bifunctional organocatalysis in the asymmetric aza-baylis...

145
Bifunctional Organocatalysis in the Asymmetric Aza-Baylis- Hillman Reaction Von der Fakultät für Mathematik, Informatik und Naturwissenschaften der Rheinisch- Westfälischen Technischen Hochschule Aachen zur Erlangung des akademischen Grades eines Doktors der Naturwissenschaften genehmigte Dissertation vorgelegt von Diplom-Chemiker Pascal Buskens aus Heerlen (Niederlande) Berichter: Universitätsprofessor Dr. rer. nat. Walter Leitner Universitätsprofessor Dr. rer. nat. Dieter Enders Tag der mündlichen Prüfung: 22. September 2006 Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek online verfügbar.

Upload: vankhuong

Post on 19-Feb-2019

229 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-

Hillman Reaction

Von der Fakultät für Mathematik, Informatik und Naturwissenschaften der Rheinisch-

Westfälischen Technischen Hochschule Aachen zur Erlangung des akademischen Grades

eines Doktors der Naturwissenschaften genehmigte Dissertation

vorgelegt von

Diplom-Chemiker

Pascal Buskens

aus Heerlen (Niederlande)

Berichter: Universitätsprofessor Dr. rer. nat. Walter Leitner Universitätsprofessor Dr. rer. nat. Dieter Enders

Tag der mündlichen Prüfung: 22. September 2006

Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek online verfügbar.

Page 2: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

I

Page 3: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

II

Selected results from this thesis were already published:

(1) Bifunctional Activation and Racemization in the Catalytic Asymmetric Aza-Baylis-

Hillman Reaction

Pascal Buskens, Jürgen Klankermayer, Walter Leitner, Journal of the American Chemical

Society 2005, 127, 16762-16763.

(2) Highly Enantioselective Aza-Baylis-Hillman Reaction in a Chiral Reaction Medium

Dedicated to Professor Dieter Enders on the occasion of his 60th birthday.

Rolf Gausepohl, Pascal Buskens, Jochen Kleinen, Angelika Bruckmann, Christian W.

Lehmann, Jürgen Klankermayer, Walter Leitner, Angewandte Chemie 2006, 118, 3772-3775

(Titelbild Ausgabe 22, Mai 2006); Angewandte Chemie International Edition 2006, 45, 3689-

3692 (cover picture Issue 22, May 2006).

Page 4: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

III

Abstract (English)

In this doctoral thesis, the mechanism of the Lewis base-mediated aza-Baylis-Hillman

reaction was examined. Intermediates of the reaction were intercepted and structurally

characterized, an extensive kinetic study was performed and the effects of Brønsted acidic co-

catalysts on the reaction mechanism were examined. In addition, other mechanistic aspects

like reversibility, stability of the stereocenter and chirality transfer from the catalytic system

to the reaction product were studied.

This mechanistic study provided useful information for the rational design of an

appropriate chiral catalytic system. Several potential catalytic systems were synthesized and

tested. The most successful system involved a reaction medium, an ionic liquid, as the source

of chirality. With this system, enantioselectivities up to 84% ee were obtained in the

triphenylphosphine-catalyzed aza-Baylis-Hillman reaction of aromatic tosylaldimines and

methylvinyl ketone.

Page 5: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

IV

Abstract (German)

Im Rahmen dieser Doktorarbeit wurde der Mechanismus der Lewis-Basen-katalysierten aza-

Baylis-Hillman-Reaktion untersucht. Reaktionsintermediate wurden abgefangen und

strukturell charakterisiert, die Kinetik der Reaktion wurde eingehend untersucht und die

Auswirkungen von Brønsted-sauren Cokatalysatoren wurden studiert. Zusätzlich wurden

weitere mechanistische Aspekte der Reaktion wie Reversibilität, Stabilität des Stereozentrums

und Chiralitätstransfer vom katalytischen System auf das Reaktionsprodukt behandelt.

Diese mechanistische Studie lieferte wertvolle Informationen für den rationellen

Entwurf eines geeigneten chiralen Katalysatorsystems. Das erfolgreichste System umfasste

als Chiralitätsquelle ein Reaktionsmedium, eine ionische Flüssigkeit. Mit diesem System

konnten in der triphenylphosphin-katalysierten aza-Baylis-Hillman-Reaktion von

aromatischen Tosylaldiminen mit Methylvinylketon Enantioselektivitäten von bis zu 84% ee

erreicht werden.

Page 6: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

V

Page 7: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

VI

Acknowledgements

First of all, I want to thank my Ph.D. supervisor, Prof. Dr. Walter Leitner, for his continuous

support and the helpful discussions during the last three years. In addition, I would like to

thank him for the outstanding working conditions and the possibility to present my work at

numerous conferences.

I would like to express my thanks to Prof. Dr. Dieter Enders for the kind acceptance of the

position of second referee.

Furthermore, I would like to thank Dr. John M. Brown for allowing me to work in his group

at the Chemistry Research Laboratory (CRL) in Oxford as a Marie-Curie-Fellow and for the

numerous scientific discussions we had during this period.

I would like to express my sincere thanks to Dr. Jürgen Klankermayer for his help with the

performance of NMR spectroscopic measurements, lots of valuable discussions and proof

reading this manuscript.

Special thanks to Dipl.-Chem. Rolf Gausepohl for the nice and successful cooperation on the

CHIRAL SOLVENT PROJECT.

In addition, I would like to thank …

Dr. Barbara Odell for her assistance with NMR spectroscopic measurements during my stay

in Oxford.

Dr. Lasse Greiner for his help with the interpretation of the results from the kinetic studies on

the aza-Baylis-Hillman reaction.

Dr. Markus Hölscher for the DFT calculations on the intermediates of the reaction.

Dipl.-Chem. Christian Böing, Dipl.-Chem. Martina Peters and Dipl.-Chem. Clemens Minnich

for proof reading this manuscript.

Page 8: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

VII

Dr. Giancarlo Franciò for a generous gift of (R)-2’-diphenylphosphanyl-[1,1’]binaphthalenyl-

2-ol and (R,S)-2-diphenylphosphanyl cyclohexane carboxylic acid methyl ester.

My lab colleagues Dr. Christian Steffens, Dr. Daniela Giunta, Dipl.-Chem. Rebekka Loschen

and Dipl.-Chem. Volker Gego for the excellent working atmosphere.

My research students Jochen Kleinen, Angelika Bruckmann, Dirk Iffland, Miriam Baumert

and Andreas Hergesell for their valuable help in the course of their research projects.

All members of the Leitner- and JMB group for the great working atmosphere.

All members of the technical staff of the ITMC and the CRL, especially Miss Jennifer Thelen,

for the pleasant working atmosphere and the valuable technical assistance.

Page 9: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

1

Contents

1. OBJECTIVE..........................................................................................4

2. INTRODUCTION ...................................................................................5

2.1. ASYMMETRIC ORGANOCATALYSIS ................................................................................... 5

2.1.1. General Aspects........................................................................................................ 5

2.1.2. Lewis Base Catalysis ................................................................................................ 8

2.1.3. Lewis Acid Catalysis .............................................................................................. 10

2.1.4. Brønsted Base Catalysis......................................................................................... 11

2.1.5. Brønsted Acid Catalysis ......................................................................................... 12

2.1.6. Bifunctional Acid-Base Catalysis........................................................................... 13

2.2. BAYLIS-HILLMAN REACTION ......................................................................................... 14

2.2.1. General Aspects...................................................................................................... 14

2.2.2. Mechanism ............................................................................................................. 15

2.2.3. Catalysts ................................................................................................................. 19

2.3. AZA-BAYLIS-HILLMAN REACTION................................................................................. 23

2.3.1. General Aspects...................................................................................................... 23

2.3.2. Mechanism ............................................................................................................. 26

2.3.3. Catalysts ................................................................................................................. 29

2.4. CHIRAL SOLVENTS FOR ASYMMETRIC SYNTHESIS.......................................................... 32

3. RESULTS AND DISCUSSION .................................................................35

3.1. MECHANISTIC INVESTIGATION OF THE AZA-BAYLIS-HILLMAN REACTION .................... 35

3.1.1. Kinetic Study .......................................................................................................... 35

3.1.1.1. Lewis Base Catalysts........................................................................................... 35

3.1.1.2. Lewis Base-Brønsted Acid Catalysts .................................................................. 40

3.1.2. Reversibility of the Aza-Baylis-Hillman Reaction.................................................. 44

3.1.3. Racemization of the Aza-Baylis-Hillman Product ................................................. 47

3.1.4. Chirality Transfer................................................................................................... 51

3.1.5. Implications for Asymmetric Catalysis .................................................................. 54

3.2. DESIGN OF NEW CATALYTIC SYSTEMS FOR THE ASYMMETRIC AZA-BAYLIS-HILLMAN

REACTION ............................................................................................................................. 55

3.2.1. QUINAP Derivatives.............................................................................................. 55

Page 10: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

2

3.2.2. Other phosphinocarboxylic acids........................................................................... 64

3.2.3. Chiral Brønsted Acids ............................................................................................ 65

3.2.4. Chiral Aminoalcohols ............................................................................................ 66

3.2.4.1. Synthesis.............................................................................................................. 66

3.2.4.2. Catalysis .............................................................................................................. 72

3.2.5. Chiral Solvents ....................................................................................................... 72

4. SUMMARY AND OUTLOOK..................................................................79

5. EXPERIMENTAL.................................................................................84

5.1. GENERAL METHODS, SOLVENTS AND REAGENTS........................................................... 84

5.1.1. Inert Gas Conditions .............................................................................................. 84

5.1.2. Solvents................................................................................................................... 84

5.1.3. Reagents ................................................................................................................. 85

5.2. DETERMINATION OF PHYSICAL DATA............................................................................. 85

5.2.1. NMR Spectroscopy ................................................................................................. 85

5.2.2. IR Spectroscopy...................................................................................................... 86

5.2.3. Mass Spectroscopy ................................................................................................. 86

5.2.4. Gas Chromatography ............................................................................................. 87

5.2.5. High Performance Liquid Chromatography .......................................................... 87

5.2.6. Thin Layer Chromatography.................................................................................. 88

5.3. PREPARATIVE COLUMN CHROMATOGRAPHY.................................................................. 88

5.4. SYNTHESES..................................................................................................................... 88

5.4.1. Aromatic Tosylaldimines........................................................................................ 88

5.4.2. Racemic Aza-Baylis-Hillman Products .................................................................. 91

5.4.3. Enantiomerically Enriched Aza-Baylis-Hillman Products .................................... 93

5.4.4. (R,S)-2-Diphenylphosphanyl cyclohexane carboxylic acid (140).......................... 94

5.4.5. (3-Oxobutyl)triphenylphosphonium malate (109).................................................. 95

5.4.6. [1-(3-Carboxylato-isoquinolin-1-yl)-naphthalen-2-yl]-(3-oxobutyl)-diphenyl-

phosphonium (131)........................................................................................................... 96

5.4.7. (R)-(1-Isoquinolin-1-yl-naphthalen-2-yl)-(3-oxobutyl)-diphenylphosphonium

diphenyl-acetate (172)...................................................................................................... 97

5.4.8. (R,R)- and (S,R)-{1-[3-(3-Hydroxy-2-phenylpropylcarbamoyl)-isoquinolin-1-yl]-

naphthalen-2-yl}-(3-oxobutyl)-diphenylphosphonium diphenylacetate (173) ................. 97

Page 11: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

3

5.4.9. [1-(3-Carboxy-isoquinolin-1-yl)-naphthalen-2-yl]-{3-oxo-2-[phenyl-(toluene-4-

sulfonylamino)-methyl]-butyl}-diphenyl-phosphonium diphenylacetate (132) ............... 98

5.4.10. Methyltrioctyl ammonium dimalatoborate (166) ................................................. 99

5.4.11. trans-4-Hydroxy-(S)-proline methylester hydrochloride (155).......................... 100

5.4.12. (2S,4R)-4-Hydroxy-1-benzoylproline methylester (156).................................... 101

5.4.13. (2S,4R)-4-{[1-(tert-Butyl)-1,1’-dimethylsilyl]-oxy}-1-benzoylproline methylester

(157) ............................................................................................................................... 102

5.4.14. (2S,4R)-4-{[1-(tert-Butyl)-1,1’-dimethylsilyl]-oxy}-1-benzyl-2-hydroxymethyl-

pyrrolidine (158) ............................................................................................................ 103

5.4.15. (3R,5R)-1-Benzyl-3-(tert-butyl-dimethylsilanyloxy)-5-chloro-piperidine (159) 104

5.4.16. (3R,5R)-1-Benzyl-5-chloro-piperidin-3-ol (160) ............................................... 105

5.5. CATALYSIS ................................................................................................................... 106

5.5.1. Screening of Lewis Base-Brønsted Acid Catalysts in the Aza-Baylis-Hillman

Reaction.......................................................................................................................... 106

5.5.2. Kinetic Experiments ............................................................................................. 107

5.5.3. Screening of Brønsted Acids for Kinetics............................................................. 109

5.5.4. Catalysis in Chiral Ionic Liquids ......................................................................... 109

5.5.5. Screening of Chiral Brønsted Acidic Additives.................................................... 110

5.6. MISCELLANEOUS .......................................................................................................... 111

5.6.1. Cross-Mannich Experiment.................................................................................. 111

5.6.2. Deuteration Experiment ....................................................................................... 112

5.6.3. Racemization Experiment..................................................................................... 112

6. APPENDICES....................................................................................113

6.1. ABBREVIATIONS ........................................................................................................... 113

6.2. LIST OF COMMERCIALLY OBTAINED CHEMICALS......................................................... 116

6.3. KINETIC STUDY WITH PHENOL AS CO-CATALYST ........................................................ 117

6.4. 1H AND 13C NMR SPECTRA OF COMPOUND 160 ........................................................... 118

6.5. HPLC CHROMATOGRAM OF AZA-BH PRODUCT 106 ................................................... 120

6.6. 31P NMR SPECTRUM OF COMPOUND 172 ..................................................................... 122

6.7. 31P NMR SPECTRUM OF COMPOUND 173 ..................................................................... 123

7. LITERATURE ...................................................................................124

Page 12: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

4

1. Objective

The objective of this Ph.D. thesis was to develop novel chiral catalytic systems for the

organo-mediated asymmetric aza-Baylis-Hillman reaction, an atom efficient C-C bond

forming reaction leading to highly functionalized chiral allylic amines.[1] The system should

be based on nitrogen- or phosphorus-containing nucleophilic Lewis base organocatalysts.

In principle two general strategies were viable to achieve this goal:

(1) A combinatorial approach involving a screening of a broad variety of potential catalyst

structures for the aza-Baylis-Hillman reaction. These compounds should either be

commercially available or attainable via a short and straightforward synthetic route.

(2) The rational design of suitable chiral catalytic systems based on insights into the

mechanism of the aza-Baylis-Hillman reaction and, more specifically, into the

mechanism of chirality transfer.

Since the first strategy had turned out to be widely unsuccessful in the search for

appropriate chiral catalysts for related organocatalytic C-C couplings like the Baylis-Hillman

reaction, the second strategy was elected.[2] There was no information available on the

mechanism of the aza-Baylis-Hillman reaction at the beginning of this Ph.D. study

(November 2003), so extensive mechanistic research in the initial stage of the project was

necessary. This involved the interception and characterization of reaction intermediates as

well as a full kinetic analysis of the reaction catalyzed by nucleophilic Lewis bases with and

without potential co-catalysts. Furthermore, aspects like chirality transfer, reversibility and

product stability were of high interest.

The implications of these mechanistic findings should provide information for the

rational design of novel chiral catalytic systems. In addition to the conventional approach of

applying chiral organocatalysts in the aza-Baylis-Hillman reaction, the use of well designed

chiral solvents as carriers of chiral information in a catalytic system should be explored in this

present work.

Page 13: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

5

2. Introduction

2.1. Asymmetric Organocatalysis

2.1.1. General Aspects

The term ORGANOCATALYSIS was introduced by MacMillan in 2000 and is used to describe a

research area involving reactions which are mediated by metal-free, small organic

compounds.[3],[4] Scheme 1 shows a selection of typical chiral organocatalysts.

NH

CO2HN

NH

O

Bn

OH

N

N

O

1 2

8

N

NNHBoc

HNO

NHO

ON

O

HN

NHO

ONH

OHN

OO

4

N

N

N

O

BF4-

7

N

F F

F

F F

F

5

HN

S

HN

N

6

HNNH

O

ON

HN

3

PPh

9

Scheme 1. A selection of chiral organocatalysts.

Page 14: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

6

Organocatalysis complements the classical domains of metal- and enzyme catalysis and

provides useful new strategies for organic synthesis. Organocatalysts have several advantages

over transition-metal catalysts and enzymes: they are usually robust, inexpensive and readily

available. Because of their robustness they often do not require demanding reaction conditions

like inert gas atmosphere and absolute solvents.

Whereas synthetic organic chemists have started to explore organocatalysis

systematically in the past decade, it is suggested that this field of catalysis has a long history:

Pizarrello and Weber reported evidence for a key role of organocatalytic processes in the

formation of important chiral prebiotic building blocks like sugars.[5] It was proposed that

aminoacids from carbonaceous meteorites, which carried the same sign L as terrestrial amino

acids, were delivered to the Earth in its early history via showering of the planet with

cometary and asteroidal material. As chiral organocatalysts they could be responsible for the

enantioselective synthesis of sugars. Furthermore, the authors reported that it is likely that

similar organocatalytic processes are responsible for the propagation of chirality in living

organisms.

The first examples of asymmetric organocatalytic processes, in which synthetically

useful enantioselectivities were obtained, date from the early sixties of the past century.

Pracejus reported the use of optically active amines like acetylquinine (12) as catalysts for the

reaction of phenylmethyl ketene (10) with alcohols[6] or amines (Scheme 2).[7]

OCH3OH+

12O

O

toluene, −111°C

10 11 13

OAc

N

N

O

93%74% ee

Scheme 2. Addition of methanol (11) to phenyl methyl ketene (10) as reported by Pracejus.

Page 15: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

7

In the early seventies, both the group of Hajos and Parrish[8] at Hoffmann-La Roche

and the group of Eder, Sauer and Wiechert[9] at Schering reported the L-proline-catalyzed

intramolecular aldol reaction, which nowadays carries their names (Scheme 3).

O

O

1 (3 - 47 mol-%)

CH3CN, r.t. - 80°C

O

O

83% - quant.71 - 93% ee

14 15O

Scheme 3. The Hajos-Parrish-Eder-Sauer-Wiechert reaction.

Despite these early successes it took a long time for synthetic organic chemists to

recognize and appreciate the broad scope of asymmetric organocatalysis. Only in the past

decade enantioselective organocatalysis became a main focus of research. In 2000, List and

Barbas reported pioneering studies on organocatalytic intermolecular aldol reactions[10]

parallelled by a report of the group of MacMillan on enantioselective organocatalytic Diels-

Alder reactions[3] in the same year (Scheme 4 and 5, respectively). These studies elucidated

the promising potential of organocatalytic transformations and resulted in a large scientific

interest in this area of research.[11]

OO

1 (30 mol-%)

DMSO, r.t.+

O OH

16 17 18

97%

96% ee

Scheme 4. L-Proline-catalyzed intermolecular aldol reaction reported by Barbas and List.

Page 16: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

8

O+

CHO

19 20 21

2 (5 mol-%)

23°C

82%94% ee (endo/exo 14:1)

Scheme 5. First enantioselective organocatalytic Diels-Alder reaction reported by the group

of MacMillan.

Several criteria can be applied to categorize the large amount of organocatalytic

reactions reported in literature in the past few years, such as the type of catalyst employed or

the type of reaction that is catalyzed.[11] The mechanistic classification by List involves a

division of the organocatalytic processes into four categories, according to their

mechanisms:[4],[12]

(1) Lewis Base Catalysis

(2) Lewis Acid Catalysis

(3) Brønsted Base Catalysis

(4) Brønsted Acid Catalysis

Despite the sometimes limited information on mechanisms of organocatalytic reactions and

the fact that not all organocatalytic processes can be described with these general reaction

mechanisms, this division will be adopted in this manuscript to give a brief survey of the large

diversity of asymmetric organocatalytic reactions reported in literature.

2.1.2. Lewis Base Catalysis

The general catalytic cycle for reactions catalyzed by Lewis bases (B:) is shown in Scheme 6.

This type of process starts with a nucleophilic addition of the Lewis base B: to the substrate S,

which leads to the adduct B+-S−. This adduct undergoes a reaction to the intermediate species

B+-P−, from which the product (P) is released. The catalyst is set free and re-enters the

catalytic cycle.

Page 17: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

9

Scheme 6. General mechanism for Lewis base-catalyzed reactions.

Prototypical examples of reactions catalyzed by Lewis bases are the reactions shown in

Scheme 4 and 5. The intermolecular aldol reaction, shown in Scheme 4, proceeds through the

enamine intermediate 24 which is formed by deprotonation of the primarily formed iminium

ion 23 (Scheme 7).[10a],[13]

OHN

COOH

+ N

COOHOH

- H2O

N

CO2

N

COOH

+ RCHO

16 1 22

2324

25aldol

product

Scheme 7. The formation of the enamine intermediate 24 (R = alkyl, aryl).

Enamine 24 undergoes an electrophilic attack from aldehyde 25 which leads to the aldol

product after subsequent hydrolysis. Further examples of so-called ENAMINE CATALYSIS are

the Hajos-Parrish-Eder-Sauer-Wiechert reaction (Scheme 3),[8],[9] Mannich reaction,[14]

Michael addition[15] and α-functionalization of ketones and aldehydes (α-amination,[16] α-

hydroxylation,[17] α-alkylation,[18] α-halogenation).[19]

Page 18: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

10

The Diels-Alder reaction in Scheme 5 proceeds through the iminium intermediate 26

(IMINIUM CATALYSIS,[13] Scheme 8).[3]

O + N

- H2O

20 2 26

NH

N

Bn

O

NBn O

Scheme 8. The formation of the iminium intermediate 26.

In case of the Diels-Alder reaction, shown in Scheme 5, the α,β-unsaturated aldehyde 20

forms an iminium coumpound with the chiral imidazolidinone catalyst 2. In comparison to the

unsaturated aldehyde 20, the LUMO energy of the iminium ion 26 as dienophile is lowered.[3]

In general, an iminium ion is more reactive than its corresponding carbonyl compound

and consequently facilitates reactions such as [3+2]-cycloadditions,[20] Friedel-Crafts

alkylation with pyrroles, indoles and benzenes,[21] and Mukayiama-Michael reactions.[22]

Other N-Lewis base-catalyzed processes mainly proceed through acyl-ammonium[23]

and ammonium-enolate intermediates[24] (e.g. Baylis-Hillman reaction, see chapter 2.2.2).

Although the majority of chiral Lewis bases for organocatalysis are N-based compounds, C-

(nucleophilic carbenes, generated from precursors like 7),[25], P-[26] and S-Lewis bases[27] are

also applied as chiral organocatalysts.

2.1.3. Lewis Acid Catalysis

Lewis acid catalysts (A) activate nucleophilic substrates (S:) which leads to the formation of

the adduct A−-S+. This adduct is converted into the adduct A−-P+, followed by elimination of

the product (P). This enables the catalyst (A) to re-enter the catalytic cycle (Scheme 9).

A large class of organocatalysts which can be referred to as Lewis acids are phase

transfer catalysts such as 5.[28] This catalyst was successfully applied to enantioselective α-

alkylations as well as aldol- and Michael reactions.[29] A further example of Lewis acid

catalysis is the asymmetric oxidation method reported by Shi.[30]

Page 19: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

11

Scheme 9. General mechanism for Lewis acid-catalyzed reactions.

2.1.4. Brønsted Base Catalysis

Organocatalytic processes, initiated by Brønsted bases (B:), start with a (partial)

deprotonation of the substrate S-H. The cycle proceeds through the intermediates B+H S− and

B+H P−. Elimination of the product (P-H) enables the catalyst (B) to re-enter the catalytic

cycle (Scheme 10).

Scheme 10. General mechanism for Brønsted base-catalyzed reactions.

Prototypical examples of reactions catalyzed by Brønsted bases are hydrocyanations such as

Strecker reactions (S = CN).[31] The dipeptide species 3 was reported as a successful catalyst

for this type of reactions.[32] It is postulated that in this reaction hydrogen cyanide interacts

with the nitrogen base to form a cyanide ion, which subsequently adds to the carbonyl or

imine compound coordinated to the chiral peptide via hydrogen bonding.

Page 20: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

12

2.1.5. Brønsted Acid Catalysis Brønsted acid-catalyzed reactions start with a (partial) protonation of the substrate (S:). The

resulting adduct A− S+H is converted into the adduct A− P+H, which undergoes an elimination

reaction to liberate the product (P) and the catalyst (A-H) (Scheme 11).[33]

Scheme 11. General mechanism for Brønsted acid-catalyzed reactions.

Jacobsen and co-workers used urea- and thiourea compounds as chiral Brønsted acids to

activate carbonyl- and imine groups in a broad variety of reactions: Strecker-,[34] Mannich-,[35]

Henry-,[36] hydrophosphonylation-,[37] and aza-Baylis-Hillman reactions.[38] These catalysts

activate carbonyl- or imine substrates by hydrogen bonding of the (thio)urea catalyst to the

carbonyl oxygen or imine nitrogen, respectively.

Rawal reported the use of TADDOL derivatives (28) as chiral catalysts for the hetero-

Diels-Alder reaction (Scheme 12).[39]

TBSO

N

O

R+

OHOH

O

O

Ar Ar

Ar Ar(20 mol-%)

toluene, −40°C - −78°C

27 25

28O

RTBSO

N

+ CH3COCl

29

30

O

RO

31

52 - 97%86 - 98% ee

Scheme 12. Hetero-Diels-Alder reaction catalyzed by TADDOL derivatives (R = aryl, alkyl).

Page 21: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

13

Protecting one or both hydroxy-groups leads to a decrease in activity and selectivity,

indicating the involvement of both OH groups in the activation of the carbonyl compound.

McDougal and Schaus reported an enantioselective variant of the Baylis-Hillman

reaction, catalyzed by BINOL derivatives.[40] This will be discussed in more detail in chapter

2.2.3.

2.1.6. Bifunctional Acid-Base Catalysis

In the past few years, several groups showed that the incorporation of a basic and an acidic

moiety in one chiral molecule can facilitate a synergistic interaction between the functional

groups and thereby lead to an efficient bifunctional chiral organocatalyst.

One of the first examples of a bifunctional chiral organocatalyst was reported by the

group of Takemoto in 2003.[41] They used catalyst 34 which contains an amine moiety and a

thiourea group (Scheme 13).

NO2

CO2Et

CO2Et+

HN

S

HNF3C

CF3

N

(10 mol-%)toluene, r.t.

NO2

CO2EtEtO2C

86%93% ee

32 33

34

35

Scheme 13. Michael addition of diethyl malonate (33) to trans-β-nitrostyrene (32).

In this system, the amino group serves as a Brønsted base and activates the nucleophile via

(partial) deprotonation. The thiourea group − a Brønsted acidic catalyst − coordinates to the

nitro function of the α,β-unsaturated olefin. The results obtained with catalyst 34 in the

Michael addition of malonates to a range of different nitroolefins are good to excellent (up to

93% ee).

Since then, bifunctional organocatalysts were successfully applied to a wide range of

reactions. Catalysts containing both a Brønsted basic- and a Brønsted acidic group were

applied to Michael additions,[42] Henry reactions,[43] aza-Henry reactions,[44] Strecker

reactions[45] and kinetic resolution of azlactones.[46] The combination of a Lewis basic

Page 22: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

14

(nucleophilic) and a Brønsted acidic group works well for Baylis-Hillman (see chapter

2.2.3.)[47] and aza-Baylis-Hillman reactions (see chapter 2.3.3.).[48]

2.2. Baylis-Hillman Reaction

2.2.1. General Aspects

The Baylis-Hillman (BH) reaction was first described by Morita et al.[49] and later, in a

German patent, by Baylis and Hillman.[50] It involves the atom-efficient coupling of an

activated alkene in its α-position with a carbonyl compound and is mediated by a nucleophilic

Lewis base. As prototypical example of the BH reaction, the coupling of benzaldehyde (36)

with methylvinyl ketone (MVK, 37) catalyzed by 1,4-diazabicyclo[2.2.2]octane (DABCO,

38) is shown in Scheme 14.

OO

+

NN

OH O

36 37 39

38

Scheme 14. BH reaction of benzaldehyde (36) with MVK (37) catalyzed by DABCO (38).

The reaction shows an enormous diversity in substrates.[2] As activated alkenes alkyl vinyl

ketones, alkyl (aryl) acrylates, acrylonitrile, vinyl sulfones, acryl amides, allenic esters, vinyl

sulphonates, vinyl phosphonates, acrolein, crotonitrile, crotonic acid esters and phenyl vinyl

sulfoxide can be used.[2] They couple with aliphatic, aromatic and heteroaromatic aldehydes,

α-keto esters, non-enolizable 1,2-diketones and fluoroketones.[2] Simple ketones require high

pressure to react as electrophiles in the BH reaction.[51] The BH coupling of activated alkenes

with imines – the aza-BH reaction – is discussed in chapter 2.3. A further variant of the BH

reaction involves the coupling of activated alkenes with sp3 hybridized electrophiles (alkyl

halogenides) – the so-called BH alkylation.[52]

The products of the BH reaction are highly functionalized chiral allylic alcohols of

type 39. They are useful intermediates for the synthesis of more complex organic structures

like anti-tumor agents,[53] anti-biotics[54] and natural products.[55] The potential of the reaction

Page 23: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

15

is further illustrated by its application in industrial processes. tert-Butyl-2-

(hydroxymethyl)propenoate, a key intermediate in Pfizer’s synthetic route to the Zn-

metallaprotease inhibitor Sampatrilat, is prepared via BH coupling on a 40 kg scale.[56]

In principle, the reaction can be mediated by various types of catalysts. In all cases, a

nucleophilic Lewis base, which can be an amine,[57] phosphine,[58] sulphide[59] or selenide,[59]

is part of the catalytic system. Lewis-[60] and Brønsted acids[40] can be added as co-catalysts.

With sulfides and selenides, it is necessary to use a Lewis acidic co-catalyst (e.g. TiCl4) to

obtain a useful catalytic activity.[59] The research described in this thesis focuses solely on

phosphine- and amine catalysts and Brønsted acids as potential co-catalysts.

