methodology for the olefination of aldehydes and ketones via the

57
e Florida State University DigiNole Commons Electronic eses, Treatises and Dissertations e Graduate School 4-2-2009 Methodology for the Olefination Of Aldehydes and Ketones Via the Meyer-Schuster Reaction Susana Sorina López Florida State University Follow this and additional works at: hp://diginole.lib.fsu.edu/etd is esis - Open Access is brought to you for free and open access by the e Graduate School at DigiNole Commons. It has been accepted for inclusion in Electronic eses, Treatises and Dissertations by an authorized administrator of DigiNole Commons. For more information, please contact [email protected]. Recommended Citation López, Susana Sorina, "Methodology for the Olefination Of Aldehydes and Ketones Via the Meyer-Schuster Reaction" (2009). Electronic eses, Treatises and Dissertations. Paper 1060.

Upload: conker4

Post on 13-Sep-2015

6 views

Category:

Documents


1 download

DESCRIPTION

i

TRANSCRIPT

  • The Florida State UniversityDigiNole Commons

    Electronic Theses, Treatises and Dissertations The Graduate School

    4-2-2009

    Methodology for the Olefination Of Aldehydesand Ketones Via the Meyer-Schuster ReactionSusana Sorina LpezFlorida State University

    Follow this and additional works at: http://diginole.lib.fsu.edu/etd

    This Thesis - Open Access is brought to you for free and open access by the The Graduate School at DigiNole Commons. It has been accepted forinclusion in Electronic Theses, Treatises and Dissertations by an authorized administrator of DigiNole Commons. For more information, please [email protected].

    Recommended CitationLpez, Susana Sorina, "Methodology for the Olefination Of Aldehydes and Ketones Via the Meyer-Schuster Reaction" (2009).Electronic Theses, Treatises and Dissertations. Paper 1060.

  • FLORIDA STATE UNIVERSITY

    COLLEGE OF ARTS AND SCIENCES

    METHODOLOGY FOR THE OLEFINATION OF ALDEHYDES AND

    KETONES VIA THE MEYER-SCHUSTER REACTION

    By

    SUSANA SORINA LPEZ

    A Thesis submitted to the Department of Chemistry and Biochemistry

    in partial fulfillment of the requirements for the degree of

    Master of Science

    Degree Awarded: Spring Semester, 2009

    Copyright 2009 Florida State University

    All Rights Reserved

  • ii

    The members of the Committee approve the Thesis of Susana Sorina Lpez defended on

    April 2nd, 2009.

    __________________________________ Gregory B. Dudley Professor Directing Thesis

    ___________________________________ Igor Alabugin Committee Member

    __________________________________ Lei Zhu Committee Member

    __________________________________

    Michael Shatruk Committee Member

    Approved: _____________________________________ Joseph Schlenoff, Chair, Arts and Sciences The Graduate School has verified and approved the above named committee members.

  • iii

    Quiero dedicar esta tesis a mis padres, Oscar y Susana Mercedes Lpez por todos los

    sacrificios que han hecho a lo largo de los aos para ayudarme a convertirme

    en la mujer que soy hoy. Sin su amor y apoyo esto no habra sido posible.

    I would also like to dedicate this to Dr. Paul I. Higgs, who has provided me with

    the encouragement and guidance that has allowed me to never give up on myself.

    Lastly, I dedicate this to Brian Ray Jacobs, who has shown me that love can provide

    strength in times of weakness.

    With all my all love and appreciation,

    (Con todo mi amor y aprecio),

    Susana

  • ii

    ACKNOWLEDGEMENTS

    I would like to express my gratitude to my major professor, Dr. Gregory B. Dudley, for

    his support and guidance during these first three years of my graduate studies. I would

    also like to thank the past and present members of the Dudley group for their friendship

    and support: Dr. Mariya V. Kozytska, David M. Jones, Jingyue Yang, Sami Tlais,

    Daniella Barker, Jumreang Tummatorn, post-docs: Dr. Philip Albiniak and Dr. Jeannie

    Jeong for their guidance during their time in our lab and Douglas A. Engel for the work

    and direction during our collaboration on the Meyer-Schuster chemistry. The members of

    my committee: Dr, Igor Alabugin, Dr. Lei Zhu and Dr. Michael Shatruk for their

    assistance and patience during the preparation of this thesis. Lastly, I would like to

    acknowledge Dr. George Fisher and Mara Tsesarskaja for exposing me to chemistry

    research for the first time as an undergraduate at Barry University.

