university of groningen rhodium-catalyzed boronic acid ......complementary rhodium-catalyzed aca of...

175
University of Groningen Rhodium-catalyzed boronic acid additions Jagt, Roelof Bauke Christiaan IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below. Document Version Publisher's PDF, also known as Version of record Publication date: 2006 Link to publication in University of Groningen/UMCG research database Citation for published version (APA): Jagt, R. B. C. (2006). Rhodium-catalyzed boronic acid additions: a combinatorial approach to homogeneous asymmetric catalysis. s.n. Copyright Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons). Take-down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum. Download date: 24-06-2021

Upload: others

Post on 04-Feb-2021

5 views

Category:

Documents


0 download

TRANSCRIPT

  • University of Groningen

    Rhodium-catalyzed boronic acid additionsJagt, Roelof Bauke Christiaan

    IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite fromit. Please check the document version below.

    Document VersionPublisher's PDF, also known as Version of record

    Publication date:2006

    Link to publication in University of Groningen/UMCG research database

    Citation for published version (APA):Jagt, R. B. C. (2006). Rhodium-catalyzed boronic acid additions: a combinatorial approach tohomogeneous asymmetric catalysis. s.n.

    CopyrightOther than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of theauthor(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

    Take-down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

    Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons thenumber of authors shown on this cover page is limited to 10 maximum.

    Download date: 24-06-2021

    https://research.rug.nl/en/publications/rhodiumcatalyzed-boronic-acid-additions(f6b6f774-c1f4-4035-a131-f1f4d52bf232).html

  • Rhodium-Catalyzed Boronic Acid Additions A Combinatorial Approach to

    Homogeneous Asymmetric Catalysis

    Richard B. C. Jagt

  • R. B. C. Jagt, “Rhodium-Catalyzed Boronic Acid Additions; A Combinatorial Approach to Homogeneous Asymmetric Catalysis”, Ph.D. Thesis, University of Groningen, The Netherlands, 2006.

    ISBN: 90-367-2710-3

    ISBN: 90-367-2711-1 (electronic version)

    The work described in this thesis was carried out at the Department of Organic and Molecular Inorganic Chemistry, Stratingh Institute, University of Groningen, The Netherlands.

    Financial support was received from DSM Pharmaceutical Products, the Ministry of Economic Affairs, and the Chemical Sciences division of the Netherlands Organization for Scientific Research (NWO/CW), administered through the NWO/CW Combichem program.

  • RIJKSUNIVERSITEIT GRONINGEN

    Rhodium-Catalyzed Boronic Acid Additions A Combinatorial Approach to

    Homogeneous Asymmetric Catalysis

    Proefschrift

    ter verkrijging van het doctoraat in de

    Wiskunde en Natuurwetenschappen

    aan de Rijksuniversiteit Groningen

    op gezag van de

    Rector Magnificus, dr. F. Zwarts,

    in het openbaar te verdedigen op

    vrijdag 8 september 2006

    om 14.45 uur

    door

    Roelof Bauke Christiaan Jagt

    geboren op 19 oktober 1977

    te Emmen

  • Promotores: Prof. Dr. B. L. Feringa

    Prof. Dr. Ir. A. J. Minnaard

    Beoordelingscommissie: Prof. Dr. J. B. F. N. Engberts

    Prof. Dr. J. N. H. Reek

    Prof. Dr. J. G. de Vries

    ISBN: 90-367-2710-3

  • Voor mijn ouders.

  • 04_Contents.doc

    Contents

    Chapter 1 Introduction

    1.1 The Impact of Molecular Chirality

    1.2 The Preparation of Enantiopure Molecules

    1.3 Asymmetric Conjugate Addition Reactions

    1.4 1,2-Arylations Using Organometallic Reagents

    1.5 A Ligand Library Approach to Asymmetric Catalysis

    1.6 Aims and Outline of this Thesis

    1.7 References and Notes

    1

    1

    2

    3

    6

    12

    16

    17

    Chapter 2 Enantioselective Synthesis of 2-Aryl-4-piperidones

    2.1 Introduction

    2.2 Results and Discussion

    2.2.1 Preliminary Studies

    2.2.2 Scope of the Reaction

    2.3 Further Developments

    2.4 Conclusions

    2.5 Experimental Section

    2.6 References and Notes

    25

    26

    29

    29

    32

    33

    34

    35

    41

    Chapter 3 Tandem Conjugate Addition/Protonation:

    Enantioselective Synthesis of α-Amino Acids

    3.1 Introduction

    3.2 Results and Discussion

    45

    46

    52

  • 04_Contents.doc

    Contents

    3.3 Conclusions

    3.4 Experimental Section

    3.5 References and Notes

    56

    57

    58

    Chapter 4 Enantioselective Synthesis of Diarylmethanols

    4.1 Introduction

    4.2 Results and Discussion

    4.2.1 Preliminary Studies and Mechanistic Considerations

    4.2.2 Ligand Library Approach

    4.2.3 Scope of the Reaction

    4.3 Further Developments

    4.4 Conclusions

    4.5 Experimental Section

    4.6 References and Notes

    63

    64

    66

    66

    69

    75

    77

    78

    78

    84

    Chapter 5 Enantioselective Synthesis of N-protected Diarylmethylamines

    5.1 Introduction

    5.2 Results and Discussion

    5.2.1 Preliminary Studies

    5.2.2 Screening of a Primary Ligand Library

    5.2.3 Introduction of the N,N-Dimethylsulfamoyl Protecting Group

    5.2.4 Screening of a Secondary Ligand Library

    5.2.5 Scope of the Reaction

    5.2.6 Removal of the N,N-Dimethylsulfamoyl Group

    5.3 Conclusions

    5.4 Experimental Section

    5.5 References and Notes

    89

    90

    92

    92

    92

    95

    97

    99

    101

    101

    102

    110

  • 04_Contents.doc

    Rhodium-Catalyzed Boronic Acid Additions

    Chapter 6 An Entry to Diversity in 3-Aryl- and

    3-Alkenyl-3-hydroxyoxindoles

    6.1 Introduction

    6.2 Results and Discussion

    6.2.1 The Synthesis of Racemic 3-Substituted-3-hydroxyoxindoles

    6.2.2 Enantioselective Synthesis of 3-Phenyl-3-hyroxyoxindoles

    6.3 Further Developments

    6.4 Conclusions

    6.5 Experimental Section

    6.6 References and Notes

    113

    114

    117

    117

    119

    124

    124

    125

    130

    Chapter 7 Enantioselective Synthesis of Trifluoromethyl

    Substituted Tertiary Alcohols

    7.1 Introduction

    7.2 Results and Discussion

    7.3 Conclusions

    7.4 Experimental Section

    7.5 References and Notes

    135

    136

    137

    139

    140

    143

    Summary

    147

    Samenvatting

    153

    Acknowledgements 159

  • 04b_Emptypage.doc

  • 1

    05_Chapter 1.doc

    Chapter 1 Introduction

    1.1 The Impact of Molecular Chirality

    Chirality1 is an important aspect of the most fundamental processes of life.2 The sugars that

    constitute DNA and RNA possess a uniform stereochemical configuration. The proteins

    encoded by these oligonucleotides, crucial for the chemical transformations in cells, consist

    of chiral α-amino acids that occur exclusively in the L-configuration. Without this chiral

    homogeneity, the biomachinery that makes up all known living organisms would not be

    able to function. Even in the most elementary forms of life, such as bacteria,3 a myriad of

    different chiral molecules are involved in complex signaling pathways. Receptor proteins

    on the cell membrane or within the cytoplasm or cell nucleus can specifically bind to one

    enantiomer of a chiral “messenger” molecule and initiate a corresponding cellular response.

    These diastereomeric interactions are the key to modern drug development.4

    The interactions between biological systems and synthetic chiral molecules has a huge

    impact on contemporary everyday life and applications range from flavors, fragrances, and

    food additives to agrochemicals and life-saving drugs. Homochirality in drugs is as old as

    the first therapeutic agents isolated from natural sources, such as quinine and morphine.

    However, as products of synthetic chemistry, until recently chiral drugs were manufactured

    as racemates. The assumption that only one enantiomer of a drug has biological activity and

    the other serves as “isomeric ballast”4 has turned out to be a rather dangerous one. The two

    enantiomers of a compound most frequently bind to different receptors, and therefore have

    completely different physiological effects. The presence of the “wrong” enantiomer has, in

    some cases, been known to cause serious side-effects. This has resulted in severe

    restrictions5 to the production of bioactive molecules and at present time “single-

  • Chapter 1

    2

    05_Chapter 1.doc

    enantiomer drugs have a commanding presence in the global pharmaceutical landscape”.6

    The development of efficient methodologies for the synthesis of the individual enantiomers

    of an asymmetric target compound is, therefore, of continuous interest to scientists in both

    industry and academia.

    1.2 The Preparation of Enantiopure Molecules

    Routes to single enantiomers of small molecules can be classified into three groups:

    1) chiral pool technology, 2) resolution of racemic mixtures, and 3) asymmetric synthesis.

    The former is the most straightforward route to enantiopure compounds and starts from

    homochiral biomolecules which are provided by natural processes such as agriculture or

    fermentation. However, the lack of availability of both enantiomers for most naturally

    occurring compounds severely limits the scope of this methodology and many chiral

    building blocks can only be obtained by synthetic procedures.7

    The “classical” resolution of racemic mixtures by diastereomeric crystallisation, to date,

    often constitutes the industrial method of choice to obtain large quantities of enantiopure

    compounds.8 However, unless it can be recycled, half of the racemic starting material (the

    “unwanted” enantiomer) is a waste-product. This intrinsic property of classical resolutions

    poses a major disadvantage from an atom-economy point of view.9 The same disadvantage

    applies to chemical or enzymatic kinetic resolutions, involving a reaction in which one of

    the two enantiomers reacts more rapidly than the other based on a difference in transition

    state Gibbs energy. Although in certain cases, (dynamic) kinetic resolution can lead to

    complete conversion of the starting material by in situ racemization, generally one

    enantiomer reacts whereas the other remains intact.

    The remaining option for the preparation of enantiopure molecules involves the

    introduction of chirality to a prochiral substrate by asymmetric induction.10 This may

    involve the use of stoichiometric amounts of chiral reagent or a chiral auxiliary followed by

    the subsequent diastereoselective introduction of a stereogenic center. However, the use of

    equimolar amounts of valuable chiral auxiliary materials makes these approaches rather

  • Introduction

    3

    05_Chapter 1.doc

    unappealing. A far more attractive form of stereoselective synthesis involves the

    application of asymmetric catalysts. A relatively small amount of enantiopure catalyst can,

    in an ideal scenario, produce large quantities of enantiopure product. Although powerful

    biocatalytic methods exist, employing enzymes or antibodies as catalysts,11 their

    biomolecular homochirality often poses a problem when the “non-natural” enantiomer of

    the product is desired. Recently, directed evolution methods have resulted in enzymes

    which produce the unnatural enantiomers in excess.12 Alternatively, chemical catalysts can

    be adapted to provide the desired enantiomer of the product by choosing the appropriate

    enantiomer of the ligand. Although asymmetric organocatalysis – based on the use of small

    organic molecules as catalysts − is an emerging field,13 in the last decades considerable

    progress has been made in the development of highly active metal-catalyzed asymmetric

    transformations based on enantiopure ligands complexed to a (transition) metal core.14 The

    pioneering work of Knowles, Noyori, and Sharpless on chirally-catalyzed hydrogenation

    and oxidation reactions, for which they received the 2001 Nobel prize in chemistry,15 has

    opened the field of homogeneous asymmetric catalysis. Asymmetric reductions and

    oxidations have been developed to an extent that they are in some cases used for industrial

    production of enantiomerically enriched compounds. However, in catalytic asymmetric

    carbon-carbon bond forming reactions high catalytic activity and enantioselectivity are less

    well established.