2.2.2. Mechanism

Several studies contribute to the understanding of the mechanism of the BH reaction. The

generally accepted reaction intermediates have been intercepted and structurally characterized

by using electrospray ionization with mass and tandem mass spectroscopy.[61] Further

experimental evidence is delivered by the group of Drewes through isolation of the second

zwitterionic intermediate (type-41, see Scheme 15) as coumarine salt.[62] The catalytic cycle

of the BH reaction based on the intermediates, which are intercepted and characterized in

these studies, is shown for the prototypical reaction of benzaldehyde (36) with MVK (37)

(Scheme 15).

The reaction starts with a 1,4-addition of the nucleophilic catalyst 38 to the activated

olefin 37, which leads to the formation of the zwitterionic 3-ammonium enolate 40. Enolate

40 is then converted to the second zwitterionic intermediate 41 by aldol-type coupling with

36. The last step involves a proton migration from the α-carbon to the alcoholate oxygen and

release of the catalyst (38) which provides the BH product 39. There is some discussion in

literature regarding the configuration of 40 and 41. Some groups report evidence for an open

isomer as the reactive species, [52b],[63] other groups postulate a (Z)-isomer with a Nu+-O−

stabilization.[64] In Scheme 15, the enolate geometry is therefore left undefined which is

indicated by a wavy bond.

Page 24: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

16

O

NN

OOH O

36

3738

39

NN

O

4041

NN

O

Ph O

O

NN

OOH O

36

3738

39

NN

O

4041

NN

O

Ph O

Scheme 15. Catalytic cycle for the BH reaction of 36 with 37 based on the intercepted and

characterized intermediates.

All steps of the catalytic cycle are potentially reversible.[65] The reaction has a high

negative activation volume which explains the increase of the reaction rate of the typically

slow BH reaction under high-pressure conditions (about 5-15 kbar) reported by several

groups.[66] In some cases, even stereo- and enantioselectivity are affected by a change in

pressure.[67],[51b]

Three mechanistic models are postulated to explain the results obtained in different

kinetic studies. The first model – the so-called CLASSICAL MODEL – was developed by Kaye

and co-workers to interpret the kinetic data obtained for the BH reaction of acrylate esters

with pyridine carboxaldehydes in CDCl3.[68] They monitored the reaction via NMR

spectroscopy and observed a first-order dependence in catalyst, aldehyde, and acrylate and

Page 25: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

17

postulated a mechanism which contained the intermediates shown in Scheme 15 with the

aldol-coupling as the rate determining step (RDS).

The second model – the so-called HEMI-ACETAL MODEL – was developed by the group

of McQuade to explain their kinetic results.[69] They investigated the reaction of various

aromatic aldehydes with methyl acrylate (42) both in non-protic polar solvents like DMSO

and THF and under protic conditions. With rate- and isotope data, they showed that the

reaction is second order in aldehyde and that the α-C-H bond breaks in the RDS. To include

the second equivalent of aldehyde, they altered the cycle shown in Scheme 15 and proposed

the formation of an additional hemi-acetal type intermediate – intermediate 43 – which

preceded the proton migration (Scheme 16). This model was further supported by the

observation of dioxanones as typical by-products of the BH reaction when the acrylate ester

was a reasonable leaving group and the aldehyde concentration was high.[70] Furthermore,

they reported a first order dependence on solvent and a solvent-induced acceleration which

roughly followed the dielectric constant of the solvent or solvent mixture. They explained

these results as a medium effect in which the ionic transition states of the BH reaction were

stabilized in the presence of polar solvents.

O

O O

Ar

O Ar

NN

H 2 eq. aldehyde1 eq. acrylate

1 eq. DABCO

C-H bond broken

43

Scheme 16. Hemi-acetal intermediate 43 as postulated by McQuade.

The third proposed mechanism is the so-called AUTOCATALYSIS MODEL.[71] Aggarwal

and Lloyd-Jones investigated the reaction between benzaldehyde (36) and ethyl acrylate (44)

catalyzed by quinuclidine (45) without the use of a solvent. In a competition experiment with

α-deuterated ([D]44) and non-deuterated ethyl acrylate (44), they observed a substantial

increase of the mole fraction of [D]44 in the early stages of the reaction (< 20% conversion).

As the reaction proceeded, the ratio 44/[D]44 reached a constant level. This was explained

with a scenario where in the early stage of the reaction, the proton migration was rate limiting

but became increasingly efficient during the course of the reaction which ultimately resulted

Page 26: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

18

in a shift of the RDS from the proton migration to the aldol-type coupling. They rationalized

these findings by a model in which the product serves as a co-catalyst through facilitation of

the proton migration from the α-keto methine to the alkoxide making the reaction

autocatalytic. They postulated a proton transfer via transition state 46 (Scheme 17).

R

O

R3N

HE

HO R

46

Scheme 17. Transition state for product-assisted proton migration according to Aggarwal and

Lloyd-Jones (product = ROH).[71]

The AUTOCATALYSIS MODEL, including the TS depicted in Scheme 17, also explained the

large rate enhancements in the BH reaction caused by water and other protic solvents or

additives.[72] At the end of their report, the authors postulated a model for enantioselective BH

reactions with bifunctional chiral catalysts containing a Lewis basic and a Brønsted acidic

moiety and discussed the implications of their mechanistic findings for asymmetric catalysis

(Scheme 18).

RCHO + OR

O

Nu

OHR

O

Nu

O

ORH

HO

D1

D2

D3

D4

fast

slow

slow

slow

type-41

***

*

Scheme 18. Model for enantioselective BH reactions by Aggarwal and Lloyd-Jones (Nu∩OH

= chiral bifunctional catalyst).

Since the second intermediate of the BH reaction (type 41) contained two centers of chirality,

four diastereoisomers could be formed with an enantiopure catalyst. The authors suggested

that all four diastereoisomers (D1-4) of the “type-41 intermediate” should be formed during

the reaction, but only one had the hydrogen-bond donor suitably positioned to allow an

Page 27: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

19

effective proton transfer. The other diastereoisomers would revert back to the starting

materials and eventually the reaction would “filter” through the pathway that leads to fast

elimination. This implied that the proton migration step determined the enantioselectivity of

the reaction and that the reaction itself could be regarded as a dynamic kinetic resolution of

the intermediate of type 41 which is shown in Scheme 18. Other models explaining the

enantioselectivity in catalytic asymmetric BH reactions for specific catalyst systems are

discussed in the next chapter.

2.2.3. Catalysts

As already mentioned, the BH reaction can be mediated by nucleophilic Lewis bases or by

combined Lewis base/Lewis acid or Lewis base/Brønsted acid systems. Only the systems

based on N- and P-Lewis bases (in combination with Brønsted acids) will be discussed here in

more detail.

Suitable N-Lewis base catalysts for the BH reaction are compounds like DABCO

(38),[73] quinuclidine (45),[74] 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU, 47),[75] imidazole

(48)[76] and 4-dimethylaminopyridine (DMAP, 49) (Scheme 19).[77]

NN

NN N

N

N

NHN

38 45 47 48 49

Scheme 19. Typical N-Lewis base catalysts for the BH reaction.

These Lewis bases can be divided into two categories. The first group contains highly

nucleophilic and unhindered Lewis bases like DABCO (38) and quinuclidine (45). Because of

their high nucleophilicity they enable the efficient formation of the 3-ammonium enolate

(type-40) which leads to a high overall reaction rate. The second group consists of catalysts

which can stabilize this 3-ammonium enolate by delocalization of the positive charge on

nitrogen (e.g. 47-49) (Scheme 20).

Page 28: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

20

NN

OO

NN

OO

50 50

Scheme 20. Stabilization of the DBU-containing 3-ammonium enolate (50) of methyl

acrylate (42).

Typical P-containing Lewis bases for the BH reaction are trialkylphosphines like P(n-

Bu)3 (51),[78] PEt3 (52) and the air-stable trialkylphosphine 1,3,5-triaza-7-phosphaadamantane

(PTA, 54) (Scheme 21).[79] Due to their reduced nucleophilicity, triarylphosphines like PPh3

(53) can only initiate the reaction in the presence of a suitable co-catalyst like phenol.[80]

NN

P

NP

P

P

51 52 53 54

Scheme 21. Typical P-Lewis base catalysts for the BH reaction.

As already mentioned before, Brønsted acids are efficient co-catalysts for the BH

reaction and lead to significant rate enhancements. Typical protic additives which are used to

increase the efficiency of BH reactions are for example methanol,[81] water,[82] (thio)ureas,[83]

and phenols.[40] Both catalytic moieties can also be incorporated in one catalyst. A simple

example is 3-quinuclidinol (55), which in general shows a much higher catalytic activity than

quinuclidine (45) itself due to the formation of hydrogen bonds in the corresponding

intermediates.[84] This concept of applying a bifunctional catalyst in BH reactions has proven

to be successful in asymmetric catalytic variants.

The first successful chiral catalyst that resulted in synthetically useful

enantioselectivities in the BH reaction has been developed only recently. 1999, Hatakeyama

and co-workers applied the cinchona alkaloid derivative 56 as catalyst for the BH reaction of

various aldehydes with 1,1,1,3,3,3-hexafluoroisopropyl acrylate (57) (Scheme 22).[47b]

Page 29: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

21

At −55°C in DMF they obtained the corresponding products (58) in enantioselectivities up to

99% ee.

RCHO +O

O

CF3

CF3

N

NO

OH

R

OH O

O

CF3

CF3

O O

R

R O

+

57

56

58 59

−55°C, DMF

yield up to 58%up to 99% ee

(10 mol-%)

25

Scheme 22. Reaction of various aldehydes with 57 under the catalytic influence of 56 (R =

C6H5, p-NO2C6H4, CH3CH2, C6H11 etc).

The explanation for the high enantioselectivity as given by the authors was based on three

features of catalyst 56:

(1) its increased nucleophilicity in comparison to related compounds like quinidine

(2) the presence of the free phenolic-OH group at the quinoline ring

(3) the anti-open conformation of catalyst 56

The authors postulated that the anti-open conformation allowed an optimal stabilization of

zwitterion 60 through intramolecular hydrogen bonding (Scheme 23).

NN

O

O H O

O

O

CF3

CF3

R

60

A

B

Scheme 23. Postulated stabilization of intermediate 60 with Hatakeyama’s catalyst 56.

Page 30: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

22

Since the chirality of the catalyst is fixed and intermediate 60 contains two chiral centers, A

and B, four diastereoisomers can be formed in this stage of the reaction. The authors

postulated that diastereoisomer 60, depicted in Scheme 23, is the most stable and has a nearly

ideal conformation for the subsequent elimination reaction. Elimination from the catalyst in

intermediate 60 leads to the BH product with the stereoconfiguration of 58. A major

disadvantage of Hatakeyama’s catalytic system is the formation of relatively large amounts of

the dioxanone by-product 59 which suppresses the yield of the BH product 58. These

dioxanone by-products are formed in opposite stereoconfigurations and worse

enantioselectivities as compared to 58. This dioxanone formation, however, contributed to the

mechanistic understanding of the BH reaction (cf. HEMI-ACETAL MODEL). Catalyst 56 was

subject of further studies to expand the scope of substrates[85] and has also been applied in the

synthesis of natural products.[86] Other bifunctional chiral catalysts for the BH reaction which

contain both catalytic moieties in one chiral molecule, are the cinchona-alkaloid derivative 61

(pseudo-enantiomer to 56),[87] N-methylprolinol (62)[88] and the amine-thiourea catalyst 63

(Scheme 24).[89]

NH

S

NH

CF3

CF3NN

OH

up to 94% ee

N

OH

N

O

up to 91% ee up to 78% ee61 62 63

Scheme 24. Recent examples for bifunctional chiral catalysts for the BH reaction.

Examples of efficient systems with chiral Brønsted acidic co-catalysts in combination

with simple achiral nucleophiles like DABCO (38) and PEt3 (52) are based on L-proline

(1),[90] BINOL derivatives (64),[40] the bis-thiourea compound 65[91] and pipecolinic acid (66)

(Scheme 25).[92]

Page 31: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

23

NH

CO2H

66

up to 84% ee (with methylimidazol)

NHS

HN

CF3

CF3

NH

S

NH

CF3

CF3

65up to 90% ee (with DABCO)

NH

CO2H

1up to 81% ee

(with methylimidazol-modified peptide)

X

OH

X

OH

64

up to 96% ee (with PEt3)

Scheme 25. Chiral co-catalysts for the BH reaction (X = H, Br, Ph etc).

There are only few non-bifunctional chiral catalyst systems for the BH reaction

reported in literature. Hayashi and co-workers reported a chiral diamine catalyst with

enantioselectivities up to 75% ee.[93] As a chiral biphosphine catalyst, BINAP resulted in

enantioselectivities up to 44% ee.[94]

2.3. Aza-Baylis-Hillman Reaction

2.3.1. General Aspects

The aza-Baylis-Hillman (aza-BH) reaction – the coupling of activated alkenes with imines –

was first reported by Perlmutter and Teo in 1984 (Scheme 26).[1] This atom-efficient C-C

bond forming reaction is, in analogy to the BH reaction, usually catalyzed by nucleophilic

Lewis bases like DABCO (38) and leads to highly functionalized chiral allylic amines (68).

NTos

+O

O

NN

NH O

O38

4467 68

Tos

Scheme 26. Aza-BH reaction of N-benzylidene-4-methylbenzenesulfonamide (67) with ethyl

acrylate (44), catalyzed by DABCO (38).

Page 32: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

24

The reaction is flexible towards the choice of starting materials. As electrophiles,

various activated aromatic aldimines like tosyl-,[48b],[93] nosyl-[38] and phosphinoylimines[95]

can be used. The reaction can also be performed as a three component reaction in which

aldehyde, activated alkene and tosylamide[96] or diphenylphosphinamide[97] are coupled in one

pot. Furthermore, sulfinimines[98] and azadicarboxylates[99] can be applied in the aza-BH

reaction.

The spectrum of Michael acceptors, which can be used for this reaction, is even

broader.[100] MVK, acrolein, phenyl acrylate, α-naphthyl acrylate, methyl acrylate, 2-

cyclopentenone and 2-cyclohexenone as well as other β-substituted activated alkenes were

reported as suitable Michael acceptors.[101] Also, activated allenes and alkynes are appropriate

substrates.[102]

In the presence of P(n-Bu)3 (51) as Lewis base the aza-BH reaction of aldimines like

67 with cyclic α,β-unsaturated ketones like 2-cyclohexenone (69) can result in the formation

of the expected aza-BH product 70 along with the bicyclic by-product 71 (Scheme 27).[103]

NTos

+NH

51

6967 70

TosO P(n-Bu)3 (20 mol-%) O+

MeOH, 40°C

O 67 O

Ph

NTos

N

O

Tos

Ph

H+

72 73 71

N

O

Tos

Ph

71

Scheme 27. Formation of the bicyclic compound 71 under aza-BH reaction conditions.

The mechanism postulated for the formation of 71 starts with a Mannich-type coupling of 72

and 67 which results in the formation of 73. In the next step, intermediate 73 undergoes

intramolecular Michael addition and subsequent protonation to form 71.

Another substrate combination which can lead to unexpected (by-)products under aza-

BH reaction conditions is 67/MVK (37) (Scheme 28).[104] In the case of PPh3 (53) as catalyst,

the expected aza-BH product 77 was the only observed reaction product. With P(n-Bu)3 (51)

Page 33: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

25

as catalyst under otherwise identical reaction conditions, the only observed products were 74

and 75. The proposed mechanism that explains the formation of these products involves the

hemi-aminal intermediate 78. In case of the BH reaction, a similar intermediate (hemi-acetal)

is postulated to explain the formation of dioxanones as by-products as well as specific kinetic

results (see chapter 2.2.2. and 2.2.3.).

NTos

+ 51

3767

O P(n-Bu)3 (20 mol-%)

THF, r.t.N HNTos

Ph

Ph

O

Ph

Ph O

74 75

O

R3P

NPhTos

76

- PPh3 (53) NH OTos

77

(R = Bu)67

PBu3

NPhTos

N

Ph

Tos

O

78

- PBu3 (51)

PBu3

NPhTos

HN

Ph

Tos

O

79

NTos Ph

NHTos

Ph

O

74 + 75

80

Scheme 28. Unexpected products for the reaction of 67 with 37.

The synthetic potential of the aza-BH reaction is illustrated by its integration into

synthetic routes. It is, for example, used for the synthesis of PEG-supported α-methylene-β-

aminoesters,[105] substituted pyrroles[106] and 2,5-dihydropyrroles.[107]

Page 34: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

26

2.3.2. Mechanism

At the beginning of this Ph.D. study (November 2003) no information was available on the

mechanism of the aza-BH reaction. It was assumed that the catalytic cycle of the reaction

contained intermediates similar to those commonly accepted for the BH reaction (Scheme 29).

NN

NH O

O

Tos

67

443868

NN

O

O

8182

NN

O

O

Ph NTos

O

O

NTos

NN

NH O

O

Tos

67

443868

NN

O

O

8182

NN

O

O

Ph NTos

O

O

NTos

Scheme 29. Assumed catalytic cycle for the aza-BH reaction.

In parallel to the work described in this thesis, Jacobsen and co-workers investigated the

reaction between the aromatic nosyl-aldimine 83 and methyl acrylate (42) (Scheme 30).[38]

They published their results almost simultaneously with our first communication on the

reaction mechanism.

Page 35: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

27

NNos

+CO2Me

38 +

N

O

tBu

NH

S

NH

BnN

HO

But tBu NHCO2Me

Nos(1 eq.) (20 mol-%)

84

xylene, r.t.

55%97% ee

83 42 85

Scheme 30. Reaction of the nosyl imine 83 with methyl acrylate (42), catalyzed by DABCO

(38) and thiourea 84.

As catalyst they used a combined system of the Lewis basic nucleophile DABCO (38) and the

Brønsted-acidic chiral thiourea 84 in xylene. They observed a yellow precipitate during the

course of the reaction and spectroscopically identified it as the second intermediate of the

catalytic cycle 86 (Scheme 31). Upon treatment with an excess of hydrochloric acid, they

could isolate and characterize the dihydrochloride salt 87 as a glassy solid.

CO2MePh

NNos

NN

HCl CO2MePh

NHNos

NN

H

2 Cl-

86 87

Scheme 31. Isolation and characterization of the dihydrochloride salt 87 of intermediate 86 by

the group of Jacobsen.

Intermediate 86 (and its corresponding dihydrochloride salt 87), which resulted from a

Mannich-type coupling of the corresponding zwitterionic 3-ammonium enolate with imine 83

was obtained in high diastereomeric purity (d.r. > 10/1) with a relative stereochemistry of the

major isomer assigned as anti. Furthermore, they disclosed that 86 could racemize under turn-

over conditions. This stereochemical lability was explained by the reversibility of the

Page 36: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

28

Mannich-type coupling step which would create a racemization pathway for 86 in which the

Brønsted acidic co-catalyst did not play a role (Scheme 32).

CO2MePh

NHNos

NN

H

2 Cl-

87

DBU (47)rac-85 + 83 42+

trace amount

Scheme 32. Racemization under turn-over conditions.

Because the reaction mixture for the reaction disclosed in Scheme 30 was not

homogeneous, kinetic experiments could not be performed for that exact system. Therefore

the group of Jacobsen studied the reaction of imine 83 with methyl acrylate (42) without co-

catalyst in CHCl3 which provided a homogeneous mixture during the complete course of the

reaction. They monitored the conversion by GC analysis and analyzed their data with the

method of initial reaction rates.[108] The reaction showed a first order kinetic dependence on

both DABCO (38) and methyl acrylate (42). In contrast to the BH reaction, they observed rate

saturation for the imine 83 as electrophile.

In kinetic isotope effect studies using α-deuterated methyl acrylate ([D]42) they

disclosed a prominent primary kinetic isotope effect (kH/kD = 3.81), which strongly suggested

that a cleavage of the α-CH(D) bond was involved in the RDS (RDS = proton

migration/elimination).

To determine the influence of the electrophile on the elimination reaction, they

investigated the DBU (47)-mediated elimination of 87 in methanol via IR spectroscopy. They

found a first-order reaction rate profile for more than 4 half-lives, which was consistent with a

rapid and irreversible deprotonation of 87 to 86, followed by intramolecular proton transfer.

Since added imine had no effect on the rate of elimination, they concluded that the

electrophile did not play a role in mediating the elimination step thereby excluding a pathway

similar to the HEMI-ACETAL MECHANISM postulated by McQuade for the BH reaction.[69]

Further mechanistic investigations were carried out by the groups of Sasai and Shi

focussing on the origin of enantioselectivity in their specific systems. The results of these

studies will be discussed in the next chapter.

Page 37: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

29

2.3.3. Catalysts

In accordance with the BH reaction, the aza-BH reaction can be mediated by nucleophilic

Lewis bases, by Lewis bases in combination with Lewis acids[109] and by combined Lewis

base/Brønsted acid systems. Only the systems based on N- and P-Lewis bases (with Brønsted

acid) will be discussed here in more detail.

In general, the achiral nucleophilic N- and P-containing Lewis bases, which were

discussed as catalysts for the BH reaction, can also be applied to mediate the aza-BH reaction.

For the asymmetric aza-BH reaction, however, different catalytic systems are applied.

One of the most successful chiral catalytic systems for the aza-BH reaction was

developed by Shi and co-workers in 2003.[110] They developed the P-based catalyst 88 which

carried a Brønsted acidic phenol moiety and tested it initially for the coupling of aromatic

tosylaldimines like 67 with simple Michael acceptors like MVK (37) and phenyl acrylate

(Scheme 33). Depending on the substitution pattern of the aromatic tosylaldimine,

enantioselectivities up to 97% ee were obtained.

NTos

O+

OHPPh2

NH OTos

67 37

88

89

(10 mol-%)

THF, −30°C, MS 4A, 24 h

83%83% ee

Scheme 33. Reaction of the aromatic tosylaldimine 67 with MVK (37), catalyzed by 88.

In further studies, the substrate scope of this catalytic system was investigated.[48b],[111]

Besides aromatic tosylaldimines like 67, aromatic mesyl-, p-chlorobenzenesulfonyl- and β-

trimethylethanesulfonyl (SES) aldimines could be used. The catalyst was inactive in reactions

with aliphatic aldimines.

As Michael acceptor, MVK, ethyl vinyl ketone (EVK), phenyl acrylate, α-naphthyl

acrylate, β-naphthyl acrylate and acrolein could be used. Reactions with methyl acrylate

Page 38: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

30

proceeded in a very sluggish way with low enantioselectivities (typically about 20% ee). β-

Substituted Michael acceptors proved inactive with 88. Replacement of the PPh2 group with a

more nucleophilic PMe2 group, however, resulted in a suitable chiral catalyst system for

reactions with this type of Michael acceptors.

In addition, the authors carried out some mechanistic investigations on their system.

The phenolic OH group turned out to be crucial for the success of the catalyst.[112] The

introduction of substituents in 3, 6 or 6’ position of 88 did not result in significant changes in

activity and selectivity. Replacement of the PPh2 group with a P(o-tolyl)2 group, however,

yielded an inactive system, which indicated that the steric bulk around the P-atom played a

key role in initiating the aza-BH reaction. Furthermore, the authors characterized the “type-81

intermediate” of 88 with 37 via NMR spectroscopy. They suggested a stabilization of this

intermediate by hydrogen bond interaction of the phenolic OH group with the enolate oxygen.

They explained the success of their chirality transfer with a model similar to that postulated

by Hatakeyama and co-workers for catalyst 56 (page 21).

A second system based on a phosphorus-containing Lewis base was developed by

Sasai and co-workers in 2006.[113] They attached a phosphine unit through an aromatic ring to

the 3-position of (S)-BINOL as a chiral Brønsted acid and tested the system for the reaction of

imine 67 with MVK (37) (Scheme 34).

NTos

O+

NH OTos

67 37

90

89

(10 mol-%)

MTBE, −20°C,9 d

97%87% ee

OH

PPh2OH

Scheme 34. Reaction of imine 67 with MVK (37) catalyzed by 90.

They obtained enantioselectivities up to 95% ee, albeit with a very slow reaction rate. When

both catalytic moieties of this bifunctional catalyst were positioned in a way that no

Page 39: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

31

synergistic interaction between the Lewis base and the Brønsted acid could occur in an

intramolecular fashion, the selectivity decreased dramatically.

Similar results were reported by the same group for a nitrogen containing bifunctional

system which was introduced in 2005.[114] They investigated the reaction of aromatic

tosylaldimines like 67 with MVK (37) catalyzed by the bifunctional catalyst 91 (Scheme 35).

NTos

O+

NH OTos

67 37

91

89

(10 mol-%)

toluene:CPME (1:9), −15°C, 168 h

93%87% ee

OHOH

NN

Scheme 35. Reaction of 67 with 37 catalyzed by the bifunctional catalyst 91.

In a mixture of toluene and cyclopentyl methyl ether (CPME) at −15°C, enantioselectivities

up to 96% ee could be obtained for aza-BH reactions between aromatic tosylaldimines and

simple Michael acceptors like MVK, EVK and acrolein. The authors showed that activity and

enantioselectivity of the catalytic system were severely dependent on the exact position of the

active units within the catalyst.[48a] Many similar catalysts which contained different spacers

between the catalytic moieties and/or different positions of the pyridine nitrogen relative to

the Brønsted acidic group were often inactive and non-selective. Only when both catalytic

functions were correctly positioned within the organocatalyst, a synergistic interaction was

possible which resulted in an efficient bifunctional chiral catalyst. In a more extensive

mechanistic study, they disclosed that the catalytically active groups of catalyst 91 are the

pyridine-N atom and the phenolic OH group in 2’-position.[48a]

A fourth bifunctional system which proved effective in mediating the aza-BH reaction

of aromatic tosylaldimines involves the BH catalyst 56. Both the group of Shi[115] and the

group of Adolfsson[116] showed that catalyst 56 did not only efficiently catalyze the BH but

also the aza-BH reaction. The substrate scope of this catalytic system was fairly broad: MVK,

Page 40: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

32

EVK, acrolein, methyl acrylate, phenyl acrylate and naphthyl acrylates could be used as

Michael acceptors. Furthermore, the reaction could be performed enantioselectively as a three

component coupling of aldehyde, tosylamide and activated alkene. With catalyst 56,

enantioselectivities up to 99% ee were obtained. This successful transfer of chirality was

explained by a mechanistic model similar to that for the BH reaction with 56 as catalyst.

In principle, it is also possible to decouple both catalytic moieties. As already partially

described in 2.3.2., Jacobsen and co-workers used chiral thiourea derivatives like 84 in

combination with simple achiral nucleophiles like DABCO (38).[38] They investigated the

coupling of aromatic nosylaldimines like 83 with methyl acrylate (42) and obtained the

corresponding reaction products in enantioselectivities up to 99% ee. The activity of this

catalytic system, however, was low to moderate. Furthermore, the observed inverse

relationship between chemical- and optical yield for this system leaves no room for

improvement. In addition, the substrate scope is, to date, very limited.

2.4. Chiral Solvents for Asymmetric Synthesis

A chemical reaction that leads to enantiomerically enriched products can in principle be

carried out with a source of chirality in any of the involved components – starting

materials,[117] catalysts,[118] or solvents.[119] The strategy of applying chiral solvents in

asymmetric synthesis was pioneered by Seebach and Oei.[120] They used a chiral aminoether

as the solvent in the electrochemical reduction of ketones but observed only low

enantioselectivities (up to 23.5% ee). Since then, all attempts to apply chiral solvents in

asymmetric synthesis have resulted in similar or lower enantioselectivities, which led to the

generally accepted conclusion that “asymmetric induction caused by chiral solvents is usually

rather small”.[119]

As ionic liquids (ILs), however, recently attracted a lot of attention as a new class of

solvents,[121] the incorporation of chiral information in these media has caught the interest of

several groups.[122] A selection of chiral ions, used for the synthesis of chiral ILs, is shown in

Scheme 36.

The first contribution to this area of research was made by the group of Seddon and

co-workers in 1999.[123] They introduced [bmim] lactate (92) as anion-chiral IL, which was

prepared by ion metathesis of sodium lactate with [bmim] chloride. Application of this IL as a

solvent for the Diels-Alder reaction of ethyl acrylate with cyclopentadiene, however, did not

result in a transfer of chirality from the reaction medium to the corresponding reaction

Page 41: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

33

product. Although this first application of a chiral IL as reaction medium was rather

unsuccessful, Seddon’s study initiated the synthesis of a broad spectrum of chiral ILs, as

shown in Scheme 36.

NO

R'RN

HO

CO2-

OHH CO2

-

R O

OP

O

OSO3

- O

92 93 94 95

Seddon, 1999 Dorta, 2005 Dorta, 2005Ohno, 2005

96 97

N N

98

OO

H

NC10H21

R

H

99

Wasserscheid, 2001 Wasserscheid, 2001 Saigo, 2002 Plaquevent, 2005

Scheme 36. A selection of chiral ions which are used for the synthesis of chiral ILs.

In 2001 Wasserscheid and co-workers developed a first set of ILs containing chiral

cations including ephedrinium based ILs (96).[124] Very recently, Vo-Thanh and co-workers

used these ILs as solvent for the DABCO-catalyzed BH reaction of benzaldehyde and methyl

acrylate.[125] With this IL, they obtained the first significant asymmetric induction caused by a

chiral IL as the only source of chirality in an asymmetric reaction. The corresponding BH

product 101 was obtained in an enantiomeric excess of up to 44% (Scheme 37).

Page 42: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

34

O O

O+38 (1 eq), 30°C, 7 d

O

O

OH

60%44% ee

36 42 101

NHO

C8H17

100OTf

Scheme 37. BH reaction in the ephedrinium IL 100.

Further examples of asymmetric induction caused by cation-chiral ILs as solvents were

reported by Armstrong and co-workers (photoisomerization of

dibenzobicyclo[2.2.2]octatriene diacid, ee up to 11%)[126] and by the group of Bao (Michael

additions of malonates to enones, ee up to 25%).[127] No significant asymmetric induction has

been reported so far for anion-chiral ILs.

The comparatively high enantioselectivities obtained in the BH reaction of

benzaldehyde and methyl acrylate indicate that the application of chiral ILs as the source of

chirality in organocatalytic processes can be a promising strategy to obtain the corresponding

products in valuable enantioselectivities. A potential reason for the comparatively high

enantioselectivity in this reaction may be a strong interaction of the zwitterionic reaction

intermediates (transition states) with the chiral ions of the solvent. Possible modes of

interaction with ILs are electrostatic attraction and, in case of IL 100, hydrogen bonding.

Page 43: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

35

3. Results and Discussion

3.1. Mechanistic Investigation of the Aza-Baylis-Hillman Reaction

3.1.1. Kinetic Study

3.1.1.1. Lewis Base Catalysts

In initial experiments, kinetic measurements were performed on the catalytic system which

contained only a Lewis base catalyst. As a model reaction, the aza-BH coupling of N-(3-

fluorobenzylidene)-4-methylbenzenesulfonamide (102) with MVK (37) in the presence of

triphenylphosphine (53) as catalyst was examined (Scheme 38). The reaction was carried out

in THF at room temperature.