  • iii

    TABLE OF CONTENTS List of Tables ............................................................................................. vi List of Figures ............................................................................................ vii List of Symbols .......................................................................................... ix Abstract .................................................................................... xiv 1. Introduction............................................................................................ 1 1.1.1 Olefination Strategies for the Synthesis of ,-Unsaturated Esters.. 1 1.1.2 Wittig Reaction ................................................................................. 3 1.1.3 Horner-Wadsworth-Emmons............................................................ 5 1.1.4 Meyer-Schuster Rearrangement of Propargyl Alcohols ................... 6 1.1.5 Mechanism .................................................................................... 8 1.1.6 Earlier Work in the Dudley Lab........................................................ 9 1.1.7 Conclusion ..................................................................................... 11 2. Results and Discussion .......................................................................... 12 2.2.1 Lewis-acid Catalyzed Rearrangement of Ethoxyalkynyl Carbinols 12 2.2.2 Gold catalyzed Meyer-Schuster Reaction of Secondary Ethoxyalkynyl Carbinols .................................................. 13 2.2.3 Substrate Scope and Stereoselectivity ....................................... 17 2.2.4 Conclusion ................................................................................. 18 2.2.5 Alternative Catalysts for the Meyer-Schuster Reaction of Secondary and Tertiary Ethoxyalkynyl Carbinols ...................................... 20 2.2.6 Screening of Alternative Catalysts ............................................. 21 2.2.7 Effects of Additives ................................................................... 23 2.2.8 Optimization of Reaction Conditions and Stereoselectivity ...... 24 2.2.9 Two-stage Olefination of Aldehydes and Ketones .................... 25 2.3.1 Mechanistic Hypothesis of the Lewis-acid Catalyzed

  • iv

    2.3.2 Meyer-Schuster Reaction ........................................................... 28 2.3.3 Conclusion ................................................................................. 31 3. Experimental ... ........................................................................ 32 REFERENCES .......................................................................................... 38 BIOGRAPHICAL SKETCH ...................................................................... 41

  • v

    LIST OF TABLES

    Table1: Catalyst screenings .................................................................................... 14 Table 2: Solvent screenings .................................................................................... 15 Table 3: Additive screenings .................................................................................. 16 Table 4: Series of representative secondary alcohol substrates .............................. 17 Table 5: Catalytic screenings of alternative Lewis-acid catalysts .......................... 21 Table 6: Effect of additives on top three Lewis-acid catalysts ............................... 22 Table 7: Ethanol as an additive vs. ethanol as a co-solvent.................................... 24 Table 8: Scandium (III) triflate catalyzed homologation of hindered ketones ....... 26 Table 9: Statiscal incorporation of n-propanol ....................................................... 30

  • vi

    LIST OF FIGURES

    Figure 1: Aldol condensation.................................................................................. 1 Figure 2: Acid-catalyzed aldol condensation.......................................................... 2 Figure 3: Base-catalyzed aldol condensation.......................................................... 2 Figure 4: Peterson olefination................................................................................. 3 Figure 5: Wittig reaction......................................................................................... 3 Figure 6: Examples of three different ylide categories........................................... 4 Figure 7: Horner-Wadsworth-Emmons reaction .................................................... 5 Figure 8: HWE stereoselectivity............................................................................. 6 Figure 9: Two possible reaction pathways of propargyl alcohols .......................... 7 Figure 10: Acetylide addition/Meyer-Schuster reaction......................................... 8 Figure 11: Lewis-acid catalyzed mechanism for activating propargyl alcohols..... 9 Figure 12: Gold-catalyzed Meyer-Schuster reactions of tertiary ethoxyalkynyl