    1.3 Asymmetric Conjugate Addition Reactions

    The asymmetric conjugate addition (ACA) of organometallic reagents to electron-deficient

    alkenes constitutes an important approach for enantioselective carbon-carbon bond

    formation.16 In recent years, numerous chiral catalysts have been introduced for this

    asymmetric reaction.17 A breakthrough in the field was achieved by our group in 1996 with

    the introduction of chiral monodentate phosphoramidite L1, which proved to be a highly

    efficient ligand in the copper-catalyzed ACA of dialkylzinc reagents to enones (e.g.

    Scheme 1.1, reaction a).18,19 Phosphoramidites comprise a class of cost-effective and easily

    tunable ligands that have since proven useful in a host of different reactions (vide infra).

  • Chapter 1

    4

    05_Chapter 1.doc

    Although the highly enantioselective copper-catalyzed conjugate addition of diphenylzinc,

    using the same catalyst, has been reported for 2-cyclohexenone,20 the lack of readily

    available diarylzinc reagents severely limits this method. A more convenient reaction for

    the formation of sp2-sp3 carbon-carbon bonds, introducing a stereogenic centre, is the

    complementary rhodium-catalyzed ACA of boronic acids pioneered by Miyaura and

    Hayashi in 1998 (e.g. Scheme 1.1, reaction b).21 Excellent enantioselectivities have been

    achieved in the addition of aryl and alkenyl groups to a broad range of both cyclic and

    acyclic α,β-unsaturated enones employing BINAP as a chiral ligand.

    O 0.5 mol% Cu(OTf)21 mol% L1

    toluene, -30 oC

    Et2Zn (1.2 equiv)

    O

    >98% ee

    OO

    P N

    (S,R,R)-L1

    a)

    b)

    O 3 mol% Rh(acac)(eth)23 mol% BINAP

    dioxane/H2O : 10/1, 100 oC

    PhB(OH)2 (3 equiv)

    O

    PPh2PPh2

    (S)-BINAP97% ee

    Scheme 1.1 Two complementary ACA reactions

    Since its introduction, this reaction has received increasing attention.22,23 Rhodium-

    catalyzed conjugate additions of arylstannane,24 arylsilicon,25 aryltitanium,26 and

    alkenylzirconium27 reagents have also been reported. However, from a practical point of

    view, boronic acids remain the most interesting arylating reagents because they are shelf

    stable, readily available, and compatible with a large variety of functional groups.28 The

    chiral ligands employed in the rhodium-catalyzed addition of arylboronic acids are mostly

    bidentate and comprise biaryl bis-phosphines,21a,29 ferrocenyl-based bis-phosphines,29c

    bis-β-naphthol (BINOL) based diphosphonites,30 amidomonophosphines,31 N-heterocyclic

    carbenes,32 a P-chiral phosphine,33 and dienes.34 Some other bis-phosphine ligands,

    although efficient in the rhodium-catalyzed hydrogenation, fail to induce high reactivity

  • Introduction

    5

    05_Chapter 1.doc

    and/or selectivity in the rhodium-catalyzed conjugate addition to enones.22,29a,c Recently, it

    was shown by our group that monodentate phosphoramidite L2 is an exceptionally efficient

    ligand for the rhodium-catalyzed ACA of arylboronic acids to enones (Scheme 1.2).35

    Enantioselectivities comparable to those obtained with BINAP have been achieved, while

    the reaction rate is greatly enhanced. Employing L2 as a ligand, using 1 mol% of catalyst,

    >99% conversion was reached within 5 min at 100 oC. Miyaura and co-workers confirmed

    the successful use of phosphoramidites in this reaction.36

    O 3 mol% Rh(acac)(eth)27.5 mol% (S)-L2

    dioxane/H2O : 10/1, 100 oC

    ArB(OH)2 (3 equiv)

    O

    OO

    P

    (S)-L2

    N

    Up to 98% ee

    R

    Scheme 1.2 Rhodium/phosphoramidite-catalyzed ACA of arylboronic acids

    The mechanism for the rhodium-catalyzed addition of boronic acids (Scheme 1.3) has been

    investigated by Hayashi et al.37 and all intermediates were identified by NMR. In order to

    displace the acetylacetonate (acac), and to generate the catalytically active hydroxy species,

    the presence of water and high temperatures are needed (step A). It was shown that, starting

    from the hydroxy species, the reaction is faster than when the “precatalyst” prepared from

    Rh(acac)(C2H4)2 and BINAP was used. A phenyl-rhodium complex and B(OH)3 are formed

    after transmetallation of the phenyl group from boron to rhodium (step B). After

    coordination of the substrate (step C), insertion of 1 into the phenyl-rhodium bond gives an

    oxa-π-allyl species (step D). Hydrolysis of this species gives the desired phenylated product

    2 and regenerates the catalytically active hydroxy species (step E). As water is involved in

    forming the active hydroxy species and acts as a proton donor facilitating the

    transmetallation, its presence is essential in order to complete the catalytic cycle. A major

    drawback of the use of water, in combination with high temperature, is hydrolysis of

    phenylboronic acid to benzene (step F). For this reason, the arylboronic acid is often added

    in excess (3-5 equiv) compared to the enone.22

  • Chapter 1

    6

    05_Chapter 1.doc

    Rh PhP*

    P*

    RhP*

    P*

    BPh B(OH)2

    B(OH)3

    FH2O C

    R'

    OB(OH)3

    +

    Ph-H

    1

    Rh OHP*

    P*Rh(acac)(eth)2

    + 2P*A

    H2O

    R'

    O

    2Ph H2O

    R'

    O

    Ph

    Rh

    P*

    P*

    Ph R'

    O

    DE

    R

    R

    R

    R

    Scheme 1.3 Proposed catalytic cycle for the rhodium-catalyzed asymmetric conjugate addition reaction

    Inspired by the catalytic cycle proposed by Hayashi et al.,37 it was recently found that the

    use of a [Rh(OH)(cod)]2 (cod = 1,5-cyclooctadiene) or [RhCl(cod)]2/KOH catalyst

    precursor allows the reaction to be carried out at room temperature with only 1.5 equiv of

    boronic acid.38

    1.4 1,2-Arylations Using Organometallic Reagents

    Over the past 20 years, enantioselective formation of chiral diarylmethanols and

    diarylmethylamines has attracted a great deal of interest.39 Enantiopure derivatives of these

    compounds are important intermediates for the synthesis of biologically active molecules.40

    In 1998, Miyaura and co-workers demonstrated the rhodium-catalyzed addition of

    arylboronic acids to aromatic aldehydes under conditions similar to those used for their

    conjugate addition to enones.41,42 In an asymmetric version of this reaction, employing

    MeO-MOP as chiral ligand, 41% ee was achieved for the addition of phenylboronic acid to

  • Introduction

    7

    05_Chapter 1.doc

    1-naphthaldehyde (Scheme 1.4, reaction a). An attempt by Frost et al. to improve the

    enantioselectivity of this reaction by using sparteine or bis-oxazolines as ligands remained

    unsuccessful (

  • Chapter 1

    8

    05_Chapter 1.doc

    reaction. Hayashi et al. reported excellent enantioselectivities in the addition of

    arylboroxines to a range of N-tosyl- and N-nosyl-benzaldimines (Scheme 1.5) employing

    L5 and L6, respectively.50

    O

    NPPh2

    L4

    Ph

    Ph

    Ph

    Ph

    L5NHBoc

    L6

    NPPh2

    PPh2

    (R,R)-DeguPHOS

    Bn

    Figure 1.1 Ligands reported in the 1,2-addition of arylboronic acids to benzaldimines

    NPG

    HR1

    ArB(OH)2 (2 equiv)

    [RhCl(C2H4)2]2 (3 mol%)

    L5 or L6 (3 mol%)

    KOH/H2O

    dioxane, 60 oC, 6 h

    HNPG

    R1

    R2

    PG = S (Tosyl)

    S NO2 (Nosyl)

    PG = Tosyl: L5, 95-99% eePG = Nosyl: L6, 95-99% ee

    O

    O

    O

    O

    Scheme 1.5 Arylboronic acid addition to N-tosyl and N-nosyl protected benzaldimines

    Next to the rhodium-catalyzed arylation of aldehydes and imines, a successful alternative

    for such reactions can be found in the addition of diarylzinc reagents employing in situ

    formed complexes of zinc with chiral aminoalcohols and diols as catalysts.51 In contrast to

    the addition of dialkylzinc reagents, however, addition of diphenylzinc to aldehydes also

    proceeds quite efficiently without the presence of a catalyst. This background reaction

    makes it difficult to achieve high enantioselectivities. In 1997, Dosa and Fu reported the

    first enantioselective catalytic addition of diphenylzinc to aldehydes employing axially

    chiral ferrocene-based ligand L7 (Figure 1.2).52 The addition of diphenylzinc to

    p-chlorobenzaldehyde provided the product alcohol with an enantioselectivity of 57% ee.

  • Introduction

    9

    05_Chapter 1.doc

    In 1999, Pu et al. reported high ee values in the addition of diphenylzinc to aldehydes using

    L8a as a ligand (Scheme 1.6).53 This chiral binaphthol was also previously employed as a

    highly enantioselective catalyst in the addition of dialkylzinc.54 Interestingly, the authors

    observed that when 20 mol% L8a was pretreated with 40 mol% of diethylzinc, the resulting

    chiral PhZnEt complex facilitated the addition of diphenylzinc with a considerable increase

    in enantioselectivity.55

    FeN

    MeMeMe

    MeMe

    L7

    OHO

    Ph

    Ph

    Figure 1.2 Ferrocene ligand employed by Dosa and Fu

    Cl

    CHO+ Ph2Zn

    OHOH

    Ar

    Ar

    OC6H13

    C6H13O

    F

    C6H13O

    F

    L8a

    L8b

    Ar =

    Cl

    Ph

    OH

    20 mol% L840 mol% Et2Zn

    Et2O, RT, 10 h:

    CH2Cl2, RT, 5 h:

    86% yield, 94% ee

    92% yield, 95% ee

    Ar =

    Scheme 1.6 Addition of diphenylzinc to aldehydes using diethylzinc as an additive

    The use of fluorinated binaphthol-type ligands, resulting in an enhanced reaction rate, was

    reported by Pu in 2000.56 Various aromatic aldehydes were converted with diphenylzinc in

    the presence of 20 mol% of (S)-L8b in dichloromethane. After 5 h at room temperature the

    products were formed with good enantioselectivities in high yields. In 1999 Bolm and

    Muñiz reported ferrocene-based ligands L9 and L10 (Figure 1.3) in the addition of

    diphenylzinc to aldehydes.57 The addition of diphenylzinc to p-chlorobenzaldehyde and

  • Chapter 1

    10

    05_Chapter 1.doc

    ferrocenecarboxaldehyde, employing ligand L9, proceeded with 90% and >96% ee,

    respectively. However, the enantioselectivity for the reaction of other aromatic or aliphatic

    aldehydes was considerably lower (3-75%) due to the background reaction. Also here, the

    use of a mixture of diphenyl- and diethylzinc (in a 1 : 2 ratio) significantly increased the

    enantioselectivity.58 For the addition of diphenylzinc to p-chlorobenzaldehyde an ee of 99%

    was achieved. Using a similar procedure, L9 catalyzed the reaction of diphenylzinc with a

    range of aldehydes with very high enantioselectivity. This methodology has been adapted

    by Bolm and Bräse to the addition of diphenylzinc to imines (Scheme 1.7).59 The addition

    of diphenylzinc to a range of in situ formed N-formylimines with different electronic and

    steric modifications proceeded with high enantioselectivities.