F

N TosO

+

PPh3, C6F6

THF, r.t.

F

NH OTos

102 103

10453

37

− 112.6 −114.1−164.0δ (19F) ppm

Scheme 38. Model reaction for kinetic study of the Lewis base-catalyzed aza-BH reaction.

The first method considered for this study was on-line IR spectroscopy. In cooperation

with Patrick Bäuerlein from the group of Dr. Lasse Greiner, the reaction could be monitored

via IR spectroscopy in an Argon-glovebox. For this purpose, an ATR-IR sensor from the

company INFRARED FIBER SENSORS was used, which measures the wave-length dependent

attenuation of the intensity of the IR signal for the total reflection at a diamond crystal.[128]

The laser beam was injected and released through a fiber optical cord. The probe and the

setup in the glovebox are depicted in Figure 1.

Page 44: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

36

Figure 1. ATR-IR probe from INFRARED FIBER SENSORS, mounted for reaction monitoring in

a glovebox.

The main advantage as compared to other potential analysis techniques (e.g. NMR

spectroscopy, vide infra) was the velocity of the measurement and the high flexibility of the

probe as a result of the attached fiber optics. A major disadvantage, however, was the

complexity of the handling and analysis of the obtained data involving operations like base

line corrections and curve fitting. Furthermore, an external reference was necessary for an

appropriate quantification. Although the data that were obtained with this technique were in

accordance with data obtained from 1H NMR spectroscopic studies, the disadvantages of this

method outweighed its advantages for the aza-BH reaction.[129]

As a second method NMR spectroscopy was considered. A general advantage of

NMR- over IR-spectroscopy was the easy data handling. An important drawback of this

technique, however, was the inability to obtain data during the first minutes of a reaction due

to experimental factors (lock/shim). Since the aza-BH reaction was a relatively slow process,

this disadvantage played only a minor role in this kinetic study. Consequently, NMR

spectroscopy was elected as the method of choice for these investigations. Furthermore, the

use of the fluorinated electrophilic coupling partner 102 allowed reaction monitoring via 19F

NMR spectroscopy. In consequence, the use of large amounts of expensive deuterated

solvents like [D8]THF which would be necessary for 1H NMR spectroscopic measurements,

could be avoided. Hexafluorobenzene (104) was applied as internal standard.

Typical concentration-time profiles that were obtained with 19F NMR spectroscopy,

are shown in Figure 2.

Page 45: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

37

0,000

0,020

0,040

0,060

0,080

0,100

0,120

0,140

0 100 200 300 400 500 600 700

t / min

c / (

mol

/l)

48 mol-%64 mol-%78 mol-%105 mol-%

Figure 2. Typical concentration-time profiles. Starting concentrations: c (102) = 0.1514

mol⋅l−1, c (37) = 0.1531 mol⋅l−1, c (104) = 0.0285 mol⋅l−1, c (53) = 0.0724 - 0.1595 mol⋅l−1.

The concentration-time profiles were fitted with a polynomial of fourth order

(Equation 1). The initial reaction rate was determined through differentiation (Equation 2).

pdtctbtattc ++++= 234)( ℜ∈pdcba ,,,, Equation 1

ddt

tdcdctbtatdt

tdc

t

=⎟⎠⎞

⎜⎝⎛⇒+++=

=0

23 )(234)( ℜ∈pdcba ,,,, Equation 2

The initial reaction rate was analyzed as a function of the starting concentration of the

individual components. The corresponding plots are shown in Figure 3a-d.

Page 46: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

38

0,000E+00

1,000E-04

2,000E-04

3,000E-04

4,000E-04

5,000E-04

6,000E-04

0,000 0,050 0,100 0,150 0,200 0,250 0,300

c (imine) / (mol / l)

r(0)

/ [m

ol /

(l*m

in)]

Figure 3a. Partial order in imine 102. Starting concentrations: c(53) = 0.0501 mol⋅l−1, c(37) =

0.1438 mol⋅l−1, c(104) = 0.0243 mol⋅l−1, c(102) = 0.0998-0.2530 mol⋅l−1.

log r(0) = 0,4685log c(imine) + 3,0243R2 = 0,9908

3,00

3,10

3,20

3,30

3,40

3,50

3,60

0,50 0,60 0,70 0,80 0,90 1,00 1,10

log c(imine)

log

r(0)

Figure 3b. Partial order in imine (double-logarithmic).

Page 47: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

39

r(0) = 0,0027*c(MVK)R2 = 0,9909

0,000E+00

1,000E-04

2,000E-04

3,000E-04

4,000E-04

5,000E-04

6,000E-04

7,000E-04

8,000E-04

9,000E-04

0,00 0,05 0,10 0,15 0,20 0,25 0,30 0,35

c(MVK) / (mol/l)

r(0)

/ [m

ol/(l

*min

)

Figure 3c. Partial order in MVK (37). Starting concentrations: c(53) = 0.0500 mol⋅l−1, c(102)

= 0.1506 mol⋅l−1, c(104) = 0.0214 mol⋅l−1, c(37) = 0.1712 – 0.3163 mol⋅l−1.

r(0) = 0,00698*c(catalyst)R2 = 0,98780

0,000E+00

2,000E-04

4,000E-04

6,000E-04

8,000E-04

1,000E-03

1,200E-03

0,00 0,02 0,04 0,06 0,08 0,10 0,12 0,14 0,16 0,18

c(catalyst) / (mol/l)

r(0)

/ [m

ol/(l

*min

)]

Figure 3d. Partial order in PPh3 (53). Starting concentrations: c(37) = 0.1531 mol⋅l−1, c(102)

= 0.1514 mol⋅l−1, c(104) = 0.0285 mol⋅l−1, c(53) = 0.0724 – 0.1595 mol⋅l−1.

Page 48: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

40

These results demonstrated that the reaction showed a first order dependence in Michael

acceptor (37) and catalyst (53). To determine the non-integer order in imine (102), the data

were analyzed by using a double-logarithmic plot (Figure 3b).[130] The gradient of this plot

corresponded to a partial reaction order of approximately 0.5 (Equation 3).

)()()(5.0 hosphinepcacceptorMichaelcimineckr obs ⋅−⋅⋅= Equation 3

A comparison of these data to the kinetic results obtained by Aggarwal and Lloyd-Jones for

the autocatalytic BH reaction revealed that the aza-BH coupling was no autocatalytic process

under the conditions applied in this kinetic study.[71] A HEMI-ACETAL MECHANISM, as

proposed by McQuade for the BH reaction in aprotic solvents, could also be ruled out.[69]

Based on the rate law in Equation 3, however, no autonomous mechanism for the Lewis base-

catalyzed aza-BH reaction could be postulated. A mechanistic interpretation of these data will

be provided in the next chapter by comparison to the kinetic data of the aza-BH reaction

mediated by a bifunctional Lewis Base-Brønsted acid catalyst system. Similar kinetic results

for the aza-BH reaction catalyzed by nucleophilic Lewis bases were independently reported

by the group of Jacobsen.[38]

3.1.1.2. Lewis Base-Brønsted Acid Catalysts

In a second set of experiments, the influence of Brønsted acidic additives (co-catalysts) on the

reaction of N-(4-bromobenzylidene)-4-methylbenzenesulfonamide (105) with MVK (37) in

[D8]THF was examined (Scheme 39).

N TosO

+

PPh3, Brønsted acid

[D8]THF, r.t.

NH OTos

Br Br37

53

105 106

Scheme 39. Reaction of imine 105 with MVK (37), catalyzed by the system PPh3

(53)/Brønsted acid.

A set of Brønsted acids with pKA values in the range of 4 to 16 was screened as potential

additives for this model reaction.[131] Approximately one equivalent of Brønsted acid per

Page 49: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

41

imine was used. A strong influence of strength of the Brønsted acid on the rate of the reaction

was observed (Table 1, Figure 4).

Table 1. Influence of Brønsted acidic additives on the reaction rate.

entry additive pKA relative initial rate 1 malic acid 3.6 0.1 2 - - 1.0 3 aza-BH product (106) n.a. 1.2 4 water 15.4 2.4 5 methanol 15.5 3.7 6 N-methyl-tosylamide 11.7 4.8 7 benzoic acid 4.2 6.5 8 phenol 9.9 7.3 9 3,5-bis(CF3)phenol 8.0 13.9

Starting concentrations: c(37) = 0.150 mol⋅l−1, c(105) = 0.125 mol⋅l−1, c(53) = 0.0125 mol⋅l−1

and 0.172 mol⋅l−1 Brønsted acid in [D8]THF.

0,0

2,0

4,0

6,0

8,0

10,0

12,0

14,0

16,0

0,0 2,0 4,0 6,0 8,0 10,0 12,0 14,0 16,0 18,0

pKa

rela

tive

initi

al r

ate

malic acid

benzoic acid

3,5-bis(CF3)phenol

phenol

N-methyl-tosylamide

water

methanol

0,0

2,0

4,0

6,0

8,0

10,0

12,0

14,0

16,0

0,0 2,0 4,0 6,0 8,0 10,0 12,0 14,0 16,0 18,0

pKa

rela

tive

initi

al r

ate

malic acid

benzoic acid

3,5-bis(CF3)phenol

phenol

N-methyl-tosylamide

water

methanol

Figure 4. Influence of Brønsted acidic additives on the reaction rate.

Page 50: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

42

A comparison of the relative initial rate of the reaction with the aza-BH product 106 as

additive (entry 3) to the relative initial rate of the reaction without co-catalyst (entry 2)

confirmed that the reaction was not autocatalytic under the applied conditions. A maximum in

reaction rate was observed with 3,5-bis(CF3)phenol (107) (pKA ≈ 8, entry 9) as additive,

corresponding to a 14-fold rate enhancement as compared to the reaction without additive

(entry 2). Less acidic compounds showed a smaller effect (entry 4-6, 8). Additives, which

were more acidic than 3,5-bis(CF3)phenol (107), ultimately even inhibited the reaction (entry

1).

NMR studies showed that this was mainly due to a protonation of the first zwitterionic

reaction intermediate and to a minor extent to protonation of the nucleophilic catalyst itself

(Scheme 40).

O+

2 HX

[D8]THF, r.t.

O

3753

2 PPh3

Ph3PX

+ HPPh3+X−

108

109 110

Scheme 40. Protonation of the nucleophilic catalyst and the first zwitterionic reaction

intermediate.

Treatment of a 1:1-mixture of PPh3 (53) with MVK (37) in [D8]THF at r.t. with one

equivalent of malic acid resulted in a mixture of the corresponding 3-phosphonium enolate in

its protonated form (109) and the triphenylphosphonium salt 110. After 5 min all PPh3 was

converted. A comparison of the signal intensities in the 31P NMR spectrum showed that the

ratio 109/100 corresponded to a value of 95/5. These results are i.a. supported by a report

from Wasserscheid and co-workers. They described a similar reaction as side reaction during

the preformation of their metal-phosphine catalytic system for acrylate dimerisation.[132]

Additionally, the rate law of the reaction was determined in the presence of phenol

(111) as prototypical additive. The reaction was monitored by 19F NMR spectroscopy. The

dependence of the initial rate on the starting concentration of imine is depicted in Figure 5.

Page 51: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

43

r(0) = 0,0111*c(imine)R2 = 0,9906

0,000E+00

2,000E-04

4,000E-04

6,000E-04

8,000E-04

1,000E-03

1,200E-03

1,400E-03

1,600E-03

1,800E-03

2,000E-03

0,0000 0,0200 0,0400 0,0600 0,0800 0,1000 0,1200 0,1400 0,1600 0,1800

c(imine) in [mol/l]

r(0)

in [m

ol/(l

*min

)]

Figure 5. Initial rate as function of starting concentration of the imine 102 in the presence of

phenol as Brønsted acidic co-catalyst. Starting concentrations: c(53) = 0.0294 mol⋅l−1, c(37) =

0.1502 mol⋅l−1, c(111) = 0.1584 mol⋅l−1, c(104) = 0.0341 mol⋅l−1, c(102) = 0.0962 – 0.1479

mol⋅l−1.

For the other curves r(0) = f(c(x)) {x = MVK (37) and phenol (111)}, see Appendix 3.

A first order dependence in all components was observed, including the imine and Brønsted

acid (Equation 4).

)()()()('' acidcphosphinecacceptorMichaelcimineckr obs ⋅⋅−⋅⋅= Equation 4

These results are in line with the results obtained by Kaye in the so-called CLASSICAL MODEL

for the amine-catalyzed BH reaction of pyridinecarboxaldehydes with acrylate esters.[68] This

indicates, that the RDS of the aza-BH reaction catalyzed by the combined system Lewis Base-

Brønsted acid is the C-C coupling step, the Mannich reaction.

A comparison of these results to the results obtained with a Lewis base catalyst alone

(Equation 3), clearly indicates that the mechanism undergoes a change upon addition of the

Page 52: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

44

Brønsted acidic co-catalyst. A potential mechanistic interpretation of the change in the rate

law – the partial order in imine changed from 0.5 to 1 – involves a shift of the RDS from the

proton migration to the Mannich-type coupling reaction. Addition of a Brønsted acid

facilitates the proton transfer which ultimately becomes so fast that it is not rate limiting

anymore. This hypothesis is supported by the data, provided by Aggarwal and Lloyd-Jones in

their mechanistic study on the BH coupling.[71] A plausible transition state, which explains the

facilitation of the proton transfer, is shown in Scheme 41. Without Brønsted acid, the proton

migration step most likely proceeds either through a 4-membered cyclic TS or in an

intermolecular fashion. Addition of a suitable Brønsted acid (ROH) enables a facilitation of

the migration step via 6-membered cyclic TS.

H O

HN

OR

Tos

PPh3

112

F

Scheme 41. Postulated transition state for the Brønsted acid-mediated proton transfer.

Furthermore, the partial order of 1 in Brønsted acid indicates that the Brønsted acid also plays

a role in the Mannich type reaction, probably through activation of the imine. At

approximately one equivalent of Brønsted acid per imine, a start of saturation was observed

(see Appendix 3).

3.1.2. Reversibility of the Aza-Baylis-Hillman Reaction

To assess the reversibility of the aza-BH reaction, both cross-Mannich (Scheme 42) and

racemization experiments (Scheme 43) with the catalytic system PPh3 (53)/phenol (111) were

performed.

Page 53: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

45

PPh3, phenol

THF, r.t.

NH OTos

Br

53

F

N Tos

+111

F

NH OTos

Br

N Tos

+

106 105102 103

Scheme 42. Cross-Mannich experiments with the catalytic system PPh3 (53)/phenol (111).

Reaction conditions: 30.0 mg (0.0735 mmol, 1.00 eq) 106, 203.7 mg (0.735 mmol, 10.0 eq)

102, 19.3 mg (0.0735 mmol, 1.00 eq) 53, 34.6 mg (0.367 mmol, 5.00 eq) 111, 2 ml (2.7

ml⋅mmol−1) THF, r.t., n h.

Figure 6 shows a direct comparison of kinetic results that were obtained with imine 102 and

105, respectively, in the reaction with MVK (37). As catalytic system, the combined system

PPh3 (53)/phenol (111) was applied.

0

0,05

0,1

0,15

0,2

0,25

0 10 20 30 40 50 60 70

t / min

conv

ersi

on in

[%]

106103

Figure 6. Kinetic results obtained with imine 103 and 106, respectively. Starting

concentrations: c(37) = 0.15 mol⋅l−1, c(103/106) = 0.13 mol⋅l−1, c(53) = 0.014 mol⋅l−1,

c(111) = 0.15 mol⋅l−1in THF.

Page 54: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

46

These results showed that the reactivity of imine 102 was higher than the reactivity of 105.

Under turn-over conditions, however, the product of cross-Mannich reaction (103) was not

observed even after prolonged reaction times (t = 72 h).

A further probe for the reversibility of the reaction was the stability of the stereocenter

of the aza-BH product 106 under turn-over conditions (Scheme 43). With the catalytic system

PPh3 (53)/phenol (111), racemization of enantiomerically enriched 106 (ee = 88% (S)) was

observed under turn-over conditions (Figure 7). The absence of cross-Mannich products under

these conditions indicated that the racemization of the aza-BH product 106 was independent

from the catalytic cycle. Further experiments to elucidate the mechanism of racemization are

described below.

PPh3, phenol

THF, r.t.

NH OTos

Br

53 111

(S)-106

NH OTos

Br

(R)-106

Scheme 43. Racemization experiments with the catalytic system PPh3 (53)/phenol (111).

Reaction conditions: 38.1 mg (0.0933 mmol, 1.00 eq) 106, 24.5 mg (0.0933 mmol, 1.00 eq)

53, 26.3 mg (0.280 mmol, 3.00 eq) 111, 0.75 ml (8.0 ml⋅mmol−1) THF, r.t., n h.

0,0

2,0

4,0

6,0

8,0

10,0

12,0

0,00 5,00 10,00 15,00 20,00 25,00

t / h

enan

tiom

eric

rat

io S

/R

Figure 7. Racemization under turn-over conditions. Starting concentrations: c(106) = 0.125

mol⋅l−1, c(53) = 0.125 mol⋅l−1, c(111) = 0.373 mol⋅l−1 in THF.

Page 55: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

47

3.1.3. Racemization of the Aza-Baylis-Hillman Product

Firstly it was investigated whether the C-C, C-H or C-N bond at the stereogenic center was

broken during the racemization process. To obtain this information, the aza-BH product rac-

106 was treated with PPh3 (53) in [D4]methanol. The result of this experiment is shown in

Scheme 44.

PPh3

CD3OD, r.t., 5h

O

Br

53O

Br

NHTos NDTosH D

D D

106 [D4]106

Scheme 44. Deuteration of rac-106 with [D4]methanol. Reaction conditions: 20.4 mg (0.050

mmol, 1.00 eq) 106, 13.1 mg (0.l050 mmol, 1.00 eq) 53, 1.0 ml (20 ml⋅mmol−1)

[D4]methanol, r.t., 5 h.

As shown in Scheme 44, a proton-deuterium exchange was observed in several positions of

106, also at the stereogenic center. This indicated that the C-H bond at the stereogenic center

broke during the racemization process and thereby confirmed that racemization of the aza-BH

product was not caused by retro-Mannich reaction.

In a further set of experiments, the role of all components in the racemization was

investigated. The results are shown in Table 2. These experiments showed that racemization

was not caused by simple deprotonation at the stereogenic center (entry 6). Also the presence

of triphenylphosphine alone did not lead to product racemization (entry 1). Besides a

nucleophilic catalyst a Brønsted acidic additive (entry 3) and/or imine (entry 4, 5) was

necessary for racemization. With Shi’s phosphinyl-BINOL catalyst 88, however, no

racemization was observed under the same conditions, even though the catalytic system

contained all components identified in the model reaction. This indicated that the spatial

arrangement of a bifunctional catalytic system is of major importance for the prevention of

racemization. This provides important information for the efficient design of successful chiral

catalyst systems

Page 56: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

48

Table 2. Role of all reaction components in the racemization process.

entry system yes no

1 PPh3 x

2 phenol x

3 PPh3 + phenol x

4 PPh3 + imine x

5 PPh3 + phenol + imine x

6 KOH x

Reaction conditions: 40.0 mg (0.107 mmol, 1.00 eq) (S)-4-methyl-N-[2-methylene-1-(4-

nitrophenyl)-3-oxobutyl]-benzenesulfonamide (114), 28.0 mg (0.107 mmol, 1.00 eq) PPh3

(53), 30.2 mg (0.321 mmol, 3.00 eq) phenol (111), 16.3 mg (0.053 mmol, 0.50 eq) N-(4-nitro-

benzylidene)-4-methylbenzenesulfonamide (113), 0.85 ml THF, r.t., 24 h. The experiment

with KOH was performed in a biphasic system water/diethylether.

To assess the influence of the Brønsted acidic additive on the rate of racemization, a

screening of additives with pKA values of 4-16 was performed.[131] The influence of the

Brønsted acidic additive on the racemization rate was determined by single point

measurements of the enantiomeric ratio. The results are shown in Figure 8.

The results depicted in Figure 8 show a clear dependency of the rate of racemization

on the pKA value of the additive. Rapid and nearly complete racemization was observed with

phenol (pKA = 9.90) as additive. Weak Brønsted acids like water and methanol caused no

significant racemization on a relevant time scale. Stronger acidic additives resulted in a

slower rate of racemization ultimately leading to a non-significant racemization.

Page 57: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

49

0

2

4

6

8

10

12

no additive 3,6 (malic acid) 4,2 (benzoic acid) 9,9 (phenol) 15,4 (methanol) 15,5 (water)

Additive

Ena

ntio

mer

ic r

atio

S/R

Figure 8. Single point measurements for comparison of the influence of different Brønsted

acidic additives on the rate of racemization. Reaction conditions: 38.1 mg (0.0933 mmol, 1.00

eq) 106 (e.r. = 11.4 (S)), 211.1 mg (0.0933 mmol, 1.00 eq) PPh3 (53), 0.2799 mmol (3.00 eq)

Brønsted acid, 0.75 ml THF, r.t., t = 24 h.

A plausible model which explains these findings involves a racemization pathway in

which both the nucleophilic catalyst and the conjugated base of the Brønsted acid play a

crucial role (Scheme 45). All steps in this model are equilibrium reactions. When the

Brønsted acid is strong (e.g. malic acid, pKA = 3.6), the conjugated base is relatively weak. It

is easily formed but not able to deprotonate the stereogenic center. In case of very weak

Brønsted acids (e.g. methanol, water, pKA ≈ 15.5), a protonation of enolate 115/116 does not

take place and consequently the conjugated base, which is necessary for deprotonation at the

stereogenic center, is not formed. Only for a small spectrum of Brønsted acids of medium

strength, the concentration and strength of the conjugated base such as 117 is suitable to

deprotonate type-118 intermediates and hence cause racemization. This base-generation

hypothesis is supported by the mechanistic work of Bergman and Toste on the phosphine-

catalyzed hydration and hydroalkylation of activated olefins.[133]

Page 58: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

50

PPh3 +OTosHN

OH

O+

H OTosHN H

PPh3

OTosHN H

PPh3

H

− PPh3, phenol

OH NHTos

X X

XX

53 (S)-106 (X = Br)(S)-114 (X = NO2)

115 (X = Br)116 (X = NO2)

111

117118 (X = Br)119 (X = NO2)

(R)-106 (X = Br)(R)-114 (X = NO2)

53 111

Scheme 45. Potential racemization pathway involving phenol (111) /phenolate (117).

The stability of the stereocenter of the product upon treatment with successful bifunctional

catalysts like Shi’s phosphinyl-BINOL catalyst 88 indicates that the spatial arrangement of a

bifunctional catalytic system is of major importance for the prevention of racemization

(Scheme 46). A possible explanation for this phenomenon involves the intramolecular

coordination of the enolate-oxygen by the phenolic OH group of catalyst 88. This may form a

stabile adduct like 120/121 and thereby prevent the formation of the conjugated phenolate

base which is postulated as a key intermediate in the racemization of the aza-BH product.

PPh2 +OTosHN H

O

TosHN H

PX X

88 (S)-106 (X = Br)(S)-114 (X = NO2)

120(X = Br)121 (X = NO2)

HO

*

OH

*

Scheme 46. Model for the prevention of racemization with Shi’s catalyst 88 (compound 88 is

represented as HO^PPh2).

In absence of a Brønsted acid, an imine-mediated racemization takes place (Table 2,

entry 4). A possible mechanism, which explains this racemization pathway, might involve a

Page 59: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

51

hemi-aminal intermediate. Similar intermediates have been suggested by McQuade and co-

workers in the so-called HEMI-ACETAL MECHANISM for the BH reaction. Furthermore, the

occurrence of cyclic products in the aza-BH reaction was explained by the formation of hemi-

aminals (page 25). There is, however, not enough direct experimental evidence available to

unequivocally identify such kind of mechanism at the present stage.

3.1.4. Chirality Transfer

Representative for the chirality transfer models that are reported in literature is the hypothesis

by Shi and co-workers involving catalyst 88.[48b] This model is depicted in Scheme 47.

OH

PPh2

OO

PPh2

OH

ArCH=NTosArCH=NTos

O

PPh2

HN

H

ArHTos

OO

PPh2

HN

H

HArTos

O

Ar

NH OTos

Ar

NH OTos

88 37 122

125124

123 123

(S)-126 (R)-126

P

O HN

ArH

Tos

O

PhPh

P

O H**

N

HAr

Tos

O

PhPh

124' 125'

+

S RRR

S R

R R

Scheme 47. Model by Shi and co-workers for the chirality transfer in the aza-BH reaction.

Page 60: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

52

They postulated that compound 88 acts as a bifunctional chiral catalyst in the aza-BH

reaction. The phosphine atom acts as a Lewis base and the phenolic OH group as a Brønsted

acid. The Lewis basic moiety initiates the aza-BH reaction whereas the Brønsted acid

stabilizes both intermediates of the catalytic cycle (122 and 124/125) through hydrogen

bonding to the enolate-oxygen and imine-nitrogen, respectively. During the Mannich-type

coupling step, 4 diastereoisomers can in principle be formed. In Shi’s model, only the (R,S,R)-

and the (R,R,R)-diastereoisomers 124 and 125 are considered. Because of the steric

repulsions between the C(O)Me group with the aromatic group and the aromatic group with

the two phenyl groups at phosphorus in intermediate 125, the authors suggest that the

activation energy for the pathway through 125’ is higher than through 124’. This kinetic

selection implements the necessity of a reversible Mannich-type coupling. In accordance with

the model reported by Aggarwal and Lloyd-Jones, the reaction would “filter” through the

elimination pathway.[71]

Shi’s model, however, comprehends several ambiguities. The first ambiguity involves

the consideration of only two of four diastereoisomeric intermediates that may result from the

Mannich-type coupling of enolate 122 and imine 123. This selection is not motivated in their

report. Therefore, the relative energies of all four diastereoisomers that may result from the

Mannich type coupling of N-benzylidene-4-methylbenzenesulfonamide (67) with enolate 122

were determined in a computational study using DFT calculations in the present work.[134],[135]

The results are shown in Figure 9.

The (R,S,R)- and (R,R,S)-diasteroisomers were significantly higher in energy (6.2 and

8.3 kcal⋅mol−1, respectively) relative to the lowest energy conformer of the (R,S,S)-form. The

energy of the (R,R,R)-diastereoisomer, however, was only 1.2 kcal⋅mol−1 above the energy of

the (R,S,S)-form. Accordingly we re-optimized the structures of the (R,S,S)- and (R,R,R)-

diastereoisomers using the larger TZVP basis set by Ahlrichs and co-workers.[136a] For these

optimizations (Turbomole 5.7)[136b] the BP86[136c,d] functional was used (Turbomole 5.7 was

used simply for efficiency considerations). Again the energy difference between (R,S,S)- and

(R,R,R)-form was small (0.5 kcal⋅mol−1), now with the (R,R,R)-diastereoisomer being the

more stable compound.

Page 61: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

53

(R,S,S) 0.0 (R,R,R) 1.2

(R,S,R) 6.2 (R,R,S) 8.3

(R,S,S) 0.0 (R,R,R) 1.2

(R,S,R) 6.2 (R,R,S) 8.3

Figure 9. Calculated structures of all four diastereoisomers (relative energies in kcal⋅mol−1).

Following Shi’s explanation, in which the stability of the intermediates correlates directly

with the relative energies of the transition states that lead to elimination of the aza-BH

product, the (R)-product should be formed in a fairly low enantiomeric excess of about

48%.[137] The observed enantioselection for this reaction, however, corresponded to an

enantiomeric excess of 83% (S) at −30°C.[48b]

A further requirement for the validity of Shi’s model is the reversibility of the

Mannich-type coupling. For the model system PPh3 (53)/phenol (111), no products from

cross-Mannich reactions were observed under turn-over conditions. Therefore, it is unlikely

that the Mannich coupling is reversible with Shi’s catalyst 88. Consequently, a model in

which the enantioselection is fully determined in the Mannich type coupling step seems a

more plausible alternative.

Furthermore, it is noteworthy that the phenolic OH group of catalyst 88 tends to

coordinate to the SO oxygen rather than to the imine nitrogen in the calculated structures

Page 62: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

54

depicted in Figure 9. This illustrates that the activating group on the imine might play a vital

role in the chirality transfer from the catalyst to the aza-BH product.

3.1.5. Implications for Asymmetric Catalysis

All successful chiral catalysts to date contain two catalytic moieties: a Lewis basic and a

Brønsted acidic one. The Brønsted acidic co-catalyst is necessary to obtain an effective

system with regard to the catalyst’s activity and selectivity. In contrast to the results obtained

by Aggarwal and Lloyd-Jones for the BH reaction, there is no evidence for autocatalysis.

Competitive catalysis by the reaction product can therefore be ruled out. The work of Sasai

and Shi shows that a suitable positioning of both catalytically active groups is necessary to

obtain a high enantiomeric excess. The Mannich-type coupling can lead to four

diastereoisomers. There are two scenarios to explain the enantioselectivity of the reaction:

(1) All four diastereoisomers are formed, and because of the spatial arrangement only one

undergoes a fast proton transfer. The others then revert back to the starting materials and

the proton transfer serves as a selective “filter” for the reaction.

(2) A second pathway involves the selective formation of one (or two) diastereoisomers

leading directly to the product with the desired configuration.

The first route is dependent on the reversibility of the C-C coupling step – the Mannich

reaction. In the cross-Mannich experiments performed in this mechanistic study no cross-

Mannich products were observed which makes this explanation very unlikely.

Considering the limitations of the model, in which the proton migration step

determines the enantioselectivity of the reaction, it seems more likely that the enantioselection

takes place in the C-C coupling step. This, however, could not yet be validated in

experimental or theoretical studies.

A potential reason for the small success rate in the design of chiral catalysts for the

aza-BH reaction may lie in the lability of the product’s stereocenter. The results obtained with

Shi’s catalyst 88 in comparison to the model system triphenylphosphine (53)/phenol (111)

show that the spatial arrangement of a bifunctional catalyst is not only important for an

efficient transfer of the chiral information but also for the prevention of racemization.