    carbinols ................................................................................................ 10 Figure 13: Activation of Lewis basic sites of electronically neutral propargyl alcohols.................................................................................................. 12 Figure 14: Single step formation of ,-unsaturated ketones ................................. 12 Figure 15: Gold(I) and silver (I) hexafluoroantimonate Meyer-Schuster rearrangement ............................................................. 13 Figure 16: Conditions for catalytic screenings ....................................................... 14 Figure 17: Conditions for solvent screenings ......................................................... 15 Figure 18: Conditions used for additive screenings................................................ 16 Figure 19: Optimized conditions for the gold (I) silver hexafluoroantimonate Meyer-Schuster rearrangement ............................................................ 17

  • vii

    Figure 20: Lewis-basic sites of propargyl alcohols ................................................ 23 Figure 21: Optimized conditions for Cu(I) and Sc(I) Meyer-Schuster rearrangement ....................................................................................... 25 Figure 22: Compatibility of the Meyer-Schuster reaction ...................................... 27 Figure 23.Hypothesized gold catalyzed Meyer-Schuster rearrangement ............... 28 Figure 24: Ratio of ethyl to n-propyl esters ............................................................ 28 Figure 25: .Reaction of ,-unsaturated ester in Meyer-Schuster conditions......... 29 Figure 26: Hypothesized Meyer-Schuster rearrangement with n-propanol............ 30 Figure 27: Conditions for the Meyer-Schuster rearrangement with n-propanol..... 30

  • viii

    LIST OF SYMBOLS Ac acetyl

    acac acetylacetonate

    AIBN 2,2-azobisisobutyronitrile

    anhyd anhydrous

    Ar aryl

    atm atmosphere(s)

    9-BBN 9-borabicyclo[3.3.1]nonyl

    Bn benzyl

    BOC tert-butoxycarbonyl

    bp boiling point

    br broad (spectral)

    Bu butyl

    i-Bu iso-butyl

    s-Bu sec-butyl

    t-Bu tert-butyl

    C degrees Celsius

    calcd calculated

    Cbz benzyloxycarbonyl

    CI chemical ionization (in mass spectrometry)

    cm centimeter(s)

    concd concentrated

    COSY correlation spectroscopy

    COT cyclooctatetraene

    Cp cyclopentadienyl

    Cy-hexyl cyclohexyl

    chemical shift in parts per million downfield from tetramethylsilane

    d day(s); doublet (spectral)

    DABCO 1,4-diazabicyclo[2.2.2]octane

    DBN 1,5-diazabicyclo[4.3.0]non-5-ene

  • ix

    DBU 1,8-diazabicyclo[5.4.0]undec-7-ene

    DCB 2,6-dichlorobenzyl

    DCC N,N-dicyclohexylcarbodiimide

    DCM dichloromethane

    DDQ 2,3-dichloro-5,6-dicyano-1,4,benzoquinone

    DEAD diethyl azodicarboxylate

    DEPT distortionless enhancement by polarization transfer

    DIBALH diisobutylaluminum hydride

    DMAP 4-(dimethylamino)pyridine

    DME 1,2-dimethoxyethane

    DMF dimethylformamide

    DMPU dimethylpropylene urea

    DMSO dimethyl sulfoxide

    E1 unimolecular elimination

    E2 bimolecular elimination

    ee enantiomeric excess

    EI electron impact (in mass spectrometry)

    Et ethyl

    FAB fast action bombardment (in mass spectrometry)

    FT Fourier transform

    g gram(s)

    GC gas chromatography

    H hours(s)

    HMO Hckel molecular orbital

    HMPA hexamethylphosphoric triamide

    HOMO highest occupied molecular orbital

    HPLC high-performance liquid chromatography

    HRMS high-resolution mass spectrometry

    Hz hertz

    IP ionization potential

    IR infrared

  • x

    J coupling constant (in NMR)

    k kilo

    KOH potassium hydroxide

    L liter(s)