    FeN

    O

    CPh2OH FeN

    O

    PhCPh2OH

    L9 L10

    Figure 1.3 Ferrocene ligands employed by Bolm and Muñiz

    SO2Tol

    HN H

    O

    R

    N H

    O

    R

    ZnEt2/ZnPh2

    ZnEt2/ZnPh2 H

    (Rp,S)-L11 (10 mol%)

    HN H

    O

    ROH

    N

    Ph

    Ph

    (Rp,S)-L11Up to 97% ee

    Scheme 1.7 Catalytic asymmetric phenyl transfer reaction onto N-formylimines

  • Introduction

    11

    05_Chapter 1.doc

    Several alternative systems for the addition of mixtures of diphenylzinc and diethylzinc to

    aldehydes have been reported.60 Triphenylborane was recently found to be an interesting

    alternative to diphenylzinc as a phenyl source.61 It is commercially available in bulk

    quantities and rather inexpensive compared to diphenylzinc. All these methods are,

    however, limited to the addition of phenyl groups. More general systems that facilitate the

    addition of a range of aryl groups, as is the case for the addition of arylboronic acids, are

    limited. Recently, however, Bolm et al.62 reported an excellent hybrid method in which the

    aryl transfer reagent was generated by mixing arylboronic acids with an excess of

    diethylzinc (Scheme 1.8).63 Subsequently, a multigram scale application was described.64

    Employing ligand L9, the addition of different aryls to a range of benzaldehydes proceeded

    with enantioselectivities from 31 to 95% ee. The presence of catalytic amounts of a

    dimethylpolyethyleneglycol (DiMPEG, Mw = 2000 g mol-1) further improved the

    enantioselectivities.65

    CHOR

    ArB(OH)2Et2Zn (7.2 equiv)

    10 mol% L910 mol% DiMPEG

    10 oC, 12 h

    ROH

    Ar

    Scheme 1.8 Use of in situ generated zinc reagents from diethylzinc and arylboronic acids

    High enantioselectivities in the arylation of benzaldehydes, using a mixture of arylboronic

    acids and diethylzinc, have recently also been achieved using chiral binaphthol

    dicarboxamides,66 hydroxy oxazolines,67 ferrocene based silanols,68 aminonaphtholes,69 a

    camphor-derived γ-amino thiol,70 and β-amino alcohol-based catalysts.63b,71 In addition to

    boronic acids, also their boroxine trimers have been used successfully in combination with

    diethylzinc.72 All of these methods, however, are dependent on the use of an excess of

    rather expensive and reactive diethylzinc. From a practical and industrial point of view, the

    development of a direct catalytic addition of boronic acids to aldehydes with high

    enantioselectivity is more desirable.

  • Chapter 1

    12

    05_Chapter 1.doc

    1.5 A Ligand Library Approach to Asymmetric Catalysis

    The identification of suitable asymmetric catalysts still poses one of the most challenging

    endeavours of contemporary organic chemistry.14 Current mechanistic knowledge is not

    sufficient to “design” catalysts which will induce high enantioselectivity. All

    stereoselective syntheses are based on the principle that the products are formed via

    diastereomeric transition states with a difference in Gibbs energy of activation (∆∆G‡).1b

    Seemingly insignificant variations in the catalyst or substrate structure and reaction

    conditions, corresponding to minute differences in transition state energy (∆∆G‡ ≅ 1-2

    kcal/mol), can cause significant changes in ee. Moreover, the degree of structural

    recognition that is a prerequisite for selective catalysts also renders them highly substrate

    dependent. In many asymmetric reactions the ligand structure must be optimized for each

    substrate class.

    The relationship between ligand structure and the chemical and physical properties of

    derived complexes is a central theme in many branches of chemistry. In essence, the

    problems involved with developing asymmetric catalysts are similar to those encountered

    by medicinal chemists when developing a new drug that is required to undergo selective

    diastereomeric interactions with a specific receptor protein. Conventional drug

    development, until recently, involved the laborious process of synthesizing and evaluating

    hundreds to thousands of organic compounds in a one-at-a-time fashion in an attempt to

    enhance biological activity and selectivity. In the 1990s pressure within the pharmaceutical

    industry to speed up the drug discovery process, combined with the increasing political and

    social pressure on drug prices, have resulted in a paradigm-shift in the field.

    “Combinatorial chemistry”73 became the new standard in drug development, dramatically

    cutting the time and costs associated with serial drug discovery. The basic principle of this

    new philosophy is the parallel synthesis of a library of compounds, followed by high-

    throughput screening in order to identify the most promising leads (Scheme 1.9). More

    focussed libraries can subsequently be tested in order to optimize the results. In contrast to

    the “classical” iterative serial approach, different compounds with a modular build-up are

    generated simultaneously, in a systematic manner, under identical reaction conditions, so

  • Introduction

    13

    05_Chapter 1.doc

    that the products of all possible combinations of a set of building blocks are obtained.

    Although combinatorial chemistry was initially met with resistance from the chemical

    community because of its “black-box” character, its practice is now widespread in chemical

    biology and medicinal chemistry as a tool to systematically investigate structural

    requirements that are beyond rational design. Mechanistic insight remains very important,

    as it helps in the selection of suitable parameters.

    Design

    TraditionalSynthesis

    ParallelSynthesis

    ParallelScreening

    SerialScreening

    Scheme 1.9 Serial and parallel approaches to synthesis and screening

    Also in the field of asymmetric homogeneous catalysis, scientists have recognized the

    power of the combinatorial chemistry approach. Currently, the development of efficient

    catalysts for asymmetric catalysis is still largely empirical and often a result of knowledge-

    based intuition or serendipity. Divergent ligand synthesis strategies, in which several

    analogues of a promising ligand-type are prepared, appears to be an especially fruitful

    strategy in this area. However, the synthetic procedures for chiral ligands are often lengthy

    and unsuitable for such an approach. Due to limited time and resources usually only a small

    set of possibilities can be explored in a serial fashion. There is a clear need for a more

    methodical approach in which the structure of the catalyst is systematically varied.

    The parallel synthesis of chiral ligands with a modular build-up,74 coupled to high-

    throughput screening,75 may effectively address the problems involved with the

    optimization of asymmetric catalysts.76 Such a diversity-based approach will also allow for

    the identification of an optimal catalyst for each particular substrate, or class of substrates,

    thus overcoming the problem of generality that arises when only a small number of chiral

    catalysts are available. Although in combinatorial approaches, focussed on biological

    activity, it is quite common to use libraries of mixtures of compounds, such strategies are

    problematic in studies related to the identification of homochiral catalysts.77 Due to the

  • Chapter 1

    14

    05_Chapter 1.doc

    extreme sensitivity of reaction rate and selectivity to small structural variations, the

    examination of mixtures of catalysts can give rise to misleading conclusions (i.e. two

    effective catalysts that afford high ee values, but with opposite configurations, will give a

    perception of low selectivity). Even though it was recently shown that catalysts derived

    from mixtures of two different monodentate ligands can give superior results to their

    respective homocombinations,78 it is still preferable to employ parallel reactions in which

    one ligand is produced per vial. This strategy offers the advantage that each ligand structure

    is available spatially addressable and pure. Impressive results have been obtained with

    solid-phase bound libraries,79 however, in general the translation of this chemistry to

    solution phase can be quite problematic. Parallel synthesis of asymmetric catalysts is,

    therefore, mostly performed in solution phase.

    Chiral ligands, employed in a parallel synthesis/screening approach, require a modular

    build-up with easily connectable components. From an industrial perspective, where

    successful catalytic procedures will be subject to scale-up, the ligands should also be cost-

    effective when produced in larger quantities.76d After a seminal report of Gilbertson et al.

    on the synthesis of modular ligand libraries of phosphane-containing polypeptides for

    asymmetric hydrogenation,80 several reports have appeared concerning the parallel

    synthesis of ligand libraries for asymmetric catalysis.76 Amino acid building blocks81 have

    proven a popular choice in initial explorations at the interface of combinatorial chemistry

    and asymmetric catalysis.82 Recently, Lefort et al. reported the automated parallel synthesis

    and in situ screening of libraries of monodentate phosphoramidite ligands in rhodium-

    catalyzed hydrogenation reactions.83 This ligand class has been proven to be highly

    successful in a wide variety of metal-catalyzed enantioselective reactions, including:

    1) rhodium-catalyzed hydrogenations84 and conjugate additions of trifluoroborates85 and

    boronic acids,35 2) a palladium-catalyzed intramolecular Heck coupling,86 3) copper-

    catalyzed ring-opening of oxabicyclic alkenes,87 desymmetrization of epoxides,88 conjugate

    additions of diorganozinc reagents,19,20 and allylic alkylations,89 and 4) iridium-catalyzed

    allylic substitutions.90 A different member of the ligand family is required for most of these

    reaction-types and often for each individual substrate class.

  • Introduction

    15

    05_Chapter 1.doc

    The protocol of Lefort et al.83 allows for the parallel preparation of a solution phase library

    of phosphoramidites in a 96-well format in one day and their subsequent in situ parallel

    screening (Scheme 1.10). Other monodentate ligands, such as phosphites and phosphinites,

    should also be highly suitable for a similar approach. Their automated parallel synthesis

    has, however, yet to be reported.

    PO

    OCl

    HN(R1)R2

    *

    PO

    ON(R1)R2*

    Et3N

    Orbital Shaker

    Parallel

    Filtration

    Metal

    Precursors

    Prochiral

    SubstratesSolution Phase Library

    96-Well Microplate

    Et3N HCl- ( )

    Crude Phosphoramidite

    96-well Oleophobic Filterplate

    Parallel

    Reactor

    OO

    P ClPO

    OCl* =

    R3

    R3R4

    R4

    , OOP Cl

    R3

    R3

    Scheme 1.10 Protocol for the synthesis of a phosphoramidite library

    Phosphochloridites can be readily prepared from the corresponding diols and an excess of

    PCl3. With the help of a liquid handling robot, stock solutions of the phosphorchloridites,

    amines, and triethylamine were dispensed directly into a 96-well oleophobic filterplate. The

    vials were vortexed for 2 h, after which parallel filtration into a 96-well titerplate − in order

    to remove the precipitated Et3N·HCl salt − gave the liquid phase ligand library. The ligand

    stock solutions can be transfered to a parallel reactor for in situ complexation to a

    (transition) metal. The methodology offers the possibility to test multiple metals and

    substrates concurrently. Duursma et al. recently used this technology in order to screen a

    library of phosphoramidites in the rhodium-catalyzed conjugate addition of

  • Chapter 1

    16

    05_Chapter 1.doc

    vinyltrifluoroborates to cyclic and acyclic enones, effectively testing 96 ligands on two

    substrates in one run.85b It was shown that the method results in quick discovery of leads for

    effective enantioselective catalysts. Considering the broad range of asymmetric reactions

    that are catalyzed by monodentate phosphorus ligands,91 and the recently introduced

    possibility of using combinations of monodentate ligands,78 the methodology at hand has an

    enormous potential in the efficient development of highly enantioselective catalysts that

    have currently remained elusive.

    1.6 Aims and Outline of this Thesis

    As described in §1.3, the rhodium-catalyzed conjugate addition of boronic acids is a highly

    convenient method for the simultaneous construction of sp2-sp3 carbon-carbon bonds and

    stereogenic centers. Monodentate phosphoramidite ligands lead to highly enantioselective

    catalysts in this reaction. In part, the aim of this thesis is to expand the scope of this

    reaction with more challenging substrates, leading to useful chiral products. In Chapter 2

    the enantioselective synthesis of 2-aryl-4-piperidones by rhodium/phosphoramidite-

    catalyzed conjugate addition of arylboronic acids is described. This class of piperidones are

    important intermediates in the preparation of both naturally occurring alkaloids and a range

    of synthetic drugs. Chapter 3 describes an investigation into the feasibility of a selective

    rhodium/phosphoramidite-catalyzed addition of arylboronic acids to dehydroalanine

    derivatives. In this “tandem-reaction” the introduction of an aryl-substituent at the

    β-position by conjugate addition needs to proceed with concomitant control of the chirality

    at the α-center. The products are unnatural α-amino acids, which are increasingly important

    in the fields of drug discovery and protein engineering.