Page 63: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

55

3.2. Design of New Catalytic Systems for the Asymmetric Aza-Baylis-Hillman Reaction

3.2.1. QUINAP Derivatives

QUINAP (127), a powerful and versatile ligand for asymmetric transition-metal catalysis, was

developed by Brown and co-workers in 1993.[138] This axially chiral aminophosphine has

been successfully applied as P,N-ligand in Rh-catalyzed hydroboration and

hydroboration/amination reactions [139],[140] and Pd-catalyzed allylic substitutions.[141]

Since the binaphthyl backbone has proven to be a valuable lead structure for chiral

bifunctional organocatalysts for the aza-BH reaction,[48] the use of QUINAP (127) and

QUINAP-type derivatives seemed interesting. The structures depicted in Scheme 48 − (R)-

QUINAP (127), rac-1-(2-diphenylphosphanylnaphthalen-1-yl)isoquinoline-3-carboxylic acid

(128) and 1-(2-diphenylphosphanylnaphthalen-1-yl)isoquinoline-3-carboxylic acid (3-

hydroxy-2-phenylpropyl)amide (129) − were tested as catalysts in the aza-BH reaction of the

aromatic aldimines 67 (unsubstituted), 102 (m-F) and 113 (p-NO2) with MVK (37), methyl

acrylate (42) and cyclopentenone (130) as Michael acceptors (Scheme 49).

N

PPh2

127

N

COOH

PPh2

128

N

PPh2

O

HN OH

Ph

129

Scheme 48. QUINAP-derivatives as catalysts for the aza-BH reaction.

A starting concentration of 0.125 mol imine per l THF was applied. The reaction was

performed on a 2 ml scale. Conversion (conv.) was determined via 1H- and/or 19F NMR

spectroscopy, the enantiomeric excess of the reaction product was analyzed via chiral HPLC

(only in case of conv. > 15%).

Page 64: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

56

N Tos

X EWGEWG

NHTos

X+

catalyst (10 mol-%)

THF, r.t.

127, 128, 129

67 (X = H) 102 (X = m-F)113(X = p-NO2)

37 (EWG = C(O)CH3)42 (EWG = C(O)OCH3)130 (cyclopentenone)

Scheme 49. Aza-BH reaction with QUINAP-derivatives.

In a first set of experiments, the scope of the catalysts towards different Michael acceptors

was tested. The results are depicted in Table 3.

Table 3. Screening of different Michael acceptors.

entry imine alkene catalyst catalyst [mol-%]

alkene [eq]

t / h conv. [%]

1 113 42 128 10 1.2 24 0 2 113 42 129 10 1.2 24 0 3 113 130 128 10 1.2 24 0 4 113 130 129 10 1.2 24 0 5 113 37 128 10 1.2 24 11 6 113 37 129 10 1.2 24 7

Reaction conditions: 38.0 mg (0.125 mmol, 1.00 eq) imine 113, 0.150 mmol (1.20 eq) alkene,

0.0125 mmol (0.10 eq) catalyst, 2.0 ml (16.0 ml⋅mmol−1 113) THF, r.t., t = 24 h.

These results showed that the QUINAP-derivatives 128 and 129 displayed a low activity in

comparison to similar bifunctional catalysts like 88.[48b] Furthermore, the scope of these

catalysts towards the activated alkenes was rather limited. The only system that displayed any

activity was obtained with MVK as Michael acceptor.

To verify whether the low activity of 128 and 129 in the aza-BH reaction was intrinsic

or due to catalyst deactivation, the reaction time was prolonged. The results for the coupling

of imine 102 with the activated alkenes 37, 42 and 130 are shown in Table 4.

Page 65: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

57

Table 4. Coupling of imine 102 with the activated alkenes MVK (37), methyl acrylate (42)

and cyclopentenone (130), catalyzed by 129.

entry imine alkene catalyst catalyst [mol-%]

alkene [eq]

t / h conv. [%]

ee [%]

1 102 37 129 10 1.2 24 10 n.d. 2 102 37 129 10 1.2 67.5 24 6.0 3 102 37 129 20 5.0 24 50 6.0 4 102 42 129 20 5.0 24 0 n.d. 5 102 130 129 20 5.0 24 0 n.d.

Reaction conditions: 69.4 mg (0.250 mmol, 1.00 eq) imine 102, 2.0 ml (8.0 ml⋅mmol−1 102)

THF, r.t.

These results show, that the catalyst was not completely deactivated after 24 h in the coupling

of imine 102 with MVK (37) (entry 1, 2). Furthermore, a reasonable yield within 24 h could

be achieved by a 4-fold increase of the amount of MVK (37) and a 2-fold increase of the

amount of catalyst 129 (entry 3). Under these reaction conditions, however, catalyst 129 still

showed no activity in the coupling of imine 102 with methyl acrylate (42) and cyclopentenone

(130) (entry 4, 5). The enantioselectivity in the aza-BH reaction of imine 102 with MVK (37)

was very poor and independent from the reaction time, the amount of catalyst, and Michael

acceptor (entry 2, 3).

To determine the relative activity and selectivity of all three catalysts, they were tested

in the coupling of the imines 67 and 102 with MVK (37) under the reaction conditions,

displayed in entry 3 of Table 4 (20 mol-% catalyst, 5.0 eq MVK). The results are shown in

Table 5.

These results showed that the activity of all three catalysts was rather low in

comparison to similar bifunctional catalysts reported in literature.[48b] The best results with

respect to catalytic activity were obtained with catalyst 128, the QUINAP-derivative with

carboxylic acid group (128 > 129 > 127). The enantioselectivity that was obtained with the

enantiopure catalysts 127 and 129 was very poor.

Page 66: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

58

Table 5. Coupling of the imines 67 and 102 with MVK (37).

entry imine catalyst Conv. [%] ee [%] 1 67 127 12 rac 2 67 128 47 - 3 67 129 34 rac 4 102 127 43 rac 5 102 128 94 - 6 102 129 50 6.0

Reaction conditions: 0.250 mmol (1.00 eq) imine, 104.1 μl (1.250 mmol, 5.00 eq) MVK,

0.050 mmol (0.20 eq) catalyst, 2.0 ml (8.0 ml⋅mmol−1 102) THF, r.t., 24 h.

QUINAP does not contain an acidic moiety which can serve as co-catalyst for the aza-

BH reaction. Therefore, its low activity and enantioselectivity was not unexpected. By

protonation of the quinoline-nitrogen, however, one would create a side group with a pKA

similar to the pKA of phenol.[132] This would result in a catalyst, which is structurally related to

Shi’s catalyst 88 and contains an acidic group with a pKA value similar to that of the phenolic-

OH group in 88. Selective N-protonation of QUINAP (127) requires a Brønsted acid of

medium strength. Strong Brønsted acids would deactivate the catalyst by protonation of the

phosphino-group. Phenol (111) was chosen as Brønsted acid to generate the protonated

QUINAP derivative in situ. The results are shown in Table 6.

Table 6. Catalysis results with N-protonated QUINAP.

entry imine catalyst Conv. [%] ee [%] 1 67 127 12 rac 2 67 127/111 35 rac 3 102 127 43 rac 4 102 127/111 88 rac

Reaction conditions: 0.250 mmol (1.00 eq) imine, 104.1 μl (1.250 mmol, 5.00 eq) MVK,

0.050 mmol catalyst (0.20 eq), 2 ml THF (8.0 ml⋅mmol−1 imine), r.t., 24 h.

These results show that the activity of the catalyst increased significantly through addition of

phenol (111) (entry 2, 4). The enantioselectivity, however, remained unchanged.

Page 67: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

59

Although the activity and selectivity of these catalysts in the asymmetric aza-BH reaction was

rather low, catalyst 128 could be used to intercept and characterize the intermediates of the

phosphine-catalyzed aza-BH reaction for the first time (Scheme 50).

N

COO

P

O

Ph Ph

N

COOH

P

O

Ph PhPh

NH

Tos

131 132

Scheme 50. The intermediates of the aza-BH reaction of imine 67 with MVK (37), catalyzed

by QUINAP-derivative 128, in their protonated forms.

The first intermediate 133 results from a Michael type addition of the phosphine

catalyst 128 to MVK (37). It could be intercepted in its protonated form (131) by addition of

one equivalent of MVK (37) to a solution of catalyst 128 in CDCl3. The solution was stirred

for 5 min at room temperature (Scheme 51).

N

COOH

P

O

Ph Ph

133

N

COOH

PPh2+ O

CDCl3, r.t., 5 min N

CO2−

P

O

Ph Ph

13137128

Scheme 51. Interception of intermediate 133 in its protonated form (131).

The protonated intermediate 131 was identified via multinuclear NMR- and high-resolution

electrospray ionization mass spectroscopy (HR-ESI-MS). The 1H- and 31P NMR spectra are

shown in Figure 10 and 11, respectively.

In the 1H NMR spectrum, the signals which correspond to the PCH2CH2 moiety were

positioned between 3 and 5 ppm. Due to the chirality of catalyst 128 this moiety consisted of

four protons in a 1:1:1:1 ratio.

Page 68: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

60

2.02.53.03.54.04.55.05.56.06.57.07.58.08.59.09.510.0 ppm

N

CO2−

P

O

Ph Ph

131

2.02.53.03.54.04.55.05.56.06.57.07.58.08.59.09.510.0 ppm

N

CO2−

P

O

Ph Ph

131

Figure 10. 1H NMR spectrum of the reaction mixture containing intermediate 131.

-10-530 25 20 15 10 5 0 ppm

N

CO2−

P

O

Ph Ph

131

N

COOH

PPh2

128

134

-10-530 25 20 15 10 5 0 ppm

N

CO2−

P

O

Ph Ph

131

N

COOH

PPh2

128

134

Figure 11. 31P NMR spectrum of the reaction mixture containing intermediate 131.

Page 69: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

61

The 31P signal of catalyst 128 at −13.7 ppm disappeared nearly completely. This

demonstrated that catalyst 128 was almost completely converted to intermediate 131 (δ = 26.9

ppm) under these reaction conditions. The remaining signal at 30.3 ppm corresponds to the

phosphine oxide (134).

The second intermediate (135), which results from a Mannich-type coupling of enolate

133 with imine 67 could in principle be obtained both from the side of the starting materials

and the reaction product. In systems with Brønsted acidic co-catalysts, however, the Mannich-

type coupling is the RDS. This implied that intermediate 135 would be formed after the RDS

and therefore most likely be difficult to intercept. Thus, starting from the reaction product in a

Michael-type reaction seemed more promising (Scheme 52).

N

COOH

PPh2

128

ONHTos

N

COOH

P

O

Ph PhPh

NH

Tos

132

+

N

COOH

P

O

Ph PhPh

NH

Tos

135

H+

77

Scheme 52. Interception of intermediate 135 in its protonated form (132).

For this reaction diphenylacetic acid was used as an additional proton source. The resulting

reaction mixture contained only a small amount of 132 (about 13% of the P-containing

species in solution). A characterization of 132, however, was possible via multinuclear NMR

and HR-ESI-MS. The (1H, 31P)-correlation spectrum, the 1H spectrum and the 31P spectrum

are shown in Figure 12, 13 and 14, respectively.

Page 70: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

62

ppm

3.03.54.04.55.05.56.06.57.07.5 ppm

19.5

20.0

20.5

21.0

21.5

22.0

22.5

23.0

23.5

24.0

132

1H

31P

ppm

3.03.54.04.55.05.56.06.57.07.5 ppm

19.5

20.0

20.5

21.0

21.5

22.0

22.5

23.0

23.5

24.0

132

1H

31P

Figure 12. (1H, 31P)-correlation spectrum of the reaction mixture with 132.

x

xxo

x

xx

x

xxo

N

COOH

P

O

Ph PhPh

NH

Tos

132

x

xxo

x

xx

x

xxo

N

COOH

P

O

Ph PhPh

NH

Tos

132

Figure 13. 1H NMR spectrum of the reaction mixture with 132 (x = from (1H,31P)-correlation

spectrum, o = from (1H,1H)-COSY and TOCSY).

The signals marked with “x” were unambiguously assigned to the aliphatic protons of

intermediate 132 in α- and β-position to the P-atom via (1H, 31P)-correlation. The fourth

signal, marked with “o”, was assigned to the proton in γ-position through (1H, 1H)-correlation

and TOCSY.

Page 71: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

63

-10-530 25 20 15 10 5 0 ppm

N

COOH

PPh2

128

N

COOH

P

O

Ph PhPh

NH

Tos

132

134

-10-530 25 20 15 10 5 0 ppm

N

COOH

PPh2

128

N

COOH

P

O

Ph PhPh

NH

Tos

132

134

Figure 14. 31P NMR spectrum of the reaction mixture with 132.

This was the first time that both intermediates of the phosphine-catalyzed aza-BH reaction

were intercepted and structurally characterized. The first intermediate (type-131) was also

obtained for various other phosphine catalysts in combination with MVK (37) and an external

acid source (Table 7). Conv. (31P) is the conversion of the starting phosphine to the

corresponding adduct (type-131), derived from the 31P NMR spectra.

Table 7. Intercepted first intermediate of the aza-BH reaction.

entry phosphine acid acid:alkene δ 31P (phosph./add.) [ppm]

Conv. (31P) [%]

1 53 malic acid 1.0 -5.4 / 25.5 95.0 2 127 diphenyl acetic acid 1.0 -14.1 / 26.4 52.6 3 128 - 0.0 -13.7 /26.9 64.3 4 129 diphenyl acetic acid 1.0 -13.4 /26.0 32.3

Reaction conditions: 1.00 eq phosphine, 1.00 eq MVK, 1.00 eq diphenylacetic acid, NMR

solvent, r.t., 5 h.

Page 72: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

64

3.2.2. Other phosphinocarboxylic acids

As a model system to verify the catalytic potential of phosphinocarboxylic acids with various

distances between both functionalities in the aza-BH reaction, 2-, 3- and 4-

diphenylphosphanyl benzoic acid were tested as catalysts (Scheme 53).

COOH

Ph2P

COOH

PPh2

COOH

PPh2

PPh3COOH

+

136 137 138 53 139

0% 70% 38% 86%

Scheme 53. Diphenylphosphanyl benzoic acids as catalysts for the aza-BH reaction of imine

102 (m-F) with MVK (37). Reaction conditions: 69.33 mg (0.250 mmol, 1.00 eq) imine 102,

7.66 mg (0.0250 mmol, 0.10 eq) diphenylphosphanyl benzoic acid, 25.0 μl (0.300 mmol, 1.20

eq) MVK, 2.0 ml (8.0 ml⋅mmol−1) THF, r.t., 24 h.

These results clearly indicated that the activity of the catalyst was dependent on its geometry:

compound 136, in which both catalytic moieties are positioned ortho to each other, was

inactive. The “meta-catalyst” 137 was the most active compound albeit its catalytic activity is

still lower than the activity of the 1:1 system PPh3 (53)/benzoic acid (139). This results in

following trend in activity: 53/139 > 137 > 138 > 136.

As chiral phosphinocarboxylic acid, (R,S)-2-diphenylphosphanyl-cyclohexane-

carboxylic acid (140) was tested as catalyst for the coupling of imine 105 with MVK (37)

(Scheme 54).

Br

N TosO

+

Br

NH OTos

COOH

PPh2

(10 mol-%)

r.t., THF, 24 h

105 37

140

106

86%rac

Scheme 54. Coupling of imine 105 with MVK (37), catalyzed by phosphinocarboxylic acid

140. Reaction conditions: 51.0 mg (0.125 mmol, 1.00 eq) imine 105, 12.5 μl (0.150 mmol,

1.20 eq) MVK, 3.90 mg (0.0125 mmol, 0.10 eq) catalyst 140, 1.0 ml (8.0 ml⋅mmol−1) THF,

r.t., 24 h.

Page 73: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

65

Although the activity of the system is higher than the activity of Shi’s catalyst 88 under the

same conditions (60% yield), the product is obtained as racemate.[48b] Racemization of

enantiomerically enriched aza-BH product 106 under the influence of catalyst 140 was

excluded in the corresponding control experiment. Consequently, the reason for the formation

of racemic product was the low intrinsic enantioselectivity of catalyst 140.

3.2.3. Chiral Brønsted Acids

The use of a combined system Lewis base/chiral Brønsted acid proved to be a valuable

strategy for obtaining high enantioselectivities in the BH- and aza-BH reaction (cf. Jacobsen’s

urea-catalysts, Schaus’ BINOL system), although the risk of subsequent racemization of the

aza-BH product is high as compared to catalysts that combine both functionalities suitably

positioned in one chiral compound.[38], [40]

As Brønsted acids, chiral alcohols were tested as co-catalysts in the PPh3-catalyzed

coupling of imine 105 (p-Br) with MVK (37) (Scheme 55). In all experiments, the

enantiomeric excess of the corresponding reaction product was very poor. The best result (ee

= 11%) was obtained with (R)-BINOL as additive.

OHOH

OHOHO

OO

O

OHHO O O

HO OHH H

141 142 143 144

97%11% ee

40%rac

31%rac

66%6% ee

Scheme 55. Chiral co-catalysts for the PPh3-catalyzed aza-BH coupling of imine 105 with

MVK (37). Reaction conditions: 51.0 mg (0.125 mol, 1.00 eq) imine 105, 12.5 μl (0.150

mmol, 1.20 eq) MVK, 3.28 mg (0.0125 mmol, 0.10 eq) PPh3, 0.0375 mmol (0.30 eq)

Brønsted acid, 1.0 ml (8.0 ml⋅mmol−1 105) THF, r.t., 24 h.

Page 74: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

66

3.2.4. Chiral Aminoalcohols

3.2.4.1. Synthesis In accordance with previous reports in literature (page 29) [38], [88], [114]-[116] and the mechanistic

investigations described above, aminoalcohols of type-145 are potential bifunctional

asymmetric catalysts for the aza-BH reaction. Besides a Lewis basic tertiary amino group,

they contain an alcohol moiety as co-catalyst. Kinetic- and racemization studies with weak

acids like alcohols in the phosphine-catalyzed aza-BH reaction showed that they are able to

serve as co-catalyst – methanol causes rate acceleration by a factor of 3.7 – but that their

acidic strength is insufficient to induce racemization of the reaction product. An efficient

access to this class of aminoalcohols must be straightforward and substitution diversity should

be possible without major changes of the synthetic route. One flexible strategy to obtain this

type of aminoalcohols is shown in Scheme 56.

This strategy involves an ex-chiral-pool synthesis starting from the chiral building

block hydroxyproline. In step A the acid group would be protected by esterification, followed

by protection of the aminogroup of aminoester 152 in step B. For step F it is required that the

aminogroup carries an electron donating group. This should be considered in the choice of an

appropriate protecting group R’. In step C, the OH group of 151 should be protected, followed

by a reduction of the ester moiety in step D. The OH group of the corresponding vic-

aminoalcohol 149 would then be transformed into a leaving group (e.g. via mesylation). In an

intramolecular SN2 reaction followed by a ring opening of the corresponding aziridinium-ion

147, compound 146 would be formed as a product of ring expansion. Deprotection of the OH

group would then lead to the desired class of aminoalcohols 145.

Page 75: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

67

NR'

HO X* *

NR'

O X* *PG

NR'

OH

OPG*

*

NR'

OR

OPG*

*NR'

OR

HO*

*

ONH

OR

HO*

*

O

B C

D

FGH

145 146 147 148

150151152

NH

OH

HO*

*

O

153

A

O

NR'

LG

OPG

* *

E

149

NR'

OPG ** LG

Scheme 56. Strategy for the synthesis of aminoalcohol 145.

For the synthesis, trans-4-hydroxy-L-proline (154) was used as starting material. In the first

step, compound 154 was converted into the corresponding methyl ester. This was isolated in

form of its hydrochloride salt 155.

NH

OH

HO

ONH2

O

HO

O Cl−

154 155

SOCl2

CH3OH, 0°C - r.t., 24 h

> 99%

Scheme 57. Synthesis of the methylester 155.

According to a procedure reported by Enders and co-workers, 155 was prepared by addition

of trans-4-hydroxy-L-proline to a solution of thionylchloride in methanol at 0°C.[142] After

warming to room temperature, the mixture was stirred for 24 h. Then, the excess SOCl2 and

methanol was removed in vacuo. Subsequently compound 155 was crystallized from a

mixture of methanol and diethylether and isolated in quantitative yield.

The second step consisted of protection of the amino group of 155 via benzoylation

(Scheme 58).[142] 155 was dissolved in dichlormethane at room temperature and treated with a

Page 76: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

68

large excess of triethylamine. After stirring the mixture for 1.5 h at room temperature, DMAP

and benzoylchloride were added. The resulting reaction mixture was stirred overnight.

Precipitation from dichloromethane/pentane yielded compound 156 as a colorless precipitate

in 80% yield.

NH2

O

HO

O Cl-NEt3, DMAP, PhC(O)Cl

DCM, r.t., 14 hN

O

HO

OO

155 15680%

Scheme 58. Benzoyl-protection of the amino group in 155.

In the third step, 156 was converted into the corresponding tert-butyldimethylsilyl

(TBS) ether 157 (Scheme 59).[142]

TBSCl, imidazol

DMF / toluene, r.t., 43 hN

O

HO

OO

156

NO

O

OO

157

Si

> 99%

Scheme 59. Conversion of 156 into its corresponding TBS-ether 157.

Amide 156 and imidazole were dissolved in DMF and stirred at room temperature for 3 h.

Then, a solution of tert-butyldimethylsilyl chloride (TBSCl) in toluene was added dropwise

and the resulting mixture was stirred at room temperature for 40 h. After removal of the

volatiles in vacuo, the residue was taken up in ether and underwent a standard organic

workup. The resulting organic residue was recrystallized from dichloromethane/pentane.

Compound 157 was obtained in quantitative yield in form of colorless crystals.

In the next step, the ester moiety was reduced to the corresponding primary alcohol

(Scheme 60). Under the employed conditions, the benzoyl group was reduced as well.[142]

Page 77: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

69

LAH

THF, rfl, 1.5 hN

O

O

OO

157

Si

NOH

O

158

Si

> 99%

Scheme 60. Reduction of 157 with lithium aluminiumhydride (LAH).

To a suspension of LAH in THF, a solution of ester 157 in THF was carefully added. The

resulting mixture was heated under reflux for approximately 1.5 h. After hydrolysis with 20-

% aqueous KOH solution and standard organic workup, the compound was obtained as

colorless oil in quantitative yield. Aminoalcohol 158 was used without further purification.

The vic-aminoalcohol 158 was treated with methanesulfonyl chloride (MsCl) to

convert the OH group to an appropriate leaving group for nucleophilic substitution (Scheme

61).

MsCl, NEt3

THF, 0°C - rfl, 12 hNOH

O

158

Si

NOMs

OSi

− HNEt3MsO

N

OSi

Cl

N

ClOSi

15957%

+ HNEt3Cl

Scheme 61. Reaction of 158 with MsCl.

MsCl and triethylamine were added to a solution of aminoalcohol 158 in THF at 0°C. After

0.5 h, the reaction mixture was heated to reflux for 12 h. Then, the suspension was cooled to

room temperature and poured into an aqueous 2.5 M-NaOH solution. After standard organic

Page 78: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

70

workup, the resulting oil was purified via column chromatography and isolated as yellow oil

in 57% yield. The NMR spectra of the resulting compound provide proof for the formation of

one single diastereoisomer. The absolute configuration was assigned via NOESY (Figure 15).

The NOESY spectrum was recorded at −40°C in CDCl3 and displayed a clear NOE contact

between the Si-CH3 group and the proton in α-position to the Cl-atom. This, in combination

with the fixed R-configuration for the stereocenter adjacent to the silyl group, unambiguously

proved the (R,R)-configuration of compound 159.

NMR-, GC- and GC-MS analysis proved that the intermediate aziridinium ion was

solely attacked by chloride anions. The corresponding amino mesylate (159, Cl = OMs) could

not be detected. Similar results were obtained by the group of Cossy.[143] They postulated the

involvement of a tight ion pair to explain this chemoselectivity.

In the last step, the TBS-ether was cleaved according to standard procedure with tetra-

n-butyl ammonium fluoride (TBAF) in THF (Scheme 62).[144]

TBAF

DCM, r.t., 14 hN

ClOSi

159

N

ClHO

16081%

Scheme 62. Cleavage of the TBS-ether 159.

TBAF was added to a solution of TBS-ether 159 in dichloromethane. The mixture was stirred

at room temperature overnight. After standard organic workup, the resulting colourless solid

was purified via column chromatography. Aminoalcohol 160 was isolated as a colourless

solid in 81% yield.

Page 79: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

ppm

4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 ppm

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

N

H

H

O

HH

HH

Cl

Ph

H

H

Si

ppm

4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 ppm

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

N

H

H

O

HH

HH

Cl

Ph

H

H

Si

Figure 15. NOESY spectrum of compound 159 in CDCl3 at −40°C.

Page 80: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

In conclusion, vic-aminoalcohol 160 could be synthesized in a 6 step-synthesis with an

overall yield of 37%. In principle, all four diastereoisomers of this compound can be

selectively produced starting from hydroxyproline precursors with different stereochemistry.

3.2.4.2. Catalysis

Both the TBS protected vic-aminoalcohol 159 and the deprotected aminoalcohol 160 were

tested as catalysts for the aza-BH coupling of imine 113 (p-NO2) with MVK (37) in THF at

room temperature. With a catalyst loading of 10 mol-% and a ratio of MVK : imine = 1.2 : 1,

no conversion was observed within 24 h. Several parameters were varied to increase the

system’s reactivity: the catalyst loading was increased up to a stoichiometric amount (100

mol-%), the ratio MVK : imine was varied in the range of 1.2 - 5.0, the temperature was

increased up to 60°C and the substitution pattern (X = p-NO2, m-F, p-Br, p-Me) of the imine

was varied. In all cases, no conversion could be detected. A possible reason for this

unexpected result could be the nucleophilicity of the amino group. In case of non-sufficient

nucleophilicity, the catalyst cannot initiate the Michael addition as first step of the catalytic

cycle and remains unactive.

A possible strategy to overcome this problem involves the introduction of different

protecting groups at the Nitrogen atom. It was, however, decided to focuss on a different

strategy to create a new enantioselective system for the aza-BH reaction – the design and

application of a chiral reaction medium. In future investigations, vic-aminoalcohol 160 as well

as its precursor 159 may serve as chiral backbone chiral for the formation of catalysts for

other organocatalytic processes as well as ligands for homogeneous transition metal catalysis.

3.2.5. Chiral Solvents

ILs, often referred to as DESIGNER SOLVENTS, offer the possibility of adapting their solvent

properties to a given chemical need through structural variations of one or both ions.[145] The

investigations on the mechanism of the aza-BH reaction provided useful information for the

design of a chiral solvent for this process. It was shown that a bifunctional interaction of the

catalyst with the zwitterionic reaction intermediates (transition states) is beneficial for

obtaining high enantioselectivity and preventing subsequent racemization of the reaction

product. In conventional systems, this can be achieved by incorporating both the nucleophilic

catalyst and the Brønsted acidic co-catalyst in one chiral molecule. ILs, however, might offer

Page 81: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

73

the possibility of obtaining a bidentate interaction by incorporating the acidic center in the

chiral solvent, allowing simple monofunctional achiral nucleophiles to be used as catalysts in

the enantioselective aza-BH reaction (Scheme 63). A long and complicated synthesis of a

suitable bidentate catalyst can then be avoided.

PRRR

OHO

R'

161

Scheme 63. Illustration for a possible bifunctional interaction of the first zwitterionic 3-

phosphoniumenolate intermediate of the aza-BH reaction with the anion of an anion-chiral IL

containing a hydrogen bond donor.

Apart from the structural requirements elucidated above, the chiral reaction

medium has to meet some general stipulations. It should be cheap, easy to prepare and stable

under reaction conditions. Furthermore, the chiral building block should preferably be

available in both enantiomeric forms to obtain a selective access to both optical antipodes of

the desired product. These considerations led to the application of an anion-chiral IL derived

from malic acid (162). In a straight-forward ex-chiral-pool synthesis developed by Rolf

Gausepohl, the desired reaction medium was obtained as a colorless room-temperature IL

(Scheme 64).[146]

The first step involved the preparation of the corresponding sodium dimalatoborate

salt (165) by stirring an aqueous solution of boric acid (164), sodium hydroxide (163) and L-

(−)-malic acid (162) in an open flask at 100°C. After complete evaporation (within

approximately 4 h), sodium dimalatoborate (165) was isolated as a colourless salt in

quantitative yield. The dimalatoborate anion contained exclusively five-membered rings and

consisted of a nearly 1:1-mixture of diastereoisomers. This resulted from the fixed chirality at

the asymmetric carbon atoms and the labile chirality at the boron center of this spiro-type

compound. Configurational lability of orthoborates has been described earlier in literature by

Boeseken and Mijs.[147]

Page 82: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

74

OHO

HO

O OH

B(OH)3 + NaOH + 2

O O

OO

O

OO O

Na

H2O, 100°C, 4h

MtOA Cl, acetone,

rt, 14h

- NaCl(s)

> 99%

B

[MtOA]

164 163 162 165

166

O

OHH

O O

OO

O

OO OB

O

OHH

Scheme 64. Two-step synthesis of methyltrioctyl ammonium dimalatoborate (166) (only the

(S,S,S)-diastereoisomer is shown).

The second step involved a cation-exchange replacing the sodium ion with a cation

typically used for the formation of ILs: the methyltrioctyl ammonium ion. The reaction took

place with Aliquat 336® (≡ methyltrioctyl ammonium chloride, 167) in acetone. Under the

precipitation of sodium chloride the desired compound 166 was isolated as a hygroscopic and

viscous colorless liquid (C = 72 ± 2%, determined from 1H NMR spectroscopy).[148]

The aza-BH reaction between MVK (37) and the aromatic tosyl-aldimines 105, 168

and 113 was performed in this new reaction medium (166) (Scheme 65).

N TosONHTos

+chiral solvent 166, r.t., 24 h

105 (X = Br) 168 (X = CH3)113 (X = NO2)

O

37

PR3 (10 mol-%)

X X

106 (X = Br) 169 (X = CH3)114 (X = NO2)

Scheme 65. Reaction between the aromatic tosylaldimines 67, 168 and 113 and MVK (37) in

methyl trioctylammonium dimalatoborate (166) (R = C6H5, o-CH3C6H4, etc).

Page 83: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

75

As nucleophilic catalysts the commercially available, achiral triarylphosphines PPh3, P(o-

tolyl)3 and (C6F5)PPh2 were applied. The best results were obtained with imine 105 (X = Br)

and triphenylphosphine (53) as catalyst. Using 1.2 eq of MVK (37) and 10 mol-% of

phosphine 53, the corresponding chiral allylic amine (R)-106 (X = Br) was obtained with a

reasonable conversion and a high enantiomeric excess.[149] In a series of four independent

experiments using various batches of IL 166, conversion varied between 34-39% and the

enantioselectivity ranged from 71-84% ee. The highest ee of 84% observed in these

experiments corresponds to an enantiomeric ratio of er = 11.5 and hence an energetic

differentiation of ΔΔG‡ = 6.05 kJ·mol-1 for the formation of the two enantiomers.[137] So far,

this is the highest level of enantioselectivity obtained with a chiral solvent as the sole source

of chirality. It exceeds the asymmetric induction reported in the work by Loupy and co-

workers for the BH reaction of benzaldehyde with methyl acrylate by a factor of 3.2.[125]

Although a variation of the achiral catalyst (PPh3, P(o-tolyl)3 and (C6F5)PPh2) did not

cause a significant change in enantioselectivity, conversion dropped with a decrease in

nucleophilicity of the phosphine (P(o-tolyl)3 ≈ PPh3 > (C6F5)PPh2) (Table 8).

Table 8. Different phosphine catalysts for the coupling of imine 105 (X = Br) with MVK (37)

in IL 166.

entry phosphine conv. [%] ee [%] 1 PPh3 37 77 (R) 2 P(o-tol)3 35 74 (R) 3 (C6F5)PPh2 9 71 (R)

Reaction conditions: 51.0 mg (0.125 mol, 1.00 eq) imine 105, 12.5 μl (0.l50 mmol, 1.20 eq)

MVK, 0.0125 mmol (0.10 eq) phosphine, 1.0 ml (8.0 ml⋅mmol−1 105) IL 166, r.t., 24 h.

In contrast, the influence of the p-substituent X of the imine on the enantioselectivity of the

reaction is very pronounced. Although increasing the electron density of the aromatic imine

by substituting the Br- for a Me group in the para-position caused no large difference in

conversion (C = 30% after 24 h) and enantioselectivity (ee = 64% (R)), a strong decrease of

the enantiomeric excess was observed upon introduction of the highly electron-withdrawing

nitro moiety (ee = 10% (R)) (Table 9).

Page 84: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

76

Table 9. Influence of the para-substituent X of the aromatic tosylaldimine on conversion and

enantioselectivity.

entry X conv. [%] ee [%] 1 Br 37 77 (R) 2 CH3 39 64 (R) 3 NO2 38 10 (R)

Reaction conditions: 0.125 mmol (1.00 eq) imine, 12.5 μl (0.l50 mmol, 1.20 eq) MVK, 3.28

mg (0.0125 mmol, 0.10 eq) PPh3, 1.0 ml (8.0 ml⋅mmol−1 imine) IL 166, r.t., 24 h.

Since the stability of the stereogenic center of the aza-BH product under turn-over

conditions is likely to decrease with the introduction of electron-withdrawing groups, one of

the possible explanations for the dramatic drop of the enantiomeric excess could be a

subsequent racemization of the allylic amine 114 (X = NO2). A control experiment with the

enantiomerically enriched reaction product 114 (X = NO2), however, excluded this possibility

indicating that the dependence of the enantioselectivity on the nature of the substituent X is a

reaction intrinsic phenomenon. A similar decrease in optical induction was described by Vo-

Thanh and co-workers for the mechanistically related BH reaction of p-nitrobenzaldehyde and

methylacrylate in ephedrinium based cation-chiral ionic liquids.[125] They explained the drop

in enantioselectivity by the formation of a hydrogen bond between the OH group of the chiral

ion and the NO2 function of the substrate.

To investigate whether methyltrioctyl ammonium dimalatoborate (166) acts as a chiral

additive rather than a chiral solvent, the MtOA-IL 166, its corresponding sodium salt 165 and

L-malic acid (162) were tested as additives for the triphenylphosphine-catalyzed aza-BH

reaction of imine 105 (X = Br) with MVK (37) in THF. In all cases, racemic product (106, X

= Br) was obtained in low conversions (Table 10).

Page 85: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

77

Table 10. PPh3-catalyzed reaction of imine 105 and MVK (37) with malic acid-based chiral

additives.

entry additive conv. [%] ee [%] 1 166 10 rac 2 165 10 rac 3 162 8 rac

Reaction conditions: 51.0 mg (0.125 mmol, 1.00 eq) imine 105, 12.5 μl (0.l50 mmol, 1.20 eq)

MVK, 3.28 mg (0.0125 mmol, 0.10 eq) PPh3, 0.0125 mmol (0.10 eq) additive, 1.0 ml (8.0

ml⋅mmol−1) THF, r.t., 24 h.

Consequently, neither the IL 166 nor any of its chiral precursors (165, 162) can be used as

additive for the aza-BH reaction in conventional organic solvents, which clearly underlines

the necessity of using the IL 166 as a solvent for this process.

In the discussion of the structural requirements for an ideal IL for the aza-BH reaction,

the necessity of a hydrogen bond donor group in the anion of the solvent was postulated for a

benificial interaction with the zwitterionic transition states of this reaction (vide supra). To

validate this hypothesis, the structurally related anion-chiral ILs 170 and 171 without a

hydrogen bond donor, prepared by Gausepohl in a procedure which is similar to the synthesis

of 166, were tested.[145] The results obtained with these solvents in the PPh3-catalyzed aza-BH

coupling of imine 67 (X = Br) with MVK (37) are depicted in Scheme 66.

[MtOA][MtOA] [MtOA]

O

O OOB

O

O

O

O

OO

O

O O O

OO

O OB

O O

OO

O

OO OB

O

OHH

170 171 16614% 15%

rac rac up to 84% ee

up to 39%

Scheme 66. ILs containing chiral anions without a Brønsted acid group. Reaction conditions:

51.0 mg (0.125 mol, 1.00 eq) imine 105, 12.5 μl (0.l50 mmol, 1.20 eq) MVK, 3.28 mg

(0.0125 mmol, 0.10 eq) PPh3, 1.0 ml (8.0 ml⋅mmol−1 105) IL, r.t., 24 h.

Page 86: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

78

Notably, the structurally related ILs 170 and 171 resulted in lower conversions (14% and 15%

respectively) and formation of racemic product 106 (X = Br), which indicates the necessity of

a Brønsted acidic moiety within the chiral ion.

In conclusion, it is shown that the strategy of applying chiral solvents in asymmetric

synthesis can be a valuable strategy for obtaining chiral products with high enantioselectivity.

With the specifically designed methyltrioctyl ammonium dimalatoborate IL 166, the aza-BH

product 106 could be obtained with enantioselectivities up to 84% ee.

Page 87: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

79

4. Summary and Outlook

In this thesis, the rational development of a new enantioselective catalytic system for the

asymmetric aza-BH reaction was described (Scheme 67).

N Tos

EWGNHTos

+ EWG *chiral catalyst/chiral solvent

X X

Scheme 67. Aza-BH reaction of aromatic tosylaldimines with activated alkenes (X = H, p-

NO2, m-F etc; EWG = C(O)CH3, C(O)OCH3 etc).

Since no mechanistic details on this reaction were known at the beginning of this work, an

extensive study on the reaction mechanism was performed. This showed that the mechanistic

details of the aza-BH reaction (see Scheme 29) did not fit into any of the existing mechanistic

hypotheses for the related BH coupling.

Detailed kinetic investigations via 19F NMR spectroscopy yielded rate law equations

for the phosphine-catalyzed aza-BH reaction with and without Brønsted acidic co-catalysts

(Scheme 68).

F

N TosO

+

PPh3, C6F6, (acid)

THF, r.t.

F

NH OTos

102 103

10453

37

− 112.6 −114.1−164.0δ (19F) ppm

Scheme 68. Model reaction for kinetic study of the aza-BH reaction.

A comparison of these equations enabled the localization of the RDS in both cases:

without co-catalyst, proton migration was involved in the RDS. Upon addition of a Brønsted

acidic co-catalyst, this reaction step was considerably facilitated leading to a shift of the RDS

to the Mannich-type coupling. Additionally, a set of Brønsted acids with pKA values in the

range of 4-16 was screened. An evaluation of these results showed that a maximum in

reaction rate was obtained with 3,5-bis(CF3)phenol (pKA ≈ 8) as additive, corresponding to a

Page 88: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

80

14-fold rate enhancement as compared to the reaction without additive. Weakly acidic

additives like methanol and water caused only a small rate enhancement; highly acidic

additives like malic acid inhibited the reaction by protonation of the nucleophilic catalyst

itself and the 3-phosphonium enolate.

In addition to this kinetic study, investigations towards the reversibility of the reaction

were performed. Cross-Mannich experiments were carried out and evaluated. In none of these

reactions, however, cross-over products were observed under turn-over conditions. This

provided strong evidence for the irreversibility of this coupling step under these conditions.

Racemization experiments showed that the stereocenter of the aza-BH product was not

always stable under turn-over conditions (Scheme 69).

PPh3, phenol

THF, r.t.

NH OTos

Br

53 111

(S)-106

NH OTos

Br

(R)-106

Scheme 69. Racemization of the aza-BH product 106 under the influence of PPh3 and phenol.

When the reaction was performed in [D4]methanol, an H/D-exchange occurred at the

stereogenic center. This demonstrated that the C-H bond at the stereogenic center was labile

under turn-over conditions and that the racemization process is independent from the catalytic

cycle. Further investigations on the rate of racemization and the influence of different reaction

parameters displayed that either a Brønsted acid or remaining starting material (imine) in

combination with the Lewis basic catalyst is necessary for product racemization. The results

from more elaborate studies on the influence of the nature of the Brønsted acid on the rate of

racemization provided information for the postulation of a mechanism, in which the

conjugated base of the Brønsted acid played a vital role.

Information on the transfer of chirality from a bifunctional catalyst to the reaction

product was obtained by studying Shi’s successful phosphine-BINOL catalyst 88 (see Scheme

47).[48b] Shi’s mechanistic proposal was proven incomplete and amended. Furthermore, it was

demonstrated via DFT calculations that bifunctional stabilization of the second intermediate

of the aza-BH reaction through coordination of the phenolic OH group to the negatively

charged imine nitrogen was very unlikely (see Figure 9). In the calculated structures the OH

group coordinated to the SO2 moiety of the activating tosylgroup. This might have wide-

Page 89: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

81

ranging consequences for the influence of the activating group on the enantioselectivity of the

reaction.

The results from these mechanistic investigations provided useful information for the

design of novel enantioselective systems. As potential chiral catalysts phosphinoarboxylic

acids, chiral Brønsted acidic co-catalysts and QUINAP derivatives were tested. Although the

obtained enantioselectivities were very poor in all cases, a more extensive study on the

QUINAP-type system enabled the interception and characterization of all commonly accepted

intermediates of the phosphine-catalyzed aza-BH reaction.

Furthermore, the chiral aminoalcohol 160 was synthesized in a six-step sequence

starting form trans-4-hydroxy-L-proline (Scheme 70). The overall yield of this successful and

modular synthesis was 37%. Evaluation of this compound as a catalyst for the aza-BH

reaction, however, displayed this vic-aminoalcohol as inactive.

NH

HO

COOH NH

HO Cl

37%

6 steps

154 160

Scheme 70. Synthesis of vic-aminoalcohol 160.

A complementary approach was chosen with the application of a solvent as the sole

source of chirality for the aza-BH reaction. An anion-chiral IL with an acidic side group was

specifically designed for this process. The PPh3-catalyzed aza-BH coupling of tosylaldimines

with MVK yielded the corresponding reaction products in enantioselectivities up to 84% ee

(er = 11.5; ΔΔG‡ = 6.05 kJ·mol-1) in IL 166 (Scheme 71).

N TosONHTos

+166, r.t., 24 h

O

37

PR3 (10 mol-%)

X Xup to 84% ee

[MtOA]

O O

OO

O

OO OB

O

OHH

166

Scheme 71. Aza-BH reaction in methyltrioctyl dimalatoborate (166) (X = CH3, Br, NO2; PR3

= PPh3, P(o-tolyl)3, (C6F5)PPh2).

Page 90: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

82

So far, this is the highest enantioselectivity obtained with a chiral solvent as the sole source of

chirality for an asymmetric reaction.

In future experiments, the racemization process should be investigated more intensively.

Especially, a detailed study towards the rate law for this process would provide valuable

additional information for processing the aza-BH reaction in an enantioselective way.

Additionally, the mechanism of chirality transfer should be readdressed. To determine the

exact influence of the activating group and fully clarify the mechanism, additional theoretical

and experimental studies, e.g. on matched-mismatched effects of succesfull catalysts with S-

chiral sulfinimines, are required.

Regarding the chiral vic-aminoalcohols based on trans-4-hydroxy-L-proline, a

variation of the protecting group on the Nitrogen atom seems a viable strategy for tuning the

nucleophilicity of these compounds. It is expected that increasing the nucleophilicity of the

aminogroup leads to an active catalyst for the aza-BH reaction. Furthermore, vic-

aminoalcohol 160 as well as its TBS-protected precursor 159 may serve as chiral backbone

chiral for the formation of catalysts for other organocatalytic processes as well as ligands for

homogeneous transition metal catalysis.

Concerning the reaction in anion-chiral ILs, a more extensive screening of substrates

would provide valuable information on the flexibility of this catalytic system. Furthermore,

mechanistic investigations on the structure and dynamics of these solvents and their

interaction with dissolved substances are inevitable for the clarification of the mechanism of

chirality transfer in these media. Useful experiments in this respect are:

1. Dilution experiments with conventional, achiral ILs: How does the activity and

enantioselectivity of the system depend on the fraction of chiral IL in the solvent

mixture?

2. Controlled addition of impurities: What are the influences of impurities like water on

the system’s activity and enantioselectivity?

3. NMR spectroscopic studies: Is there a signal shift upon dissolution of a compound in

the IL and does this correlate to the coordination strength of the IL?

In summary, a detailed insight into the mechanism of the aza-BH reaction under bifunctional

activation was obtained. This provided useful information for the development of the first

efficient enantioselective system in which the solvent is used as the sole source of chirality.

Page 91: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

83

Based on these results, future synthetic and mechanistic work in this research area seems very

promising.

Page 92: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

84

5. Experimental

5.1. General Methods, Solvents and Reagents

5.1.1. Inert Gas Conditions

All reactions involving air-sensitive chemicals were performed under Argon atmosphere

using standard Schlenk techniques. Glassware was heated in vacuo and subsequently purged

with Argon. The addition of reagents and solvents was carried out with polypropylene

syringes equipped with V2A steel needles under an Argon stream. Sensitive chemicals like

MVK were degassed and stored under Argon at −30°C.

5.1.2. Solvents

The following solvents were dried and purified by distillation under Argon according to

standard procedures (Table 11).[150] For the purification of the corresponding deuterated

solvents, identical procedures were applied.

Table 11. Purification conditions for solvents.

entry solvent conditions 1 tetrahydrofurane heated under reflux over sodium benzophenone ketyl,

followed by distillation under Argon 2 diethylether heated under reflux over sodium benzophenone ketyl,

followed by distillation under Argon 3 toluene heated under reflux over sodium benzophenone ketyl,

followed by distillation under Argon 4 benzene heated under reflux over sodium benzophenone ketyl,

followed by distillation under Argon 5 chloroform heated under reflux over calcium hydride,

followed by distillation under Argon 6 dichloromethane heated under reflux over calcium hydride,

followed by distillation under Argon 7 pentane heated under reflux over calcium hydride,

followed by distillation under Argon 8 acetone heated under reflux over calcium chloride,

followed by distillation under Argon

Page 93: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

85

N,N-Dimethylformamide (DMF) and dimethylsulfoxide (DMSO) were purchased from

Aldrich in following quality:

(a) DMF: puriss, absolute, over MS 4Å (H2O ≤ 50 ppm) under nitrogen

(b) DMSO: water-free (H2O ≤ 50 ppm) under nitrogen

Furthermore, all solvents for column chromatography, recrystallization and reaction workup

were distilled prior to use.

5.1.3. Reagents

The commercially obtained reagents which were used in this Ph.D. project were obtained

from various suppliers and in different qualities. Appendix 2 contains a detailed list with

reagent, supplier and quality.

(R)-2’-diphenylphosphanyl-[1,1’]binaphthalenyl-2-ol (88) and (R,S)-2-diphenylphos-

phanyl cyclohexane carboxylic acid methyl ester were provided by Dr. Giancarlo Franciò.

(R)-QUINAP (127) and its derivatives 128 and 129 were donated by Dr. John M. Brown FRS.

The diphenylphosphinocarboxylic acids 136, 137 and 138 were kindly allocated by Dr.

Christian Steffens. The chiral ILs 170 and 171 were donated by Dipl.-Chem. Rolf Gausepohl.

5.2. Determination of Physical Data

5.2.1. NMR Spectroscopy

1H-, 13C-, 31P-, 19F- and 11B NMR spectra were recorded on Bruker DPX 200, DPX 300, AV

400, AV 500 and AV 600 spectrometers. The operating frequencies of these spectrometers for

NMR measurements are listed in Table 12. For 1H- and 13C NMR spectroscopy, the chemical

shifts δ are given in ppm using the residual solvent signals as internal standard. For 31P- and 19F-NMR spectroscopy the chemical shift values δ are given in ppm relative to 85%

phosphoric acid and hexafluorobenzene as external standard, respectively. The multiplicity of

the signals was assigned assuming spectra of first order. The coupling constant J is given in

Hertz. For the description of the multiplicity of the signal following abbreviations are used: s

Page 94: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

86

= singulet, d = doublet, t = triplet, q = quadruplet, m = multiplet, br = broad. The proton

which belongs to the signal is underlined in the added structure section. Unless indicated

otherwise, the spectra are recorded at room temperature.

Table 12. Operating Frequencies of NMR spectrometers.

entry spectrometer operating frequency /

MHz

1H 13C 31P 19F 11B 1 DPX 200 200.1 50.3 81.0 188.3 64.2 2 DPX 300 300.1 75.5 121.5 282.4 96.3 3 AV 400 400.1 100.6 162.0 376.5 128.4 4 AV 500 500.1 125.8 202.5 470.6 160.5 5 AV 600 600.1 150.9 242.9 564.7 192.5

5.2.2. IR Spectroscopy

IR spectra were recorded with a Perkin-Elmer spectrometer (PE 1720X, PE 1760 FT). The

measurements were carried out on a film between two plates of KBr for liquids, in CHCl3 for

oils and KBr pills for solids. The positions of the absorption signals are listed in cm−1. Only

signals with a transmission T ≤ 80% are reported. The peak intensity is characterized using

the following abbreviations: vs (very strong, 0% ≤ T ≤ 20%), s (strong, 20% < T ≤ 40%), m

(medium, 40% < T ≤ 60%), w (weak, 60% < T ≤ 80%).

5.2.3. Mass Spectroscopy

Mass spectroscopic (MS) measurements were performed on a Finnigan SSQ 7000 (EI 70 eV,

CI 100 eV) or Fisons Platform spectrometer (Electrospray mode). High-resolution MS

measurements were carried out on a Finnigan MAT 95 or Micromass QTOF Micro

spectrometer (Electrospray mode).

The mass of fragments (m/z) is given using a number without dimension with the

relative peak intensity added in brackets. For high-resolution MS, the mass of the molecule

ion is given. The error Δ is given in ppm (Equation 5).

Page 95: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

87

calc

calcobs

mmm −

=Δ 610 Equation 5

Δ = error (in ppm)

mobs = observed mass

mcalc = calculated mass

5.2.4. Gas Chromatography

GC analysis was performed on a HP5890 (serie 2) using a Si 54 column of 25 m length.

Following temperature program was applied: 50°C (1 min) – 8°C/min – 250°C (5 min).

5.2.5. High Performance Liquid Chromatography

HPLC analyses of the aza-BH products 106 (p-Br), 169 (p-CH3), 114 (p-NO2), 89 (H) and

103 (m-F) were performed on Diacel Chiralcel OD-H, AD-H or Chiralpack IA column at

20°C.

106 (p-Br): 223 nm, heptane/isopropanol = 85/15, flow rate 0.5 ml/min, t(S) = 31.4 min, t(R)

= 35.1 min (Diacel Chiralcel OD-H)

169 (p-CH3): 223 nm, heptane/ethanol = 90/10, flow rate 0.5 ml/min, t(S) = 52.1 min, t(R) =

60.7 min (Chiralpack IA)

114 (p-NO2): 223 nm, heptane/ethanol = 80/20, flow rate 0.7 ml/min, t(S) = 39.6 min, t(R) =

61.1 min (Chiralpack IA)

89 (H): 223 nm, heptane/isopropanol = 85/15, flow rate 0.5 ml/min, t1 = 42.6 min, t2 = 49.0

min (Chiralcel AD-H)

103 (m-F): 223 nm, heptane/isopropanol = 85/15, flow rate 0.5 ml/min, t1 = 38.0 min, t2 =

46.5 min (Chiralcel AD-H)

Page 96: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

88

5.2.6. Thin Layer Chromatography

For Thin Layer Chromatography (TLC), POLYGRAM SIL G/UV254 plates from Macherey

and Nagel with a silica layer of 0.2 mm were used. Vizualization was realized through UV or

permanganate treatment.

5.3. Preparative Column Chromatography

For preparative column chromatography, KIESELGEL 60 from Merck with a particle size in

the range of 0.040 – 0.063 mm was used. The chromatographic purification was carried out

according to the general procedure by Still and co-workers.[151]

5.4. Syntheses

5.4.1. Aromatic Tosylaldimines

N SO

ON S

O

O

Br

N SO

O

N SO

O

O2N

N SO

O

F

67 168 105

113102

Synthesis of N-Benzylidene-4-methylbenzenesulfonamide (67) as prototypical procedure:

According to Love and co-workers,[152] tosylamide (8.56 g, 50.0 mmol, 1.00 eq) was added to

a mixture of benzaldehyde (5.07 ml, 50.0 mmol, 1.00 eq) and tetraethoxysilane (11.7 ml, 52.5

mmol, 1.05 eq) under inert gas conditions. The resulting suspension was heated in a

Page 97: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

89

distillation apparatus to 170°C and stirred overnight. Ethanol was continuously removed by

distillation. The product was isolated by crystallization from a mixture of ethyl acetate and

pentane.

Yield: 9.72 g (37.5 mmol, 75%)

1H NMR (300 MHz, CDCl3): δ = 9.05 (s, 1H, imine-H), 7.96-7.89 (m, 4H, Ar), 7.63 (m, 1H,

Ar), 7.50 (dd {3JHH = 7.7 Hz, 3JHH = 7.3 Hz}, 2H, Ar), 7.37 (d {3JHH = 8.6 Hz}, 2H, Ar), 2.45

(s, 3H, tos-CH3) ppm.

13C NMR (75.5 MHz, CDCl3): δ = 170.2, 144.7, 135.1, 135.0, 132.4, 131.3, 129.8, 129.2,

128.1, 21.7 ppm.

TLC: Rf = 0.40 (pentane/ethyl acetate = 5/1)

N-(4-Methylbenzylidene)-4-methylbenzenesulfonamide (168)

Yield: 58%

1H NMR (300 MHz, CDCl3): δ = 9.00 (s, 1H, imine-H), 7.90 (d {3JHH = 8.2 Hz}, 2H, Ar), 7.83

(d {3JHH = 8.2 Hz}, 2H, Ar), 7.35 (d {3JHH = 8.2 Hz}, 2H, Ar), 7.30 (d {3JHH = 8.2 Hz}, 2H,

Ar), 2.45 (s, 3H, CH3), 2.44 (s, 3H, CH3) ppm.

13C NMR (75.5 MHz, CDCl3): δ = 170.0, 146.4, 144.5, 135.4, 131.4, 130.0, 129.9, 129.8,

128.0, 22.0, 21.7 ppm.

TLC: Rf = 0.37 (pentane/ethyl acetate = 5/1)

N-(4-Bromobenzylidene)-4-methylbenzenesulfonamide (105)

Yield: 65%

Page 98: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

90

1H NMR (300 MHz, CDCl3): δ = 8.90 (s, 1H, imine-H), 7.80 (d {3JHH = 8.3 Hz}, 2H, Ar), 7.70

(d {3JHH = 8.5 Hz}, 2H, Ar), 7.55 (d {3JHH = 8.5 Hz}, 2H, Ar), 7.27 (d {3JHH = 8.3 Hz}, 2H,

Ar), 2.36 (s, 3H, tos-CH3) ppm.

13C NMR (75.5 MHz, CDCl3): δ = 168.8, 144.8, 134.8, 132.6, 131.2, 130.3, 129.9, 128.2, 21.7

ppm.

TLC: Rf = 0.43 (pentane/ethyl acetate = 5/1)

N-(3-Fluorobenzylidene)-4-methylbenzenesulfonamide (102)

Yield: 79%

1H NMR (300 MHz, CDCl3): δ = 9.03 (s, 1H, imine-H), 7.91 (d {3JHH = 8.4 Hz}, 2H, Ar), 7.83

(d {3JHH = 8.4 Hz}, 1H, Ar), 7.71-7.65 (m, 2H, Ar), 7.49 (m, 1H, Ar), 7.38 (d {3JHH = 8.4

Hz}, 2H, Ar), 2.46 (s, 3H, tos-CH3) ppm.

13C NMR (75.5 MHz, CDCl3): δ = 168.8 (d {JCF =3.1 Hz}), 162.8 (d {JCF = 249.1}), 145.0,

134.7, 134.5 (d {JCF = 7.6 Hz}), 130.9 (d {JCF = 8.0}), 129.9, 128.2, 127.9 (d{JCF = 3.1 Hz}),

121.9 (d {JCF = 21.8 Hz}), 116.6 (d {JCF = 22.5 Hz}), 21.7 ppm.

19F NMR (282.4 MHz, CDCl3): δ = −112.6 ppm.

TLC: Rf = 0.35 (pentane/ethyl acetate = 5/1)

N-(4-Nitrobenzylidene)-4-methylbenzenesulfonamide (113)

Yield: 86%

1H NMR (300 MHz, CDCl3): δ = 9.04 (s, 1H, imine-H), 8.26 (d {3JHH = 8.7 Hz}, 2H, Ar), 8.04

(d {3JHH = 8.7 Hz}, 2H, Ar), 7.84 (d {3JHH = 8.3 Hz}, 2H, Ar), 7.31 (d {3JHH = 8.3 Hz}, 2H,

Ar), 2.39 (s, 3H, tos-CH3) ppm.

Page 99: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

91

13C NMR (75.5 MHz, CDCl3): δ = 167.3, 145.4, 137.4, 134.1, 131.9, 130.0, 129.9, 128.4,

124.2, 21.7 ppm.

TLC: Rf = 0.26 (pentane/ethyl acetate = 5/1)

5.4.2. Racemic Aza-Baylis-Hillman Products

NH OSO

O NH OSO

ONH OS

O

O

Br

NH OSO

O

O2N

NH OSO

O

F

89 169 106

103 114

Synthesis of 4-Methyl-N-(2-methylene-3-oxo-1-phenylbutyl)-4-methylbenzenesulfon-

amide (89) as prototypical procedure:

N-Benzylidene-4-methylbenzenesulfonamide (67) (1.647 g, 5.00 mmol, 1.00 eq), DABCO

(56.1 mg, 0.50 mmol, 0.10 eq) and phenol (47.1 mg, 0.50 mmol, 0.10 eq) were dissolved in

dry THF (10 ml, 2 ml⋅mmol−1). Then, MVK (0.499 ml, 6.00 mmol, 1.20 eq) was added to this

solution. The reaction was stirred at room temperature. Conversion of the starting material 67

to the corresponding aza-BH product 89 was monitored via TLC (pentane/ethylacetate = 5/1).

After complete conversion, all volatiles were removed in vacuo and the solid residue was

purified via column chromatography (pentane/ethyl acetate = 4/1 – 2/1).

1H NMR (300 MHz, CDCl3): δ = 7.57 (d {3JHH = 8.3 Hz}, 2H, Ar), 7.11 (m, 5H, Ar), 7.02 (m,

2H, Ar), 6.02 (s, 1H, CHH), 6.01 (s, 1H, CHH), 5.73 (d {3JHH = 8.7 Hz}, 1H,CHNH), 5.21 (d

{3JHH = 8.7 Hz}, 1H,CHNH), 2.33 (s, 3H, tos-CH3), 2.07 (s, 3H, C(O)CH3) ppm.

13C NMR (75.5 MHz, CDCl3): δ =198.8, 146.5, 143.3, 138.8, 137.4, 129.5, 128.5, 128.2,

127.6, 127.3, 126.4, 58.7, 26.3, 21.5 ppm.

Page 100: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

92

TLC: Rf = 0.14 (pentane/ethyl acetate = 5/1)

4-Methyl-N-(2-methylene-3-oxo-1-p-tolyl-butyl)-benzenesulfonamide (169)

1H NMR (300 MHz, CDCl3): δ = 7.58 (d {3JHH = 8.3 Hz}, 2H, Ar), 7.15 (d {3JHH = 8.6 Hz},

2H, Ar), 6.92 (m, 4H, Ar), 6.03 (s, 1H, CHH), 6.02 (s, 1H, CHH), 5.48 (d {3JHH = 8.4 Hz},

1H, CHNH), 5.16 (d {3JHH = 8.4 Hz}, 1H, CHNH), 2.34 (s, 3H, tos-CH3), 2.19 (s, 3H, p-

CH3), 2.09 (s, 3H, C(O)CH3) ppm.

13C NMR (75.5 MHz, CDCl3): δ = 198.9, 146.6, 143.3, 137.4, 135.9, 129.5, 129.2, 128.0,

127.3, 126.3, 58.7, 26.4, 21.5, 21.0 ppm.

TLC: Rf = 0.12 (pentane/ethyl acetate = 5/1)

N-[1-(4-Bromophenyl)-2-methylene-3-oxobutyl]-4-methylbenzenesulfonamide (106)

1H NMR (300 MHz, CDCl3): δ = 7.55 (d {3JHH = 8.0 Hz}, 2H, Ar), 7.22 (d {3JHH = 7.8 Hz},

2H, Ar), 7.16 (d {3JHH = 8.0 Hz}, 2H, Ar), 6.88 (d {3JHH = 7.8 Hz}, 2H, Ar), 6.00 (m, 3H,

CHH and CHNH), 5.16 (d {3JHH = 8.5 Hz}, 1H, CHNH), 2.33 (s, 3H, tos-CH3), 2.05 (s, 3H,

C(O)CH3) ppm.

13C NMR (75.5 MHz, CDCl3): δ = 198.8, 146.2, 143.6, 138.1, 137.5, 135.9, 131.3, 129.6,

128.6, 128.3, 127.2, 121.5, 58.0, 26.3, 21.5 ppm.

TLC: Rf = 0.12 (pentane/ethyl acetate = 5/1)

N-[1-(3-Fluorophenyl)-2-methylene-3-oxobutyl]-4-methylbenzenesulfonamide (103)

1H NMR (300 MHz, CDCl3): δ = 7.58 (d {3JHH = 8.3 Hz}, 2H, Ar), 7.16 (d {3JHH = 8.3 Hz},

2H, Ar), 7.10 (m, 1H, Ar), 6.80 (m, 3H, Ar), 6.02 (s, 1H, CHH), 6.01 (s, 1H, CHH), 5.76 (d

{3JHH = 8.8 Hz}, 1H, CHNH), 5.19 (d {3JHH = 8.8 Hz}, 1H, CHNH), 2.35 (s, 3H, tos-CH3),

2.09 (s, 3H, C(O)CH3) ppm.

Page 101: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

93

13C NMR (75.5 MHz, CDCl3): δ = 198.8, 162.6 (d {JCF = 245 Hz}), 146.0, 143.6, 141.5 (d

{JCF = 6.8 Hz}), 137.4, 130.0 (d{JCF = 8.4 Hz}), 129.6, 128.9, 127.3, 122.0 (d {JCF = 2.8

Hz}), 114.2 (d {JCF = 21.2 Hz}), 113.7 (d {JCF = 22.8 Hz}), 58.6, 26.3, 21.5 ppm.

19F NMR (282.4 MHz, CDCl3): δ = −114.1 ppm.

TLC: Rf = 0.14 (pentane/ethyl acetate = 5/1)

4-Methyl-N-[2-methylene-1-(4-nitrophenyl)-3-oxobutyl]-benzenesulfonamide (114)

1H NMR (300 MHz, CDCl3): δ = 8.09 (d {3JHH = 8.7 Hz}, 2H, Ar), 7.68 (d {3JHH = 8.3 Hz},

2H, Ar), 7.39 (d {3JHH = 8.7 Hz}, 2H, Ar), 7.28 (d {3JHH = 8.3 Hz}, 2H, Ar), 6.10 (m, 3H,

CHNH and CHH), 5.35 (d {2JHH = 9.4 Hz}, 1H, CHNH), 2.45 (s, 3H, tos-CH3), 2.18 (s, 3H,

C(O)CH3) ppm.

13C NMR (300 MHz, CDCl3): δ = 198.7, 147.2, 146.3, 145.5, 143.9, 137.3, 129.7, 129.6,

127.3, 127.2, 123.6, 58.7, 26.2, 21.5 ppm.

TLC: Rf = 0.03 (pentane/ethyl acetate = 5/1)

5.4.3. Enantiomerically Enriched Aza-Baylis-Hillman Products

NH OSO

ONH OS

O

O

O2NBr(S)-114(S)-106

Synthesis of (S)-N-[1-(4-bromophenyl)-2-methylene-3-oxobutyl]-4-methylbenzene-

sulfonamide (106) as prototypical procedure:

N-(4-Bromobenzylidene)-4-methylbenzenesulfonamide (105) (338.2 mg, 1.00 mmol, 1.00 eq)

and (R)-2’-diphenylphosphanyl-[1,1’]binaphthalenyl-2-ol (88) (45.4 mg, 0.10 mmol, 0.10 eq)

were dissolved in THF (8.0 ml, 8.0 ml⋅mmol−1) under inert gas conditions. After addition of

Page 102: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

94

MVK (99.9 μl, 1.20 mmol, 1.20 eq), the reaction mixture was stirred at room temperature for

24 h. Then, all volatiles were removed in vacuo and the residue was purified via column

chromatography (pentane/ethyl acetate = 4/1 – 2/1).

Yield: 60%

Enantiomeric excess: 88% (S)

NMR data: see rac-106

(S)-4-Methyl-N-[2-methylene-1-(4-nitrophenyl)-3-oxobutyl]-benzenesulfonamide (114)

Yield: 98%

Enantiomeric excess: 74% (S)

NMR data: see rac-114

5.4.4. (R,S)-2-Diphenylphosphanyl cyclohexane carboxylic acid (140)

COOH

PPh2

140

(R,S)-2-diphenylphosphanyl cyclohexane carboxylic acid methyl ester (326.0 mg, 1.00 mmol,

1.00 eq) was dissolved in ethanol (4 ml, 4 ml⋅mmol−1 ester) under Argon at room temperature.

Then, a solution of KOH (98.2 mg, 1.75 mmol, 1.75 eq) in degassed water (3.5 ml, 2

ml⋅mmol−1 KOH) was added dropwise and the resulting mixture was heated under reflux for 4

h. After removal of all volatiles in vacuo, degassed water was added (4 ml, 4 ml⋅mmol−1 ester)

and the aqueous phase was extracted with dichloromethane (3 x 8 ml, 3 x 2 ml⋅mmol−1 ester)

under Argon. The organic phase was separated and dried over Na2SO4. All volatiles were

removed in vacuo and (R,S)-2-diphenylphosphanyl cyclohexane carboxylic acid (140) was

isolated as a colorless crystalline solid.

Page 103: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

95

Yield: 309 mg (0.99 mmol, 99%)

1H NMR (300 MHz, [D8]THF): δ = 7.58 (m, 4H, Ar), 7.42 (m, 6H, Ar), 2.85 (m, 1H,

CHPPh2), 2.35 (m, 1H, CHCOOH), 2.10 (m, 1H, CHHCHCOOH), 1.85 (m, 1H, CHHCHP),

1.75 (m, 1H, CHHCHCOOH), 1.50 (m, 1H, CHHCHHCHP), 1.42 (m, 1H,

CHHCHHCHCOOH), 1.35 (m, 1H, CHHCHHCHP), 1.22 (m, 1H, CHHCHP), 1.10 (m, 1H,

CHHCHHCHCOOH) ppm.

13C NMR (75.5 MHz, [D8]THF): δ = 174.8 (d {JPC = 5.4 Hz}), 136.8 (d {JPC = 15.0 Hz}),

135.8 (d {JPC = 18.0 Hz}), 134.2 (d {JPC = 21.1 Hz}), 132.0 (d {JPC = 17.5 Hz}), 128.3, 127.7

(d {JPC = 5.6 Hz), 127.5 (d {JPC = 7.5 Hz}), 127.3, 44.2 (d {JPC = 19.9 Hz}), 34.6 (d {JPC =

15.1}), 28.2 (d {JPC = 10.5 Hz}), 25.6 (d {JPC = 3.4 Hz}), 24.3, 24.1 (d {JPC = 0.9 Hz}) ppm.

31P NMR (121.5 MHz, [D8]THF): δ = −6.11 ppm.

5.4.5. (3-Oxobutyl)triphenylphosphonium malate (109)

O

Ph3P

O

OHO

O

OH

109

Triphenylphosphine (15.7 mg, 0.060 mmol, 1.00 eq) and MVK (4.99 μl, 0.060 mmol, 1.00 eq)

were dissolved in [D8]THF under Argon. After addition of L-(−)-malic acid (8.0 mg, 0.060

mmol, 1.00 eq), the mixture was stirred at room temperature for 5 minutes and transferred

into an NMR tube under Argon. The mixture was analyzed via NMR spectroscopy.

Conversion: 95%

1H NMR (300 MHz, [D8]THF): δ = 9.62 (br. s., 1H, COOH), 7.78 (m, 3H, Ar), 7.63 (m, 12H,

Ar), 3.98 (dd {2JHH = 9.3 Hz, 2JHH = 4.9 Hz}, 1H, CHOH), 3.55 (dt {2JPH = 12.9 Hz , 3JHH =

7.3 Hz}, 2H, PCH2CH2), 2.95 (dt {3JPH = 12.9, 3JHH = 7.3 Hz}, 2H, PCH2CH2), 2.57 (d {3JHH

= 4.9 Hz} 1H, CHHCOOH), 2.55 (d {3JHH = 9.3 Hz}, 1H, CHHCOOH), 2.05 (s, 3H,

C(O)CH3) ppm.

Page 104: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

96

13C NMR (75.5 MHz, [D8]THF): δ = 203.9 (d {JPC = 10.8 Hz}), 178.1, 173.9, 133.4 (d {JPC =

10.0 Hz}), 130.7 (d {JPC = 12.7 Hz}), 118.3, 117.2, 66.2, 41.3, 35.4 (d {JPC = 3.1 Hz}), 29.6,

21.6 (d {JPC = 56.4}) ppm.

31P NMR (121.5 MHz, [D8]THF): δ = 25.3 ppm.

5.4.6. [1-(3-Carboxylato-isoquinolin-1-yl)-naphthalen-2-yl]-(3-oxobutyl)-diphenyl-

phosphonium (131)

N

P

O

Ph Ph

O

O131

MVK (4.16 μl, 0.050 mmol, 1.00 eq) was added to a solution of rac-1-(2-

diphenylphosphanylnaphthalen-1-yl)-isoquinoline-3-carboxylic acid (128) (24.0 mg, 0.050

mmol, 1.00 eq) in CDCl3 (0.7 ml, 14 ml⋅mmol−1). The mixture was stirred at room

temperature for 5 min. After filtration, the resulting solution was analyzed via NMR

spectroscopy and High-Resolution MS.

Conversion: 64%

1H NMR (300 MHz, CDCl3): δ = 8.05 (dd {3JHH = 8.8 Hz, 3JHP = 2.7 Hz}, 1H, Ar), 7.94 (m,

2H, Ar), 7.84 (m, 2H, Ar), 7.72 (m, 2H, Ar), 7.60 (m, 4H, Ar), 7.49 (m, 4H, Ar), 7.20 (m, 6H,

Ar), 4.76 (m, 1H, PCHHCHH), 3.94 (m, 1H, PCHHCHH), 3.30 (m, 1H, PCHHCHH), 3.09

(m, 1H, PCHHCHH), 2.02 (s, 3H, C(O)CH3) ppm

13C NMR (75.5 MHz, CDCl3): δ = 37.4 (d {JPC = 4.0}), 29.7, 16.8 (d {JPC = 55.7 Hz})

ppm.[153]

31P NMR (162.0 MHz, CDCl3): δ = 26.8 ppm.

Page 105: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

97

HR-MS (ESI+): (m/z)obs = 554.1866 (Δ = 2.4 ppm)

5.4.7. (R)-(1-Isoquinolin-1-yl-naphthalen-2-yl)-(3-oxobutyl)-diphenylphosphonium diphenyl-

acetate (172)

N

PPh2

O

O

O

172

MVK (4.2 μl, 0.050 mmol, 1.00 eq) was added to a solution of (R)-QUINAP (127) (22.0 mg,

0.050 mmol, 1.00 eq) and diphenyl acetic acid (10.7 mg, 0.050 mmol, 1.00 eq) in CDCl3 (0.5

ml) under Argon. The reaction mixture was stirred at room temperature for 1 h. After

filtration of the resulting suspension, the solution was analyzed via NMR spectroscopy.[154]

Conversion: 15%

1H NMR (400 MHz, CDCl3): δ = 5.06 (s, 1H, Ph2CH), 3.36 (m, 2H, PCH2CH2), 3.22 (m, 1H,

PCH2CHH), 2.64 (m, 1H, PCH2CHH), 2.29 (s, 3H, C(O)CH3) ppm.

31P NMR (162.0 MHz, CDCl3): δ = 26.43 ppm.

5.4.8. (R,R)- and (S,R)-{1-[3-(3-Hydroxy-2-phenylpropylcarbamoyl)-isoquinolin-1-yl]-

naphthalen-2-yl}-(3-oxobutyl)-diphenylphosphonium diphenylacetate (173)

N

PPh2

O

HN OH

PhO

O

O

173

A 1:1-mixture of (R,R)- and (S,R)-1-(2-diphenylphosphanylnaphthalen-1-yl)-isoquinoline-3-

carboxylic acid (3-hydroxy-2-phenylpropyl)-amide (129) (60.3 mg, 0.100 mmol, 1.00 eq)

and MVK (8.32 μl, 0.100 mmol, 1.00 eq) were dissolved in [D6]benzene (1.5 ml) under

Page 106: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

98

Argon. After addition of diphenyl acetic acid (21.3 mg, 0.100 mmol, 1.00 eq), the mixture

was stirred at room temperature for 5 min, filtered and transferred into a NMR tube under

inert gas. The resulting solution was analyzed via NMR spectroscopy.[154]

Conversion: 32%

1H NMR (400 MHz, [D6]benzene): δ = 3.48 (m, 2H, PCH2CH2), 3.00 (m, 2H, PCH2CH2),

2.02 (s, 3H, C(O)CH3) ppm.

31P NMR (162.0 MHz, [D6]benzene): δ = 26.0 ppm.

5.4.9. [1-(3-Carboxy-isoquinolin-1-yl)-naphthalen-2-yl]-{3-oxo-2-[phenyl-(toluene-4-

sulfonylamino)-methyl]-butyl}-diphenyl-phosphonium diphenylacetate (132)

N

COOH

P

O

Ph PhPh

NH

Tos O

O

132

rac-1-(2-Diphenylphosphanylnaphthalen-1-yl)-isoquinoline-3-carboxylic acid (128) (30.2 mg,

0.050 mmol, 1.00 eq), rac-4-Methyl-N-(2-methylene-3-oxo-1-phenylbutyl)-4-methylbenzene-

sulfonamide (89) (16.5 mg, 0.050 mmol, 1.00 eq) and diphenylacetic acid (10.7 mg, 0.050

mmol, 1.00 eq) were dissolved in 0.7 ml [D8]toluene. The mixture was stirred at room

temperature for 8 h, filtered and subsequently analyzed via NMR spectroscopy and High-

Resolution MS.[153]

Conversion: 13%

1H NMR (400 MHz, [D8]toluene): δ = 5.93 (m, 1H, PCHHCH), 4.90 (m, 1H, PCHHCHCH),

4.68 (m, 1H, PCHHCH), 4.17 (m, 1H, PCHHCH) ppm.

31P NMR (162.0 MHz, [D8]toluene): δ = 22.2 ppm.

Page 107: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

99

HR-MS (ESI+): (m/z)obs = 813.2544 (Δ = 0.3 ppm)

5.4.10. Methyltrioctyl ammonium dimalatoborate (166)

[MtOA]

O O

OO

O

OO OB

O

OHH

166

The synthesis was performed according to the procedure developed by Gausepohl.[146]

Sodium dimalatoborate (165): L-(−)-malic acid (10.73 g, 80.0 mmol, 2.00 eq) was dissolved

in water (10 ml, 0.125 ml⋅mmol−1 malic acid) and treated with boric acid (2.47 g, 40.0 mmol,

1.00 eq). Then, a solution of sodium hydroxide (1.60 g, 40.0 mmol, 1.00 eq) in water (30 ml,

0.75 ml⋅mmol−1 NaOH) was added. The reaction mixture was stirred in an open flask at

100°C. After evaporation of the water (ca. 4 h), sodium dimalatoborate (165) was isolated as a

colorless powder.

Yield: 11.92 g (40.0 mmol, > 99%)

1H NMR (600 MHz, [D6]DMSO): δ = 12.24 (s, 2H, COOH), 4.32 and 4.31 (dd {3JHH,cis = 3.3

Hz, 3JHH,trans = 8.8 Hz}, 2H, CHOB), 2.53 and 2.49 (dd {3JHH,cis = 3.3 Hz, 2JHH = 15.8 Hz},

2H, CHH), 2.27 and 2.25 (dd {3JHH,trans = 8.8 Hz, 2JHH = 15.8 Hz}, 2H, CHH) ppm.

13C NMR (150 MHz, [D6]DMSO): δ = 178.3 and 178.2, 172.6, 72.7 and 72.4, 39.6 and 39.5

ppm.

11B NMR (192.5 MHz, D2O): δ = 9.85 ppm.

Methyltrioctyl ammonium dimalatoborate (166): Methyltrioctyl ammonium chloride

(Aliquat 336®, 3.54 g, 8.75 mmol, 1.00 eq) was dissolved in acetone (20 ml, 2.3 ml⋅mmol−1)

and treated with a solution of the sodium borate salt 165 (2.61 g, 8.75 mmol, 1.00 eq) in

Page 108: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

100

acetone (20 ml, 2.3 ml⋅mmol−1). The reaction mixture was stirred at room temperature

overnight. During this period sodium chloride precipitated as a colorless solid. The reaction

mixture was filtered and the solvent was removed in vacuo to yield the IL 166 as a highly

viscous, hygroscopic colorless liquid.

Conversion: 72 ± 2% (determined from 1H NMR)

1H NMR (600 MHz, [D6]DMSO): δ = 12.24 (s, 2H, COOH), 4.32 and 4.31 (dd {3JHH,cis = 3.3

Hz, 3JHH,trans = 8.8 Hz}, 2H, CHOB), 3.22 (tt {3JHH = 8.2, 4JHH = 4.6 Hz}, 6H, C1-H2), 2.96 (s,

3H, C9-H3), 2.54 and 2.50 (dd {3JHH,cis = 3.3 Hz, 2JHH = 15.8 Hz}, 2H, CHH), 2.28 and 2.26

(dd {3JHH,trans = 8.8 Hz, 2JHH = 15.8 Hz}, 2H, CHH), 1.62 (m, 6H, C6-H2), 1.26 (m, 30H, [C2-

C5+C7]-H2) , 0.87 (tt {3JHH = 6.7 Hz, 4JHH = 2.2 Hz}, 9H, C8-H3) ppm. [155]

13C NMR (150 MHz, [D6]DMSO): δ = 178.3 and 178.2, 172.6, 72.7 and 72.4, 60.4, 47.9, 39.6

and 39.5, 31.1, 28.7, 28.7, 25.7, 22.0, 21.3, 13.9 ppm.

11B NMR (192.5 MHz, [D6]DMSO): δ = 9.85 ppm.

5.4.11. trans-4-Hydroxy-(S)-proline methylester hydrochloride (155)

NH2 O

O

HO

Cl−

155

Thionylchloride (48.33 ml, 662.6 mmol, 3.13 eq) was carefully dissolved in methanol (250

ml) at 0°C. Then, trans-4-hydroxy-L-proline (24.33 g, 211.3 mmol, 1.00 eq) was added to this

solution in small portions. The resulting mixture was stirred at 0°C for 0.5 h. After warming

to room temperature, the mixture was stirred for additional 24 h and subsequently

concentrated in vacuo. The residue was taken up in methanol and diethyl ether was added

until the product precipitated as a white solid. The solid was collected on a filter and washed

with small portions of ether. The mother liquor was concentrated and a second crystallization

furnished the product in quantitative yield.

Page 109: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

101

Yield: 38.00 g (209.2 mmol, 99%)

1H NMR (600 MHz, [D6]DMSO): δ = 9.95 (br. s, 2H, NH2), 5.67 (br. s, 1H, OH), 4.42 (m, 2H,

CHOH and CHC(O)OCH3), 3.73 (s, 3H, C(O)OCH3), 3.35 (dd {2JHH = 12.1 Hz, 3JHH = 3.6

Hz}, 1H, NH2CHHCH(OH)), 3.03 (br. d {2JHH = 12.1 Hz}, 1H, NH2CHHCH(OH)), 2.18 (m,

1H, CH(OH)CHHCHC(O)OCH3), 2.07 (m, 1H, CH(OH)CHHCHC(O)OCH3) ppm.

13C NMR (150 MHz, [D6]DMSO): δ = 169.5, 69.8, 58.4, 53.7, 53.0, 37.2 ppm.

5.4.12. (2S,4R)-4-Hydroxy-1-benzoylproline methylester (156)

NO

O

HO

O

156

Triethylamine (29.0 ml, 206.5 mmol, 2.50 eq) was added to a solution of trans-4-hydroxy-(S)-

proline methylester hydrochloride (155) (15.00 g, 82.6 mmol, 1.00 eq) in dichloromethane

(90 ml, 6 ml⋅g−1 ester). The resulting mixture was stirred for 1.5 h at room temperature. After

addition of DMAP (0.463 g, 4.1 mmol, 5 mol-%), benzoylchloride (9.52 ml, 82.6 mmol, 1.00

eq) was added dropwise. The reaction mixture was stirred for 12 h at room temperature. The

resulting suspension was washed with water (30 ml), hydrochloric acid (5%, 20 ml), 0.5 M

NaOH (20 ml) and again water (30 ml). The organic phase was dried over Na2SO4 and

concentrated in vacuo. The resulting residue was taken up in a small amount of

dichloromethane. From this solution, the product was isolated as a white solid through

precipitation with pentane.

Yield: 16.48 g (66.1 mmol, 80%)

1H NMR (300 MHz, CDCl3): δ = 7.55 (m, 2H, o-H), 7.40 (m, 3H, m-H and p-H), 4.82 (m, 1H,

CHCO(O)CH3), 4.39 (m, 1H, CH(OH)), 3.75 (s, 3H, C(O)CH3), 3.50 (m, 2H,

Page 110: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

102

NCHHCH(OH)), 3.27 (s, 1H, CH(OH)), 2.35 (m, 1H, CH(OH)CHHCHC(O)OCH3), 2.10 (m,

1H, CH(OH)CHHCHC(O)OCH3) ppm.

13C NMR (75.5 MHz, CDCl3): δ = 172.9, 170.3, 135.5, 130.4, 128.3, 127.5, 70.2, 58.0, 52.4,

37.7 ppm.

5.4.13. (2S,4R)-4-{[1-(tert-Butyl)-1,1’-dimethylsilyl]-oxy}-1-benzoylproline methylester (157)

NO

O

O

O

Si

157

(2S,4R)-4-Hydroxy-1-benzoylproline methylester (156) (10.00 g, 40.12 mmol, 1.00 eq) and

imidazole (6.83 g, 100.3 mmol, 2.50 eq) were dissolved in DMF (50 ml, 5 ml⋅g−1 alcohol) and

stirred at room temperature for 3 h. Then, a solution of tert-butyl-dimethylsilyl chloride in

toluene (50 wt-%, 17.5 ml, 50.15 mmol, 1.25 eq) was added dropwise. The resulting mixture

was stirred for 40 h at room temperature. After concentration in vacuo, the residue was

dissolved in ether (60 ml) and washed with water (30 ml), diluted acetic acid (1M, 30 ml),

saturated NaHCO3 solution (30 ml) and again water (2 x 30 ml). After that, the organic phase

was dried over Na2SO4 and the solvent was removed in vacuo. The obtained colorless residue

was used in the next step without further purification.

Yield: 14.37 g (39.53 mmol, 99%)

1H NMR (300 MHz, CDCl3): δ = 7.52 (m, 2H, Ar), 7.36 (m, 3H, Ar), 4.78 (t {3JHH = 8.2 Hz},

1H, CHC(O)OCH3), 4.40 (m, 1H, CHOSi), 3.76 (s, 3H, C(O)OCH3), 3.72 (m, 1H,

NCHHCHOSi), 3.37 (m, 1H, NCHHCHOSi), 2.25 (m, 1H, SiOCHCHHCHC(O)OCH3), 2.05

(m, 1H, SiOCHCHHCHC(O)OCH3), 0.79 (s, 9H, SiC(CH3)3), 0.00 (s, 3H, SiCH3), −0.08 (s,

3H, SiCH3) ppm.

Page 111: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

103

13C NMR (75.5 MHz, CDCl3): δ = 172.9, 170.2, 135.9, 130.3, 128.3, 127.4, 70.7, 58.1, 57.9,

52.3, 38.4, 25.6, 17.8, −4.8, −5.0 ppm.

5.4.14. (2S,4R)-4-{[1-(tert-Butyl)-1,1’-dimethylsilyl]-oxy}-1-benzyl-2-hydroxymethyl-

pyrrolidine (158)

NOH

OSi

158

A suspension of LAH (3.445 g, 90.77 mmol, 3.30 eq) in THF (55 ml, 16 5 ml⋅g−1 LAH) was

heated under reflux for 15 min. Then, a solution of (2S,4R)-4-{[1-(tert-butyl)-1,1’-

dimethylsilyl]-oxy}-1-benzoylproline methylester (157) (10.00 g, 27.51 mmol, 1.00 eq) in

THF (15 ml, 1.5 ml⋅g−1 ester) was added at a rate keeping the mixture under gentle reflux.

After complete addition, the mixture was heated under reflux for 1 h, cooled to room

temperature and carefully hydrolyzed with 20%-KOH solution (15 ml, 1.5 ml⋅g−1 157). The

solid was filtered off and extracted with ether under reflux for 1 h. The combined organic

phase was concentrated and ether (50 ml, 5 ml⋅g−1 ester) was added. After drying over

Na2SO4 and evaporation of the volatiles under reduced pressure, the obtained colorless oil

was used in the next step without further purification.

Yield: 8.75 g (27.41 mmol, 99%)

1H NMR (300 MHz, CDCl3): δ = 7.28 (m, 5H, Ar), 4.25 (m, 1H, CHOSi), 3.95 (d {2JHH = 12.9

Hz}, 1H, NCHHPh), 3.63 (dd {2JHH = 11.1 Hz, 3JHH = 3.4 Hz}, 1H, CHH(OH)), 3.45 (d {2JHH

= 12.9 Hz}, 1H, NCHHPh), 3.36 (dd {2JHH = 11.1 Hz, 3JHH = 1.5 Hz}, 1H, CHHOH), 3.11

(dd {2JHH = 9.8 Hz, 3JHH = 5.5 Hz}, 1H, NCHHCHOSi), 3.05 (m, 1H, NCHCH2), 2.34 (dd

{2JHH = 9.8 Hz, 3JHH = 5.7 Hz}, 1H, NCHHCHOSi), 2.05 (m, 1H, NCHCHH), 1.82 (m, 1H,

NCHCHH), 0.86 (s, 9H, SiC(CH3)3), 0.02 (s, 3H, SiCH3), 0.00 (s, 3H, SiCH3) ppm.

Page 112: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

104

13C NMR (75.5 MHz, CDCl3): δ = 139.1, 128.7, 128.4, 127.1, 70.7, 68.0, 63.4, 62.3, 61.0,

58.7, 37.9, 25.8, 18.0, −4.7, −4.8 ppm.

5.4.15. (3R,5R)-1-Benzyl-3-(tert-butyl-dimethylsilanyloxy)-5-chloro-piperidine (159)

N

ClOSi

159

Methanesulfonylchloride (0.904 ml, 11.68 mmol, 1.10 eq) and triethylamine (5.93 ml, 42.48

mmol, 4.00 eq) were added dropwise to a solution of (2S,4R)-4-{[1-(tert-butyl)-1,1’-

dimethylsilyl]-oxy}-1-benzyl-2-hydroxymethylpyrrolidine (158) (3.415 g, 10.62 mmol, 1.00

eq) in THF (87 ml, 8.2 ml⋅mmol−1 alcohol) at 0°C. The mixture was heated under reflux for

12 h. After cooling to room temperature, the reaction mixture was poured into 2.5 M NaOH

solution (30 ml). After extraction with dichloromethane, the combined organic phase was

dried over Na2SO4 and concentrated in vacuo. The resulting dark brown oil was purified via

column chromatography with pentane/ethyl acetate = 3/1.

Yield: 2.058 g (6.05 mmol, 57%)

1H NMR (300 MHz, CDCl3): δ = 7.28 (m, 5H, Ar), 4.34 (m, 1H, CHOSi), 4.09 (m, 1H,

CHCl), 3.68 (d {2JHH = 13.7 Hz}, 1H, NCHHPh), 3.50 (d {2JHH = 13.7 Hz}, 1H, NCHHPh),

2.68 (dd {2JHH = 11.7 Hz, 3JHH = 2.7 Hz}, 1H, NCHHCHOSi), 2.55 (m, 2H, NCHHCHOSi

and NCHHCHCl), 2.33 (dd {2JHH = 11.3 Hz, 3JHH = 6.1 Hz}, 1H, NCHHCHCl), 1.92 (m, 2H,

SiOCHCHHCHCl), 0.86 (s, 9H, SiC(CH3)3), 0.02 (s, 3H, SiCH3), 0.00 (s, 3H, SiCH3) ppm.

13C NMR (75.5 MHz, CDCl3): δ = 137.9, 128.8, 128.2, 127.0, 65.9, 62.0, 59.7, 59.6, 54.8,

42.3, 25.8, 18.1, −4.7, −4.8 ppm.

GC: t = 12.96 min

MS (CI): m/z (%) = 340 (M+1, 100), 304 (49), 282 (17).

Page 113: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

105

MS (EI): m/z (%) = 339 (M, 14), 304 (77), 296 (17), 290 (12), 282 (95), 134 (17), 91 (100),

73 (12).

HR-MS: (m/z)obs = 339.1785 (Δ = 0.2 ppm)

IR (kap): ν = 2943 (vs), 2794 (w), 2352 (m), 1645 (m), 1461 (m), 1369 (w), 1258 (s), 1096

(vs), 1038 (vs), 836 (vs), 782 (m) cm−1.

5.4.16. (3R,5R)-1-Benzyl-5-chloro-piperidin-3-ol (160)

N

ClHO

160

(3R,5R)-1-benzyl-3-(tert-butyl-dimethylsilanyloxy)-5-chloro-piperidine (159) (339.9 mg, 1.00

mmol, 1.00 eq) was dissolved in THF (15 ml, 15 ml⋅mmol−1 159) at room temperature. Then,

a solution of tetra-n-butylammonium fluoride in dichloromethane (1 M, 2.00 ml, 2.00 mmol,

2.00 eq) was added dropwise. After stirring at room temperature overnight, the reaction

solution was poured into a biphasic mixture of aqueous NH4Cl and ether. The organic phase

was separated and the resulting aqueous phase was extracted with ether. Then, the combined

organic phases were washed with water and dried over Na2SO4. After removal of the solvent

in vacuo, the remaining residue was purified via column chromatography with pentane/ethyl

acetate = 1/1.

Yield: 183.1 mg (0.811 mmol, 81%)

1H NMR (300 MHz, CDCl3): δ = 7.33 (m, 5H, Ar), 4.25 (m, 1H, CHCl), 4.02 (m, 1H, OH),

3.62 (s, 2H, NCH2Ph), 3.13 (m, 1H, CHOH), 2.81 (m, 1H, HOCHCHHCHCl), 2.30 (m, 4H,

NCH2CHOH and NCH2CHCl), 1.72 (m, 1H, HOCHCHHCHCl) ppm.

Page 114: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

106

13C NMR (75.5 MHz, CDCl3): δ = 137.2, 129.0, 128.5, 127.5, 66.0, 62.1, 60.9, 58.6, 52.5, 41.6

ppm.

MS (CI): m/z (%) = 226 (M+1, 100), 190 (16).

MS (EI): m/z (%) = 225 (M, 66), 190 (13), 148 (16), 136 (15), 134 (58), 120 (10), 91 (100).

HR-MS: (m/z)obs = 225.0921 (Δ = 0.4 ppm)

IR (KBr): ν = 3177 (vs), 2932 (vs), 2840 (w), 2779 (vs), 2365 (s), 2342 (m), 1443 (s), 1353

(w), 1293 (vs), 1142 (m), 1075 (m), 1000 (s), 880 (w), 794 (m), 746 (s), 687 (vs), 600 (w),

482 (s) cm−1.

5.5. Catalysis

5.5.1. Screening of Lewis Base-Brønsted Acid Catalysts in the Aza-Baylis-Hillman Reaction

As prototypical example, the screening of (R,S)-2-diphenylphosphanyl-

cyclohexanecarboxylic acid (140) in the coupling of N-(4-Bromobenzylidene)-4-

methylbenzenesulfonamide (105) with MVK (37) is described:

Br

N TosO

+

Br

NH OTos

COOH

PPh2

r.t., THF, 24 h

105 37

140

106

N-(4-Bromobenzylidene)-4-methylbenzenesulfonamide (105) (51.0 mg, 0.125 mmol, 1.00 eq)

and (R,S)-2-diphenylphosphanyl-cyclohexanecarboxylic acid (140) (3.90 mg, 0.0125 mmol,

0.10 eq) were dissolved in THF under inert gas conditions. After addition of MVK (37) (12.5

μl, 0.150 mmol, 1.20 eq), the reaction solution was stirred at room temperature for 24 h. Then,

all volatiles were removed in vacuo. The composition of the resulting residue was analyzed

via 1H NMR spectroscopy (conversion) and HPLC (enantiomeric excess).

Page 115: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

107

Conversion: 86%

Enantiomeric excess: rac

5.5.2. Kinetic Experiments

For all kinetic experiments, stock solutions were prepared to minimize the weighing error.

They were stored under Argon at −10°C.

a) Without Brønsted acid

F

N TosO

+

PPh3, C6F6

THF, r.t.

F

NH OTos

102 37 103

53 104

Prototypical example:

N-(3-Fluorobenzylidene)-4-methylbenzenesulfonamide (102) (16.60 mg, 59.9 μmol, 1.00 eq),

PPh3 (53) (7.88 mg, 30.1 μmol, 0.50 eq) and C6F6 (104) (2.71 mg, 14.6 μmol, 0.24 eq) were

dissolved in THF (0.60 ml) at room temperature. The reaction was performed in a sealed

NMR tube under Argon. After addition of MVK (37) (7.18 μl, 86.3 μmol, 1.44 eq),

conversion was monitored via 19F NMR spectroscopy for approximately 10 h.

Starting concentrations:

Partial order imine 102: c(53) = 0.0501 mol⋅l−1, c(37) = 0.1438 mol⋅l−1, c(104) = 0.0243

mol⋅l−1, c(102) = 0.0998-0.2530 mol⋅l−1.

Partial order MVK (37): c(53) = 0.0500 mol⋅l−1, c(102) = 0.1506 mol⋅l−1, c(104) = 0.0214

mol⋅l−1, c(37) = 0.1712 – 0.3163 mol⋅l−1.

Page 116: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

108

Partial order PPh3 (53): c(37) = 0.1531 mol⋅l−1, c(102) = 0.1514 mol⋅l−1, c(104) = 0.0285

mol⋅l−1, c(53) = 0.0724 – 0.1595 mol⋅l−1.

b) With Brønsted acid = phenol

F

N TosO

+

PPh3, phenol, C6F6

THF, r.t.F

NH OTos

102 37 103

53 111 104

Prototypical example:

N-(3-Fluorobenzylidene)-4-methylbenzenesulfonamide (102) (16.01 mg, 57.7 μmol, 1.00 eq),

PPh3 (53) (4.63 mg, 17.6 μmol, 0.31 eq), C6F6 (104) (3.81 mg, 20.5 μmol, 0.36 eq) and phenol

(111) (8.94 mg, 95.0 μmol, 1.65 eq) were dissolved in THF (0.60 ml) at room temperature.

The reaction was performed in a sealed NMR tube under Argon. After addition of MVK (37)

(7.50 μl, 90.1 μmol, 1.56 eq), conversion was monitored on-line via 19F NMR spectroscopy

for approximately 10 h.

Starting concentrations:

Partial order imine 102: c(53) = 0.0294 mol⋅l−1, c(37) = 0.1502 mol⋅l−1, c(111) = 0.1584

mol⋅l−1, c(104) = 0.0341 mol⋅l−1, c(102) = 0.0962 – 0.1479 mol⋅l−1.

Partial order MVK (37): c(53) = 0.0153 mol⋅l−1, c(111) = 0.1622 mol⋅l−1, c(104) = 0.0234

mol⋅l−1, c(102) = 0.1532 mol⋅l−1, c(37) = 0.1474 – 0.2211 mol⋅l−1.

Partial order phenol (111): c(102) = 0.1264 mol⋅l−1, c(37) = 0.1527 mol⋅l−1, c(104) = 0.0350

mol⋅l−1, c(53) = 0.0154 mol⋅l−1, c(111) = 0.087 – 0.166 mol⋅l−1.

Page 117: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

109

5.5.3. Screening of Brønsted Acids for Kinetics

For all kinetic experiments, stock solutions were prepared to minimize the weighing error.

They were stored under Argon at −10°C.

As prototypical example, the evaluation of benzoic acid as co-catalyst in the PPh3-

catalyzed coupling of N-(4-bromobenzylidene)-4-methylbenzenesulfonamide (105) with

methyl vinyl ketone (37) in [D8]THF is described:

N TosO

+

PPh3, benzoic acid

[D8] THF, r.t.

NH OTos

Br Br

105 10637

53

Benzoic acid (12.6 mg, 0.103 mmol, 1.375 eq) was added to a mixture of N-(4-

bromobenzylidene)-4-methylbenzenesulfonamide (105) (25.4 mg, 0.075 mmol, 1.00 eq) and

PPh3 (53) (2.0 mg, 0.0075 mmol, 0.10 eq) in deuterated THF (0.60 ml) at room temperature.

The reaction was performed in a sealed NMR tube under Argon. After addition of MVK (37)

(7.5 μl, 0.090 mmol, 1.20 eq), conversion was monitored online via 1H NMR spectroscopy for

700 min.

The following compounds were also tested as additives: N-[1-(4-bromophenyl)-2-methylene-

3-oxobutyl]-4-methylbenzenesulfonamide (42.1 mg), methanol (3.3 mg), phenol (9.7 mg), N-

methyltosylamide (19.1 mg), bis-3,5-(trifluormethyl)phenol (23.7 mg), water (1.9 mg), malic

acid (13.8 mg).

5.5.4. Catalysis in Chiral Ionic Liquids

As prototypical example, the PPh3-catalyzed coupling of N-(4-bromobenzylidene)-4-

methylbenzenesulfonamide (105) with MVK (37) in methyltrioctyl ammonium

dimalatoborate (166) is described:

Page 118: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

110

Br

N TosO

+

Br

NH OTos

166, r.t., 24 h

PPh3

53

105 10637

N-(4-Bromobenzylidene)-4-methylbenzenesulfonamide (105) (42.3 mg, 0.125 mmol, 1.00 eq)

and catalyst 3 (3.3 mg, 12.5 μmol, 0.10 eq) were dissolved in methyltrioctyl ammonium

dimalatoborate (166) (1.0 ml, 6.7 ml⋅mmol−1 imine). After addition of MVK (37) (12.5 μl,

0.150 mmol, 1.20 eq), the reaction was stirred at room temperature for 24 h. After removing

excess MVK (37) in vacuo, a small sample of the reaction mixture was dissolved in

[D6]DMSO to determine the conversion by 1H NMR spectroscopy. The main part of the

reaction mixture was dissolved in a 1:1 mixture of water and methanol (5 ml). Further

methanol was added to obtain a clear solution of the reaction mixture. After (partly)

exchanging the cation on an ion-exchange resin (contact time app. 5 min), all volatiles were

removed in vacuo.[156]The resulting solid was extracted with heptane/iPrOH (85:15, 1 ml) and

the solution was directly analyzed by chiral HPLC to determine the ee of the product.

Conversion: 34 – 39%

Enantiomeric excess: 71 – 84% ee

5.5.5. Screening of Chiral Brønsted Acidic Additives

As prototypical example, the PPh3-catalyzed coupling of N-(4-bromobenzylidene)-4-

methylbenzenesulfonamide (105) with MVK (37) using L-(−)-malic acid (162) as additive

is presented:

Br

N TosO

+

Br

NH OTos

THF, r.t., 24 h

PPh3, L-(−)-malic acid

105 37 106

53 162

N-(4-Bromobenzylidene)-4-methylbenzenesulfonamide (105) (42.3 mg, 0.125 mmol, 1.00

eq), triphenylphosphine (53) (3.3 mg, 0.0125 mmol, 0.10 eq) and L-(−)-malic acid (162) (1.7

Page 119: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

111

mg, 0.0125 mmol, 0.10 eq) were dissolved in THF (1.0 ml, 6.7 ml⋅mmol−1 105). After

addition of MVK (37) (12.5 μl, 10.5 mg, 0.150 mmol, 1.20 eq), the reaction mixture was

stirred at room temperature for 24 h. After the reaction, all volatile components were removed

in vacuo. The remaining solid was analyzed by 1H NMR (conversion) and chiral HPLC

(enantiomeric excess).

Conversion: 8%

Enantiomeric excess: ee = rac

5.6. Miscellaneous

5.6.1. Cross-Mannich Experiment

PPh3, phenol

THF, r.t.

NH OTos

Br F

N Tos

+

F

NH OTos

Br

N Tos

+

106 105 103102

53 111

N-[1-(4-Bromophenyl)-2-methylene-3-oxobutyl]-4-methylbenzenesulfonamide (106) (30.0

mg, 0.0735 mmol, 1.00 eq) and N-(3-fluorobenzylidene)-4-methylbenzenesulfonamide (102)

(203.7 mg, 0.735 mmol, 10.0 eq) were dissolved in THF (2 ml, 2.7 ml⋅mmol−1 102) under

Argon. After addition of triphenylphosphine (53) (19.3 mg, 0.0735 mmol, 1.00 eq) and

phenol (111) (34.6 mg, 0.367 mmol, 5.00 eq), the mixture was stirred at room temperature for

n h. The mixture was analyzed by 1H and 19F NMR spectroscopy.

n = 24: no cross-Mannich product (103)

n = 72: no cross-Mannich product (103)

Page 120: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

112

5.6.2. Deuteration Experiment

PPh3

CD3OD, r.t., 24h

O

Br

O

Br

NHTos NDTosH D

D D

106 [D4]106

53

N-[1-(4-Bromophenyl)-2-methylene-3-oxobutyl]-4-methylbenzenesulfonamide (106) (20.4

mg, 0.050 mmol, 1.00 eq) and triphenylphosphine (53) (13.1 mg, 0.050 mmol, 1.00 eq) were

dissolved in [D4]methanol (1.0 ml, 20 ml⋅mmol−1) under Argon. After stirring the reaction

mixture at room temperature for 24 h, all volatiles were removed in vacuo. The residue was

dissolved in CDCl3 and characterized via NMR spectroscopy.

1H NMR (300 MHz, CDCl3): δ = 7.55 (d {3JHH = 8.0 Hz}, 2H, Ar), 7.22 (d {3JHH = 7.8 Hz},

2H, Ar), 7.16 (d {3JHH = 8.0 Hz}, 2H, Ar), 6.88 (d {3JHH = 7.8 Hz}, 2H, Ar), 6.00 (m, 3H,

CHH and CHNH), 5.16 (m, 1H, CHNH), 2.33 (s, 3H, tos-CH3), 2.05 (s, 3H, C(O)CH3) ppm.

The bold red signals showed a decrease in intensity relative to the other peaks.

5.6.3. Racemization Experiment

PPh3, phenol

THF, r.t.

NH OTos

Br

NH OTos

Br(S)-106 (R)-106

53 111

As prototypical example, the racemization of (S)-N-[1-(4-bromophenyl)-2-methylene-3-

oxobutyl]-4-methylbenzenesulfonamide (106) with the combined system PPh3

(53)/phenol (111) is presented:

To a mixture of (S)-N-[1-(4-bromophenyl)-2-methylene-3-oxobutyl]-4-methylbenzene-

sulfonamide (106) (38.1 mg, 0.0933 mmol, 1.00 eq) and triphenylphosphine (53) (24.5 mg,

0.0933 mmol, 1.00 eq) in THF (0.75 ml, 8.0 ml⋅mmol−1), phenol (111) (26.3 mg, 0.280 mmol,

3.00 eq) was added. The resulting solution was stirred at room temperature for n h and

samples were analyzed by chiral HPLC.

Page 121: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

113

6. Appendices

6.1. Abbreviations

A acid

Ac acetyl, acetate

Ar aryl

B base

BH Baylis-Hillman

BINAP 2,2’-Bis(diphenylphosphino)-1,1’-binaphthyl

BINOL 1,1’-Bi-2-naphthol

Bmim butylmethyl imidazolium

Bn benzyl

Boc tert-butoxycarbonyl

Bu butyl

c concentration

conv conversion

calc calculated

cat. catalyst, catalytic

CI chemical ionization

COSY Correlated Spectroscopy, correlation spectrum

d doublet, day(s)

DABCO 1,4-diazabicyclo[2.2.2]octane

DBU 1,8-diazabicyclo[5.4.0]undec-7-ene

DCM dichloromethane

DFT Density Functional Theory

DMAP 4-dimethylaminopyridine

DMF N,N-dimethylformamide

DMSO dimethylsulfoxide

ee enantiomeric excess

EI electron ionisation

er enantiomeric ratio

Et ethyl

eq equivalent(s)

Page 122: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

114

EVK ethylvinyl ketone

EWG electron withdrawing group

GC gas chromatography

h hour(s)

HPLC High Performance Liquid Chromatography

HR-MS High Resolution Mass Spectroscopy

HX acid

Hz Hertz

IL ionic liquid

IR infrared

J coupling constant

LAH lithium aluminiumhydride

Nos p-nitrobenzenesulfonyl

Nu nucleophile

m multiplet, medium

m meta

M molar

M+ molecule ion

Me methyl

Mes methanesulfonyl

min minute(s)

MVK methylvinyl ketone

MS Mass Spectroscopy, mass spectrum, molecular sieves

MTOA methyltrioctyl ammonium

n normal

n.a. not available

NMR Nuclear Magnetic Resonance

NOE Nuclear Overhauser Effect

NOESY Nuclaer Overhauser Effect Spectroscopy

o ortho

obs observed

p para

P product

PG protecting group

Page 123: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

115

Ph phenyl

ppm parts per million

PTA 1,3,5-triaza-7-phosphaadamantane

QUINAP 1-(2-Diphenylphosphino-1-naphthyl)isoquinoline

q quartet

quant. quantitative

r reaction rate

r(0) initial reaction rate

R organic rest

RDS rate determining step

Rf ratio of fronts (TLC)

rfl. reflux

r.t. room temperature

s singulet, strong

S substrate

SES β-trimethylethanesulfonyl

t time, triplet

t tertiary

T transmission

TADDOL trans-α,α'-(Dimethyl-1,3-dioxolane-4,5-diyl)bis(diphenylmethanol)

TBS tert-butyldimethylsilyl

Tf trifluormethanesulfonyl

THF tetrahydrofurane

TLC Thin Layer Chromatography

TOCSY Total Correlated Spectroscopy

Tos p-toluenesulfonate

TS transition state

UV ultraviolet

v very

vs very strong

w weak

X halide

y yield

* symbol for a stereogenic center

Page 124: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

116

6.2. List of Commercially Obtained Chemicals

Reagent Supplier Quality 3,5-bis-(trifluoromethyl)phenol Aldrich 95% Aliquat 336 Acros unknown benzaldehyde Avocado 99+% benzoic acid Avocado 99% benzoylchloride Avocado 99% boric acid Merck 99% p-bromobenzaldehyde Aldrich 99% 2-cyclopentene-1-one Avocado 98%

1,4-diazabicyclo[2.2.2]octane Avocado 97% p-dimethylaminopyridine Aldrich puriss, >99.0% diphenylacetic acid Avocado 99% m-fluorobenzaldehyde Aldrich 97% hexafluorobenzene Aldrich 99.9% imidazole Avocado 99% lithiumaluminium hydride Aldrich 95%, powder L-(-)-malic acid Aldrich 97% methanesulfonyl chloride Aldrich >99.7% methyl acrylate Avocado 99%

methylvinyl ketone Aldrich 99% p-nitrobenzaldehyde Aldrich 98% (pentafluorophenyl)diphenyl phosphine

Aldrich 96%

phenol Aldrich >99% tert-butyldimethylsilyl chloride Aldrich solution, 50% (w/w) in

toluene tetraethoxysilane Aldrich >99% tetra-n-butylammonium fluoride Aldrich solution, 1M in THF thionylchloride Aldrich 99% p-tolualdehyde Aldrich 97% p-toluenesulfonamide Avocado 98+% trans-4-hydroxy-L-proline Acros 99+% triethylamine Avocado 99% triphenylphosphine Acros 99% tris-(o-tolyl)phosphine Strem 99%

Page 125: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

117

6.3. Kinetic Study with Phenol as Co-Catalyst

Teilordnung MVK_Add

r(0) = 0,0047c(0)R2 = 0,9664

0,00E+00

2,00E-04

4,00E-04

6,00E-04

8,00E-04

1,00E-03

1,20E-03

1,40E-03

0,000 0,050 0,100 0,150 0,200 0,250 0,300

c(0) in [mol/l]

r(0)

in [m

ol/(l

*min

)]

Teilordnung MVK_Add

r(0) = 0,0047c(0)R2 = 0,9664

0,00E+00

2,00E-04

4,00E-04

6,00E-04

8,00E-04

1,00E-03

1,20E-03

1,40E-03

0,000 0,050 0,100 0,150 0,200 0,250 0,300

c(0) in [mol/l]

r(0)

in [m

ol/(l

*min

)]

Teilordnung MVK_Add

r(0) = 0,0047c(0)R2 = 0,9664

0,00E+00

2,00E-04

4,00E-04

6,00E-04

8,00E-04

1,00E-03

1,20E-03

1,40E-03

0,000 0,050 0,100 0,150 0,200 0,250 0,300

c(0) in [mol/l]

r(0)

in [m

ol/(l

*min

)]

Teilordnung MVK_Add

r(0) = 0,0047c(0)R2 = 0,9664

0,00E+00

2,00E-04

4,00E-04

6,00E-04

8,00E-04

1,00E-03

1,20E-03

1,40E-03

0,000 0,050 0,100 0,150 0,200 0,250 0,300

c(0) in [mol/l]

r(0)

in [m

ol/(l

*min

)]

Phenol

r(0)= 0,0045c(0)R2 = 0,9861

0,000E+00

2,000E-04

4,000E-04

6,000E-04

8,000E-04

1,000E-03

1,200E-03

1,400E-03

0,000 0,050 0,100 0,150 0,200 0,250 0,300

c(0) in [mol/l]

r(0)

in [m

ol/l]

Page 126: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

6.4. 1H and 13C NMR Spectra of Compound 160

1.52.02.53.03.54.04.55.05.56.06.57.07.5 ppm

N

ClHO

160

1.52.02.53.03.54.04.55.05.56.06.57.07.5 ppm

N

ClHO

160

Page 127: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

119

4550556065707580859095100105110115120125130135 ppm

41.6

4

52.5

4

58.5

960

.87

62.1

1

65.9

8

76.6

177

.04

77.4

6

127.

5112

8.48

129.

00

137.

19

Page 128: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

6.5. HPLC Chromatogram of Aza-BH Product 106

a) Racemate

Page 129: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

121

b) Authentic sample (reaction in IL 166)

Page 130: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

6.6. 31P NMR Spectrum of Compound 172

-10-525 20 15 10 5 0 ppm

-14.

19-1

4.13

-14.

07

26.1

626

.25

26.4

3

Page 131: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

123

6.7. 31P NMR Spectrum of Compound 173

-10-530 25 20 15 10 5 0 ppm

-13.

7-1

3.6

-13.

4-1

3.4

-13.

3

20.0

22.3

24.9

26.0

26.2

27.8

28.7

29.1

30.9

Page 132: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

7. Literature

1 P. Perlmutter, C. C. Teo, Tetrahedron Lett. 1984, 25, 5951-5952. 2 D. Basavaiah, A. J. Rao, T. Satayanarayana, Chem. Rev. 2003, 103, 811-891 and

references cited therein. 3 K. A. Ahrendt, C. J. Borths, D. W. C. MacMillan, J. Am. Chem. Soc. 2000, 122, 4243-

4244. 4 For a recent review on organocatalysis, see: P. I. Dalko, L. Moisan, Angew. Chem. 2004,

116, 5248-5286; Angew. Chem. Int. Ed. 2004, 43, 5138-5175. 5 S. Pizzarello, A. L. Weber, Science 2004, 303, 1151. 6 H. Prajecus, Justus Liebigs Ann. Chem. 1960, 634, 9-22. 7 H. Prajecus, Justus Liebigs Ann. Chem. 1960, 634, 23-29. 8 Z. G. Hajos, D. R. J. Parrish, J. Org. Chem. 1974, 39, 1615-1621. 9 U. Eder, G. Sauer, R. Wiechert, Angew. Chem. 1971, 83, 492-493; Angew. Chem. Int. Ed.

Engl. 1971, 10, 496-497. 10 (a) B. List, R. A. Lerner, C. F. Barbas III, J. Am. Chem. Soc. 2000, 122, 2395-2396;

(b) B. List, Tetrahedron 2002, 58, 5573-5590. 11 A. Berkessel, H. Gröger, Asymmetric Organocatalysis – From Biomimetic Concepts to

Applications in Asymmetric Synthesis, Wiley-VCH, Weinheim, 2005. 12 J. Seayad, B. List, Org. Biomol. Chem. 2005, 3, 719-724. 13 For a recent review on enamine/iminium-catalysis, see: B. List, Chem. Commun. 2006,

819-824. 14 For example: (a) B. List, J. Am. Chem. Soc. 2000, 122, 9336-9337; (b) Y. Hayashi, W.

Tsuboi, M. Shoji, N. Suzuki, J. Am. Chem. Soc. 2003, 125, 11208-11209; (c) W. Notz, F.

Tanaka, S. Watanabe, N. S. Chaudhari, J. M. Turner, R. Thayumanavan, C. F. Barbas III,

J. Org. Chem. 2003, 68, 9624-9634. 15 For example: B. List, P. Pojarliev, H. J. Martin, Org. Lett. 2001, 3, 2423-2425; (b) D.

Enders, A. Seki, Synlett 2002, 26-28; (c) N. Halland, T. Hansen, K. A. Jørgensen, Angew.

Chem. 2003, 115, 5105-5107; Angew. Chem. Int. Ed. 2003, 42, 4955-4957. (d) A. J. A.

Cobb, D. A. Longbottom, D. M. Shaw, S. V. Ley, Chem. Commun. 2004, 1808-1809; (e)

T. J. Peelen, Y. Chi, S. H. Gellman, J. Am. Chem. Soc. 2005, 127, 11598-11599. 16 For example: (a) B. List, J. Am. Chem. Soc. 2002, 124, 5656-5657; (b) N.

Kumaragurubaran, K. Juhl, W. Zhuang, A. Bøgevig, K. A. Jørgensen, J. Am. Chem. Soc.

Page 133: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

125

2002, 124, 6254-6255; (c) A. Bøgevig, K. Juhl, N. Kumaragurubaran, W. Zhuang, K. A.

Jørgensen, Angew. Chem. 2002, 114, 1868-1871; Angew. Chem. Int. Ed. 2002, 41, 1790-

1792. 17 For example: (a) S. P. Brown, M. P. Brochu, C. J. Sinz, D. W. C. MacMillan, J. Am. Chem.

Soc. 2003, 125, 10808-10809; (b) G. Zhong, Angew. Chem. 2003, 115, 4379-4382;

Angew. Chem. Int. Ed. 2003, 42, 4247-4250; (c) Y. Hayashi, J. Yamaguchi, T. Sumiya,

M. Shoji, Angew. Chem. 2004, 116, 1132-1135; Angew. Chem. Int. Ed. 2004, 43, 1112-

1115; (d) A. Bøgevig, H. Sundéen, A. Córdova, Angew. Chem. 2004, 116, 1129-1132;

Angew. Chem. Int. Ed. 2004, 43, 1109-1112. 18 For example: (a) N. Vignola, B. List, J. Am. Chem. Soc. 2004, 126, 450-451; (b) A. Fu, B.

List, W. Thiel, J. Org. Chem. 2006, 71, 320-326; (c) I. Ibrahem, A. Cordova, Angew.

Chem. 2006, 118, 1986-1990; Angew. Chem. Int. Ed. 2006, 45, 1952-1956. 19 For example: (a) M. P. Brochu, S. P. Brown, D. W. C. MacMillan, J. Am. Chem. Soc.

2004, 126, 4108-4109; (b) N. Halland, A. Braunton, S. Bachmann, M. Marigo, K. A.

Jørgensen, J. Am. Chem. Soc. 2004, 126, 4790-4791; (c) M. Marigo, S. Bachmann, N.

Halland, A. Braunton, K. A. Jørgensen, Angew. Chem. 2004, 116, 5623-5626; Angew.

Chem. Int. Ed. 2004, 43, 5507-5510. 20 For example: (a) W. S. Jen, J. J. M. Wiener, D. W. C. MacMillan, J. Am. Chem. Soc.

2000, 122, 9874-9875; (b) S. Karlsson, H. E. Hoegberg, Eur. J. Org. Chem. 2003, 2782-

2791; (c) A. Puglisi, M. Benaglia, M. Cinquini, F. Cozzi, G. Cementano, Eur. J. Org.

Chem. 2004, 567-573. 21 For example: N. A. Paras, D. W. C. MacMillan, J. Am. Chem. Soc. 2001, 123, 4370-4371;

(b) J. F. Austin, D. W. C. MacMillan, J. Am. Chem. Soc. 2002, 124, 1172-1173; (c) N. A.

Paras, D. W. C. MacMillan, J. Am. Chem. Soc. 2002, 124, 7894-7895. 22 For example: (a) S. P. Brown, N. C. Goodwin, D. W. C. MacMillan, J. Am. Chem. Soc.

2003, 125, 1192-1194; (b) W. Wang, H. Li, J. Wang, Org. Lett. 2005, 7, 1637-1639; (c) J.

Rochibaud, F. Tremblay, Org. Lett. 2006, 8, 597-600. 23 For example: (a) J. C. Ruble, J. Tweddell, G. C. Fu, J. Org. Chem. 1998, 63, 2794-2795;

(b) S. J. Miller, C. T. Copeland, N. Papaioannou, T. E. Horstmann, E. M. Ruel, J. Am.

Chem. Soc. 1998, 120, 1629-1630; (c) S. J. Miller, Acc. Chem. Res. 2004, 37, 601-610. 24 For example: (a) M. A. Calter, R. K. Orr, W. Song, Org. Lett. 2003, 5, 4745; (b) N.

Bremeyer, S. C. Smith, S. V. Ley, M. J. Gaunt, Angew. Chem. 2004, 116, 2735-2738;

Angew. Chem. Int. Ed. 2004, 43, 2681-2684.

Page 134: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

126

25 For example: (a) J. S. Johnson, Angew. Chem. 2004, 116, 1348-1350; Angew. Chem. Int.

Ed. 2004, 43, 1326-1328; (b) D. Enders, T. Balensiefer, Acc. Chem. Res. 2004, 37, 534-

541; (c) D. Enders, U. Kalfass, Angew. Chem. 2002, 114, 1822-1824; Angew. Chem. Int.

Ed. 2002, 41, 1743-1744; (d) D. Enders, O. Niemeyer, T. Balensiefer, Angew. Chem.

2006, 118, 1491-1495; Angew. Chem. Int. Ed. 2006, 45, 1463-1467. 26 For example: (a) E. Vedesj, O. Daugulis, S. T. Diver, J. Org. Chem. 1996, 61, 430-431;

(b) G. Zhu, Z. Chen, Q. Jiang, D. Xiao, P. Cao, X. Zhang, J. Am. Chem. Soc. 1997, 119,

3836-3837; (c) J. L. Methot, W. R. Roush, Adv. Synth. Catal. 2004, 346, 1035-1050. 27 For example: (a) J. Zanardi, C. Leriverend, D. Aubert, K. Julienne, P. Metzner, J. Org.

Chem. 2001, 66, 5620-5623; (b) V. K. Aggarwal, C. L. Winn, Acc. Chem. Res. 2004, 37,

611-620. 28 For example: (a) U.-H. Dolling, P. Davis, E. J. J. Grabowski, J. Am. Chem. Soc. 1984, 118,

446-447; (b) M. J. O’Donnel, Acc. Chem. Res. 2004, 37, 506-517; (c) E. J. Corey, F. Xu,

M. C. Noe, J. Am. Chem. Soc. 1997, 119, 12414-12415. 29 For example: K. Moruoka, T. Ooi, Chem. Rev. 2003, 103, 3013-3028; (b) T. Ooi, K.

Moruoka, Acc. Chem. Res. 2004, 37, 526-533. 30 For example: (a) Y. Tu, Z.-X. Wang, Y. Shi, J. Am. Chem. Soc. 1996, 118, 9806-9807; (b)

H. Tian, X. She, J. Xu, Y. Shi, Org. Lett. 2001, 3, 1929-1931. (b) X.-Y. Wu, X. She, J.

Xu, Y. Shi, J. Am. Chem. Soc. 2002, 124, 8792-8793. 31 For recent reviews, see: (a) H. Gröger, Chem. Rev. 2003, 103, 2795-2827; (b) M. North,

Tetrahedron: Asymmetry 2003, 14, 147-176; (c) E. R. Jarvo, S. J. Miller, Tetrahedron

2002, 58, 2481-2495. 32 K. Tanaka, A. Mori, S. Inoue, J. Org. Chem. 1990, 55, 181-185. 33 For recent reviews on Brønsted acid catalysis, see: (a) P. R. Schreiner, Chem. Soc. Rev.

2003, 32, 289-296; (b) P. M. Pihko, Angew. Chem. 2004, 116, 2110-2113; Angew. Chem.

Int. Ed. 2004, 43, 2062-2064; (c) M. S. Taylor, E. N. Jacobsen, Angew. Chem. 2006, 118,

1550-1573; Angew. Chem. Int. Ed. 2006, 45, 1520-1543. 34 (a) P. Vachal, E. N. Jacobsen, Org. Lett. 2000, 2, 867-870; (b) M. S. Sigman, P. Vachal,

E. N. Jacobsen, Angew. Chem. 2000, 112, 1336-1338; Angew. Chem. Int. Ed. 2000, 39,

1279-1281; (c) P. Vachal, E. N. Jacobsen, J. Am. Chem. Soc. 2002, 124, 10012-10014 35 (a) A. G. Wenzel, E. N. Jacobsen, J. Am. Chem. Soc. 2002, 124, 12964-12965; (b) M. S.

Taylor, N. Tokunaga, E. N. Jacobsen, Angew. Chem. 2005, 117, 6858-6862; Angew.

Chem. Int. Ed. 2005, 44, 6700-6704.

Page 135: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

127

36 P. T. Yoon, E. N. Jacobsen, Angew. Chem. 2005, 117, 7493-7495; Angew. Chem. Int. Ed.

2005, 44, 466-468. 37 G. D. Joly, E. N. Jacobsen, J. Am. Chem. Soc. 2004, 126, 4102-4103. 38 I. T. Raheem, E. N. Jacobsen, Adv. Synth. Catal. 2005, 347, 1701-1708. 39 Y. Huang, K. A. Unni, A. N. Thadani, V. H. Rawal, Nature (London), 2003, 424, 146. 40 (a) N. T. McDougal, S. E. Schaus, J. Am. Chem. Soc. 2003, 125, 12094-12095; (b) N. T.

McDougal, W. L. Trevellini, S. A. Rodgen, L. T. Kliman, S. E. Schaus, Adv. Synth. Catal.

2004, 346, 1231-1240. 41 T. Okino, Y. Hoashi, Y. Takemoto, J. Am. Chem. Soc. 2003, 125, 12672-12673. 42 For example: (a) Y. Hoashi, T. Yabuta, P. Yuan, H. Miyabe, Y. Takemoto, Tetrahedron

2006, 62, 365-374; (b) S. H. McCooey, S. J. Connon, Angew. Chem. 2005, 117, 6525-

6528; Angew. Chem. Int. Ed. 2005, 44, 6367-6370. 43 For example: (a) T. Marcelli, R. N. S. van der Haas, J. H. van Maarseveen, H. Hiemstra,

Synlett 2005, 2817-2819; (b) T. Marcelli, R. N. S. van der Haas, J. H. Van Maarseveen, H.

Hiemstra, Angew. Chem. 2006, 118, 943-945; Angew. Chem. Int. Ed. 2006, 45, 929-931;

(c) Y. Sohtome, Y. Hashimoto, K. Nagasawa, Adv. Synth. Catal. 2005, 347, 1643-1648. 44 For example: (a) T. Okino, S. Nakamura, T. Furukawa, Y. Takemoto, Org. Lett. 2004, 6,

625-627; (b) X. Xu, T. Furukawa, T. Okino, H. Miyabe, Y. Takemoto, Chem. Eur. J.

2006, 12, 466-476. 45 For example: (a) S. B. Tsogoeva, D. A. Yalalov, M. J. Hateley, C. Weckbecker, K.

Huthmacher, Eur. J. Org. Chem. 2005, 4995-5000. 46 A. Berkessel, F. Cleemann, S. Mukherjee, T. N. Müller, J. Lex, Angew. Chem. 2005, 117,

817-821; Angew. Chem. Int. Ed. 2005, 44, 807-811. 47 For example: (a) J. Wang, H. Li, X. Yu, L. Zu, W. Wang, Org. Lett. 2005, 7, 4293-4296;

(b) Y. Iwabuchi, M. Nakatani, N. Yokoyama, S. Hatakeyama, J. Am. Chem. Soc. 1999,

121, 10219-10220. 48 For example: (a) K. Matsui, K. Tanaka, A. Horii, S. Takizawa, H. Sasai, Tetrahedron:

Asymmetry 2006, 17, 578-583; (b) M. Shi, L.-H. Chen, C.-Q. Li, J. Am. Chem. Soc. 2005,

127, 3790-3800. 49 K. Morita, Z. Suzuki, H. Hirose, Bull. Chem. Soc. Jpn. 1968, 41, 2815.

50 A. B. Baylis, M. E. D. Hillman, German Patent 2155113, 1972.

51 (a) J. S. Hill, N. S. Isaacs, Tetrahedron Lett. 1986, 27, 5007-5010; (b) J. S. Hill, N. S.

Isaacs, J. Chem. Res. 1988, 330.

Page 136: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

128

52 (a) M. E. Krafft, K. A. Seibert, T. F. N. Haxell, C. Hirosawa, Chem. Commun. 2005, 5274-

5276; (b) M. E. Krafft, T. F. N. Haxell, K. A. Seibert, K. A. Abboud, J. Am. Chem. Soc.

2006, 128, 4174-4175. 53 For example: (a) A. Ho, K. Cyrus, K.-B. Kim, Eur. J. Org. Chem. 2005, 4829-4834; (b)

M. Dadwal, R. Mohan, D. Panda, S. M. Mobin, I. N. N. Namboothiri, Chem. Commun.

2006, 338-340. 54 For example: (a) B. M. Trost, O. R. Thiel, H.-C. Tsui, J. Am. Chem. Soc. 2002, 124,

11616-11617; (b) W. P. Hong, K.-J. Lee, Synthesis 2006, 963-968. 55 For example: (a) D. Basavaiah, M. Bakthadoss, S. Pandiaraju, Chem. Commun. 1998,

1639-1640; (b) L. R. Reddy, P. Saravanan, E. J. Corey, J. Am. Chem. Soc. 2004, 126,

6230-6231; (c) B. M. Trost, M. H. Machacek, H. C. Tsui, J. Am. Chem. Soc. 2005, 127,

7014-7024; (d) D. Basavaiah, T. Satayanarayana, Chem. Commun. 2004, 32-33; (e) D. K.

O’Dell, K. M. Nicholas, J. Org. Chem. 2003, 68, 6427-6430. 56 P. J. Dunn, M. L. Hughes, P. M. Searle, A. S. Wood, Org. Proc. Res. Dev. 2003, 7, 244-

253. 57 For example: (a) R. J. W. Schuurman, A. van der Linden, R. P. F. Grimbergen, R. J. M.

Nolte, H. W. Scheeren, Tetrahedron 1996, 52, 8307-8314; (b) N. E. Leadbeater, C. van

der Pol, J. Chem. Soc. Perk. Trans. 1, 2001, 2831-2835. 58 For example: (a) S. Rafel, J. W. Leahy, J. Org. Chem. 1997, 62, 1521-1522. 59 For example: (a) T. Kataoka, T. Iwama, S.-I. Tsjujiyama, Chem. Commun. 1998, 197, (b)

T. Kataoka, T. Iwama, S.-i. Tsjujiyama, T. Iwamura, S.-I. Watanabe, Tetrahedron, 1998,

54, 11813. 60 For example: (a) M. Shi, J.-K. Jiang, Y.-S. Feng, Org. Lett. 2000, 2, 2397-2400; (b) M.

Shi, S.C. Cui, Chin. J. Chem. 2002, 20, 277-285. 61 L. S. Santos, C. H. Pavam, W. P. Almeida, F. Coelho, M. N. Eberlin, Angew. Chem. 2004,

116, 4430-4433; Angew. Chem. Int. Ed. 2004, 43, 4330-4331. 62 S. E. Drewes, O. L. Njamela, N. D. Emslie, N. Ramesar, J. S. Field, Synth. Commun. 1993,

23, 2807. 63 (a) M. E. Krafft, T. F. N. Haxell, K. A. Seibert, K. A. Abboud, J. Am. Chem. Soc.,

published on the web; (b) M. Kawamura, S. Kobayashi, Tetrahedron Lett. 1999, 40, 1539-

1542; (c) N. T. McDougal, S. E. Schaus, J. Am. Chem. Soc. 2003, 125, 12094-12095; (d)

X. Wang, S. Li, Y. Jiang, J. Phys. Chem. A. 2005, 109, 10770-10775.

Page 137: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

129

64 (a) R. K. Thalji, W. R. Roush, J. Am. Chem. Soc. 2005, 127, 16778-16779; (b) M. Shi, L.-

H. Chen, C.-Q. Li, J. Am. Chem. Soc. 2005, 127, 3790-3800; (c) K. Matsui, K. Tanaka, A.

Horii, S. Takizawa, H. Sasai, Tetrahedron: Asymmetry 2006, published on the web. 65 (a) Y. Forth, M. C. Berthe, P. Caubere, Tetrahedron 1992, 28, 6371-6384 and references

cited therein; (b) J. S. Hill, N. S. Isaacs, J. Phys. Org. Chem. 1990, 3, 285-288; (c) H. M.

R. Hoffman, J. Rabe, Angew. Chem. 1983, 95, 796-797; Angew. Chem. Int. Ed. Engl.

1983, 22, 795. 66 (a) J. S. Hill, N. S. Isaacs, J. Chem. Res. 1988, 330-331; (b) A. Gilbert, T. W. Heritage, N.

S. Isaacs, Tetrahedron: Asymmetry 1991, 2, 969-972; (c) P. M. Rose, A. A. Clifford, C.

M. Rayner, Chem. Commun. 2002, 968-969; (d) Y. Hayashi, K. Okado, I. Ashimini, M.

Shoji, Tetrahedron Lett. 2002, 43, 8683-8686. 67 (a) E. L. M. van Rozendaal, B. M. W. Voss, H. W. Scheeren, Tetrahedron 1993, 49, 6931-

6936; (b) T. Oishi, H. Oguri, M. Hirama, Tetrahedron: Asymmetry 1995, 6, 1241-1244;

(c) I. E. Marko, Paul R. Giles, N. J. Hindley, Tetrahedron 1997, 53, 1015-1024. 68 M. L. Bode, P. T. Kaye, Tetrahedron Lett. 1991, 32, 5611-5614. 69 (a) K. E. Price, S. J. Broadwater, M. H. Jung, D. T. McQuade, Org. Lett. 2005, 7, 147-

150; (b) K. E. Price, S. J. Broadwater, B. J. Walker, D. T. McQuade, J. Org. Chem. 2005,

70, 3980-3987. 70 (a) S. E. Drewes, N. D. Emslie, N. Karodia, A. A. Khan, Chem. Ber. 1990, 123, 1447-

1448; (b) A. A. Khan, N. D. Emslie, S. E. Drewes, J. S. Field, N. Ramesar, Chem. Ber.

Recl. 1993, 126, 1477-1480; (c) S. E. Drewes, N. D. Emslie, J. S. Field, A. A. Khan, N.

Ramesar, Tetrahedron Lett. 1993, 34, 1205-1208. 71 V. K. Aggarwal, S. Y. Fulford, G. C. Lloyd-Jones, Angew. Chem. 2005, 117, 1734-1736;

Angew. Chem. Int. Ed. 2005, 44, 1706-1708. 72 For example: (a) J. Augé, N. Lubin, A. Lubineau, Tetrahedron Lett. 1994, 35, 7947-7948;

(b) V. K. Aggarwal, D. K. Dean, A. Mereu, R. Williams, J. Org. Chem. 2002, 67, 510-

514; (c) J. Cai, Z. Zhou, G. Zhao, C. Tang, Org. Lett. 2002, 4, 4723-4725; (d) C. Yu, B.

Liu, L. Hu, J. Org. Chem. 2001, 66, 5413-5418; (e) Y. M. A. Yamada, S. Ikegami,

Tetrahedron Lett. 2000, 41, 2165-2169. 73 For example: (a) D. Basavaiah, V. V. L. Gowriswari, P. K. S. Sarma, D. P. Rao,

Tetrahedron Lett. 1990, 31, 1621-1624; (b) S. E. Drewes, A. A. Khan, K. Rowland, Synth.

Commun. 1993, 23, 183-188; (c) E. P. Kündig, L. H. Xu, P. Romanens, G. Bernardelli,

Tetrahedron Lett. 1993, 34, 7049-7054; (d) P. R. Krishna, P. S. Reddy, N. M. Srivinas, B.

Sateesh, G. Sastri, Synlett 2006, 595-599.

Page 138: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

130

74 For example: (a) V. K. Aggarwal, I. Emme, S. Y. Fulford, J. Org. Chem. 2003, 68, 692-

700; (b) P. Ribière, N. Yadav-Bhatnagar, J. Martinez, F. Lamaty, QSAR Comb. Sci. 2004,

23, 911-914. 75 For example: (a) V. K. Aggarwal, A. Mereu, Chem. Commun. 1999, 2311-2312; (b) D.

Basavaiah, D. S. Sharada, N. Kumaragurubaran, R. M. Reddy, J. Org. Chem. 2002, 67,

7135-7137. 76 For example: (a) M. Shi, J. K. Jiang, C.-Q. Li, Tetrahedron Lett. 2002, 43, 127-130; (b) R.

Gatri, M. M. El Gayed, Tetrahedron Lett. 2002, 43, 7835-7836. 77 For example: (a) F. Rezgui, M. M. El Gayed, Tetrahedron Lett. 1998, 39, 5965-5966; (b)

M. Shi, C.-Q. Li, J. K. Jiang, Chem. Commun. 2001, 833-834. 78 (a) Y. M. A. Yamada, S. Ikegami, Tetrahedron Lett. 2000, 41, 2165-2169; (b) M. R.

Netherton, G. C. Fu, Org. Lett. 2001, 3, 4295-4298. 79 Z. He, X. Tang, Y. Chen, Z. He, Adv. Synth. Catal. 2006, 348, 413-417. 80 M. Shi, Y.-H. Liu, Org. Biomol. Chem. 2006, 4, 1486-1470. 81 For example: S. Luo, X. Mi, H. Xu, P. G. Wang, J.-P. Cheng, J. Org. Chem. 2004, 69,

8413-8422. 82 For example: (a) C. Yu, B. Liu, L. Hu, J. Org. Chem. 2001, 66, 5413; (b) J. Cai, Z. Zhou,

G. Zhao, C. Tang, Org. Lett. 2002, 4, 4723-4725. 83 For example: (a) D. J. Maher, S. J. Connon, Tetrahedron Lett. 2004, 45, 1301-1305; (b) Y.

Sotohme, A. Tanatani, Y. Hashimoto, K. Nagasawa, Tetrahedron Lett. 2004, 45, 5589-

5592. 84 For example: (a) H. Richter, G. Jung, Mol. Divers. 1997, 3, 191-194; (b) D. Basavaiah, P.

Dharma Rao, R. Sugna Hyma, Tetrahedron, 1996, 52, 8001-8062. 85 (a) M. Shi, J. K. Jiang, Tetrahedron: Asymmetry 2002, 13, 1941-1947; (b) A. Nakano, S.

Kawahara, S. Akamatsu, K. Morokuma, M. Nakatani, Y. Iwabuchi, K. Takahashi, J.

Ishihara, S. Hatakeyama, Tetrahedron 2006, 62, 381-389. 86 (a) Y. Iwabuchi, M. Furukawa, T. Esumi, S. Hatakeyama, Chem. Commun. 2001, 2030-

2031. 87 (a) A. Nakano, K. Takahashi, J. Ishihara, S. Hatakeyama, Heterocycles 2005, 66, 371-383;

(b) A. Nakano, M. Ushiyama, Y. Iwabuchi, S. Hatakeyama, Adv. Synth. Catal. 2005, 347,

1790-1796. 88 P. R. Krishna, V. Kannan, P. V. N. Reddy, Adv. Synth. Catal. 2004, 346, 603-606. 89 J. Wang, H. Li, X. Yu. L. Zu, W. Wang, Org. Lett. 2005, 7, 4293-4296. 90 J. E. Imbriglio, M. M. Vasbinder, S. J. Miller, Org. Lett. 2003, 5, 3741-3743.

Page 139: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

131

91 Y. Sotohme, A. Tanatani, Y. Hashimoto, K. Nagasawa, Tetrahedron Lett. 2004, 45, 5589-

5592. 92 C. E. Aroyan, M. M. Vasbinder, S. J. Miller, Org. Lett. 2005, 7, 3849-3851. 93 Y. Hayashi, T. Tamura, M. Shoji, Adv. Synth. Catal. 2004, 346, 1106-1110. 94 T. Hayase, T. Shibata, K. Soai, Y. Wakatsuki, Chem. Commun. 1998, 1271-1272. 95 M. Shi, G.-L. Zhao, Adv. Synth. Catal. 2004, 346, 1205-1219. 96 D. Balan, H. Adolfsson, J. Org. Chem. 2001, 66, 6498-6501. 97 M. Shi, L. G. Zhao, Tetrahdron Lett. 2002, 43, 9171-9174. 98 (a) M. Shi, Y.-M. Xu, Tetrahedron: Asymmetry 2002, 13, 1195-1200; (b) V. K. Aggarwal,

A. M. M. Castro, A. Mereu, H. Adams, Tetrahdron Lett. 2002, 43, 1577-1581. 99 (a) A. Kamimura, Y. Gunjigake, H. Mitsuder, S. Yokoyama, Tetrahedron Lett. 1998, 39,

7323-7324; (b) M. Shi, G.-L. Zhao, Tetrahedron 2004, 60, 2083-2089. 100 Y.-M. Xu, M. Shi, J. Org. Chem. 2004, 69, 417-425. 101 (a) Y.-L. Shi, Y.-M. Xu, M. Shi, Adv. Synth. Catal. 2004, 346, 1220-1230; (b) Y.-L- Shi,

M. Shi, Tetrahedron 2006, 62, 461-475. 102 (a) G.-L. Zhao, J.-W. Huang, M. Shi, Org. Lett. 2003, 5, 4737-4739; (b) G.-L. Zhao, M.

Shi, J. Org. Chem. 2005, 70, 9975-9984. 103 (a) M. Shi, Y.-M. Xu, Chem. Commun. 2001, 1876-1877; (b) M. Shi, Y.-M. Xu, G.-L.

Zhao, W.-F. Wu, Eur. J. Org. Chem. 2002, 3666-3679. 104 M. Shi, Y.-M. Xu, Eur. J. Org. Chem. 2002, 696-701. 105 (a) P. Ribière, N. Yadav-Bhatnagar, J. Martinez, F. Lamaty, QSAR Comb. Sci. 2004, 23,

911-914; (b) P. Ribiere, C. Enjalbal, J.-L. Aubagnac, N. Yadav-Bhatnagar, J. Martinez, F.

Lamaty, J. Comb. Chem. 2004, 6, 464-467. 106 V. Declerck, P. Ribiere, J. Martinez, F. Lamaty, J. Org. Chem. 2004, 69, 8372-8381. 107 D. Balan, H. Adolfsson, Tetrahedron Lett. 2004, 45, 3089-3092. 108 R. G. Wilkins, Kinetics and Mechanism of Reactions of Transition Metal Complexes,

VCH, Weinheim, 1991. 109 D. Balan, H. Adolfsson, J. Org. Chem. 2002, 67, 2329-2334. 110 M. Shi, L.-H. Chen, Chem. Commun. 2003, 1310-1311. 111 (a) M. Shi, L.-H. Chen, W.-D. Teng, Adv. Synth. Catal. 2005, 347, 1781-1789; (b) M. Shi,

C.-Q- Li, Tetrahedron: Asymmetry 2005, 16, 1385-1391. 112 The introduction of multiple phenol-groups can lead to an increased enantioselectivity:

Y.-H. Liu, L.-H. Chen, M. Shi, Adv. Synth. Catal. 2006, 348, 973-979 113 K. Matsui, S. Takizawa, H. Sasai, Synlett 2006, 761-765.

Page 140: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

132

114 K. Matsui, S. Takizawa, H. Sasai, J. Am. Chem. Soc. 2005, 127, 3680-3681. 115 (a) M. Shi, Y.-M. Xu, Angew. Chem. 2002, 114, 4689-4692; Angew. Chem. Int. Ed. 2002,

41, 4507-4510; (b) M: Shi, Y.-M. Xu, Y.-L. Shi, Chem. Eur. J. 2005, 11, 1794-1802. 116 D. Balan, H. Adolfsson, Tetrahedron Lett. 2003, 44, 2521-2524. 117 G. H. Roos, Compendium of Chiral Auxilliary Applications, Academic Press, San Diego,

2001. 118 E. N. Jacobsen, A. Pfaltz, H. Yamamoto, Comprehensive Asymmetric Catalysis, Springer

Verlag, Berlin, 1999. 119 C. Reichardt, Solvents and Solvent Effects in Organic Chemistry, Wiley-VCH, Weinheim,

2003. 120 D. Seebach, H. A. Oei, Angew. Chem. 1975, 87, 629-630; Angew. Chem. Int. Ed. Engl.

1975, 14, 634-635. 121 (a) P. Wasserscheid, W. Keim Angew. Chem. 2000, 112, 3926-3945; Angew. Chem. Int.

Ed. 2000, 39, 3772-3789; b) P. Wasserscheid, T. Welton, Ionic Liquids in Synthesis,

Wiley-VCH, Weinheim, 2003. 122 For recent reviews on chiral ILs, see: (a) C. Baudequin, D. Brégeon, J. Levillain, F.

Guillen, J.-C. Plaquevent, A.-C. Gaumont, Tetrahedron: Asymmetry, 2005, 16, 3921-

3945; (b) J. Ding, D. W. Armstrong, Chirality 2005, 17, 281-292. 123 M. J. Earle, P. B. McCormac, K. R. Seddon, Green Chem. 1999, 1, 23. 124 (a) A. Bösmann, P. Wasserscheid, C. Bolm, W. Keim, DE10003708, 2001; (b) P.

Wasserscheid, A. Bösmann, C. Bolm, Chem. Commun. 2002, 200-201. 125 B. Pégot, G. Vo-Thanh, A. Loupy, Tetrahedron Lett. 2004, 45, 6425-6428. 126 J. Ding, V. Desikan, X. Han, T. L. Xiao, R. Ding, W. S. Jenks, D. W. Armstrong, Org.

Lett. 2005, 7, 335-337. 127 Z. Wang, Q. Wang, Y. Zhang, W. Bao, Tetrahedron Lett. 2005, 46, 4657-4660. 128 For general information on ATR-IR spectroscopy, see: (a) H. Günzler, H.-U. Gremlich,

IR-Spektroskopie – Eine Einführung, 4rd revised edition, Wiley-VCH, Weinheim, 2003,

184 ff. For the specific setup used in these studies, see: (b) H. M. Heise, L. Küpper, L. N.

Butvina, Spectrochim. Acta B 2002, 57, 1649-1663; (c) H. M. Heise, L. Küpper, L. N.

Butvina, J. Mol. Struct. 2003, 661-662, 381-389; (d) C. Minnich, P. Buskens, W. Leitner,

M. A. Liauw, L. Greiner, manuscript in preparation. 129 P. S. Bäuerlein, Diploma thesis, RWTH Aachen University, Aachen, 2005. 130 imine,0imine,0 loglog)0(log)0( cxYrYcr y

xY +=⇔= with x = partial order in imine.

Page 141: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

133

131 Evan’s pKA table: http://daecr1.harvard.edu/pKa/pka.html (in water). 132 J. Zimmermann, I. Tkatchenko, P. Wasserscheid, Adv. Synth. Catal. 2003, 345, 402-409. 133 I. C. Stewart, R. G. Bergman, F. D. Toste, J. Am. Chem. Soc. 2003, 125, 8696-8697. 134 The four diastereomers were at first optimized using the B3LYP[135a-d] hybrid functional

using the Gaussian03[135e] program series with the 6-31G(d) basis set. [135f-j] 135 (a) A. D. Becke, J. Chem. Phys. 1993, 98, 5648-5652; (b) C. Lee, W. Yang, R. G. Parr,

Phys. Rev. 1988, 37, 785-789; (c) S. H. Vosko, L. Wilk, M. Nusair, Can. J. Phys. 1980, 58

1200-1211; (d) P. J. Stephens, F. J. Delvin, C. F. Chabalowski, M. J. Frisch, J. Phys.

Chem. 1994, 98, 11623-11627; (e) Gaussian 03, Revision B.03, M. J. Frisch, G. W.

Trucks, H. B. Schlegel, G. E.Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery,

Jr., T. Vreven, K. N. Kudin, J. C.Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V.

Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M.

Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,

O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, C. Adamo,

J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C.

Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G.A. Voth, P. Salvador, J. J.

Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C.Strain, O. Farkas, D. K.

Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz,Q. Cui, A. G. Baboul,

S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi,

R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M.

Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, and J. A.

Pople, Gaussian, Inc., Pittsburgh PA, 2003; (f) R. Ditchfield, W. J. Hehre, J. A. Pople, J.

Chem. Phys. 1971, 54, 724-728; (g) W. J. Hehre, R. Ditchfield, J. A. Pople, J. Chem.

Phys. 1972, 56, 2257-2261; (h) P. C. Hariharan, J. A. Pople, Mol. Phys. 1974, 27, 209-

214; (i) M. S. Gordon, Chem. Phys. Lett. 1980, 76, 163; (j) P. C. Hariharan, J. A. Pople,

Theo. Chim. Acta 1973, 28, 213-222. 136 (a) A. Schäfer, C. Huber, R. Ahlrichs, J. Chem. Phys. 1994, 100, 5829-5835; (b)

Turbomole 5.7 The Quantum Chemistry Group, University of Karlsruhe, 1988-2005:

Reinhart Ahlrichs, Michael Bär, Hans-Peter Baron, Rüdiger Bauernschmitt, Stephan

Böcker, Peter Deglmann, Michael Ehrig, Karin Eichkorn, Simon Elliott, Filipp Furche,

Frank Haase, Marco Häser, Hans Horn, Christof Hättig, Christian Huber, Uwe Huniar,

Marco Kattannek, Andreas Köhn, Christoph Kölmel, Markus Kollwitz, Klaus May,

Christian Ochsenfeld, Holger Öhm, Holger Patzelt, Oliver Rubner, Ansgar Schäfer, Uwe

Schneider, Marek Sierka, Oliver Treutler, Barbara Unterreiner, Malte von Arnim, Florian

Page 142: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

134

Weigend, Patrick Weis, Horst Weiss; (c) A. D. Becke, Phys. Rev. A, 1988, 38, 3098-3100;

(d) J. P. Perdew, Phys. Rev. B. 1986, 33, 8822.

137 ⎟⎠⎞

⎜⎝⎛−=ΔΔ

SRRTG ln

138 N. W. Alcock, J. M. Brown, D. I. Hulmes, Tetrahedron: Asymmetry 1993, 4, 743-756. 139 For example: (a) J. M. Brown, D. I. Hulmes, T. P. Layzell, J. Chem. Soc., Chem. Commun.

1993, 1673-1674; (b) H. Doucet, E. Fernandez, T. P. Layzell, J. M. Brown, Chem. Eur. J.

1999, 5, 1320-1330. 140 E. Fernandez, M. W. Hooper, F. I. Knight, J. M. Brown, J. Chem. Soc., Chem. Commun.

1997, 173-174. 141 (a) J. M. Brown, D. I. Hulmes, P. J. Guiry, Tetrahedron 1994, 50, 4493-4506; (b) J. M.

Valk, T. D. W. Claridge, J. M. Brown, D. Hibbs, M. B. Hursthouse, Tetrahedron:

Asymmetry 1995, 6, 2597-2610. 142 D. Enders, J. H. Kirchhoff, J. Köbberling, T. H. Peiffer, Org. Lett. 2001, 3, 1241-1244. 143 J. Cossy, C. Dumas, D. G. Pardo, Eur. J. Org. Chem. 1999, 1693-1699. 144 T. W. Greene, P. G. M. Wuts, Protective groups in organic synthesis, John Wiley & Sons,

New York, 1991. 145 (a) M. Fremantle, Chem. Eng. News 1998, 76, 32; (b1) Z. Fei, T. J. Geldbach, D. Zhao, P.

J. Dyson, Chem. Eur. J. 2006, DOI: 10.1002/chem.200500581 146 R. Gausepohl, Ph.D. thesis, RWTH Aachen University, Aachen, 2006. 147 J. Boeseken, J. A. Mijs, Recueil des Travaux Chimiques des Pays-Bas et de la Belgique

1925, 44, 758. 148 Impurities: octanol (from Aliquat 336 ®), water (200-2800 ppm, determined from Karl-

Fischer Titration). 149 By comparison with the work from Shi and co-workers[48d], the absolute configuration of

106 was determined as R. 150 D. D. Perrin, W. L. F. Amarego, D. R. Perri, Purification of Laboratory Chemicals,

Pergamon Press: Oxford, 2nd edition, 1980. 151 W. C. Still, M. Kahn, A. Mitra, J. Org. Chem. 1978, 43, 2923-2925. 152 B. E. Love, P. S. Raje, T. C. Williams, Synlett 1994, 493-494. 153 Due to the complexity of the mixture, a full characterization of the aromatic C-signals was

not possible. 154 Due to the complexity of the mixture, a full characterization of the aromatic proton signals

and the C-signals was not possible.

Page 143: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

135

155 For the assignment of the carbon signals, following nomenclature was used:

(C9)N+(C1C2C3C4C5C6C7C8)3 156 Preparation of the ion-exchange resin: the H-form of Marathon ® MSC (supplied by

Aldrich) was stirred with an excess of NaOH in aqueous solution. After 1h, the NaOH

solution was removed by decantation and the resin was washed with water until the

resulting aqueous phase showed a neutral pH.

Page 144: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

136

CURRICULUM VITAE

Name: Pascal Buskens

Date of birth (place): 5th February 1980 (Heerlen, Netherlands)

Nationality: Dutch

Education

2003 – today Ph.D. study

RWTH Aachen University, Aachen (Germany)

Institute of Technical and Macromolecular Chemistry

Group of Prof. Dr. Walter Leitner

Subject: Bifunctional Organocatalysis in the Asymmetric

Aza-Baylis-Hillman Reaction

October 2004 – today Member and holder of a scholarship of the Graduate School

440 Methods in Asymmetric Synthesis

September – November 2005 Research project at the University of Oxford, Oxford (UK)

Department of Organic Chemistry

Group of Dr. John M. Brown FRS

Subject: QUINAP as Organocatalyst

2000 – 2003 Chemistry, Diplom

RWTH Aachen University, Aachen (Germany)

May – October 2003 Diploma thesis

RWTH Aachen University, Aachen (Germany)

Institute of Technical and Macromolecular Chemistry

Group of Prof. Dr. Walter Leitner

Subject: Activation and Deactivation of a Novel Non-

Classical Rutheniumhydride Catalyst for the Murai Reaction

Page 145: Bifunctional Organocatalysis in the Asymmetric Aza-Baylis ...darwin.bth.rwth-aachen.de/opus3/volltexte/2006/1664/pdf/Buskens... · Bifunctional Organocatalysis in the Asymmetric Aza-Baylis-Hillman

137

January – March 2002 Research project at the University of York, York (UK)

Department of Inorganic Chemistry, Green Chemistry Group

Group of Dr. Karen Wilson

Subject: Solid Sulphonic Acid Catalysts for Esterification

Reactions

1998 – 2000 RWTH Aachen University, Aachen (Germany)

Chemistry, Vordiplom (pre-degree)

1992 – 1998 College Rolduc, Kerkrade (Netherlands)

Gymnasium