Org. Chem. 2019, 84, 11014.

(19) Scheeff, S.; Menche, D. Org. Lett. 2019, 21, 271.

(20) Lee, S.; Kang, G.; Chung, G.; Kim, D.; Lee, H.-Y.; Han, S. Angew.

Chem. Int. Ed. 2020, 59, 6894.

(21) Gajula, S.; Reddy, A. V. V.; Reddy, D. P.; Yadav, J. S.; Mohapatra,

D. K. ACS Omega 2020, 5, 10217.

(22) AnkiReddy, P.; AnkiReddy, S.; Gowravaram, S. Org. Biomol.

Chem. 2018, 16, 4191.

(23) Boeckman, R. K. Jr.; Wang, H.; Rugg, K. W.; Genung, N. E.; Chen,

K.; Ryder, T. R. Org. Lett. 2016, 18, 6136.

(24) Raji Reddy, C.; Mohammed, S. Z.; Gaddam, K.; Prapurna, Y. L.

Synth. Commun. 2019, 49, 1153.

(25) Kim, J. H.; Kim, I.; Song, Y.; Kim, M. J.; Kim, S. Angew. Chem. Int.

Ed. 2019, 58, 11018.

(26) Boyko, Y. D.; Huck, C. J.; Sarlah, D. J. Am. Chem. Soc. 2019, 141,

14131.

(27) Poock, C.; Kalesse, M. Org. Lett. 2017, 19, 4536.

(28) Bruckner, S.; Weise, M.; Schobert, R. J. Org. Chem. 2018, 83,

10805.

(29) Das, S.; Goswami, R. K. Org. Biomol. Chem. 2017, 15, 4842.

(30) Sakai, T.; Fukuta, A.; Nakamura, K.; Nakano, M.; Mori, Y. J. Org.

Chem. 2016, 81, 3799.

(31) Brumley, D. A.; Gunasekera, S. P.; Chen, Q.-Y.; Paul, V. J.; Luesch,

H. Org. Lett. 2020, 22, 4235.

(32) Dash, U.; Sengupta, S.; Sim, T. Eur. J. Org. Chem. 2015, 2015,

3963.

(33) Yang, L.; Lin, Z.; Shao, S.; Zhao, Q.; Hong, R. Angew. Chem. Int. Ed.

2018, 57, 16200.

(34) Long, X.; Ding, Y.; Deng, J. Angew. Chem. Int. Ed. 2018, 57, 14221.

(35) Nicolaou, K.; Shelke, Y. G.; Dherange, B. D.; Kempema, A.; Lin, B.;

Gu, C.; Sandoval, J.; Hammond, M.; Aujay, M.; Gavrilyuk, J. J. Org.

Chem. 2020, 85, 2865.

(36) Saha, D.; Guchhait, S.; Goswami, R. K. Org. Lett. 2020, 22, 745.

(37) Pakhare, D.; Kusurkar, R. New J. Chem. 2016, 40, 5428.

(38) Chinthakindi, P. K.; Nandi, G. C.; Govender, T.; Kruger, H. G.;

Naicker, T.; Arvidsson, P. I. Synlett 2016, 27, 1423.

(39) (a) Stork, G.; Nakamura, E. J. Org. Chem. 1979, 44, 4010.

(b) Nangia, A.; Prasuna, G.; Bheema Rao, P. Tetrahedron Lett.

1994, 35, 3755.

(40) Stork, G.; Matthews, R. J. Chem. Soc. D 1970, 445.

(41) (a) Nicolaou, K. C.; Seitz, S. P.; Pavia, M. R.; Petasis, N. A. J. Org.

Chem. 1979, 44, 4011. (b) Still, W. C.; Novack, V. J. Am. Chem. Soc.

1984, 106, 1148. (c) Still, W. C.; Gennari, C.; Noguez, J. A.;

Pearson, D. A. J. Am. Chem. Soc. 1984, 106, 260. (d) Morin-Fox, M.

L.; Lipton, M. A. Tetrahedron Lett. 1993, 34, 7899.

(42) Gourves, J.-P.; Couthon, H.; Sturtz, G. Eur. J. Org. Chem. 1999,

1999, 3489.

(43) (a) Meyers, A. I.; Comins, D. L.; Roland, D. M.; Henning, R.;

Shimizu, K. J. Am. Chem. Soc. 1979, 101, 7104. (b) Tius, M. A.;

Fauq, A. H. J. Am. Chem. Soc. 1986, 108, 1035. (c) Tius, M. A.;

Fauq, A. J. Am. Chem. Soc. 1986, 108, 6389. (d) Marshall, J. A.;

DeHoff, B. S. Tetrahedron 1987, 43, 4849. (e) Smith, A. B.; Rano,

T. A.; Chida, N.; Sulikowski, G. A.; Wood, J. L. J. Am. Chem. Soc.

1992, 114, 8008. (f) Zhang, J.; Xu, X. Tetrahedron Lett. 2000, 41,

941.

(44) (a) Wessjohann, L. A.; Ruijter, E.; Garcia-Rivera, D.; Brandt, W.

Mol. Diversity 2005, 9, 171. (b) Swinney, D. C.; Anthony, J. Nat.

Rev. Drug Discovery 2011, 10, 507. (c) Madsen, C. M.; Clausen, M.

H. Eur. J. Org. Chem. 2011, 2011, 3107. (d) Mallinson, J.; Collins, I.

Future Med. Chem. 2012, 4, 1409. (e) Yu, X.; Sun, D. Molecules

2013, 18, 6230. (f) Giordanetto, F.; Kihlberg, J. J. Med. Chem.

2014, 57, 278. (g) Newman, D. J.; Cragg, G. M. Bioactive Macro-

cycles from Nature, In Macrocycles in Drug Discovery; Chap. 1;

RSC: Cambridge, 2015, 1. (h) Mortensen, K. T.; Osberger, T. J.;

King, T. A.; Sore, H. F.; Spring, D. R. Chem. Rev. 2019, 119, 10288.

(45) Scherlach, K.; Boettger, D.; Remme, N.; Hertweck, C. Nat. Prod.

Rep. 2010, 27, 869.

(46) (a) Schofield, J. Nat. New Biol. 1971, 234, 215. (b) Turner, W. B.;

Carter, S. B. Biochem. J. 1972, 127, 1P. (c) Crivello, J. F.; Jefcoate,

C. R. Biochim. Biophys. Acta, Gen. Subj. 1978, 542, 315. (d) Hu, Y.;

Zhang, W.; Zhang, P.; Ruan, W.; Zhu, X. J. Agric. Food Chem.

2013, 61, 41. (e) Hua, C.; Yang, Y.; Sun, L.; Dou, H.; Tan, R.; Hou,

Y. Immunobiology 2013, 218, 292.

(47) Stork, G.; Nakamura, E. J. Am. Chem. Soc. 1983, 105, 5510.

Synthesis 2021, 53, 2713–2739

Page 26: Applications of the Horner–Wadsworth–Emmons Olefination in

2738

D. Roman et al. ReviewSynthesis

(48) Stork, G.; Nakahara, Y.; Nakahara, Y.; Greenlee, W. J. Am. Chem.

Soc. 1978, 100, 7775.

(49) Thomas, E. J.; Whitehead, J. W. F. J. Chem. Soc., Chem. Commun.

1986, 727.

(50) Merifield, E.; Thomas, E. J. J. Chem. Soc., Perkin Trans. 1 1999,

3269.

(51) Thomas, E. J.; Willis, M. Org. Biomol. Chem. 2014, 12, 7537.

(52) Trost, B. M.; Ohmori, M.; Boyd, S. A.; Okawara, H.; Brickner, S. J.

J. Am. Chem. Soc. 1989, 111, 8281.

(53) Reyes, J. R.; Winter, N.; Spessert, L.; Trauner, D. Angew. Chem.

Int. Ed. 2018, 57, 15587.

(54) Draskovits, M.; Stanetty, C.; Baxendale, I. R.; Mihovilovic, M. D.

J. Org. Chem. 2018, 83, 2647.

(55) Enders, D.; Geibel, G.; Osborne, S. Chem. Eur. J. 2000, 6, 1302.