    It becomes apparent in §1.4 that catalysts for the conjugate addition of arylboronic acids to

    alkenes are often also suitable for their 1,2-addition to benzaldehydes and N-protected

    benzaldimines. The products of such reactions are important building blocks for a range of

    organic compounds with pharmaceutical properties. The powerful method of parallel

    synthesis and in situ screening of phosphoramidite ligands, as described in §1.5, offers the

    potential to find the most efficient catalyst for both of these substrate classes. The second

  • Introduction

    17

    05_Chapter 1.doc

    aim of this thesis is the development of new rhodium/phosphoramidite-catalysts for the 1,2-

    addition of arylboronic acids to aldehydes and imines. In Chapter 4 the enantioselective

    synthesis of diarylmethanols through the rhodium/phosphoramidite-catalyzed addition of

    arylboronic acids to aldehydes is described. Chapter 5 describes the asymmetric synthesis

    of N-protected diarylmethylamines from their corresponding imines. In addition, the

    possibility of using activated ketones as substrates for this reaction has been investigated. In

    Chapter 6 the synthesis of 3-aryl-3-hydroxyoxindoles through 1,2-addition of arylboronic

    acids to isatins will be described. Finally, in Chapter 7 the addition of arylboronic acids to

    2,2,2-trifluoroacetophenones is discussed.

    1.7 References and Notes

    1. Chirality (Greek, handedness, derived from the word stem χειρ~, ch[e]ir~, hand~) is a “dissymmetry” property important in several branches of science. An object or a system is called chiral if it differs from its mirror image. The term chirality was coined by Lord Kelvin: (a) W. T. Kelvin, Baltimore Lectures on Molecular Dynamics and the Wave Theory of Light, C. J. Clay and Sons: London, 1904. For further stereochemical definitions, see (b) E. L. Eliel, S. H. Wilen, Stereochemistry of Organic Compounds, Wiley: New York, 1994.

    2. (a) F. Krick, Life Itself, McDonald & Co.: London, 1981. (b) M. Gardner, The Ambidextrous Universe, 2nd Ed., C. Scribner: New York, Harmondsworth: UK, 1982.

    3. For a review on chemical signaling among bacteria, see: G. J. Lyon, T. W. Muir, Chem. Biol. 2003, 10, 1007.

    4. E. J. Ariens, W. Soudijin, P. B. M. W. M. Timmermans (Eds.), Stereochemistry and Biological Activity of Drugs, Blackwell Scientific Publications: Oxford, 1983.

    5. Food and Drug Administration, Fed. Reg. 1992, 22:249.

    6. A. M. Rouhi, Chem. Eng. News 2003, 81, 45.

    7. S. Hanessian, Total Synthesis of Natural Products: the “Chiron” approach, Pergamon Press: Oxford, 1983.

    8. (a) J. Jacques, A. Collet, S. H. Wilen, Enantiomers, racemates, and resolutions, Wiley: New York, 1981. (b) A. N. Collins, G. N. Sheldrake, J. Crosby (Eds.), Chirality in Industry II, Wiley: Chichester, 1997. (c) T. Vries, H. Wynberg, E. van Echten, J. Koek, W. ten Hoeve, R. M. Kellogg, Q. B. Boxterman, A. J. Minnaard,

  • Chapter 1

    18

    05_Chapter 1.doc

    B. Kaptein, S. van der Sluis, L. Hulshof, J. Kooistra, Angew. Chem. Int. Ed. 1998, 37, 2349.

    9. B. M. Trost, Angew. Chem. Int. Ed. Engl. 1995, 34, 259.

    10. J. Seyden-Penne, Chiral Auxiliaries and Ligands in Asymmetric Catalysis, Wiley: New York, 1995.

    11. (a) J. Wagner, R. A. Lerner, C. F. Barbaras III, Science 1995, 270, 1797. (b) A. M. Klibanov, Nature 2001, 409, 241.

    12. For reviews on the directed evolution of enantioselective enzymes, see: (a) M. T. Reetz, Tetrahedron 2002, 58, 6595; (b) M. T. Reetz, Proc. Natl. Acad. Sci. USA 2004, 101, 5716; (b) M. T. Reetz In Methods in Enzymology, Vol. 388, D. E. Robertson, J. P. Noel (Eds.), Elsevier: San Diego, 2004, 238. See also: (d) N. J. Turner, Trends Biotechnol. 2003, 21, 474.

    13. A. Berkessel, H. Gröger (Eds.), D. MacMillan, Asymmetric Organocatalysis: From Biomimetic Synthesis to Applications in Asymmetric Synthesis, Wiley-VCH: Weinheim, 2005.

    14. E. N. Jacobsen, A. Pfaltz, H. Yamamoto (Eds.), Comprehensive Asymmetric Catalysis, Vol. 1-3, Springer: Berlin, 1999.

    15. (a) W. S. Knowles, Angew. Chem. Int. Ed. 2002, 41, 1998. (b) R. Noyori, Angew. Chem. Int. Ed. 2002, 41, 2008. (c) K. B. Sharpless, Angew. Chem. Int. Ed. 2002, 41, 2024.

    16. K. Tomioka, Y. Nagaoka In Comprehensive Asymmetric Catalysis, Vol. 3, E. N. Jacobsen, A. Pfaltz, H. Yamamoto (Eds.), Springer: Berlin, 1999, §31.1.

    17. (a) N. Krause, A. Gerold, Angew. Chem. Int. Ed. 1997, 36, 186. (b) N. Krause, Angew. Chem. Int. Ed. 1998, 37, 283. (c) B. L. Feringa, Acc. Chem. Res. 2000, 33, 346. (d) N. Krause, A. Hoffman-Röder, Synthesis 2001, 171. (e) B. L. Feringa, R. Naasz, R. Imbos, L. A. Arnold In Modern Organocopper Chemistry, N. Krause (Ed.), Wiley-VCH: Weinheim, 2002, Chapter 7, 224. (f) A. Alexakis, C. Benhaim, Eur. J. Org. Chem. 2002, 3221. (f) A. Alexakis, D. Polet, S. Rosset, S. March, J. Org. Chem. 2004, 69, 5660.

    18. For the first example of chiral phosphine ligands in this reaction (30% ee), see: A. Alexakis, J. Frutos, P. Mangeney, Tetrahedron: Asymmetry 1993, 4, 2427.

    19. (a) A. H. M. de Vries, A. Meetsma, B. L. Feringa, Angew. Chem. Int. Ed. 1996, 35, 2374. (b) B. L. Feringa, M. Pineschi, L. A. Arnold, R. Imbos, A. H. M. de Vries, Angew. Chem. Int. Ed. 1997, 36, 2620. (c) R. Naasz, L. A. Arnold, A. J. Minnaard, B. L. Feringa, Angew. Chem. Int. Ed. 2001, 40, 927. (d) F. Bertozzi, P. Crotti, F. Macchia, M. Pineschi, B. L. Feringa, Angew. Chem. Int. Ed. 2001, 40, 930. (e) L. A. Arnold, R. Naasz, A. J. Minnaard, B. L. Feringa, J. Am. Chem. Soc. 2003,

  • Introduction

    19

    05_Chapter 1.doc

    125, 3700. (f) R. Šebesta, M. G. Pizzuti, A. J. Boersma, A. J. Minnaard, B. L. Feringa, Chem. Commun. 2005, 1711.

    20. D. Peña, F. Lopez, S. R. Harutyunyan, A. J. Minnaard, B. L. Feringa, Chem. Commun. 2004, 1836.

    21. (a) Y. Takaya, M. Ogasawara, T. Hayashi, M. Sakai, N. Miyaura, J. Am. Chem. Soc. 1998, 120, 5579. For the first (non-asymmetric) rhodium-catalyzed conjugate addition of aryl- and alkenylboronic acids, see: (b) M. Sakai, H. Hayashi, N. Miyaura, Organometallics 1997, 16, 4229.

    22. For a recent review, see: (a) T. Hayashi, K. Yamasaki, Chem. Rev. 2003, 103, 2829; (b) K. Fagnou, M. Lautens, Chem. Rev. 2003, 103, 169.

    23. For some recent results of conjugate additions to other α,β-unsaturated systems, see: (a) R. Shintani, K. Ueyama, I. Yamada, T. Hayashi, Org. Lett. 2004, 6, 3425 for fumaric and maleic substrates; (b) J. F. Paquin, C. Defieber, C. R. J. Stephenson, E. M. Careira, J. Am. Chem. Soc. 2005, 127, 10850 and (c) T. Hayashi, N. Tokunaga, K. Okamoto, R. Shintani, Chem. Lett. 2005, 34, 1480 for enals; (d) J. F. Paquin, C. R. J. Stephenson, C. Defieber, E. M. Carreira, Org. Lett. 2005, 7, 3821 for acrylic esters; (e) R. Shintani, W. L. Duan, T. Nagano, A. Okada, T. Hayashi, Angew. Chem. Int. Ed. 2005, 44, 4611 for maleimides; (f) R. Shintani,, T. Kimura, T. Hayashi, Chem. Commun. 2005, 25, 3212 for Weinreb amides; (g) G. Chen, N. Tokunaga, T. Hayashi, Org. Lett. 2005, 7, 2285 and references therein for coumarins.

    24. (a) S. Oi, M. Moro, S. Ono, Y. Inoue, Chem. Lett. 1998, 27, 83. (b) T. Hayashi, M. Ishigedani, J. Am. Chem. Soc. 2000, 122, 976. (c) S. Oi, M. Moro, H. Ito, Y. Honma, S. Miyano, Y. Inoue, Tetrahedron 2002, 58, 91. (d) M. Dziedzic, M. Malecka, B. Furman, Org. Lett. 2005, 7, 1725.

    25. S. Oi, Y. Honma, Y. Inoue, Org. Lett. 2002, 4, 667.

    26. (a) T. Hayashi, N. Tokunaga, K. Yoshida, J. W. Han, J. Am. Chem. Soc. 2002, 124, 12102. (b) N. Tokunaga, K. Yoshida, T. Hayashi, Proc. Natl. Acad. Sci. U. S . A. 2004, 101, 5445.

    27. K. C. Nicolaou, W. Tang, P. Dagneau, R. Faraoni, Angew. Chem. Int. Ed. 2005, 44, 3874.

    28. (a) N. Miyaura, Z. Suzuki, Chem. Rev. 1995, 95, 2457. (b) D. G. Hall (Ed.), Boronic Acids: Preparation and Applications in Organic Synthesis and Medicine, Wiley-VCH: Weinheim, 2005.

    29. (a) R. Amengual, V. Michelet, J.-P. Genêt, Synlett 2002, 1791. (b) M. Pucheault, S. Darses, J.-P. Genêt, Eur. J. Org. Chem. 2002, 3552. (c) Q. Shi. L. Xu, X. Li, X. Jia, R. Wang, T. T.-L. Au-Yeung, A. S. C. Chan, T. Hayashi, R. Cao, M. Hong, Tetrahedron Lett. 2003, 44, 6505. (d) J. Madec, G. Michaud, J.-P. Genêt, A. Marinetti, Tetrahedron: Asymmetry 2004, 15, 2253. (e) W.-Y. Yuan, L.-F. Chun,

  • Chapter 1

    20

    05_Chapter 1.doc

    L.-Z. Gong, A.-Q. Mi, Y.-Z. Jiang, Tetrahedron Lett. 2005, 46, 509. (f) M. Ogasawara, H. L. Ngo, T. Sakamoto, T. Takahashi, W. Lin, Org. Lett. 2005, 7, 2881. (g) K. Vandyck, B. Matthys, M. Willen, K. Robeyns, L. van Meervelt, J. van der Eycken, Org. Lett. 2006, 8, 363.

    30. M. T. Reetz, D. Moulin, A. Gosberg, Org. Lett. 2001, 3, 4083.

    31. M. Kuriyama, K. Nagai, K. Yamada, Y. Miwa, T. Taga, K. Tomioka, J. Am. Chem. Soc. 2002, 124, 8932.

    32. Y. Ma, S. Song, C. Ma, Z. Sun, Q. Chai, M. B. Andrus, Angew. Chem. Int. Ed. 2003, 42, 5871.

    33. T. Imamoto, K. Sugita, K. Yoshida, J. Am. Chem. Soc. 2005, 127, 11934.

    34. (a) T. Hayashi, K. Ueyama, N. Tokunaga, K. Yoshida, J. Am. Chem. Soc. 2003, 125, 11508. (b) C. Defieber, J.-F. Paquin, S. Serna, E. M. Carreira, Org. Lett. 2004, 6, 3873. (c) Y. Otomaru, A. Kina, R. Shintani, T. Hayashi, Tetrahedron: Asymmetry 2005, 16, 1673. (d) Y. Otomaru, K. Okamoto, R. Shintani, T. Hayashi, J. Org. Chem. 2005, 70, 2503. (e) F. Läng, F. Breher, D. Stein, H. Grützmacher, Organometallics 2005, 24, 2997. (f) R. Shintani, K. Okamoto, Y. Otomaru, K. Ueyama, T. Hayashi, J. Am. Chem. Soc. 2005, 127, 54. (g) R. Shintani, K. Okamoto, T. Hayashi, Chem. Lett. 2005, 34, 1294. (h) F.-X. Chen, A. Kina, T. Hayashi, Org. Lett. 2006, 8, 341.

    35. (a) J.-G. Boiteau, R. Imbos, A. J. Minnaard, B. L. Feringa, Org. Lett. 2003, 5, 681; see also: Org. Lett. 2003, 5, 1385. (b) J.-G. Boiteau, A. J. Minnaard, B. L. Feringa, J. Org. Chem. 2003, 68, 9481. (c) A. Duursma, R. Hoen, J. Schuppan, R. Hulst, A. J. Minnaard, B. L. Feringa, Org. Lett. 2003, 5, 3111.

    36. Y. Iguchi, R. Itooka, N. Miyaura, Synlett. 2003, 1040.

    37. T. Hayashi, M. Takahashi, Y. Takaya, M. Ogasawara, J. Am. Chem. Soc. 2002, 124, 5052.

    38. (a) R. Itooka, Y. Iguchi, N. Miyaura, J. Org. Chem. 2003, 68, 6000. (b) S. L. X. Martina, A. J. Minnaard, B. Hessen, B. L. Feringa, Tetrahedron Lett. 2005, 46, 7159.

    39. For a recent review on catalytic asymmetric approaches towards enantiomerically enriched diarylmethanols and diarylmethylamines, see: F. Schmidt, R. T. Stemmler, J. Rudolph, C. Bolm, Chem. Soc. Rev. 2006, 35, 454.

    40. A. F. Harms, W. Hespe, W. T. Nauta, R. F. Rekker, Drug Design 1976, 6, 1.

    41. M. Saikai, M. Ueda, N. Miyaura, Angew. Chem. Int. Ed. 1998, 37, 3279.

    42. For the addition of arylboronic acids to aldehydes leading to racemic products, see: reference 41 and: (a) M. Ueda, N. Miyaura, J. Org. Chem. 2000, 65, 4450; (b) A. Fürstner, H. Krause, Adv. Synth. Catal. 2001, 343, 343.

  • Introduction

    21

    05_Chapter 1.doc

    43. C. Moreau, C. Hague, A. S. Weller, C. G. Frost, Tetrahedron Lett. 2001, 42, 6957.

    44. A. Fürstner, H. Krause, Adv. Synth. Catal. 2001, 42, 6957.

    45. T. Focken, J. Rudolph, C. Bolm, Synthesis 2005, 429.

    46. (a) T. Hayashi, M. Ishigedani, J. Am. Chem. Soc. 2000, 122, 976. (b) M. Ueda, N. Miyaura, J. Organomet. Chem. 2000, 595, 31.

    47. T. Hayashi, M. Kawai, N. Tokunaga, Angew. Chem. Int. Ed. 2004, 43, 6125.

    48. M. Kuriyama, T. Soeta, X. Hao, Q. Chen, K. Tomioka, J. Am. Chem. Soc. 2004, 126, 8128.

    49. D. J. Weix, Y. Shi, J. A. Ellman, J. Am. Chem. Soc. 2005, 127, 1092.

    50. (a) N. Tokunaga, Y. Otomaru, K. Okamoto, K. Ueyama, R. Shintani, T. Hayashi, J. Am. Chem. Soc. 2004, 126, 13584. (b) Y. Otomaru, N. Tokunaga, R. Shintani, T. Hayashi, Org. Lett. 2005, 7, 307.

    51. For reviews on the addition of zinc reagents to aldehydes, see : (a) R. Noyori, M. Kitamura, Angew. Chem. Int. Ed. 1991, 30, 49; (b) K. Soai, S. Niwa, Chem. Rev. 1992, 92, 883; (c) L. Pu, H.-B. Yu, Chem. Rev. 2001, 101, 757.

    52. (a) P. I. Dosa, J. C. Ruble, G. C. Fu, J. Org. Chem. 1997, 62, 444. (b) P. I. Dosa, G. C. Fu, J. Am. Chem. Soc. 1998, 120, 445.

    53. W.-S. Huang, L. Pu, J. Org. Chem. 1999, 64, 4222.

    54. (a) W.-S. Huang, Q.-S. Hu, L. Pu, J. Org. Chem. 1998, 63, 1364. (b) W.-S. Huang, Q.-S. Hu, L. Pu, J. Org. Chem. 1999, 64, 7940.

    55. This phenomenon was also observed by Blacker, see: J. Blacker In Proceedings of the 3rd International Conference on the Scale Up of Chemical Processes, T. Laird (Ed.), Scientific Update Mayfield: East Sussex, 1998, 74.

    56. W.-S. Huang, L. Pu, Tetrahedron Lett. 2000, 41, 145.

    57. C. Bolm, K. Muñiz, Chem. Commun. 1999, 1295.

    58. (a) C. Bolm, N. Hermans, J. P. Hildebrand, K. Muñiz, Angew. Chem. Int. Ed. 2000, 39, 3465. (b) J. Rudolph, C. Bolm, P. O. Norrby, J. Am. Chem. Soc. 2005, 127, 1548.

    59. N. Hermanns, S. Dahmen, C. Bolm, S. Bräse, Angew. Chem. Int. Ed. 2002, 41, 3692.

    60. For a review, see reference 39. (a) C. Bolm, M. Kesselgruber, A. Grenz, N. Hermanns, A. J. P. Hildebrand, New. J. Chem. 2001, 25, 13. (b) C. Bolm, M. Kesselgruber, N. Hermanns, J. P. Hildebrand, G. Raabe, Angew. Chem. Int. Ed. 2001, 40, 1488. (c) G. Zhao, X.-G. Li, X.-R. Wang, Tetrahedron: Asymmetry 2001, 12, 399. (d) D.-H. Ko, K. H. Kim, D.-C. Ha, Org. Lett. 2002, 4, 3759. (e) M. G.

  • Chapter 1

    22

    05_Chapter 1.doc

    Pizzuti, S. Superchi, Tetrahedron: Asymmetry 2005, 16, 2263. (f) M. Fontes, X. Verdaguer, L. Solà, M. A. Pericàs, A. Riera, J. Org. Chem. 2004, 69, 2532. (g) M. Hatano, T. Hiyamoto, K. Ishihara, Adv. Synth. Catal. 2005, 347, 1561. (h) Y.-C. Qin, L. Pu, Angew. Chem. Int. Ed. 2006, 45, 273. (i) I. Schiffers, T. Rantanen, F. Schmidt, W. Bergmans, L. Zani, C. Bolm, J. Org. Chem. 2006, 71, 2320.

    61. (a) J. Rudolph, F. Schmidt, C. Bolm, Adv. Synth. Catal. 2004, 346, 867. (b) S. Dahmen, M. Lormann, Org. Lett. 2005, 7, 4597.

    62. C. Bolm, J. Rudolph, J. Am. Chem. Soc. 2002, 124, 14850.

    63. For recent examples, see: (a) J.-X. Ji, J. Wu, T. T.-L. Au-Yeung, C.-W. Yip, R. K. Haynes, A. S. C. Chan, J. Org. Chem. 2005, 70, 1095; (b) A. L. Braga, D. S. Lüdtke, P. H. Schneider, F. Vargas, A. Schneider, L. A. Wessjohann, M. W. Paixão Tetrahedron Lett. 2005, 46, 7827 and references cited in these articles.

    64. J. Rudolph, F. Schmidt, C. Bolm, Synthesis 2005, 840.

    65. For further studies on the effect of additives in this reaction, see: (a) J. Rudolph, N. Hermanns, C. Bolm, J. Org. Chem. 2004, 69, 3997; (b) J. Rudolph, M. Lormann, C. Bolm, S. Dahmen, Adv. Synth. Catal. 2005, 347, 1361.

    66. K. Ito, Y. Tomita, T. Katsuki, Tetrahedron Lett. 2005, 46, 6083.

    67. C. Bolm, F. Schmidt, L. Zani, Tetrahedron: Asymmetry 2005, 16, 2299.

    68. S. Özçubukçu, F. Schmidt, C. Bolm, Org. Lett. 2005, 7, 1407.

    69. J.-X. Ji, J. Wu, T. T.-L. Au-Yeung, C.-W. Yip, R. K. Haynes, A. S. C. Chan, J. Org. Chem. 2005, 70, 1093.

    70. P.-Y. Wu, H.-L. Wu, B.-J. Uang, J. Org. Chem. 2006, 71, 833.

    71. A. L. Braga, D. S. Lüdtke, F. Vargas, M. W. Paixão, Chem. Commun. 2005, 2512.

    72. X. Wu, X. Liu, G. Zhao, Tetrahedron: Asymmetry 2005, 16, 2299.

    73. For a brief history of combinatorial chemistry, see: K. C. Nicolaou, R. Hanko, W. Hartwig, In Handbook of Combinatorial Chemistry, Vol. 1, K. C. Nicolaou, R. Hanko, W. Hartwig (Eds.), Wiley-VCH: Weinheim, 2002, 3.

    74. (a) B. Jandeleit, D. Schaefer, T. S. Powers, H. W. Turner, W. H. Weinberg, Angew. Chem. Int. Ed. 1999, 38, 2494. (b) M. T. Reetz, Angew. Chem. Int. Ed. 2001, 40, 284.

    75. (a) A. Duursma, A. J. Minnaard, B. L. Feringa, Tetrahedron 2002, 58, 5773. (b) M. T. Reetz, Angew. Chem. Int. Ed. 2002, 41, 1335.

    76. For reviews on the field of catalyst development by combinatorial chemistry in general, see reference 74a and: (a) A. H. Hoveyda, Chem. Biol. 1998, 5, R187; (b) B. Archibald, O. Brümmer, M. Devenney, S. Gorer, B. Jandeleit, T. Uno, W. H.

  • Introduction

    23

    05_Chapter 1.doc

    Weinberg, T. Weskamp, In Handbook of Combinatorial Chemistry, Vol. 2, K. C. Nicolaou, R. Hanko, W. Hartwig (Eds.), Wiley-VCH: Weinheim, 2002, 885. For reviews on the field of asymmetric catalyst development by combinatorial chemistry, see: (c) A. H. Hoveyda, In Handbook of Combinatorial Chemistry, Vol. 2, K. C. Nicolaou, R. Hanko, W. Hartwig (Eds.), Wiley-VCH: Weinheim, 2002, 991; (d) J. G. de Vries, A. H. M. de Vries, Eur. J. Org. Chem. 2003, 799; (e) C. Gennari, U. Piarulli, Chem. Rev. 2003, 103, 3071.

    77. There are exceptions, see: (a) C. Hinderling, P. Chen, Angew. Chem. Int. Ed. 1999, 38, 2253; (b) P. Krattiger, C. McCarthy, A. Pfaltz, H. Wennemers, Angew. Chem. Int. Ed. 2003, 42, 1722.

    78. For selected examples, see: (a) D. Peña, A. J. Minnaard, J. A. F. Boogers, A. H. M. de Vries, J. G. de Vries, B. L. Feringa, Org. Biomol. Chem. 2003, 1, 1087; (b) M. T. Reetz, T. Sell, A. Meiswinkel, G. Mehler, Angew. Chem. Int. Ed. 2003, 42, 790; (c) C. Monti, C. Gennari, U. Piarulli, J. G. de Vries, A. H. M. de Vries, L. Lefort, Chem. Eur. J. 2005, 11, 6701; (d) R. Hoen, J. A. F. Boogers, H. Bernsmann, A. J. Minnaard, A. Meetsma, T. D. Tiemersma-Wegman, A. H. M. de Vries, J. G. de Vries, B. L. Feringa, Angew. Chem. Int. Ed. 2005, 44, 4209.

    79. N. E. Leadbeater, M. Marco, Chem. Rev. 2002, 102, 3217.

    80. S. R. Gilbertson, X. Wang, Tetrahedron Lett. 1996, 37, 6475.

    81. G. Liu, J. A. Ellmann, J. Org. Chem. 1995, 60, 7712.

    82. (a) B. M. Cole, K. D. Schimizu, C. A. Krueger, J. P. A. Harrity, M. L. Snapper, A. H. Hoveyda, Angew. Chem. Int. Ed. 1996, 35, 1668. (b) K. D. Shimizu, B. M. Cole, C. A. Krueger, K. W. Kuntz, M. L. Snapper, A. H. Hoveyda, Angew. Chem. Int. Ed. 1997, 36, 1704. (c) M. S. Sigman, E. N. Jacobsen, J. Am. Chem. Soc. 1998, 120, 4901. (d) M. S. Sigman, E. N. Jacobsen, J. Am. Chem. Soc. 1998, 120, 5315. (e) C. Gennari, S. Ceccarelli, U. Piarulli, C. A. G. N. Montalbetti, R. F. W. Jackson, J. Org. Chem. 1998, 63, 5312. (f) A. J. Brouwer, H. J. van der Linden, R. M. J. Liskamp, J. Org. Chem. 2000, 65, 1750. (g) I. Chataigner, C. Gennari, U. Piarulli, S. Ceccarelli, Angew. Chem. Int. Ed. 2000, 39, 916.

    83. (a) L. Lefort, J. A. F. Boogers, A. H. M. de Vries, J. G. de Vries, Org. Lett. 2004, 6, 1733. (b) M. van den Berg, D. Peña, A. J. Minnaard, B. L. Feringa, L. Lefort, J. A. F. Boogers, A. H. M. de Vries, J. G. de Vries, Chim. Oggi 2004, 22, detachable insert: Chiral Catalysis-Asymmetric Hydrogenation, 18. (c) J. G. de Vries, L. Lefort, Chem. Eur. J. 2006, 12, 4722.

    84. (a) M. van den Berg, A. J. Minnaard, E. P. Schudde, J. van Esch, A. H. M. de Vries, J. G. de Vries, B. L. Feringa, J. Am. Chem. Soc. 2000, 122, 11539. (b) D. Peña, A. J. Minnaard, J. G. de Vries, B. L. Feringa, J. Am. Chem. Soc. 2002, 124, 14552. (c) M. van den Berg, A. J. Minnaard, R. M. Haak, M. Leeman, E. P. Schudde, A. Meetsma, B. L. Feringa, A. H. M. de Vries, C. E. P. Maljaars, C. E. Willans, D. Hyett, J. A. F.

  • Chapter 1

    24

    05_Chapter 1.doc

    Boogers, H. J. W. Henderickx, J. G. de Vries, Adv. Synth. Catal. 2003, 345, 308. (d) L. Panella, B. L. Feringa, J. G. de Vries, A. J. Minnaard, Org. Lett. 2005, 7, 4177. (e) L. Panella, A. Marco Aleixandre, G. J. Kruidhof, J. Robertus, B. L. Feringa, J. G. de Vries, A. J. Minnaard, J. Org. Chem. 2006, 71, 2026.

    85. (a) A. Duursma, J.-G. Boiteau, L. Lefort, J. A. F. Boogers, A. H. M. de Vries, J. G. de Vries, A. J. Minnaard, B. L. Feringa, J. Org. Chem. 2004, 69, 8045. (b) A. Duursma, L. Lefort, J. A. F. Boogers, A. H. M. de Vries, J. G. de Vries, A. J. Minnaard, B. L. Feringa, Org. Biomol. Chem. 2004, 2, 1682.

    86. (a) R. Imbos, A. J. Minnaard, B. L. Feringa, J. Am. Chem. Soc. 2002, 124, 184. (b) R. Imbos, A. J. Minnaard, B. L. Feringa, J. Chem. Soc., Dalton Trans. 2003, 2017.

    87. F. Bertozzi, M. Pineschi, F. Macchia, L. A. Arnold, A. J. Minnaard, B. L. Feringa, Org. Lett. 2002, 4, 2703.

    88. F. Del Moro, P. Crotti, V. Di Bussolo, F. Macchia, M. Pineschi, Org. Lett. 2003, 5, 1971.

    89. (a) H. Malda, A. W. van Zijl, L. A. Arnold, B. L. Feringa, Org. Lett. 2001, 3, 1169. (b) A. W. van Zijl, L. A. Arnold, A. J. Minnaard, B. L. Feringa, Adv. Synth. Catal. 2004, 346, 413.

    90. For iridium-catalyzed allylic alkylation reactions, see: (a) B. Bartels, C. Garcia-Yebra, G. Helmchen, Eur. J. Org. Chem. 2003, 1097. For iridium-catalyzed allylic amination reactions, see: (b) T. Ohmura, J. F. Hartwig, J. Am. Chem. Soc. 2002, 124, 15164; (c) G. Lipowsky, G. Helmchen, Chem. Commun. 2004, 116. For iridium-catalyzed allylic etherification reactions, see: (d) F. Lopez, T. Ohmura, J. F. Hartwig, J. Am. Chem. Soc. 2003, 125, 3426.

    91. See references 17, 19, 35, 84-90, and: (a) R. Imbos, Catalytic Asymmetric Conjugate Additions and Heck Reactions, Ph.D. Thesis, University of Groningen, 2002, 1-24. (b) A. Duursma, Asymmetric Catalysis with Chiral Monodentate Phosphoramidite Ligands, Ph.D. Thesis, University of Groningen, 2002, 6-17.

  • 06_Chapter_2.doc

    Chapter 2 Enantioselective Synthesis of 2-Aryl-4-piperidones

    NCO2Bn

    O

    NCO2Bn

    ORh(acac)(C2H4)2 (3 mol%)phosphoramidite (7.5 mol%)

    (ArBO)31,4-dioxane/H2O, 100 ºC

    R

    Up to 99% eeHigh yields

    Piperidones are important intermediates in the preparation of natural occurring alkaloids

    and synthetic pharmacophores. In this chapter the highly enantioselective synthesis of

    2-aryl-4-piperidones by rhodium/phosphoramidite-catalyzed conjugate addition of

    arylboroxines to 2,3-dihydro-4-pyridones is described.1 A variety of products with

    sterically and electronically different R-substituents have been obtained in high isolated

    yield and with excellent ee’s up to 99%.

    Part of this chapter has been published: R. B. C. Jagt, J. G. de Vries, B. L. Feringa, A. J.

    Minnaard, “Enantioselective Synthesis of 2-Aryl-4-piperidones via Rhodium/

    Phosphoramidite-Catalyzed Conjugate Addition of Arylboroxines”, Org. Lett. 2005, 7,

    2433.

  • 26

    06_Chapter_2.doc

    Chapter 2

    2.1 Introduction

    The piperidine ring system (1, Figure 2.1) is a frequently encountered heterocyclic unit in

    natural compounds and drug candidates.2 Naturally occurring piperidine alkaloids (e.g.

    2-11) and their synthetic analogues are of great interest to the pharmaceutical industry.

    NH

    NH

    CO2H NH

    piperidine 1 pipecolic acid 2 coniine 3

    NH

    HO

    Me (CH2)7CO2H

    carpamic acid 4

    NH

    HN O

    OHN

    NH

    Cl

    Cl

    NH

    OH

    DKP593A 5 histrionicotoxin 6

    NH

    N

    NH

    N

    N NCH3

    HOO OH

    O OH

    H3C

    anabasine 8 anabasamine 9 rohitukine 11

    NC2H5

    O

    O

    O

    campedine 10

    OHHO O

    NCH3

    morphine 7

    Figure 2.1 Naturally occurring piperidine alkaloids

    Considering the extensive range of biological activities these compounds exhibit, it is not

    surprising that (according to an assertion of Watson et al.) between 1988 and 1998

    thousands of piperidine-derived compounds were mentioned in clinical and preclinical

    studies.3 Piperidones serve an important role en route to substituted piperidines4 and can

    also be found as a part of more complex biologically active compounds.5 Therefore the

    development of short, enantioselective routes to substituted piperidones is a major goal.6

  • 27

    06_Chapter_2.doc

    Enantioselective Synthesis of 2-Aryl-4-piperidones

    Although few naturally occurring piperidine alkaloids contain the arylpiperidine moiety

    (8-11), the structural unit is an integral part of many polycyclic alkaloids (e.g. morphine, 7).

    Simple synthetic arylpiperidines have received increasing attention due to their biological

    activities, often resembling those of more complicated natural alkaloids. Many 3-aryl- and

    4-arylpiperidines are potent opioid receptor antagonists and can be regarded as structurally

    simplified forms of morphine.7 2-Arylpiperidines are of noteworthy interest as being

    integrated in biologically active benzo[a]- and indolo[2,3-a]quinolizidine compounds.8

    Recent biological studies of 2-arylpiperidines show a range of biological activities for this

    class of structures, revealing their potential use in the treatment of mental and

    cardiovascular diseases.9 An excellent example is found in a class of antidepressants, which

    have a common structural pattern based on a 2-arylpiperidine moiety (Figure 2.2).10, 11

    NH

    Ph

    HNAr

    12

    N

    NR2

    Me

    F NR2O13

    Figure 2.2 A class of antidepressants with a 2-arylpiperidine moiety

    Compound 13, developed by Glaxo Group Ltd., UK, is particularly useful for the treatment

    or prevention of depressive states and/or anxiety. It owes its unique pharmacological

    properties to a strong affinity with the NK1 tachykinin receptor, one of three tachykinin

    receptors identified (tachykinins are a family of peptide neurotransmitters).11 As a

    consequence of the significance of 2-arylpiperidines, the development of efficient methods

    for their enantioselective synthesis is an important objective. However, most of the existing

    methods rely on the use of a stoichiometric amount of chiral reagents.12

    Dihydropyridones of the type 14 (Figure 2.3) are versatile synthetic building blocks: they

    are easy to prepare, air-stable, and their functionality allows a variety of chemical

    transformations.13 Short, stereocontrolled synthesis of piperidine, indolizidine, quinolizine,

    and cis- and trans-decahydroquinoline alkaloids have been reported using 2,3-dihydro-4-

    pyridones as chiral building blocks.14 An attractive catalytic route toward enantiopure

  • 28

    06_Chapter_2.doc

    Chapter 2

    2-substituted piperidones is based on the enantioselective conjugate addition to readily

    available N-protected 2,3-dihydro-4-pyridones (14).15 Until recently, however, no suitable

    procedures have been developed.16,17,18

    N

    O

    PG 1,4-addition or reduction

    [2 + 2] photocycloaddition

    1,2-addition

    enolate alkylationor acetoxylation

    14

    Figure 2.3 Versatile 2,3-dihydro-4-pyridone building block (PG = Protecting Group)

    Very few catalytic enantioselective routes for the synthesis of 2-substituted piperidines

    have been reported.19 Recently, our group reported the synthesis of 2-alkyl-4-piperidones

    with high enantiomeric excess applying the copper-catalyzed conjugate addition of

    dialkylzinc reagents to 14 employing phosphoramidite ligand L1 (Scheme 2.1).17 It was

    noted that this type of substrate is less reactive towards 1,4-addition than cyclic enones, e.g.