    LAH lithium aluminum hydride

    LDA lithium diisopropylamide

    LHMDS lithium hexamethyldisilazane

    LTMP lithium 2,2,6,6-tetramethylpiperidide

    LUMO lowest occupied molecular orbital

    micro

    m multiplet (spectral), meter(s), milli

    M moles per liter

    MBH Morita-Baylis-Hillman

    m-CPBA m-chloroperoxybenzoic acid

    m/e mass to charge ratio (in mass spectrometry)

    Me methyl

    MEM (2-methoxyethoxy)methyl

    Mes mesityl, 2,4,6-trimethylphenyl

    MHz megahertz

    min minute(s)

    mM millimoles per liter

    MO molecular orbital

    mol mole(s)

    MOM methoxymethyl

    mp melting point

    Ms Methanesulfonyl (mesyl)

    MS mass spectrometry

    MVK methyl vinyl ketone

    m/z mass to charge ratio (in mass spectrometry)

    NBS N-bromosuccinimide

    NCS N-chlorosuccinimide

  • xi

    NMO N-methylmorpholine-N-oxide

    NMR nuclear magnetic resonance

    NOE nuclear Overhauser effect

    Nu nucleophile

    OD optical density

    ORD optical rotary dispersion

    PCC pyridinium chlorochromate

    PDC pyridinium dichromate

    PEG polyethylene glycol

    Ph phenyl

    PMB p-methoxybenzyl

    PPA polyphosphoric acid

    ppm parts per million (in NMR)

    PPTS pyridinium p-toluenesulfonate

    Pr propyl

    i-Pr isopropyl

    q quartet (spectral)

    re rectus (stereochemistry)

    Rf retention factor (in chromatography)

    rt room temperature

    s singlet (spectral); second(s)

    si sinister (stereochemistry)

    SN1 unimolecular nucleophilic substitution

    SN2 bimolecular nucleophilic substitution

    SN nucleophilic substitution with allylic rearrangement

    t triplet (spectral)

    TBAB tetrabutylammonium bromide

    TBDMS tert-butyldimethylsilyl

    Tf trifluoromethanesulfonyl (triflyl)

    TFA trifluoroacetic acid

    TFAA trifluoroacetic anhydride

  • xii

    THF tetrahydrofuran

    THP tetrahydropyran

    TIPS triisopropylsilyl

    TLC thin layer chromatography

    TMEDA N,N,N,N-tetramethyl-1,2-ethylenediamine

    TMS trimethysilyl, tetramethylsilane

    Tr triphenylmethyl (trityl)

    Ts tosyl, p-toluenesulfonyl

    TS transition state

    tR retention time (in chromatography)

    UV ultraviolet

  • xiii

    ABSTRACT

    Our lab was faced with a synthetic challenge during studies towards the total synthesis of

    the anti-malaria drug, artemisinin. Known methods such as the Aldol condensation, the Horner-

    Wadsworth-Emmons and the Wittig reactions were ineffective for the olefination of hindered

    ketones. We were required to find an alternative approach of olefination that would not be

    restricted by steric constraints. In 2006, reported on a two-step strategy for the HWE-type

    olefination of hindered ketones: (1) addition of ethoxyacetylide, then (2) Au3+ catalyzed Meyer

    Schuster rearrangement.

    Alkyne addition to carbonyl groups is relatively insensitive to sterics, whereas the

    resulting congested tertiary ethoxyalkynyl carbinols are sterically and electronically primed for

    rearrangement. Having identified this important two-stage synthetic application, we focused our

    attention on step two; the MeyerSchuster rearrangement. The Meyer-Schuster reaction is a

    little-known but potentially powerful rearrangement that converts propargyl alcohols into , -

    unsaturated carbonyl compounds.

    In our earlier study, which featured highly reactive tertiary propargyl alcohol substrates,

    rearrangement occurred immediately upon addition of the gold catalyst. In 2007, we expanded

    our scope and reported a new reaction protocol and important observations with respect to the

    rearrangement of secondary alcohol substrates. We found that secondary ethoxyalkynyl carbinols

    could be converted into the corresponding ethyl trans-, -unsaturated esters with moderate to

    good stereocontrol using a mixed catalyst system of gold (I) chloride and silver (I)

    hexafluoroantimonate.