(56) Zaghouani, M.; Kunz, C.; Guédon, L.; Blanchard, F.; Nay, B. Chem.

Eur. J. 2016, 22, 15257.

(57) Forsyth, C. J.; Ahmed, F.; Cink, R. D.; Lee, C. S. J. Am. Chem. Soc.

1998, 120, 5597.

(58) Haidle, A. M.; Myers, A. G. Proc. Natl. Acad. Sci. U.S.A. 2004, 101,

12048.

(59) Larsen, B. J.; Sun, Z.; Nagorny, P. Org. Lett. 2013, 15, 2998.

(60) (a) Zhanel, G. G.; Dueck, M.; Hoban, D. J.; Vercaigne, L. M.;

Embil, J. M.; Gin, A. S.; Karlowsky, J. A. Drugs 2001, 61, 443.

(b) Zhanel, G. G.; Walters, M.; Noreddin, A.; Vercaigne, L. M.;

Wierzbowski, A.; Embil, J. M.; Gin, A. S.; Douthwaite, S.; Hoban,

D. J. Drugs 2002, 62, 1771. (c) Cui, W.; Ma, S. Mini-Rev. Med.

Chem. 2011, 11, 1009. (d) Arsic, B.; Barber, J.; Čikoš, A.;

Mladenovic, M.; Stankovic, N.; Novak, P. Int. J. Antimicrob. Agents

2018, 51, 283.

(61) Liu, S.; Dai, H.; Makhloufi, G.; Heering, C.; Janiak, C.; Hartmann,

R.; Mándi, A.; Kurtán, T.; Müller, W. E.; Kassack, M. U.; Lin, W.;

Liu, Z.; Proksch, P. J. Nat. Prod. 2016, 79, 2332.

(62) Paul, D.; Saha, S.; Goswami, R. K. Org. Lett. 2018, 20, 4606.

(63) Dias, L. C.; de Lucca, E. C. J. Org. Chem. 2017, 82, 3019.

(64) (a) Takai, K.; Nitta, K.; Utimoto, K. J. Am. Chem. Soc. 1986, 108,

7408. (b) Okazoe, T.; Takai, K.; Utimoto, K. J. Am. Chem. Soc.

1987, 109, 51.

(65) (a) Inanaga, J.; Hirata, K.; Saeki, H.; Katsuki, T.; Yamaguchi, M.

Bull. Chem. Soc. Jpn. 1979, 52, 1989. (b) Kawanami, Y.; Dainobu,

Y.; Inanaga, J.; Katsuki, T.; Yamaguchi, M. Bull. Chem. Soc. Jpn.

1981, 54, 943.

(66) Clarke, A. K.; Unsworth, W. P. Chem. Sci. 2020, 11, 2876.

(67) Hubert, J. G.; Furkert, D. P.; Brimble, M. A. J. Org. Chem. 2015, 80,

2231.

(68) (a) Wang, W.; Mun, B.; Lee, Y.; Venkat Reddy, M.; Park, Y.; Lee,

J.; Kim, H.; Hahn, D.; Chin, J.; Ekins, M.; Nam, S.-J.; Kang, H.

J. Nat. Prod. 2013, 76, 170. (b) Hubert, J. G.; Furkert, D. P.;

Brimble, M. A. J. Org. Chem. 2015, 80, 2715.

(69) Saya, J. M.; Vos, K.; Kleinnijenhuis, R. A.; van Maarseveen, J. H.;

Ingemann, S.; Hiemstra, H. Org. Lett. 2015, 17, 3892.

(70) (a) Jeon, S.; Han, S. J. Am. Chem. Soc. 2017, 139, 6302. (b) Wang,

Y.; Chen, B.; He, X.; Gui, J. J. Am. Chem. Soc. 2020, 142, 5007.

(71) (a) Chirkin, E.; Atkatlian, W.; Porée, F.-H. The Securinega Alka-

loids, In The Alkaloids: Chemistry and Biology, Vol. 74, Chap. 1;

Knölker, H.-J., Ed.; Academic Press: San Diego, 2015, 1.

(b) Zhang, H.; Zhang, C.-R.; Han, Y.-S.; Wainberg, M. A.; Yue, J.-

M. RSC Adv. 2015, 5, 107045. (c) Park, K. J.; Kim, C. S.; Khan, Z.;

Oh, J.; Kim, S. Y.; Choi, S. U.; Lee, K. R. J. Nat. Prod. 2019, 82,

1345.

(72) Xie, P.; Huang, Y. Eur. J. Org. Chem. 2013, 28, 6213.

(73) (a) Cimmino, A.; Andolfi, A.; Fondevilla, S.; Abouzeid, M. A.;

Rubiales, D.; Evidente, A. J. Agric. Food Chem. 2012, 60, 5273.

(b) Yadav, J. S.; Avuluri, S.; Kattela, S. S.; Das, S. Eur. J. Org. Chem.

2013, 2013, 6967. (c) Jangili, P.; Kumar, C. G.; Poornachandra,

Y.; Das, B. Synthesis 2015, 47, 653.

(74) Shelke, A. M.; Suryavanshi, G. Tetrahedron: Asymmetry 2016, 27,

142.

(75) Kobayashi, J.-i.; Watanabe, D.; Kawasaki, N.; Tsuda, M. J. J. Org.

Chem. 1997, 62, 9236.

(76) (a) Sakai, R.; Higa, T.; Jefford, C. W.; Bernardinelli, G. J. J. Am.

Chem. Soc. 1986, 108, 6404. (b) Ashok, P.; Ganguly, S.;

Murugesan, S. Drug Discovery Today 2014, 19, 1781. (c) Ashok,

P.; Lathiya, H.; Murugesan, S. Eur. J. Med. Chem. 2015, 97, 928.

(d) Kondo, K.; Shigemori, H.; Kikuchi, Y.; Ishibashi, M.; Sasaki,

T.; Kobayashi, J. J. Org. Chem. 1992, 57, 2480.

(77) Hovey, M. T.; Cohen, D. T.; Walden, D. M.; Cheong, P. H.-Y.;

Scheidt, K. A. Angew. Chem. Int. Ed. 2017, 56, 9864.

(78) (a) Lumyong, K.; Kongkathip, B.; Chuanopparat, N.; Kongkathip,

N. Tetrahedron 2019, 75, 533. (b) Fitch, R. W.; Garraffo, H. M.;

Spande, T. F.; Yeh, H. J.; Daly, J. W. J. Nat. Prod. 2003, 66, 1345.

(79) (a) Raluy, E.; Pàmies, O.; Dieguez, M. Adv. Synth. Catal. 2009,

351, 1648. (b) Kongkathip, B.; Akkarasamiyo, S.; Kongkathip, N.

Tetrahedron 2015, 71, 2393.

(80) Fulton, M. G.; Bertron, J. L.; Reed, C. W.; Lindsley, C. W. J. Org.

Chem. 2019, 84, 12187.

(81) Zhang, D.; Tao, X.; Chen, R.; Liu, J.; Li, L.; Fang, X.; Yu, L.; Dai, J.

Org. Lett. 2015, 17, 4304.

(82) Lücke, D.; Linne, Y.; Hempel, K.; Kalesse, M. Org. Lett. 2018, 20,

4475.

(83) Gao, S.-S.; Li, X.-M.; Du, F.-Y.; Li, C.-S.; Proksch, P.; Wang, B.-G.

Mar. Drugs 2011, 9, 59.

(84) Sato, S.; Iwata, F.; Mukai, T.; Yamada, S.; Takeo, J.; Abe, A.;

Kawahara, H. J. Org. Chem. 2009, 74, 5502.

(85) Hu, N.; Dong, C.; Zhang, C.; Liang, G. Angew. Chem. Int. Ed. 2019,

58, 6659.

(86) Nicolaou, K.; Wang, Y.; Lu, M.; Mandal, D.; Pattanayak, M. R.; Yu,

R.; Shah, A. A.; Chen, J. S.; Zhang, H.; Crawford, J. J. J. Am. Chem.

Soc. 2016, 138, 8235.

(87) (a) Ley, S. V.; Baxendale, I. R.; Bream, R. N.; Jackson, P. S.; Leach,

A. G.; Longbottom, D. A.; Nesi, M.; Scott, J. S.; Storer, R. I.; Taylor,

S. J. J. Chem. Soc., Perkin Trans. 1 2000, 3815. (b) Solinas, A.;

Taddei, M. Synthesis 2007, 2409.

(88) Johnson, C. R.; Zhang, B. Tetrahedron Lett. 1995, 36, 9253.

(89) Barrett, A. G. M.; Cramp, S. M.; Roberts, R. S.; Zecri, F. J. Org. Lett.

1999, 1, 579.

(90) Nicolaou, K. C.; Pastor, J.; Winssinger, N.; Murphy, F. J. Am.

Chem. Soc. 1998, 120, 5132.

(91) Salvino, J. M.; Kiesow, T. J.; Darnbrough, S.; Labaudiniere, R.

J. Comb. Chem. 1999, 1, 134.

(92) (a) Wipf, P.; Henninger, T. C. J. Org. Chem. 1997, 62, 1586.

(b) Boldi, A. M.; Johnson, C. R.; Eissa, H. O. Tetrahedron Lett.

1999, 40, 619.

(93) (a) Martina, S. L. X.; Taylor, R. J. K. Tetrahedron Lett. 2004, 45,

3279. (b) Ando, K.; Suzuki, Y. Tetrahedron Lett. 2010, 51, 2323.

(94) Jin, Y. Z.; Yasuda, N.; Inanaga, J. Green Chem. 2002, 4, 498.

(95) Popa, A.; Avram, E.; Lisa, G.; Visa, A.; Iliescu, S.; Parvulescu, V.;

Ilia, G. Polym. Eng. Sci. 2012, 52, 352.

(96) Roman, D.; Raguž, L.; Keiff, F.; Meyer, F.; Barthels, F.;

Schirmeister, T.; Kloss, F.; Beemelmanns, C. Org. Lett. 2020, 22,

3744.

(97) Rischer, M.; Raguž, L.; Guo, H.; Keiff, F.; Diekert, G.; Goris, T.;

Beemelmanns, C. ACS Chem. Biol. 2018, 13, 1990.

(98) Fuson, R. C. Chem. Rev. 1935, 16, 1.

Synthesis 2021, 53, 2713–2739

Page 27: Applications of the Horner–Wadsworth–Emmons Olefination in

2739

D. Roman et al. ReviewSynthesis

(99) (a) Masuda, T.; Osako, K.; Shimizu, T.; Nakata, T. Org. Lett. 1999,

1, 941. (b) Barth, R.; Mulzer, J. Angew. Chem. Int. Ed. 2007, 46,

5791. (c) Denmark, S. E.; Regens, C. S.; Kobayashi, T. J. Am. Chem.

Soc. 2007, 129, 2774. (d) Sirasani, G.; Paul, T.; Andrade, R. B.

J. Org. Chem. 2008, 73, 6386.

(100) (a) Fadeyi, O. O.; Lindsley, C. W. Org. Lett. 2009, 11, 3950.

(b) Takamura, H.; Kikuchi, S.; Nakamura, Y.; Yamagami, Y.;

Kishi, T.; Kadota, I.; Yamamoto, Y. Org. Lett. 2009, 11, 2531.

(c) Chakor, N. S.; Musso, L.; Dallavalle, S. J. Org. Chem. 2009, 74,

844. (d) Geerdink, D.; Buter, J.; van Beek, T. A.; Minnaard, A. J.

Beilstein J. Org. Chem. 2014, 10, 761.

(101) Schmidt, K.; Riese, U.; Li, Z.; Hamburger, M. J. Nat. Prod. 2003,

66, 378.

(102) Zhang, J.; Meng, L.-L.; Wei, J.-J.; Fan, P.; Liu, S.-S.; Yuan, W.-Y.;

Zhao, Y.-X.; Luo, D.-Q. Molecules 2017, 22, 2058.

(103) Loscher, S.; Schobert, R. Chem. Eur. J. 2013, 19, 10619.

(104) Gerstmann, L.; Kalesse, M. Chem. Eur. J. 2016, 22, 11210.

(105) Herrmann, A. T.; Martinez, S. R.; Zakarian, A. Org. Lett. 2011, 13,

3636.

(106) Gehringer, M.; Mäder, P.; Gersbach, P.; Pfeiffer, B.; Scherr, N.;

Dangy, J.-P.; Pluschke, G.; Altmann, K.-H. Org. Lett. 2019, 21,

5853.

(107) Bieri, R.; Scherr, N.; Ruf, M.-T.; Dangy, J.-P.; Gersbach, P.;

Gehringer, M.; Altmann, K.-H.; Pluschke, G. ACS Chem. Biol.

2017, 12, 1297.

(108) McKenna, M.; Simmonds, R. E.; High, S. J. Cell Sci. 2016, 129,

1404.

(109) Chany, A.-C.; Veyron-Churlet, R.; Tresse, C.; Mayau, V.;

Casarotto, V.; Le Chevalier, F.; Guenin-Macé, L.; Demangel, C.;

Blanchard, N. J. Med. Chem. 2014, 57, 7382.

(110) Marion, E.; Song, O.-R.; Christophe, T.; Babonneau, J.; Fenistein,

D.; Eyer, J.; Letournel, F.; Henrion, D.; Clere, N.; Paille, V. Cell

2014, 157, 1565.

(111) Zhang, Z.; He, X.; Zhang, G.; Che, Q.; Zhu, T.; Gu, Q.; Li, D. J. Nat.

Prod. 2017, 80, 3167.

(112) Mandal, G. H.; Saha, D.; Goswami, R. K. Org. Biomol. Chem. 2020,

18, 2346.

(113) Blakemore, P. R.; Cole, W. J.; Kocienski, P.; Morley, A. Synlett

1998, 26.

(114) Seki, M.; Mori, K. Eur. J. Org. Chem. 2001, 2001, 503.

(115) Bae, M.; Kim, H.; Shin, Y.; Kim, B. Y.; Lee, S. K.; Oh, K.-B.; Shin, J.;

Oh, D.-C. Mar. Drugs 2013, 11, 2882.

(116) (a) Angelov, P.; Chau, Y. K. S.; Fryer, P. J.; Moloney, M. G.;

Thompson, A. L.; Trippier, P. C. Org. Biomol. Chem. 2012, 10,

3472. (b) Josa-Culleré, L.; Towers, C.; Willenbrock, F.; Macaulay,

V. M.; Christensen, K. E.; Moloney, M. G. Org. Biomol. Chem.

2017, 15, 5373.

(117) Onyango, E. O.; Tsurumoto, J.; Imai, N.; Takahashi, K.; Ishihara,

J.; Hatakeyama, S. Angew. Chem. Int. Ed. 2007, 46, 6703.

(118) Nakamura, Y.; Kato, H.; Nishikawa, T.; Iwasaki, N.; Suwa, Y.;

Rotinsulu, H.; Losung, F.; Maarisit, W.; Mangindaan, R. E. P.;

Morioka, H.; Yokosawa, H.; Tsukamoto, S. Org. Lett. 2013, 15,

322.

(119) Yoshida, M.; Saito, K.; Kato, H.; Tsukamoto, S.; Doi, T. Angew.

Chem. Int. Ed. 2018, 57, 5147.

(120) (a) Tischer, M.; Pradel, G.; Ohlsen, K.; Holzgrabe, U. ChemMed-

Chem 2012, 7, 22. (b) Petkowski, J. J.; Bains, W.; Seager, S. J. Nat.

Prod. 2018, 81, 423.

(121) Scherlach, K.; Partida-Martinez, L. P.; Dahse, H.-M.; Hertweck,

C. J. Am. Chem. Soc. 2006, 128, 11529.

(122) (a) Iwasaki, S.; Kobayashi, H.; Furukawa, J.; Namikoshi, M.;

Okuda, S.; Sato, Z.; Matsuda, I.; Noda, T. J. Antibiot. 1984, 37, 354.

(b) Tsuruo, T.; Oh-hara, T.; Iida, H.; Tsukagoshi, S.; Sato, Z.;

Matsuda, I.; Iwasaki, S.; Okuda, S.; Shimizu, F.; Sasagawa, K.

Cancer Res. 1986, 46, 381. (c) Takahashi, M.; Iwasaki, S.;

Kobayashi, H.; Okuda, S.; Murai, T.; Sato, Y.; Haraguchi-Hiraoka,

T.; Nagano, H. J. Antibiot. 1987, 40, 66.

(123) Karier, P.; Ungeheuer, F.; Ahlers, A.; Anderl, F.; Wille, C.;

Fürstner, A. Angew. Chem. Int. Ed. 2019, 58, 248.

(124) Lu, C.-K.; Chen, Y.-M.; Wang, S.-H.; Wu, Y.-Y.; Cheng, Y.-M. Tetra-

hedron Lett. 2009, 50, 1825.

(125) (a) Srinivas, B.; Reddy, D. S.; Mallampudi, N. A.; Mohapatra, D. K.

Org. Lett. 2018, 20, 6910. (b) Mallampudi, N. A.; Srinivas, B.;

Reddy, J. G.; Mohapatra, D. K. Org. Lett. 2019, 21, 5952.

(126) Ohta, S.; Uy, M. M.; Yanai, M.; Ohta, E.; Hirata, T.; Ikegami, S. Tet-

rahedron Lett. 2006, 47, 1957.

(127) (a) Crane, E. A.; Zabawa, T. P.; Farmer, R. L.; Scheidt, K. A. Angew.

Chem. Int. Ed. 2011, 50, 9112. (b) Fuwa, H.; Suzuki, T.; Kubo, H.;

Yamori, T.; Sasaki, M. Chem. Eur. J. 2011, 17, 2678. (c) Fuwa, H.;

Mizunuma, K.; Sasaki, M.; Suzuki, T.; Kubo, H. Org. Biomol.

Chem. 2013, 11, 3442.

(128) Fuwa, H.; Sasaki, M. Org. Lett. 2010, 12, 584.

(129) (a) Cook, C.; Guinchard, X.; Liron, F.; Roulland, E. Org. Lett. 2010,

12, 744. (b) Reddy, C. R.; Rao, N. N. Tetrahedron Lett. 2010, 51,

5840. (c) Cook, C.; Liron, F.; Guinchard, X.; Roulland, E. J. Org.

Chem. 2012, 77, 6728. (d) Zhang, Z.; Xie, H.; Li, H.; Gao, L.; Song,

Z. Org. Lett. 2015, 17, 4706. (e) Zhang, Y.; Lu, J.; Li, H.; Sun, X.;

Gao, L.; Song, Z. Synlett 2019, 30, 753.

(130) Oka, K.; Fuchi, S.; Komine, K.; Fukuda, H.; Hatakeyama, S.;

Ishihara, J. Chem. Eur. J. 2020, 26, 12862.

(131) Baker, B. A.; Bošković, Ž. V.; Lipshutz, B. H. Org. Lett. 2008, 10,

289.

(132) Lewis, M. D.; Cha, J. K.; Kishi, Y. J. Am. Chem. Soc. 1982, 104,

4976.

(133) Guo, L.-D.; Hou, J.; Tu, W.; Zhang, Y.; Zhang, Y.; Chen, L.; Xu, J.

J. Am. Chem. Soc. 2019, 141, 11713.

(134) Guo, L.-D.; Hu, J.; Zhang, Y.; Tu, W.; Zhang, Y.; Pu, F.; Xu, J. J. Am.

Chem. Soc. 2019, 141, 13043.

(135) Choi, J. H.; Lim, H. J. Org. Biomol. Chem. 2015, 13, 5131.

(136) Iwanejko, J.; Sowiński, M.; Wojaczyńska, E.; Olszewski, T. K.;

Górecki, M. RSC Adv. 2020, 10, 14618.

Synthesis 2021, 53, 2713–2739