    2-cyclohexenone.

    N

    O

    PGN

    O

    PGR

    Cu(OTf)2 (5 mol%)

    L1 (10 mol%)

    R2Zn

    PG = CO2Bn

    PG = CO2Ph

    PG = CO2Me

    PG = CO2Et

    PG = CO2tBu

    PG = p-Tosyl14

    R = Me, R = Et,

    R = iPr, R = Bu 15

    Up to 97% ee

    Up to 87% yield

    OO

    P N

    Ph

    PhL1

    14a: 14e: 14c: 14d: 14e: 14f:

    Scheme 2.1 Catalytic enantioselective addition of zinc reagents to N-substituted

    2,3-dihydro-4-pyridones

    Although highly enantioselective 1,4-addition of diphenylzinc, using the same catalyst, has

    been reported for 2-cyclohexenone,20 the lack of readily available diarylzinc reagents

    severely limits this method. A more convenient method for the introduction of aryl and

    alkenyl moieties is the asymmetric rhodium-catalyzed conjugate addition of boronic acids

  • 29

    06_Chapter_2.doc

    Enantioselective Synthesis of 2-Aryl-4-piperidones

    pioneered by Hayashi and Miyaura.21,22 Recently we have demonstrated that rhodium-

    catalyzed conjugate additions of arylboronic acids to enones can be achieved with high

    efficiency and excellent enantioselectivities employing monodentate phosphoramidite L2

    (Figure 2.4).23

    OO P N

    L2

    Figure 2.4 Chiral monodentate phosphoramidite ligand L2

    During our studies, Hayashi et al. reported the enantioselective addition of arylzinc

    chlorides to 2,3-dihydro-4-pyridones with excellent enantioselectivities.24 In that study, it

    was confirmed that this type of substrate is less reactive toward 1,4-addition compared to

    enones. The rhodium/BINAP-catalyzed conjugate addition of phenylboronic acid failed to

    proceed to full conversion, although the enantioselectivity was excellent. We envisioned

    that introduction of aryl groups in N-protected 2,3-dihydro-4-pyridones, using the

    rhodium/phosphoramidite-catalyzed conjugate addition of arylboronic acids, could provide

    a pathway to 2-substituted 4-piperidones that is complementary to our work with

    dialkylzinc reagents.

    2.2 Results and Discussion

    2.2.1 Preliminary Studies

    Substrate 14a was prepared in three steps from 4-piperidone monohydrate hydrochloride

    and benzyl chloroformate (Scheme 2.2, route a), following a literature procedure of Park

    et al.15a for the synthesis of ethyl 3,4-dihydro-4-oxo-1-(2H)-pyridinecarboxylate. The

    desired product was obtained as a white solid in an overall yield of 64%. The alternative

    one-pot procedure of Comins et al. (Scheme 2.2, route b)15b starting from

    4-methoxypyridine suffered from low reproducibility, providing 14a in 10% to 63% yield.

  • 30

    06_Chapter_2.doc

    Chapter 2

    N

    OMe

    NH

    O 1) ClCO2Bn, NaOH

    H2O, Et2O

    2) Br2, (HOCH2)2

    . HClNCO2Bn

    OOBr 1) DBU, Me2SO, 85 oC

    2) 3M HCl, MeOH NCO2Bn

    O

    a)

    b)1) K(i-PrO)3BH, THF, i-PrOH

    2) ClCO2Bn

    3) 3M HClNCO2Bn

    O

    14a

    14a

    10-63% yield

    64% overall yield

    (3 steps)

    Scheme 2.2 Preparation of 14a according to a) Park and b) Comins

    Initial screening of the conjugate addition of phenylboronic acid to 14a was performed

    under standard conditions in a mixture of 1,4-dioxane/water (10/1) at 100 oC with a catalyst

    generated from 3 mol% Rh(acac)(C2H4)2 and 7.5 mol% L2. As in the report of Hayashi,

    with 3 equiv of phenylboronic acid, the reaction did not go to completion according to 1H-NMR (Table 2.1, entry 1). The enantioselectivity was, however, excellent (96% ee).

    Water acts as a proton donor and facilitates the transmetallation, therefore the use of water

    is essential in order to complete the catalytic cycle (see Chapter 1, §1.3). A drawback of the

    use of water, especially in combination with high temperature, is the hydrolysis of

    phenylboronic acid to benzene. The arylboronic acid is therefore commonly added in

    excess (3-5 equiv) compared to the enone. Milder conditions can be provided by generating

    phenylboronic acid in situ from phenylboroxine ((PhBO)3) and one equiv of water with

    respect to boron (entries 2-6). Phenylboronic acid exists in equilibrium with its trimeric

    species and is released in the course of the reaction (Figure 2.5). Hayashi has demonstrated

    that the use of arylboroxines has a beneficial effect on both conversion and

    enantioselectivity in the conjugate addition to highly deactivated 1-alkenylphosphonates.25

    Dehydration of arylboronic acids results in the corresponding boroxines in quantitative

    yield by azeotropic removal of water from their xylene solution or heating neat in vacuo at

    145 oC.26 The use of boroxines did not immediately improve the conversion, but did

    improve the enantioselectivity to an excellent ee of 99%. Upon slow addition of water,

  • 31

    06_Chapter_2.doc

    Enantioselective Synthesis of 2-Aryl-4-piperidones

    thereby preventing premature hydrolysis of the boroxine, the reaction could be driven to

    84% conversion using 1 equiv of boroxine (entry 4) and to full conversion using 3 equiv of

    the reagent (entry 6), retaining 99% ee.

    Table 2.1 Optimization of the reaction conditions for the rhodium-catalyzed conjugate

    addition to 14a

    N

    O

    CO2Bn

    Rh(acac)(C2H4)2 (3 mol%)

    (R)-L2 (7.5 mol%)

    "PhB"

    1,4-dioxane/H2O, 100 oCN

    O

    CO2Bn

    14a 15a

    entry “PhB” (equiv) condition a conv. (%) b ee (%)c

    1 PhB(OH)2 (3.0) A 80 96

    2 (PhBO)3 (1.0) B 60 99

    3 (PhBO)3 (3.0) B 75 99

    4 (PhBO)3 (1.0) C 84 99

    5 (PhBO)3 (2.0) C 92 99

    6 (PhBO)3 (3.0) C >99 99

    a All reactions were performed on 0.2 mmol scale with 3 mol% Rh(acac)(C2H4)2 and 7.5 mol% (R)-L at 100 ºC for

    2 h. Condition A: 0.55 mL of 1,4-dioxane/H2O (10/1). Condition B: 0.5 mL of 1,4-dioxane, 1 equiv of H2O to

    boron. Condition C: 0.5 mL of 1,4-dioxane, slow addition of water by syringe pump. b Determined by 1H NMR. c

    Determined by chiral HPLC.

    Ph BOH

    OHB

    O BO

    BOPh

    Ph

    Ph

    -3 H2O

    +3 H2O

    Figure 2.5 Equilibrium between phenylboronic acid and phenylboroxine

  • 32

    06_Chapter_2.doc

    Chapter 2

    2.2.2 Scope of the Reaction

    With these optimized conditions in hand, the scope of the asymmetric conjugate addition of

    arylboroxines to 14a was investigated. High ee values could be obtained with a variety of

    sterically and electronically diverse arylboroxines (Table 2.2, entries 1-9). Meta- and para-

    tolyl groups could be introduced with high yield and high enantioselectivity (entries 3

    and 4).

    Table 2.2 Scope of arylboroxines in the rhodium-catalyzed asymmetric 1,4-addition to 14a

    NCO2Bn

    O

    NCO2Bn

    ORh(acac)(C2H4)2 (3 mol%)

    (R)-L2 (7.5 mol%)

    (ArBO)3 (3 equiv)

    1,4-dioxane, 100 oC

    slow addition of H2O14a 15R

    entrya Ar product yield (%)b ee (%)c,d

    1 Ph 15a 86 e 99 (R)

    2 o-(Me)C6H4 15b 82 e 24 (R)

    3 m-(Me)C6H4 15c 92 e 98 (R)

    4 p-(Me)C6H4 15d 86 e 95 (R)

    5 p-(MeO)C6H4 15e 85 e 96 (R)

    6 m,p-(MeO)2C6H3 15f 86 e 98 (R)

    7 p-(F)C6H4 15g 71 94 (R)

    8 p-(F),o-(Me)C6H3 15h 75 47 (R)

    9 p-(Cl)C6H4 15i 55 96 (R)

    a All reactions were performed in duplicate with both enantiomers of the ligand on 0.2 mmol scale with 3 mol% Rh(acac)(C2H4)2 and 7.5 mol% L2 at 100 ºC for 2 h. b Isolated yield. c Determined by chiral HPLC. d The absolute

    configuration was established by comparison of the optical rotation with literature values or by analogy (see

    experimental section). e Thin layer chromatography shows a spot-to-spot conversion in 2 h.

  • 33

    06_Chapter_2.doc

    Enantioselective Synthesis of 2-Aryl-4-piperidones

    A dramatic drop in enantioselectivity was observed for the introduction of more sterically

    demanding ortho-substituted aryl groups (entry 2 and 8), illustrating a possible limitation of

    the catalytic method. Products with one or two electron-donating substituents on the aryl

    were obtained in high yield with high enantioselectivity (entries 5 and 6). However,

    electron-withdrawing chloro and fluoro groups resulted in slower reaction, leading to

    incomplete conversions (entries 7, 8, 9). Despite this observation, the enantioselectivity is

    largely independent of the electronic nature of the substituents. All para- and meta-

    substituted products were obtained with excellent ee values between 94 and 98%.

    NCO2Bn

    O

    NH

    O

    Rh(acac)(C2H4)2 (3 mol%)

    (S)-L2 (7.5 mol%)1,4-dioxane, 100 oC

    slow addition of H2O

    14a (S)-16

    1)

    (PhBO)3

    (3equiv)

    +2) 10 mol% Pd/C

    H2 (Ca. 1 atm)

    MeOH, 20 oC, 4h79% yield, 99% ee

    (2 steps)

    0.5 g

    Scheme 2.3 Conjugate addition on a 0.5 gram scale and subsequent deprotection

    To show the applicability of this reaction for synthesis on a laboratory scale, it was

    performed on a 0.5 gram (2.2 mmol) scale (Scheme 2.3). After flash chromatography, the

    product was isolated in 86% yield with 99% ee. Subsequent removal of the

    benzyloxycarbonyl group by hydrogenation using palladium on carbon, according to a

    literature procedure,24 afforded piperidone (S)-16 in 92% isolated yield and 79% overall

    yield over the two steps.

    2.3 Further Developments

    Cationic palladium(II) complexes show relative high rates for transmetallation of

    organoboron compounds, a process which is generally slow for transition metals.27 This has

    stimulated research toward their use in conjugate addition reactions where the

  • 34

    06_Chapter_2.doc

    Chapter 2

    transmetallation is a critical step. Miyaura et al. reported both palladium-based catalyst-

    systems that are able to transfer arylboronic acids28 and the asymmetric palladium-

    catalyzed conjugate addition of aryltrifluoroborates.29 The asymmetric conjugate addition

    of arylboronic acids had, however, been elusive. Shortly after publication of our results

    regarding this arylation,30 Gini et al. developed a Pd(O2CCF3)2/Me-DuPHOS (Figure 2.6)

    catalyst for the efficient and highly enantioselective addition of arylboronic acids to a

    variety of α,β-unsaturated enones in a THF/H2O (10/1) mixture.31

    NCO2Bn

    O

    NCO2Bn

    O

    14a (R)-15a

    60% yield, 99% ee

    Pd(OCCF3)2 (5 mol%)

    PP

    (R,R)-Me-DuPHOS (5.5 mol%)

    THF/H2O: 10/1, 70 oC, 22 h

    PhB(OH)2 (3 equiv)

    (R,R)-Me-DuPHOS

    Figure 2.6 Palladium-catalyzed asymmetric conjugate addition of phenylboronic acid

    Addition of phenylboronic acid to 14a with this system at 70 oC proceeds with essentially

    complete enantioselectivity (>99% ee). Although full conversion of the starting material

    was reached within 22 h, 15a was obtained in a relatively low yield of 60%.

    2.4 Conclusions

    In summary, we have shown that conjugate addition of arylboroxines with a rhodium/

    phosphoramidite catalyst can be used to prepare 2-aryl-4-piperidones in high isolated yield

    (82–92%) and with excellent enantioselectivity (up to 99% ee).

  • 35

    06_Chapter_2.doc

    Enantioselective Synthesis of 2-Aryl-4-piperidones

    2.5 Experimental Section

    General remarks. All air- and moisture-sensitive manipulations were carried out under a

    dry nitrogen atmosphere using standard Schlenk techniques. 1,4-Dioxane was distilled from

    sodium before use. 1H-NMR and 13C-NMR spectra were recorded on a Varian 300

    (300 and 75 MHz, respectively) in CDCl3 unless stated otherwise. Mass spectra (HRMS)

    were recorded on an AEI MS-902. Optical rotations were measured on a Schmidt and

    Haensch Polartronic MH8. Rh(acac)(C2H4)2 was purchased from Strem and used without

    further purification. All other chemicals were purchased from Acros and used as received.

    Flash chromatography was performed using silica gel 60 Å (Merck, 230-400 mesh).

    Phosphoramidite ligands (S)-L2 and (R)-L2 were prepared from the corresponding

    H8-bis-β-naphthol, PCl3, and diethylamine according to a previously reported procedure.23

    All arylboroxines were prepared according to a modified literature procedure from the

    corresponding arylboronic acids by heating in vacuo overnight at 145 °C in a drying

    pistol.25

    Benzyl 3,4-dihydro-4-oxo-1-(2H)-pyridinecarboxylate (14a). This compound was

    synthesized from 4-piperidone monohydrate hydrochloride and benzyl

    chloroformate following the literature procedure for ethyl 3,4-dihydro-4-oxo-1-

    (2H)-pyridinecarboxylate.15a The product was obtained as a white solid in 64%

    yield. 1H NMR δ = 7.85 (bs, 1H), 7.40-7.37 (m, 5H), 5.35 (bs, 1H), 5.27 (s, 2H), 4.05 (t, J = 7.3 Hz, 2H), 2.56 (t, J = 7.3 Hz, 2H); 13C NMR δ = 193.2, 152.5, 143.3,

    134.8, 128.7, 128.4, 107.7, 69.0, 42.5, 35.6. Physical and spectral data were in full

    agreement with the literature.24

    General Procedure for the rhodium/phosphoramidite-catalyzed asymmetric conjugate

    addition to 2,3-dihydro-4-pyridones (15). In a flame dried Schlenk tube flushed with

    nitrogen, 1.55 mg (6.0 µmol, 3 mol%) of Rh(acac)(C2H4)2 and 5.93 mg (15.0 µmol, 7.5

    mol%) of phosphoramidite L2 were dissolved in 0.5 mL of 1,4-dioxane. After stirring for

    15 min at room temperature, 46.2 mg (0.2 mmol) of substrate 1 and 0.6 mmol of the

    arylboroxine were added and the resulting mixture was stirred at reflux conditions with

    N

    O

    CO2Bn

  • 36

    06_Chapter_2.doc

    Chapter 2

    N

    O

    CO2Bn

    slow addition of a 20 vol% solution of water in 1,4-dioxane by syringe pump (0.1 mL/h)

    during 2 h. The reaction mixture was subsequently cooled to RT, diluted with 2 mL of

    ether, and passed through a pad of silica gel. The solvent was removed in vacuo. Reactions

    were performed in duplicate, using both enantiomers of the ligand. Enantioselectivity did

    not differ more than 1% between duplos.

    (S)- and (R)-Benzyl-4-oxo-2-phenylpiperidine-1-carboxylate (15a). The crude product

    was purified by flash column chromatography (n-pentane/Et2O: 1/2) to

    give (R)-15a, in the case of the (R)-L2 ligand, as a white solid in 86%

    isolated yield with 99% ee (Table 2.2, entry 1). The ee was determined on

    a Chiralcel OD-H column with n-heptane/2-propanol: 90/10, flow = 0.5

    mL/min. Retention times: 27.3 min [(S)-enantiomer], 29.8 min [(R)-enantiomer]. 1H NMR

    δ = 7.36 (m, 7H), 7.29-7.24 (m, 3H), 5.84 (bs, 1H), 5.25 (d, J = 12.3 Hz, 1H), 5.2 (d, J =

    12.3 Hz, 1H), 2.86 (dd, J = 15.5, 7.0 Hz, 1H), 2.57-2.50 (m, 1H), 2.37 (d, J = 15.8 Hz, 1H); 13C NMR δ = 207.0, 155.3, 139.6, 136.2, 128.7, 128.5, 128.1, 127.9, 127.6, 126.6, 67.7,

    54.5, 44.1, 40.4, 38.8. Physical and spectral properties were in full agreement with the

    literature.16

    (S)- and (R)-Benzyl-4-oxo-2-o-tolylpiperidine-1-carboxylate (15b). The crude product

    was purified by flash chromatography (n-pentane/Et2O: 1/2) to give (R)-

    15b, in the case of the (R)-L2 ligand, as a clear oil in 82% isolated yield

    with 24% ee (Table 2.2, entry 2). The ee was determined on a Chiralcel

    OJ column with n-heptane/2-propanol: 90/10, flow = 1.0 mL/min.

    Retention times: 14.7 min [(S)-enantiomer], 23.5 min [(R)-enantiomer]. 1H NMR δ = 7.35-

    7.29 (m, 3H), 7.23-7.18 (m, 6H), 5.75 (bs, 1H), 5.18 (d, J = 12.0 Hz, 1H), 5.14 (d, J = 12.2

    Hz, 1H), 4.28 (bs, 1H), 3.25 (bs, 1H), 2.86 (dd, J = 15.6, 5.5 Hz, 1H), 2.80 (dd, J = 15.4,

    5.5 Hz, 1H), 2.58-2.46 (m, 1H), 2.46 (d, J = 17.1 Hz, 1H), 2.26 (s, 3H); 13C NMR δ =

    207.8, 155.1, 138.6, 136.4, 136.1, 128.5, 128.1, 127.9, 127.8, 126.2, 126.1, 67.7, 52.8, 44.7,

    40.8, 38.9, 19.2. Physical and spectral properties were in full agreement with the

    literature.16

    N

    O

    CO2Bn

  • 37

    06_Chapter_2.doc

    Enantioselective Synthesis of 2-Aryl-4-piperidones

    N

    O

    CO2Bn

    N

    O

    CO2Bn

    N

    O

    CO2BnO

    (S)- and (R)-Benzyl-4-oxo-2-m-tolylpiperidine-1-carboxylate. (15c). The crude product

    was purified by flash chromatography (n-pentane/Et2O: 1/2) to give

    (R)-15c, in the case of the (R)-L2 ligand, as a clear oil in 92% isolated

    yield with 98% ee (Table 2.2, entry 3). The ee was determined on a

    Chiralcel OD-H column with n-heptane/2-propanol: 90/10, flow = 0.5

    mL/min. Retention times: 25.1 min [(S)-enantiomer], 29.0 min [(R)-enantiomer]. The

    absolute configuration was assigned by analogy. [α]D = +90.9° (c = 0.52, CHCl3, 98% ee); 1H NMR δ = 7.02-7.39 (m, 5H), 7.14-7.21 (m, 1H), 6.95-7.03 (m, 3H), 5.75 (bs, 1H), 5.21

    (d, J = 12.5 Hz, 1H), 5.14 (d, J = 12.5 Hz, 1H), 4.23 (m, 1H), 3.15 (m, 1H), 2.94 (d, J =

    15.4 Hz, 1H), 2.79 (dd, J = 15.4, 7.0 Hz, 1H), 2.49 (m, 1H), 2.32 (m, 1H), 2.26 (s, 3H); 13C

    NMR δ = 207.3, 155.4, 139.6, 138.5, 136.2, 128.6, 128.4, 128.2, 127.9, 127.3, 123.6, 67.7,

    54.5, 44.1, 40.5, 38.9, 29.6, 21.4; HRMS calcd for C20H21NO3 323.1521 found 323.1517.

    (S)- and (R)-Benzyl-4-oxo-2-p-tolylpiperidine-1-carboxylate (15d). The crude product

    was purified by flash chromatography (n-pentane/Et2O: 1/2) to give

    (R)-15d, in the case of the (R)-L2 ligand, as a clear oil in 86% isolated

    yield with 95% ee (Table 2.2, entry 4). The ee was determined on a

    Chiralcel OJ column with n-heptane/2-propanol: 90/10, flow = 1.0

    mL/min. Retention times: 27.3 min [(S)-enantiomer], 30.8 min [(R)-enantiomer]. The

    absolute configuration was assigned by analogy. [α]D = +81.3° (c = 0.64, CHCl3, 95% ee);

    1H NMR δ = 7.28-7.42 (m, 5H), 7.07-7.20 (m, 4H), 5.82 (bs, 1H), 5.26 (d, J = 12.2 Hz,

    1H), 5.19 (d, J = 12.2 Hz, 1H), 4.25 (m, 1H), 3.18 (m, 1H), 2.98 (td, J = 15.6, 3.2, 1.2

    Hz,1H), 2.83 (dd, J = 15.4, 6.6 Hz , 1H), 2.51 (m, 2H), 2.33 (s, 3H); 13C NMR δ = 207.3,

    155.4, 137.4, 136.5, 136.2, 129.4, 128.5, 128.2, 127.9, 126.6, 67.7, 54.3, 44.1, 40.5, 38.8,

    20.9; HRMS calcd for C20H21NO3 323.1521 found 323.1515.

    (S)- and (R)-Benzyl-2-(4-methoxyphenyl)-4-oxopiperidine-1-carboxylate (15e). The

    crude product was purified by flash chromatography (n-pentane/Et2O:

    1/2) to give (R)-15e, using the (R)-L2 ligand, as a clear oil in 85%

    isolated yield with 96% ee (Table 2.2, entry 5). The ee was

    determined on a Chiralcel OD-H column with n-heptane/2-propanol:

  • 38

    06_Chapter_2.doc

    Chapter 2

    N

    O

    CO2BnO

    O

    N

    O

    CO2BnF

    90/10, flow = 0.5 mL/min. Retention times: 35.8 min [(S)-enantiomer], 39.9 min [(R)-

    enanti