    Recent advances in our methodology for the olefination of aldehydes and ketones using

    the MeyerSchuster reaction of ethoxyacetylenes focused on four key points: (1) seeking

    alternative catalysts that are more economical and widely available than gold or silver salts, (2)

    lowering the catalyst loadings more than our previously reported methods using gold and silver

    salts, (3) obtain excellent stereoselectivity in the formation of the E-alkene isomer for most

    disubstituted alkenes, and (4) examine new mechanistic data suggesting that the higher

    stereoselectivity associated with the new catalysts may stem from a subtle alteration of the

    reaction mechanism.

  • 1

    CHAPTER I

    INTRODUCTION 1.1.1 Olefination Strategies for the Synthesis of , -Unsaturated Esters

    O

    O

    H R

    O

    O

    O

    R+

    +H2O

    Figure 1. Aldol reaction

    The homologation of aldehydes and ketones to ,-unsaturated esters (Fig. 1), an indispensable

    tool for generating carboncarbon bonds, is typically achieved using aldol condensation1, Wittig,

    HornerWadsworthEmmons (HWE), or other olefination methods2,3. Of these, the aldol condensation

    is most attractive from an atom economy4 standpoint in that water is the only by-product of the reaction.

    In the presence of dilute sodium hydroxide at room temperature, acetaldehyde undergoes an base-

    catalyzed dimerization reaction to produce 3-hydroxybutanal.

    Reactions of this nature are commonly referred to as aldol additions because that 3-

    hydroxybutanal is both an alcohol and an aldehyde. The initial protocol involved the use of a Brnsted

    acid or base as the catalyst; however, this caused problematic and undesirable side reactions. Although

    the reaction was efficient, there was room for improvement to the methodology. The classical acid

    catalyzed aldol reaction (Fig. 2) is a reversible reaction in which the electrophile is activated via

    protonation and under goes nucleophilic attack by an enol. In contrast, the base catalyzed reaction (Fig.

    3) involves the formation of an enolate via deprotonation, which then adds to the carbonyl forming the

    addition product.

  • 2

    Figure 2. Acid-catalyzed aldol dehydration

    Figure 3. Base-catalyzed aldol dehydration

    Dehydration of the aldol addition product gives rise to an ,-unsaturated carbonyl compound.

    Thus, one can achieve the two-step homologation of aldehydes to ,-unsaturated esters by aldol

    addition of the ester enolate, followed by elimination. This two-step condensation is an important

    transformation in organic synthesis, but it has key limitations in scope, stereoselectivity, and functional

    group tolerance. Consequently, alternative protocols for achieving the homologation of aldehydes and

    ketones to , -unsaturated esters have emerged over the years.

    The aldol addition of a -silyl ester to an aldehyde can be followed by facile elimination of the

    silanol in a variant of the Peterson olefination (Fig. 4). The Peterson olefination uses -trimethylsilyl-

    substituted organometallic compounds which convert carbonyl compounds to alkenes via a -

    silylcarbinol.

  • 3

    R1Si

    M

    R1R1

    R2R3 R4

    O+

    R3Si

    OHR1

    R4R2

    R3

    R3Si

    OHR1

    R3R2

    R4

    base

    base

    acid

    R2

    R1 R3

    R4

    R2

    R1 R4

    R3

    Figure 4. Peterson olefination

    The -silyl carbanions can be prepared by various methods but the subsequent addition to the

    carbonyl compound gives a diastereomeric mixture of -silylcarbinols which depending on R2

    substituent may or may not be easily separated.

    1.1.2 Wittig Reaction

    The Wittig reaction1 (Fig. 5) between aldehydes or ketones with phosphoranes is a valuable

    method for the synthesis of olefins. This method provides for the synthesis of alkenes from carbonyl

    compounds by replacing the oxygen of a carbonyl with an alkylidene group. The phosphorus ylides that

    serve as the active reagents are prepared combining triphenylphosphine first with a primary or secondary

    alkyl halide and subsequently with an appropriate base. Although a strong base is typically used (eg.

    alkyllithium), if the salt is sufficiently acidic, then a mild base, such as sodium bicarbonate may be used

    for the deprotonation step.

    O

    OPPh3 H R

    O

    O

    O

    R

    ++ Ph3P O

    Figure 5. Wittig reaction

    R1= alkyl, aryl; R2=alkyl, aryl, CO2R, CN, CONR2, CH=NR, SR, SOR, SO2R, SeR, SiR3, OR, BO2R2; R3, R4=alkyl, aryl, H

  • 4

    Phosphorus ylides are prepared before the reaction or in-situ and precautions must be taken due

    to their sensitivity to moisture and air. The carbanion of the ylide is the characteristic component that

    allows for nucleophilic attack on the carbonyl carbon. The ylides have been found to demonstrate faster

    reaction rates with aldehydes than they do with ketone substrates.

    The reactivity of the ylide is dependent on its substituents. Ylides are classified into three

    different categories (Fig 6). The first category is the stabilized ylides. These ylides possess at least one

    strong electron withdrawing group which stabilizes the negative charge on the carbanion. In regards to

    the stereoselectivity when reacted with aldehydes, these stabilized ylides will yield the (E)-alkene. It is

    noteworthy to mention that ester and ketone stabilized ylides react with aldehydes to give aldol

    condensation type products. Ester-stabilized ylides are employed for homologation of aldehydes to ,-

    unsaturated esters. The reaction of ester stabilized ylides with ketones is rare.

    XR2

    R3 (R1)3P

    X= Cl, Br, I, OTs R

    2

    R3(R1)3P X

    R2

    R3(R1)3P

    R2

    R3(R1)3P

    R4 R5

    O

    -(R1)3P O

    base

    alkyl halide phsophonium salt phosphorus ylide (phosphorane) R4, R5= alkyl, aryl, alkynyl, H

    R5

    R4 R2

    R3

    olefin

    Figure 6. Examples of three different ylide categories

    On the other extreme there are the non-stabilized ylides which contain only alkyl substituents

    which do not stabilized the negative charge on the carbon. When a base is used in the absence of lithium

    halides (salt-free conditions) and polar, aprotic solvents these ylides provide a high selectivity for the

    (Z)-alkene.

    non-stabilized ylide R1= aryl and R2,R3= alkyl, H semi-stabilized ylide R1= aryl and R2,R3= alkyl, alkenyl, benzyl, allyl, H stabilized ylide R1= aryl and R2,R3= -CO2R, -SO2R, -CN, -COR

  • 5

    The third category is semi-stabilized ylides. These ylides contain at least one aryl or alkenyl

    group which is less stabilizing when compared to the structure of the stabilized ylide. In contrast to the

    stabilized and non stabilized ylides, the semi stabilized ylides have poor stereo selectivity. Some other

    considerations that influence the stereochemical outcome of the reactions are the type of carbonyl

    compound which is used, the solvent and the counter ion that is used for formation of the ylide.

    1.1.3 Horner-Wadsworth-Emmons

    The Horner-Wadsworth-Emmons (HWE) reaction1 is a variant of the Wittig designed

    specifically to overcome limitations in the reactivity of stabilized phosphorus ylides. The reaction (Fig.

    7) takes place when an aldehyde or ketone reacts with a phosphonate as opposed to a phosphorane. The

    HWE is an improvement over the Wittig ylides since phosphonate anions are more reactive than the

    phosphorus ylides. These alkylphosphonates are easier to prepare and less costly than the phosphonium

    salts.

    O

    O

    H R

    O

    O

    O

    R

    ++P OEt

    OEt

    OPOEt

    O

    HO OEt

    Figure 7. Horner-Wadsworth-Emmons reaction

    A significant advantage of HWE reagents over phosphoranes is that HWE phosphonates can

    react with ketone substrates, whereas phosphoranes do not. Another feature that makes the HWE

    advantageous is that it can give desired stereoselectivity depending on the substituent that is placed in

    the R position. Bulkier groups, such as tert-butyl, will favor the (E)-olefin and smaller groups such as

    methyl will give rise to the (Z)-olefin as the product (Fig. 8).

  • 6

    Figure 8. HWE stereoselectivity

    The Wittig and HWE reaction both use stoichiometric phosphines, phosphine oxides, or

    phosphonates to provide , -unsaturated ester products. However, these reactions produce phosphorus

    by-products that can interfere with the isolation of the desired products. Whether using designer

    olefination reagents or a traditional aldol condensation protocol, these homologation reactions are

    sensitive to steric congestion around the carbonyl, such that olefination of hindered ketones can be

    problematic. In fact, homologation reactions of hindered ketones to , -unsaturated esters were a

    largely unsolved problem at the onset of this work.

    1.1.4 Meyer-Schuster Rearrangement of Propargyl Alcohols

    Propargyl alcohols are readily available, versatile tools in organic synthesis, providing access

    through different reaction pathways to desirous products such as alkenes, allenes, alkynes, ketones,

    etc.5,6 For example, hydrometalation (syn or anti), substitution (at the - or -centers), hydration,

    oxidation, hydrogenation, and deoxygenation all may be accomplished through selective activation of

    propargyl alcohol substrates. One such pathway is the Meyer-Schuster rearrangement. This reaction

    converts propargyl alcohols into , -unsaturated carbonyls.

    The Meyer-Schuster rearrangement involves the acid catalyzed isomerization of secondary and

    tertiary propargyl alcohols. A formal [1, 3] shift of the hydroxyl group and tautomerization gives , -

    unsaturated carbonyl, probably via a propargyl cation. The reaction may be catalyzed with Lewis or

    protic acids and is not sensitive to moisture in that it may be conducted in either aqueous or anhydrous

  • 7

    conditions. However, the Meyer-Schuster rearrangement is but one possible fate of the propargyl cation

    and selecting for the Meyer-Schuster pathway has been a long lasting challenge. The most significant

    competing pathway is the Rupe rearrangement 7 (Fig 9).

    Figure 9. Two possible reaction pathways of propargyl alcohols

    Under the original Meyer-Schuster conditions, most propargyl alcohols in fact show a preference for

    reacting along the Rupe pathway.

    The Rupe and Meyer-Schuster rearrangements (Fig. 9) are not often used in chemical synthesis

    due to harsh conditions and poor selectivity. The Meyer-Schuster products (path b) are especially rare

    because the dehydration that leads into the Rupe pathway (path a) generally takes precedence under

    traditional modes of activation that target the substrate through the alcohol moiety (i.e., acidic catalysts).

  • 8

    Methods for the synthesis of propargyl alcohols from aldehydes and ketones in combination with

    the Meyer-Schuster rearrangement provide two-step routes for the olefination of , -unsaturated esters.

    A major advantage of using the acetylide addition/MeyerSchuster reaction strategy for the olefination

    of aldehydes and ketones (Fig. 10) is the efficiency of the initial carboncarbon bond-forming reaction:

    alkyne addition.

    Figure 10. Acetylide addition/MeyerSchuster reaction

    However, the second stage of this strategythe MeyerSchuster reactionis generally limiting.

    The reaction protocol uses high temperatures and acidic conditions which limit the reaction scope.

    Therefore, advances in the MeyerSchuster reaction translate directly into advances in olefination

    methods.

    1.1.5 Mechanism

    Coordination of the alkyne using soft, late-transition-metal Lewis acids, 8 including cationic gold

    catalysts 9,10, provides a fundamentally different mechanism for activating propargyl alcohols (Fig. 11)

    Also, sensitive functionalities may be more tolerant of soft alkyne activation than hard' activation of

    the oxygen atom, providing complementary selectivity.

    acetylideaddition

    MeyerSchusterrearrangementR2

    R1OHR

    3R1 R2

    O

    HR3 R3 R1

    R2

    H

    O

  • 9

    Figure 11. