Modern Alkyne Chemistry -  - E-Book

Modern Alkyne Chemistry E-Book

0,0
153,99 €

-100%
Sammeln Sie Punkte in unserem Gutscheinprogramm und kaufen Sie E-Books und Hörbücher mit bis zu 100% Rabatt.
Mehr erfahren.
Beschreibung

A comprehensive and up-to-date overview of alkyne chemistry, taking into account the progress made over the last two decades. The experienced editors are renowned world leaders in the field, while the list of contributors reads like a "Who's Who" of synthetic organic chemistry. The result is a valuable reference not only for organic chemists at universities and in the chemical industry, but also for biologists and material scientists involved in the modern synthesis of organic compounds and materials.

Sie lesen das E-Book in den Legimi-Apps auf:

Android
iOS
von Legimi
zertifizierten E-Readern

Seitenzahl: 608

Veröffentlichungsjahr: 2014

Bewertungen
0,0
0
0
0
0
0
Mehr Informationen
Mehr Informationen
Legimi prüft nicht, ob Rezensionen von Nutzern stammen, die den betreffenden Titel tatsächlich gekauft oder gelesen/gehört haben. Wir entfernen aber gefälschte Rezensionen.



Table of Contents

Cover

Related Titles

Title Page

Copyright

List of Contributors

Preface

Chapter 1: Introduction

1.1 History of Alkynes

1.2 Structure and Properties of Alkynes

1.3 Classical Reactions of Alkynes

1.4 Modern Reactions

1.5 Conclusion

References

Part I: Catalytic Isomerization of Alkynes

Chapter 2: Redox Isomerization of Propargyl Alcohols to Enones

2.1 Introduction

2.2 Base Catalysis

2.3 Ru Catalyzed

2.4 Rh Catalysis

2.5 Palladium Catalysis

2.6 Miscellaneous

2.7 Conclusions

References

Chapter 3: Carbophilic Cycloisomerization Reactions of Enynesand Domino Processes

3.1 Introduction and Reactivity Principles

3.2 Skeletal Rearrangement Reactions in the Absence of Nucleophiles

3.3 Enyne Domino Processes

3.4 Conclusion

References

Chapter 4: Alkyne Metathesis in Organic Synthesis

4.1 Introduction

4.2 Mechanistic Background and Classical Catalyst Systems

4.3 State-of-the-Art Catalysts

4.4 Basic Reaction Formats and Substrate Scope

4.5 Selected Applications

4.6 Conclusions

References

Part II: Catalytic Cycloaddition Reactions

Chapter 5: Alkyne–Azide Reactions

5.1 Introduction

5.2 Reviews on Cu-Catalyzed Azide–Alkyne Cycloaddition

5.3 Mechanistic Considerations on the Cu(1) Catalysis

5.4 The Substrates for CuAAC

5.5 The Environment

5.6 Modified 1,2,3-Triazoles and CuAAC Side Reactions

5.7 The Catalyst

5.8 Optimizing Conditions for CuAAC Reactions

5.9 CuAAC in Biological Applications

5.10 Biocompatibility of the CuAAC Reaction

References

Chapter 6: Catalytic Cycloaddition Reactions

6.1 Introduction

6.2 (2 + 2) Cycloaddition

6.3 (3 + 2) and (2 + 1) Cycloaddition

6.4 (4 + 2) Cycloaddition

6.5 (5 + 1) and (4 + 1) Cycloadditions

6.6 (5 + 2) Cycloaddition

6.7 (6 + 2) Cycloaddition

6.8 (2 + 2 + 1) Cycloaddition

6.9 (2 + 2 + 2) Cycloaddition

6.10 (3 + 2 + 1) Cycloaddition

6.11 (3 + 2 + 2) Cycloaddition

6.12 (4 + 2 + 1) and (4 + 2 + 2) Cycloaddition

6.13 (4 + 3 + 2) Cycloaddition

6.14 (5 + 2 + 1) and (5 + 1 + 2 + 1) Cycloadditions

6.15 (2 + 2 + 1 + 1) and (2 + 2 + 2 + 1) Cycloadditions

6.16 (2 + 2 + 2 + 2) Cycloaddition

6.17 Conclusions

References

Part III: Catalytic Nucleophilic Additions and Substitutions

Chapter 7: Catalytic Conjugate Additions of Alkynes

7.1 Introduction

7.2 Metal Alkynylides as Nucleophiles

7.3 Direct Use of Terminal Alkynes as Pronucleophiles

7.4 Summary and Conclusions

References

Chapter 8: Catalytic Enantioselective Addition of Terminal Alkynes to Carbonyls

8.1 Introduction

8.2 Metallation of Terminal Alkynes: Formation of Alkynyl Nucleophiles

8.3 Ligand-Catalyzed Alkyne Additions with Stoichiometric Quantities of Metal

8.4 Alkyne Additions with Catalytic Amounts of Metal

8.5 Concluding Remarks

References

Chapter 9: Catalytic Nucleophilic Addition of Alkynes to Imines: The A3 (Aldehyde–Alkyne–Amine) Coupling

9.1 A

3

Couplings Involving Primary Amines

9.2 A

3

Couplings Involving Secondary Amines

9.3 Alkyne Additions with Reusable Catalysts

9.4 Asymmetric Alkyne Addition Reactions

9.5 Alkyne Additions to Imines in Tandem Reactions

9.6 Conclusion

References

Chapter 10: The Sonogashira Reaction

10.1 Introduction

10.2 Palladium–Phosphorous Catalysts

10.3 Palladium–Nitrogen Catalysts

10.4 N-Heterocyclic Carbene (NHC)-Palladium Catalysts

10.5 Palladacycles as Catalysts

10.6 Ligand-Free Palladium Salts as Catalysts

10.7 Palladium Nanoparticles as Catalysts

10.8 Non-Palladium-Based Catalysts

10.9 Mechanistic Considerations

10.10 Summary and Conclusions

References

Part IV: Other Reactions

Chapter 11: Catalytic Dimerization of Alkynes

11.1 Introduction

11.2 Dimerization of Alkynes Catalyzed by Iron, Ruthenium, and Osmium Complexes

11.3 Dimerization of Alkynes Catalyzed by Cobalt, Rhodium, and Iridium Complexes

11.4 Dimerization of Alkynes Catalyzed by Nickel, Palladium, and Platinum Complexes

11.5 Dimerization of Alkynes Catalyzed by Group 3, Lanthanide, and Actinide Complexes

11.6 Dimerization of Alkynes Catalyzed by Titanium, Zirconium, and Hafnium Complexes

11.7 Dimerization of Alkynes Catalyzed by Other Compounds

11.8 Summary and Conclusions

Acknowledgments

References

Chapter 12: The Oxidative Dimerization of Acetylenes and Related Reactions: Synthesis and Applications of Conjugated 1,3-Diynes

12.1 Introduction

12.2 Syntheses of Conjugated 1,3-Diynes

12.3 Scope and Limitation of the Alkyne Dimerization Reaction

12.4 Scope and Limitation of Copper-Catalyzed Hetero-Coupling Reactions

12.5 The Cadiot–Chodkiewicz Reaction

12.6 Palladium-Catalyzed Acetylenic Coupling Reactions

12.7 Alternative Methods for the Synthesis of Diynes

12.8 Mechanism of Alkyne Homo-Coupling Reactions

12.9 Mechanism of Alkyne Hetero-Coupling Reactions

12.10 Utility of 1,3-Diynes in the Synthesis of Natural Products

12.11 Synthetic Utility of Conjugated 1,3-Diynes

12.12 Utility of 1,3-Diynes in Materials Science

12.13 Conclusion

References

Chapter 13: The Alkyne Zipper Reaction in Asymmetric Synthesis

13.1 Introduction

13.2 Mechanism of KNH

2

/NH

3

Isomerization

13.3 Mechanism of KAPA Isomerization

13.4 Applications in Natural Products

13.5 Conclusion

References

Index

End User License Agreement

Pages

XIII

XIV

XV

XVII

1

2

3

4

5

6

7

8

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

107

108

109

110

111

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

235

236

237

239

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

328

329

330

331

332

333

334

335

336

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

352

354

353

355

356

357

358

359

360

361

362

363

365

366

367

368

369

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

9

113

171

299

395

396

397

398

399

400

401

402

Guide

Cover

Table of Contents

Preface

Part I: Catalytic Isomerization of Alkynes

Chapter 1: Introduction

List of Illustrations

Figure 2.1

Scheme 2.1

Scheme 2.2

Scheme 3.1

Scheme 3.2

Scheme 3.3

Scheme 3.4

Scheme 3.5

Scheme 3.6

Scheme 3.7

Scheme 3.8

Scheme 3.9

Scheme 3.10

Scheme 3.11

Scheme 3.12

Scheme 3.13

Scheme 3.14

Scheme 3.15

Scheme 3.16

Scheme 3.17

Scheme 3.18

Scheme 3.19

Scheme 3.20

Scheme 3.21

Scheme 3.22

Scheme 3.23

Scheme 3.24

Scheme 3.25

Scheme 3.26

Scheme 3.27

Scheme 3.28

Scheme 3.29

Scheme 3.30

Scheme 3.31

Scheme 3.32

Scheme 3.33

Scheme 3.34

Scheme 3.35

Scheme 3.36

Scheme 3.37

Scheme 3.38

Scheme 3.39

Scheme 3.40

Scheme 3.41

Scheme 4.1

Scheme 4.2

Scheme 4.3

Scheme 4.4

Scheme 4.5

Scheme 4.6

Figure 4.1

Scheme 4.7

Scheme 4.8

Scheme 4.9

Scheme 4.10

Scheme 4.11

Scheme 4.12

Scheme 4.13

Scheme 4.14

Scheme 4.15

Scheme 4.16

Scheme 4.17

Scheme 4.18

Scheme 4.19

Scheme 4.20

Scheme 4.21

Scheme 4.22

Scheme 4.23

Scheme 4.24

Scheme 4.25

Scheme 4.26

Scheme 4.27

Scheme 4.28

Scheme 4.29

Scheme 4.30

Scheme 4.31

Scheme 4.32

Scheme 4.33

Figure 5.1

Figure 5.2

Scheme 5.1

Figure 5.3

Figure 5.4

Figure 5.5

Figure 5.6

Figure 5.7

Figure 5.8

Figure 5.9

Figure 5.10

Scheme 6.1

Scheme 6.2

Scheme 6.3

Scheme 6.4

Scheme 6.5

Scheme 6.6

Scheme 6.7

Scheme 6.8

Scheme 6.9

Scheme 6.10

Scheme 6.11

Scheme 6.12

Scheme 6.13

Scheme 6.14

Scheme 6.15

Scheme 6.16

Scheme 6.17

Scheme 6.18

Scheme 6.19

Scheme 6.20

Scheme 6.21

Scheme 6.22

Scheme 6.23

Scheme 6.24

Scheme 6.26

Scheme 6.25

Scheme 6.27

Scheme 6.28

Scheme 6.29

Scheme 6.30

Scheme 6.31

Scheme 6.32

Scheme 7.1

Figure 7.1

Scheme 7.2

Scheme 7.3

Scheme 7.4

Scheme 7.5

Scheme 7.6

Scheme 7.7

Scheme 7.8

Scheme 7.9

Scheme 7.10

Scheme 7.11

Scheme 7.12

Scheme 7.13

Scheme 7.14

Scheme 7.15

Scheme 7.16

Scheme 7.17

Scheme 7.18

Scheme 7.19

Scheme 7.20

Scheme 7.21

Scheme 7.22

Scheme 7.23

Scheme 7.24

Scheme 7.25

Scheme 7.26

Scheme 7.27

Scheme 7.28

Figure 7.2

Figure 8.1

Figure 8.2

Figure 8.3

Scheme 8.1

Figure 8.4

Scheme 8.2

Scheme 8.3

Scheme 8.4

Scheme 8.5

Scheme 8.6

Scheme 8.7

Scheme 8.8

Scheme 8.9

Scheme 8.10

Figure 8.5

Scheme 8.11

Scheme 8.12

Scheme 8.13

Scheme 8.14

Scheme 8.15

Scheme 8.17

Scheme 8.16

Scheme 8.18

Scheme 8.19

Scheme 8.20

Scheme 8.21

Scheme 8.22

Scheme 8.23

Scheme 8.24

Scheme 8.25

Scheme 8.26

Scheme 8.27

Scheme 8.28

Scheme 8.29

Scheme 8.30

Scheme 8.31

Scheme 8.32

Scheme 8.33

Scheme 8.34

Scheme 8.35

Scheme 8.36

Scheme 8.37

Scheme 8.38

Scheme 8.39

Scheme 8.40

Scheme 8.41

Scheme 8.42

Scheme 8.43

Scheme 8.44

Scheme 8.45

Scheme 9.1

Scheme 9.2

Scheme 9.3

Scheme 9.4

Scheme 9.5

Scheme 9.6

Scheme 9.7

Scheme 9.8

Scheme 9.9

Scheme 9.10

Scheme 9.11

Scheme 9.12

Scheme 9.13

Scheme 9.14

Scheme 9.15

Scheme 9.16

Scheme 9.17

Scheme 9.18

Scheme 9.19

Scheme 9.20

Scheme 9.21

Scheme 9.22

Scheme 9.23

Scheme 9.24

Scheme 9.25

Scheme 9.26

Scheme 10.1

Scheme 10.2

Scheme 10.3

Scheme 10.4

Scheme 10.5

Scheme 10.6

Scheme 10.7

Scheme 10.8

Scheme 10.9

Scheme 10.10

Scheme 10.11

Scheme 10.12

Scheme 10.13

Scheme 10.14

Scheme 10.15

Scheme 10.16

Scheme 10.17

Scheme 10.18

Scheme 10.19

Scheme 10.20

Scheme 10.21

Scheme 10.22

Scheme 10.23

Scheme 10.24

Scheme 10.25

Figure 10.1

Figure 10.2

Figure 11.1

Scheme 11.1

Scheme 11.2

Scheme 11.3

Scheme 11.4

Scheme 11.5

Scheme 11.6

Scheme 11.7

Scheme 11.8

Scheme 11.9

Scheme 11.10

Scheme 11.11

Scheme 11.12

Scheme 11.13

Scheme 11.14

Scheme 11.15

Scheme 11.16

Scheme 11.17

Scheme 11.18

Scheme 11.19

Scheme 11.20

Scheme 11.21

Scheme 11.22

Figure 11.2

Scheme 11.23

Scheme 11.24

Scheme 11.25

Scheme 11.26

Scheme 11.27

Scheme 11.28

Scheme 11.29

Scheme 11.30

Scheme 12.1

Scheme 12.2

Scheme 12.3

Scheme 12.4

Scheme 12.5

Scheme 12.6

Scheme 12.7

Scheme 12.8

Scheme 12.9

Scheme 12.10

Scheme 12.11

Scheme 12.12

Scheme 12.13

Figure 12.1

Scheme 12.14

Scheme 12.15

Scheme 12.16

Scheme 12.17

Scheme 12.18

Scheme 12.19

Scheme 12.20

Scheme 12.21

Scheme 12.22

Scheme 12.23

Scheme 12.24

Scheme 12.25

Scheme 12.26

Scheme 12.27

Scheme 13.1

Scheme 13.2

Scheme 13.3

Scheme 13.4

Scheme 13.5

Scheme 13.6

Scheme 13.7

Scheme 13.8

Scheme 13.9

Scheme 13.10

Scheme 13.11

Scheme 13.12

Scheme 13.13

Scheme 13.14

Scheme 13.15

Scheme 13.16

Scheme 13.17

Scheme 13.18

Scheme 13.19

Scheme 13.20

Scheme 13.21

Schemes 13.22

Figure 13.23

Scheme 13.24

Scheme 13.25

Scheme 13.26

Scheme 13.27

Scheme 13.28

Scheme 13.29

Scheme 13.30

Scheme 13.31

Scheme 13.32

List of Tables

Table 3.1

Table 4.1

Table 4.2

Table 4.3

Table 9.1

Table 9.2

Table 9.3

Related Titles

Georgakilas, V. (ed.)

Transition-Metal-Mediated Aromatic Ring Construction

2013

Print ISBN: 978-1-118-14892-1

(Also available in a variety of digital formats)

Hashmi, A.S., Toste, F.D. (eds.)

Modern Gold Catalyzed Synthesis

2012

Print ISBN: 978-3-527-31952-7

(Also available in a variety of digital formats)

Rios Torres, R. (ed.)

The Pauson-Khand Reaction - Scope, Variations and Applications

2012

Print ISBN: 978-0-470-97076-8

(Also available in a variety of digital formats)

Behr, A., Neubert, P.

Applied Homogeneous Catalysis

2012

Print ISBN: 978-3-527-32641-9

Bullock, R.M. (ed.)

Catalysis without Precious Metals

2010

Print ISBN: 978-3-527-32354-8

(Also available in a variety of digital formats)

Nugent, T.C. (ed.)

Chiral Amine Synthesis

Methods, Developments and Applications

2010

Print ISBN: 978-3-527-32509-2

(Also available in a variety of digital formats)

Edited by

Barry M. Trost and Chao-Jun Li

Modern Alkyne Chemistry

Catalytic and Atom-Economic Transformations

The Editors

Prof. Dr. Barry M. Trost

Stanford University

Department of Chemistry

330 Roth Way

Stanford

CA 94305-5080

USA

Prof. Dr. Chao-Jun Li

McGill University

Department of Chemistry

801 Sherbrook Street West

Montreal

Quebec H3A 0B8

Canada

All books published by Wiley-VCH are carefully produced. Nevertheless, authors, editors, and publisher do not warrant the information contained in these books, including this book, to be free of errors. Readers are advised to keep in mind that statements, data, illustrations, procedural details or other items may inadvertently be inaccurate.

Library of Congress Card No.: applied for

British Library Cataloguing-in-Publication Data

A catalogue record for this book is available from the British Library.

Bibliographic information published by the Deutsche Nationalbibliothek

The Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie; detailed bibliographic data are available on the Internet at <http://dnb.d-nb.de>.

© 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Boschstr. 12, 69469 Weinheim, Germany

All rights reserved (including those of translation into other languages). No part of this book may be reproduced in any form – by photoprinting, microfilm, or any other means – nor transmitted or translated into a machine language without written permission from the publishers. Registered names, trademarks, etc. used in this book, even when not specifically marked as such, are not to be considered unprotected by law.

Print ISBN: 978-3-527-33505-3

ePDF ISBN: 978-3-527-67790-0

ePub ISBN: 978-3-527-67791-7

Mobi ISBN: 978-3-527-67792-4

oBook ISBN: 978-3-527-67789-4

List of Contributors

Kenneth Avocetien

Northeastern University

Department of Chemistry and Chemical Biology

Hurtig Hall

Huntington Ave

Boston

MA 02115

USA

Mark J. Bartlett

University of California Berkeley

Department of Chemistry

Berkeley

CA 94720-1460

USA

Rafael Chinchilla

University of Alicante

Department of Organic Chemistry

Faculty of Sciences and Institute of Organic Synthesis

ctra San Vicente s/n

Alicante

Spain

Alois Fürstner

Max-Planck-Institut für Kohlenforschung

Kaiser-Wilhelm-Platz 1

Mülheim an der Ruhr

Germany

Sergio E. García-Garrido

Universidad de Oviedo

Laboratorio de Compuestos Organometálicos y Catálisis

Red ORFEO-CINQA - Centro de Innovación en Química Avanzada

IUQOEM, Facultad de Química

C/Julián Clavería 8

Oviedo

Spain

Jean-Pierre Genet

PSL Research University

Chimie ParisTech - CNRS

Institut de Recherche de Chimie Paris

rue P. et M. Curie

Paris Cedex 05

France

Naoya Kumagai

Institute of Microbial Chemistry (Bikaken)

Laboratory of Synthetic Organic Chemistry

3-14-23 Kamiosaki

Shinagawa-ku

Tokyo 141-0021

Japan

Chao-Jun Li

McGill University

Department of Chemistry

Sherbrook Street West

Montreal

Quebec H3A 0B8

Canada

Yu Li

Northeastern University

Department of Chemistry and Chemical Biology

Hurtig Hall

Huntington Ave

Boston

MA 02115

USA

Jean-Philip Lumb

McGill University

Department of Chemistry

Sherbrooke Street West, Room 322

Montreal

Quebec H3A 0B8

Canada

Giovanni Maestri

Institut de Chimie des Substances Naturelles

CNRS-UPR 2301

1, avenue de la Terrasse-Bât.27

Gif/Yvette Cedex

France

Max Malacria

Institut de Chimie des Substances Naturelles

CNRS-UPR 2301

1, avenue de la Terrasse-Bât.27

Gif/Yvette Cedex

France

Morten Meldal

University of Copenhagen

Department of Chemistry

Center for Evolutionary Chemical Biology (CECB)

Universitetsparken 5

Copenhagen

Denmark

Véronique Michelet

PSL Research University

Chimie ParisTech - CNRS

Institut de Recherche de Chimie Paris

rue P. et M. Curie

Paris Cedex 05

France

Carmen Nájera

University of Alicante

Department of Organic Chemistry

Faculty of Sciences and Institute of Organic Synthesis

ctra San Vicente s/n

Alicante

Spain

George A. O'Doherty

Northeastern University

Department of Chemistry and Chemical Biology

Hurtig Hall

Huntington Ave

Boston

MA 02115

USA

Raphael Rodriguez

Institut de Chimie des Substances Naturelles

CNRS-UPR 2301

1, avenue de la Terrasse-Bât.27

Gif/Yvette Cedex

France

Sanne Schoffelen

University of Copenhagen

Department of Chemistry

Center for Evolutionary Chemical Biology (CECB)

Universitetsparken 5

Copenhagen

Denmark

Masakatsu Shibasaki

Institute of Microbial Chemistry (Bikaken)

Laboratory of Synthetic Organic Chemistry

3-14-23 Kamiosaki

Shinagawa-ku

Tokyo 141-0021

Japan

Patrick Y. Toullec

PSL Research University

Chimie ParisTech - CNRS

Institut de Recherche de Chimie Paris

rue P. et M. Curie

Paris Cedex 05

France

Barry M. Trost

Stanford University

Department of Chemistry

Roth Way

Stanford

CA 94305-5080

USA

Fiona R. Truscott

Institut de Chimie des Substances Naturelles

CNRS-UPR 2301

1, avenue de la Terrasse-Bât.27

Gif/Yvette Cedex

France

Nick Uhlig

McGill University

Department of Chemistry

Sherbrook Street West

Room 322

Montreal

Quebec H3A 2K6

Canada

Woo-Jin Yoo

McGill University

Department of Chemistry

Sherbrook Street West

Room 322

Montreal

Quebec H3A 2K6

Canada

Liang Zhao

McGill University

Department of Chemistry

Sherbrook Street West

Room 322

Montreal

Quebec H3A 2K6

Canada

Preface

Alkyne is a basic functionality with “relatively low thermodynamic reactivities” in the classical text of organic chemistry. These classical alkyne reactions often require stoichiometric reagents, which result in low efficiency in chemical syntheses, and “harsh” reaction conditions that cannot tolerate the presence of the various “more reactive functional groups”. The pursuit of synthetic efficiency combined with the recent emphasis of “future sustainability and Green Chemistry,” and the pressing desire for new chemical tools in synthetic biology inspire chemists to uncover new reactions that are catalytic in nature (rather than consuming stoichiometric reagents), occur under ambient conditions (including milder temperature and aqueous media), can tolerate various functional groups, and render “dial-up” reactivity when needed. Alkynes provide the most ideal candidate for such features. While being relatively inert under “classical” conditions, alkynes can be readily “activated” selectively, in the presence of other functional groups and under mild conditions, via transition-metal catalysis through either selective alkyne carbon-carbon triple bond reactions or terminal alkyne C-H bond reactions. Such a unique reactivity allow alkynes to be embedded and be “dialed-up” whenever needed. For the past few decades, modern alkyne chemistry has thus been developed rapidly to feature these characteristics. These developments further focus on atom-economic transformations where minimal or no theoretical by-products are formed. Furthermore, many of these catalytic transformations are orthogonal to biological conditions. These modern catalytic alkyne reactions are much more resource-, time-, and manpower-efficient, and provide an alternative to classical stoichiometric alkyne chemistry. This book comprises a collection of contributions from leading experts and covers various modern catalytic reactions of alkynes. We hope that this focused book will be very helpful not only to students and researchers in chemistry but also to those in material and biological studies and will provide them with tools and opportunities unavailable with classical alkyne chemistry.

Stanford

Montreal

August 2014

Barry M. Trost

Chao-Jun Li

1Introduction

Chao-Jun Li and Barry M. Trost

1.1 History of Alkynes

Alkyne is one of the fundamental functional groups that established the foundation of organic chemistry [1]. The smallest member of this family, acetylene, was first discovered in 1836 by Edmund Davy [2]. It was rediscovered and named “acetylene” by Marcellin Berthelot in 1860 by passing vapors of organic compounds through a red-hot tube or sparking electricity through a mixture of cyanogen and hydrogen gas. Acetylene is a moderately common chemical in the universe [3], often in the atmosphere of gas giants. In 1862, Friedrich Wöhler discovered the generation of acetylene from the hydrolysis of calcium carbide (Equation 1.1). Acetylene produced by this reaction was the main source of organic chemicals in the coal-based chemical industry era. When petroleum replaced coal as the chief source of carbon in the 1950s, partial combustion of methane (Equation 1.2) or formation as a side product of hydrocarbon cracking became the prevalent industrial manufacturing processes for acetylene. The next member of the family, propyne, is also mainly prepared by the thermal cracking of hydrocarbons. The first naturally occurring acetylic compound, dehydromatricaria ester (1), was isolated in 1826 [4] from an Artemisia species. Well over 1000 alkyne-containing natural products have been isolated since then, among which many are polyyne-containing natural products isolated from plants, fungi, bacteria, marine sponges, and corals [5].

1.1
1.2

The higher members of alkynes are generally derived from the smaller homologs via alkyne homologation processes of the terminal alkynes (see Equation 1.8, below), while some alkynes are generated through elimination reactions with organic halides under basic conditions (Equation 1.3) [1]. A search in Sci Finder shows that >70 000 terminal alkynes and >10 000 internal alkynes are now commercially available from various sources.

1.3

1.2 Structure and Properties of Alkynes

Alkynes contain a tripe bond, composed of a σ-covalent bond formed from two sp-hybridized carbons and two π-bonds resulted from the overlapping of two orthogonal unhybridized p-orbitals on each carbon (2) [1]. Consequently, alkynes are generally rod-like. Cyclic alkynes are less common with benzyne as an important reactive intermediate in organic chemistry [6]. Acetylene is linear and intrinsically unstable under pressure due to its high compressibility as well as its propensity to undergo exothermic self addition reactions. Consequently, acetylene itself can explode violently at high pressure and the safe limit for acetylene is 103 kPa. Thus, acetylene is generally shipped in acetone or dimethyl formamide (DMF) solutions or contained in a gas cylinder with porous filling [7]. Acetylene has been used as a burning fuel and for illumination purposes in the late nineteenth century and early twentieth century [8]. In modern times, alkynes have found a wide range of applications ranging from organic electronic materials, metal-organic frame works (MOF), pharmaceutical agents, and others [9]. The linearity of the alkyne creates strain when an alkyne is part of a ring [10]. In spite of this fact, cyclopentyne, cyclohexyne, and cycloheptyne can be generated at least fleetingly, their existence being confirmed by in situ trapping, notably by 1,3-dipolar cycloadditions [11]. Cyclooctyne is still highly strained but has sufficient stability to be isolated and used in click chemistry to study biological processes [12].

1.3 Classical Reactions of Alkynes

The higher degree of unsaturation of alkynes compared to alkenes increases their reactivity toward addition to both alkenes and alkynes. In particular, virtually all additions of HX and RX to alkynes are exothermic. Consequently, these stoichiometric addition reactions have been the basis of most reactions in the classical alkyne chemistry (Equation 1.4) [1]. These classical alkyne addition reactions include the additions of hydrogen, halogens, water, hydrogen halides, halohydrins, hydroborations, and others. With a stoichiometric amount of a strong oxidizing reagent such as KMnO4, the addition may be followed by C–C cleavage to give the corresponding acids (Equation 1.5). Less reactive reagents can also be added through the use of a transition-metal catalyst. The unique electronic character of alkynes wherein their HOMO–LUMO gap is rather small makes them especially effective as coordinators to transition metals. Thus, they function as chemoselective functional groups for catalytic transformations. For example, catalytic addition of dihydrogen to alkynes can proceed to either alkenes or alkanes depending on the choice of the catalysts (Equation 1.6) [13]. Further, the hydroalumination [14], hydrosilylation [15], hydrostannylation [16], as well as carboalumination [17] represent important modern advances of the alkyne addition reactions.

1.4
1.5
1.6

A second class of reactions pertains to terminal alkynes. Due to the increased s-character, the alkynyl C–H bonds (pKa = 25) are much more acidic than the corresponding alkenyl C–H bonds (pKa = 43) and alkyl C–H bonds (pKa > 50) [18]. Thus, base-promoted additions of terminal alkynes to carbonyl compounds can occur under different basic conditions, a process discovered over a century ago (Equation 1.7). Treatment of terminal alkynes with bases such as lithium amide, butyllithium, or Grignard or zinc reagents generates metal acetylides stoichiometrically, which can then react with different carbon-based electrophiles to produce various higher alkyne homologs in the classical synthetic chemistry (Equation 1.8) [1]. Such processes can be catalyzed to permit deprotonation with much weaker bases as in the coupling with aryl halides under Pd/Cu catalysis (Sonogashira reaction, see Equation 1.13).

1.7
1.8

1.4 Modern Reactions

Although the classical stoichiometric addition reactions, alkyne cleavage reactions, and homologation reactions have established the foundation of alkyne chemistry, a rebirth of interest derives from recent concerns regarding societal and ecological sustainability under the mantra of Green Chemistry [19], which emphasizes chemical transformations that are more atom economic [20] and chemoselective, thereby minimizing the use of protecting groups [21]. Furthermore, rapid developments in the field of chemical biology demand chemical transformations that are orthogonal to biological conditions and functionalities in bioorganisms and which can work efficiently under both in vitro and in vivo biological conditions [22]. Alkynes, being both good π-donors and π-acceptors for transition metals as well as being energy rich, can be effectively activated by a catalyst thereby lowering the energy barrier to proceed to the more stable products while being unreactive toward various biological elements. At the same time, they can be chemoselectively activated in the presence of most typical functional groups (e.g., hydroxyl and carbonyl groups as well as alkenes) and in protic solvents including water [23]. Such triggered reactivities are orthogonal to the classical reactivities and can be tuned to target specific desired reaction sites while maintaining tolerance toward other functionalities through the discrete choice of catalyst, which will greatly simplify the syntheses of complex compounds and allow direct modifications of biomolecules in their native states and ambient environment. Modern developments, in view of atom economy, can be represented by three major classes: (i) catalytic cyclization reactions, (ii) catalytic homologations of terminal alkynes, and (iii) catalytic isomerization reactions of alkyne.

Although alkyne oligomerization was known at a high temperature since the late nineteenth century [2], various cyclization reactions of alkynes catalyzed by transition metals are among the most important developments in modern alkyne chemistry. The most well-known examples include the transition-metal-catalyzed [2 + 2 + 2] cycloaddition reactions (Equation 1.9) [24], the Pauson–Khand-type reaction of alkyne–alkene–carbon monoxide (Equation 1.10) [25], the enyne cyclization reactions (Equation 1.11) [26], and the 1,3-dipolar cycloaddition such as that with azides (the archetypical Click reaction) (Equation 1.12) [27].

1.9
1.10
1.11
1.12

The second major class of modern alkyne reactions is the catalytic transformation of terminal alkyne C–H bonds. Although homologation of terminal alkynes through the reactions of metal acetylides with organic halide is a classical alkyne reaction, such a reaction cannot be applied to aryl and vinyl halides due to their inert nature in nucleophilic substitution reactions. The development of catalytic coupling of terminal alkynes with aryl and vinyl halides (the Sonogashira reaction) has overcome this classical challenge and opened up a new reactivity mode in alkyne homologation (Equation 1.13) [28]. Complimentary to the classical Favorskii reaction (Equation 1.7), the modern development of catalytic direct addition of terminal alkynes to aldehydes provides great opportunities in generating optically active propargyl alcohols (Equation 1.14) [29]. The catalytic direct additions of terminal alkynes to imines (and derivatives) (Equation 1.15) [30] and conjugate addition to unsaturated carbonyl compounds (Equation 1.16) [31] represent other major achievements in modern alkyne reactions. On the other hand, the catalytic oxidative dimerization (Glaser–Hay coupling) [32] and simple alkyne dimerization (Equation 1.17) [33] which date from the late 1800s have become increasingly important in modern synthetic chemistry.

1.13
1.14
1.15
1.16
1.17

Two additional processes that have much unrealized potential in synthetic chemistry are the alkyne disproportionation (metathesis) and the alkyne redox isomerization reactions. Like the alkene metathesis, the catalytic alkyne–alkyne metathesis reaction retains all functionalities by switching the groups attached to the alkynes (Equation 1.18) [34]. Another unique atom-economic reaction of alkynes that is currently under-utilized but will have a great potential for future development is the “alkyne-zipper reaction” (Equation 1.19) [35]. Such reactions shift readily accessible internal alkyne triple bond to terminal positions for further homologations. A different type of “retaining functionality is found in the redox isomerization of propargyl alcohols to generate conjugated ketones” (Equation 1.20) [36].

1.18
1.19
1.20

1.5 Conclusion

With the recent emphasis on sustainability and the ever increasing needs in synthetic efficiency, alkynes provide a truly unique functionality that is orthogonal to other functional groups, biological conditions, and ambient environment, yet can be selectively triggered to occur in a specific reaction mode with the absence of protecting groups or anhydrous conditions. Such reactions will have great potential to simplify synthetic chemistry and will find wide applications in chemical biology and organic materials. This book, comprising experts on related subjects, provides an overview of developments of modern alkyne reactions. Due to the limit of space, many other important developments in modern alkyne chemistry such as various catalytic conversions of alkyne triple bonds [37] and alkyne polymerizations [38] have not been covered in this book.

References

1. (a) McMurry, J. (2007)

Organic Chemistry-A Biological Approach

, Thomson Brooks/Cole, Belmont, CA;(b) Bruce, P.Y. (2004)

Organic Chemistry

, 4th edn, Pearson Education, Upper Saddle River, NJ;(c) Wade, L.G. (2006)

Organic Chemistry

, 5th edn, Pearson Education, Upper Saddle River, NJ;(d) Solomons, T.W.G. and Fryhle, C.B. (2011)

Organic Chemistry

, 10th edn, John Wiley & Sons, Inc., New York;(e) Vollhardt, K.P.C. and Schore, N.E. (2009)

Organic Chemistry

, 6th edn, W.H. Freeman & Company;(f) Fox, M.A. and Whitesell, J.K. (2004)

Organic Chemistry

, 3rd edn, Jones & Bartlett Publishers.

2. For historical information of alkyne chemistry, please see: (a) Nieuwland, J.A. and Vogt, R.R. (1945)

Chemistry of Acetylenes

, Reinhold Publishing, New York;(b) Viehe, H.G. (ed.) (1969)

Chemistry of Acetylenes

, Marcel Dekker, New York.

3. Kwok, S. (2001)

Organic Matter in the Universe

, John Wiley & Sons, Inc., New York.

4. Bohlmann, F., Burkhardt, H., and Zdero, C. (1973)

Naturally Occurring Acetylenes

, Academic Press, New York.

5. Shi Shun, A.L.K. and Tykwinski, R.R. (2006)

Angew. Chem. Int. Ed.

,

45

, 1034.

6. Heaney, H. (1962)

Chem. Rev.

,

62

, 81.

7. Compressed Gases Association (1999)

Handbook of Compressed Gases

, Springer, London.

8. Holloway, A. (1898)

Acetylene as a Means of Public and Private Illumination

, Cradley Heath.

9. Diederich, F., Stang, P.J., and Tykwinski, R.R. (eds) (2005)

Acetylene Chemistry

, Wiley-VCH Verlag GmbH, Weinheim.

10. Bach, R.D. (2009)

J. Am. Chem. Soc.

,

131

, 5233.

11. Wittig, G. (1962)

Rev. Chim. Populaire Roum.

,

7

, 1393.

12. Sletten, E.M. and Bertozzi, C.R. (2011)

Acc. Chem. Res.

,

44

, 666.

13. Smith, G.V. and Notheisz, F. (1989)

Heterogeneous Catalysis in Organic Chemistry

, Academic Press.

14. Yamamoto, H. and Oshima, K. (2006)

Main Group Metals in Organic Synthesis

, John Wiley & Sons, Inc.

15. Marciniec, B. (2008)

Hydrosilation: A Comprehensive Review on Recent Advances

, Springer.

16. Davis, A.G. (2006)

Organotin Chemistry

, John Wiley & Sons, Inc.

17. Negishi, E. (1981)

Pure Appl. Chem.

,

53

, 2333.

18. Smith, J.G. (2008)

Organic Chemistry

, 2nd edn, McGraw-Hill, Boston, MA.

19. Anastas, P.T. and Warner, J.C. (1998)

Green Chemistry: Theory and Practice

, Oxford University Press, Oxford.

20. Trost, B.M. (1995)

Angew. Chem., Int. Ed. Engl.

,

34

, 259.

21. Li, C.-J. and Trost, B.M. (2008)

Proc. Natl. Acad. Sci. U.S.A.

,

105

, 13197.

22. (a) Zorn, J.A. and Wells, J.A. (2010)

Nat. Chem. Biol.

,

6

, 179;(b) Maeyer, G. (2009)

Angew. Chem. Int. Ed.

,

48

, 2672.

23. Uhlig, N. and Li, C.-J. (2011)

Chem. Sci.

,

2

, 1241.

24. Funk, R.L. and Vollhardt, K.P.C. (1980)

J. Am. Chem. Soc.

,

102

, 5253.

25. Blanco-Urgoiti, J., Añorbe, L., Pérez-Serrano, L., Domínguez, G., and Pérez-Castells, J. (2004)

Chem. Soc. Rev.

,

33

, 32.

26. Lu, X., Zhu, G., Wang, Z., Ma, S., Ji, J., and Zhang, Z. (1997)

Pure Appl. Chem.

,

69

, 553.

27. Kolb, H.C., Finn, M.G., and Sharpless, K.B. (2001)

Angew. Chem. Int. Ed.

,

40

, 2004.

28. Chinchilla, R. and Najera, C. (2011)

Chem. Soc. Rev.

,

40

, 5084.

29. For reviews, see: (a) Trost, B.M. and Weiss, A.H. (2009)

Adv. Synth. Catal.

,

351

, 963;(b) Pu, L. (2003)

Tetrahedron

,

59

, 9873;(c) Cozzi, P.G., Hilgraf, R., and Zimmermann, N. (2004)

Eur. J. Org. Chem.

, 4095;(d) Lu, G., Li, Y.-M., Li, X.-S., and Chan, A.S.C. (2005)

Coord. Chem. Rev.

,

249

, 1736.

30. For reviews, see: (a) Yoo, W.-J., Zhao, L., and Li, C.-J. (2011)

Aldrichim. Acta

,

44

, 43;(b) Wei, C., Li, Z., and Li, C.-J. (2004)

Synlett

, 1472;(c) Zani, L. and Bolm, C. (2006)

Chem. Commun.

, 4263;(d) Peshkov, V.A., Pereshivko, O.P., and Van der Eycken, E.V. (2012)

Chem. Soc. Rev.

,

41

, 3790.

31. Yazaki, R., Kumagai, N., and Shibasaki, M. (2010)

J. Am. Chem. Soc.

,

132

, 10275.

32. Cadiot, P. and Chodkiewicz, W. (1969) in

Chemistry of Acetylenes

(ed. Viehe, G.H.), Marcel Dekker, New York, p. 597.

33. Trost, B.M., Sorum, M.T., Chan, C., Harms, A.E., and Rühter, G. (1997)

J. Am. Chem. Soc.

,

119

, 698.

34. Fürstner, A. and Davies, P.W. (2005)

Chem. Commun.

, 2307.

35. Brown, C.A. and Yamashita, A. (1975)

J. Am. Chem. Soc.

,

97

, 891.

36. Trost, B.M. and Livingston, R.C. (1995)

J. Am. Chem. Soc.

,

117

, 9586.

37. Gorin, D.J. and Toste, F.D. (2007)

Nature

,

446

, 395.

38. Buchmeiser, M.R. (2005)

Adv. Polym. Sci.

,

176

, 89.

Part ICatalytic Isomerization of Alkynes

2Redox Isomerization of Propargyl Alcohols to Enones

Barry M. Trost

2.1 Introduction

The synthesis of enones has classically relied upon aldol condensation (Figure 2.1) [1]. Its strength lies in the ready availability of the substrates and its high atom economy. It suffers from issues of chemo- and regioselectivity. Self-condensation of the aldehyde and the ability to form two regioisomeric enone products has led to numerous variations to minimize such issues. One solution employs olefination protocols which suffer from poor atom economy. A particularly interesting strategy recognizes that propargyl alcohols are isomeric with enones as shown in Scheme 2.1. The availability of propargyl alcohols by a simple addition of a terminal acetylene to an aldehyde then can make enones readily available by an atom economic sequence of simple addition followed by isomerization. In 1922, Meyer and Schuster [2] described the rearrangement of the oxidation pattern of propargyl alcohols wherein the hydroxyl group undergoes the equivalent of a 1,3-shift to form the rearranged enone after tautomerization (Scheme 2.1, path a). The Meyer–Schuster rearrangement has been well reviewed and will not be a subject of this chapter.

Figure 2.1 Enone synthesis via aldol condensation.

Scheme 2.1 Redox neutral processes for enone synthesis.

An alternative which maintains the positional integrity of the hydroxyl group involves the shift of two hydrogens (Scheme 2.1, path b) which may be referred to as a redox isomerization. Mechanistically, such a rearrangement of hydrogens is not straightforward. Classically, this transformation typically was performed by a stoichiometric reduction of the triple bond followed by stoichiometric oxidation of the alcohol (or vice versa). While early strides revealed base catalysis could be effective for a certain very limited type of structure, the importance of making synthesis more environmentally benign stimulated efforts to broaden the generality of the process, especially to include nonactivated types of propargyl alcohols. In this chapter, an overview of redox isomerization is presented organized along the line of catalysts.

2.2 Base Catalysis

In 1949, Ninaham and Raphael [3] reported the isomerization of a γ-hydroxybutynoate to an E-γ-ketobutenoate in the presence of triethylamine. In 1954, Vaitiekunas and Nord [4] extended this isomerization of a γ-hydroxybutynoate (1) to an E-γ-ketobutenoate (3) under similar conditions (Equation 2.1) in 85% yield when

2.1

Ar = 2-thienyl. The facility of the process derives from the stability of the supposed enolate intermediate 2. Studies in 2007 revealed that the Z-enoate 4 (Ar = Ph or 2-furyl) can result from such a process, depending upon choice of base. Using bicarbonate in dimethylsulfoxide (DMSO) gives the Z isomers, presumably resulting from a kinetic protonation which occurs from the least hindered face to deliver the Z-alkene 4 (Ar = Ph or 2-furyl) [5]. On the other hand, 1,4-diazabicyclo[2.2.2.]octane (DABCO) in DMSO gives the E-isomer 4 (Ar = Ph or 2-furyl) [6]. It is possible that a tertiary amine base isomerizes the initial isomer to the thermodynamically more stable E isomer. The electron-withdrawing group can be an electron-deficient heterocycle as in 5 (Equation 2.2) [7]. The electron-withdrawing group that enhances the activity of the propargylic proton need not be directly attached to the alkyne. Thus, the alkyne 6 undergoes redox isomerization with triethylamine (Equation 2.3) [8]. Since such substrates can be produced by the Sonogashira coupling in the presence of

2.2
2.3

triethylamine, the intermediate propargyl alcohol may undergo redox isomerization during the coupling step although some role for the Pd in the redox isomerization cannot be ruled out [8].

An alternative strategy places the anion stabilizing group at the propargylic center. Thus, a substrate bearing a 2-pyridyl substituent at this position as in 7 requires both acid and base catalysis to effect isomerization as shown in Equation (2.4) [9].

2.4

Using an inductively strong electron-withdrawing group such as polyfluoro alkyl as in 8 also allows base catalyzed isomerization. While a simple tertiary amine suffices (Equation 2.5) [10], the isomerization also proceeds under

2.5

Mitsunobu conditions via a mechanism that is obscure at best (see below) [11]. Diyne carbinols 9 wherein a second alkyne serves as an adequate anion stabilizing group (Equation 2.6) also isomerizes with a somewhat stronger base such as 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) at an elevated temperature (Equation (2.6)) [12]. Surprisingly, 1,3-diarylpropargyl alcohol suffices. In one such case,

2.6
2.7

the juxtaposition of a γ-hydroxyl group as in 10 led to trapping the intermediate enone 11 to form a benzodihydrofuran 12 under the reaction conditions (Equation 2.7) [13]. The most useful variant is employing an o-aminophenyl group as the alkyne-bound aryl group [14]. Again, since the Sonogashira coupling proceeds under basic conditions, both redox isomerization and cyclization to a quinolone 13 may occur to provide a reasonably efficient atom economic reaction wherein only 1 equiv of hydroxide base and 2 equiv of water

2.8

result along with an iodide salt (Equation 2.8) [15]. Most recently, even only one aryl ring at the propargylic carbinol was sufficient to promote base-catalyzed redox isomerization. The ability of the enone products to serve as Michael acceptors allows an atom and step economic approach to β-substituted ketones such as pyrazole adduct 14 (Equation 2.9) [16].

2.9

2.3 Ru Catalyzed

The first example of a transition-metal-catalyzed isomerization of primary propargyl alcohols to enals was reported by Ma and Lu [[17] in 1989 using a Ru complex 15 which required long reaction times at high temperatures. The reaction was envisioned to proceed by hydrometallation–dehydrometallation to form an allenol

2.10

which tautomerized to the enone (Equation 2.10). An improvement was reported in 1995 using η5-indenyl-bis-(triphenylphosphine) ruthenium chloride (16) as catalyst and indium trichloride as a cocatalyst [18]. Subsequently,

2.11

a significant further improvement used indium triflate as catalyst wherein 1–3 mol% of Ru and In complexes sufficed and reactions occurred in 0.5–3 h [18]. Both primary (Equation 2.11) and secondary (Equation 2.12) alcohols participate. The excellent chemoselectivity is illustrated by the examples of propargyl alcohols 17–19 (Equations (2.12) [18], (2.13) [18], (2.14) [19]).

2.12
2.13
2.14

The dienal 20 served as a key intermediate in the synthesis of a leukotriene. Particularly noteworthy is the chemoselective bis redox isomerization of 19 to bis dienal 21 on the way to the polyacetylenic natural product adociacetylene. The chemoselectivity demonstrated in these examples is inconsistent with a hydrometallation–dehydrometallation mechanism. The ready availability of butynediol as a building block is further enhanced by its chemoselective redox isomerization. Indeed, the accessibility of the functional crotonaldehyde 22 via

2.15

Ru-catalyzed isomerization (Equation 2.15) allowed it to provide easy access to the sphingofungins. In this case, a more coordinatively unsaturated indenyl ruthenium complex was employed as depicted in Equation (2.14). Such cycloocta-1,5-diene (COD) ligands bound to Ru have been shown to be reacted off by a [2 + 2 + 2] cycloaddition with an alkyne to free two open coordination sites [18].

Deuterium labeling studies revealed that a 1,2-hydride shift occurred generating a Ru carbenoid intermediate 23 (Scheme 2.2) [18]. Further evidence for this mechanism was the interception of the Ru carbenoid by intramolecular cyclopropanation of a tethered olefin (e.g., 24) as shown in Equation (2.16) [20]. This reaction was a key step in

2.16

a concise synthesis of echinopine A, 25 (Equation 2.17) [21]. A C–C bond rather than H can migrate when a cyclopropyl ring is annealed to the propargyl carbon. Depending upon the nature of the substituent on the alkyne terminus,

2.17

either a 1,2 shift occurs to give the alkylidenecyclobutanone 26 (Equation 2.18) or a 1,3 shift occurs to form

2.18
2.19

cyclopentenone 27 (Equation 2.19) [22]. This selectivity may result from the preference to coordinate the alkyne with the hydroxyl group in the former but with the cyclopropyl C–C bond in the latter.

Scheme 2.2 Hydride shift mechanism for the Ru-catalyzed redox isomerization.

The in situ formation of a Michael acceptor via redox isomerization sets the stage for a cascade. Thus, juxtaposition of a suitable oxygen as in 28 (Equation 2.20) [23], nitrogen as in 29 (Equation 2.21) [24], or even carbon as in 30 (Equation 2.22) [25] nucleophiles provide easy access to the corresponding heterocycles 31 and 32 or carbocycle 34. In the first two cases, cyclization to the heterocycles occurred in tandem with the redox isomerization. On the other hand, under

2.20
2.21
2.22

the acidic conditions of the redox isomerization, cyclization of the carbon pronucleophile 33 did not occur. Thus, upon completion of the redox isomerization, a chiral organocatalyst was added to effect the asymmetric cyclization of 33 to give 34 in high ee. Switching from the exo type of cyclization as in Equation (2.21) to an endo geometry slowed the rate of cyclization too. Thus, in the case of the nitrogen Michael donor, redox isomerization of the sulfonamide 35 (Equation 2.23) was not accompanied by cyclization via Michael addition [24]. For completion of the

2.23

cascade, simple addition of potassium carbonate in methanol to the initial reaction mixture allowed Michael addition to form the piperidinone 36.

The vinyl ketones are also electrophilic enough in the presence of the redox catalyst system, notably the presence of the In Lewis acid, that electrophilic aromatic substitution can occur as shown in Equation (2.24) [26]. A similar

2.24
2.25

reaction was effected using a cyclopentadienone Ru complex 37 although higher temperatures were required (Equation 2.25) [27]. The chemoselectivity of the Michael addition to the divinyl ketone intermediate is also noteworthy. A

dinuclear Ru complex 38 has also been reported to be effective under milder conditions (10% NH4PF6, 1,2-dichloroethane (DCE), 60°) although the yields were considerably lower [28].

While a hydride shift mechanism appears to account for the above examples, a hydrometallation–dehydrometallation mechanism is more likely in the case of (Ph3P)3Ru(H2)CO (39) as the catalyst [29]. Using complex 39, a 1,4-dihydroxy-2-alkyne 40 can be initially isomerized to a γ-hydroxyenone which can undergo a second redox isomerization of the remaining allyl alcohol to a 1,4-diketone (Equation 2.26). At the high temperature of the reaction in the presence of acid, cyclodehydration occurs to generate the aromatic 2,5-disubstituted furan. If the redox isomerization is performed under neutral conditions, the reaction stops at the 1,4-diketone. Such

2.26

1,4-diketones are precursors to pyrroles by simple addition of a primary amine (Equation 2.27) [29]. Alternatively, base treatment can provide cyclopentenone.

2.27

2.4 Rh Catalysis

In 1995, Sarah and Pellicciari [30], motivated by the synthesis of peptide isosteres (Equation 2.41), examined the use of Wilkinson's complex for the redox isomerization of γ-hydroxy-α,β-acetylenic esters such as 39 (Equation 2.28) using a substrate derived from phenyl alanine. While the mechanism of this process has not been established, a

2.28

reasonable possibility is depicted in Equation (2.28). It invokes a coordinatively unsaturated Rh complex bound to oxygen undergoing a β-hydrogen elimination forming a rhodium hydride. Hydrometallation of the alkyne followed by protonation would complete the catalytic cycle.

Evidence for such a mechanism derives from the development of a more active catalyst as shown in Equation (2.29) [31]. This catalyst works well for aryl- and vinyl- substituted propargyl alcohols as illustrated with the

2.29

cyclohexenyl substituent as in 41 in Equation (2.29). This catalyst also proved effective to convert 2-butyne-1,4-diols 42 to 1,4-diketones 43 (Equation 2.30) even with saturated alkyl groups. Interestingly, the monomethyl ether delivered the

2.302.31

furan 44, presumably because the intermediate γ-keto-α,β-unsaturated enone preferably eliminated methanol to aromatize (Equation 2.31). An interesting application of this process is its use to effect a kinetic resolution of propargylic alcohols (Equation 2.33) [32]. While the desired product is the unchanged alcohol enantioenriched, it does show a different dimension of the process.

2.32

2.5 Palladium Catalysis

A major mechanism for Pd-catalyzed processes involves hydropalladation and dehydropalladation such as in the migration of double bonds. In 1988, this mechanism was used to effect a redox isomerization as shown

2.33

in Equation (2.33) [33]. In this case, formate serves as the hydride source to generate the [Pd–H] species. Lu and coworkers [34] reported the use of a similar mechanism for the redox isomerization of 2-butyne-1,4-diol-type compounds, 45, wherein the [Pd–H] species presumably derives by protonation of a Pd(0) by the diol (Equation 2.34). This example

2.34

highlights the chemoselectivity of the process given the highly reactive nature of a bis-enone as a product. It should be noted that these conditions failed with a mono-propargyl alcohol. Using a stronger acid for such a bis-isomerization served not only as a source for the Pd–H by protonation of Pd(0) but also effected tandem cyclodehydration to form furans (Equation 2.35) [35]. By using an external diol to generate the [PdH] species, the scope of

2.35

the reaction now allowed it to proceed with a simpler propargyl alcohol (Equation 2.36) [36]. Unfortunately, equilibration occurred under the reaction conditions to give a mixture of α,β- and β,γ-unsaturated enones. The advantage of a

2.36

bidentate phosphine, dppe, was noted in the synthesis of prostaglandin analogs [37]. The redox isomerization provided access to the flavor and perfume ingredient damascene 46 (Equation 2.37) [38]. A particularly effective way to generate an active form of the [Pd–H] species is by exposing a typical heterogeneous palladium hydrogenation

2.37

catalyst to hydrogen gas. For example, employment of Pearlman's catalyst pretreated with a small amount of hydrogen effected redox isomerization at room temperature (Equation 2.38) [39] in work directed toward the synthesis of the amphidinolides.

2.38

2.6 Miscellaneous

Lu and coworkers [40] established the effectiveness of an Ir complex to effect redox isomerization via a hydrometallation–dehydrometallation mechanism. This catalyst was equally effective for redox isomerization of a simple propargyl alcohol (Equation 2.39) as well as a 2-butyne-1,4-diol system (Equation 2.40). Unfortunately,

2.39
2.40

isomerization of the conjugated double bond to the β,γ position accompanied the redox isomerization.

While Pt complexes have not been described to perform a redox isomerization of the type discussed herein, there is a silyl version (Equation (2.41)) [41]. The reaction appears to be initiated by a facile 1,2-silyl shift promoted

2.41

by coordination of the alkyne by the coordinatively unsaturated platinum. The utility of the resultant vinylsilane to access geometrically defined trisubstituted olefins makes this version of the redox isomerization a useful variation.

The most unusual set of conditions for the redox isomerization of the propargyl alcohol 47 is the use of

2.42

the conditions of the Mitsunobu reaction (Equation 2.42) [11]. Mechanistically how this process proceeds is not defined nor even apparent.

2.7 Conclusions

While the field of redox isomerization is just emerging, its potential in improving both atom and step economy is already apparent. At this stage, Ru complexes have progressed the most as appropriate catalysts. Indeed, reasonably mild conditions give promise for good chemoselectivity. At the same time, prospecting for other catalysts has barely begun. Equally exciting is the merging of redox isomerizations with other addition reactions leading to tandem or cascade events. The combined prospects for future development are immense. At this time, redox isomerization is a great complement to the Meyer–Schuster rearrangement and thus we can tune the regioselectivity of the resultant oxidation pattern from the same precursor by just a simple change in catalyst.

References

1. (a) Minbiole, K.P.C. (2009)

Sci. Synth.

,

46

, 147;(b) Nilesen, A.T. and Houlihan, W.J. (1968)

Org. React.

,

16

, 438.

2. (a) Meyer, K.H. and Schuster, K. (1922)

Chem. Ber.

,

55

, 819;For reviews see (b) Engel, D.A. and Dudley, G.B. (2009)

Org. Biomol. Chem.

,

7

, 4149;(c) Caierno, V., Crochet, P., Garcia-Garrido, S.E., and Gimeno, J. (2010)

Dalton Trans.

,

39

, 4015;(d) Bauer, E.B. (2012)

Synthesis

,

44

, 1131.

3. Nineham, A.W. and Raphael, R.A. (1949)

J. Chem. Soc.

, 118.

4. Vaitekunas, A. and Nord, F.F. (1954)

J. Am. Chem. Soc.

,

76

, 2737.

5. Sonye, J.P. and Koide, K. (2007)

J. Org. Chem.

,

72

, 1846.

6. Sonye, J.P. and Koide, K. (2006)

J. Org. Chem.

,

71

, 6254.

7. Guo, C. and Lu, X. (1993)

J. Chem. Soc., Perkin Trans. 1

,

16

, 1921.

8. Müller, T.J.J., Ansorge, M., and Aktah, D. (2000)

Angew. Chem. Int. Ed.

,

39

, 1253.

9. Erenler, R. and Biellmann, J.-F. (2005)

Tetrahedron Lett.

,

46

, 5683.

10. Yamasaka, T., Kawasaki-Takasuka, T., Furata, A., and Sakamoto, S. (2009)

Tetrahedron

,

65

, 5945.

11. Watanabe, Y. and Yamazaki, T. (2011)

J. Org. Chem.

,

76

, 1957.

12. Wang, Y.-H., Liu, H., Zhu, L.-L., Li, X.-X., and Chen, Z. (2011)

Adv. Synth. Catal.

,

353

, 707.

13. Cho, C.S. and Shim, S.C. (2006)

Bull. Korean Chem. Soc.

,

27

, 776.

14. Cho, C.S., Lee, N.Y., Kim, T.-J., and Shim, S.C. (2004)

J. Heterocycl. Chem.

,

41

, 409.

15. Cho, C.S. (2005)

J. Organomet. Chem.

,

690

, 4094.

16. Bhamichandra, M., Rao, M., and Sahoo, A.K. (2012)

Org. Biomol. Chem.

,

10

, 3538.

17. Ma, D. and Lu, X. (1989)

Chem. Commun.

, 890.

18. (a) Trost, B.M. and Livingston, R.C. (1995)

J. Am. Chem. Soc.

,

117

, 9586;(b) Trost, B.M. and Livingston, R.C. (2008)

J. Am. Chem. Soc.

,

130

, 11970;(c) Trost, B.M., Imi, K., and Indolese, A.F. (1993)

J. Am. Chem. Soc.

,

115

, 8831.

19. Trost, B.M. and Weiss, A.H. (2006)

Org. Lett.

,

8

, 4461.

20. Trost, B.M., Breder, A., O'Keefe, B.M., Rao, M., and Franz, A.W. (2011)

J. Am. Chem. Soc.

,

133

, 4766.

21. Peixoto, P.A., Richard, J.A., Severin, R., and Chen, D.Y.-K. (2011)

Org. Lett.

,

13

, 5724.

22. Trost, B.M., Xie, J., and Maulide, N. (2008)

J. Am. Chem. Soc.

,

130

, 17258.

23. Trost, B.M., Gutierrez, A.C., and Livingston, R.C. (2009)

Org. Lett.

,

11

, 2539.

24. Trost, B.M., Maulide, N., and Livingston, R.C. (2008)

J. Am. Chem. Soc.

,

130

, 16502.

25. Trost, B.M., Breder, A., and Kai, B. (2012)

Org. Lett.

,

14

, 1708.

26. Trost, B.M. and Breder, A. (2011)

Org. Lett.

,

13

, 398.

27. Thies, N., Cristian, C.G., and Haak, E. (2012)

Chem. Eur. J.

,

18

, 6302.

28. Miyake, Y., Endo, S., Nomagachi, Y., Yuki, M., and Nishibayashi, Y. (2008)

Organometallics

,

27

, 4017.

29. Pridmore, S.J., Stratford, P.A., Taylor, J.E., Whittlesey, M.K., and Williams, J.M.J. (1995)

Tetrahedron Lett.

,

36

, 4497.

30. Saiah, M.K.E. and Pellicciari, R. (1995)

Tetrahedron Lett.

,

36

, 4497.

31. Tanaka, K., Shoji, T., and Hirano, M. (2007)

Eur. J. Org. Chem.

,

2007

, 2687.

32. Tanaka, K. and Shoji, T. (2005)

Org. Lett.

,

7

, 3561.

33. Arcadi, A., Cacchi, S., Marinelli, F., and Misiti, D. (1988)

Tetrahedron Lett.

,

29

, 1457.

34. Lu, X., Ji, J., Ma, D., and Shen, W. (1991)

J. Org. Chem.

,

56

, 5774.

35. Ji, J. and Lu, X. (1993)

Chem. Commun.

, 764.

36. Lu, X., Ji, J., Guo, C., and Shen, W. (1992)

J. Organomet. Chem.

,

428

, 259.

37. Nakazawa, M., Naora, K., and Yatagai, M. (1992) Preparation of α,β-enones as intermediates for prostaglandins. Jpn. Kokai Tokkyo Koho JP 041 03557, Apr. 06, 1992.

38. Watanabe, K. (2001) Method for preparation of damascone or damascenone by simultaneous catalytic oxidation and reduction of damascol and damascenol. Jpn Kokai Tokkyo Koho JP 2001247504 11, Sept. 2001.

39. Sabitha, G., Reddy, A.Y., Nayak, S., and Yadav, J.S. (2012)

Synthesis

,

44

, 1657.

40. Ma, D. and Lu, X. (1989)

Tetrahedron Lett.

,

30

, 2109.

41. Rooke, D.A. and Carreira, E.M. (2010)

J. Am. Chem. Soc.

,

132

, 11926.

3Carbophilic Cycloisomerization Reactions of Enynesand Domino Processes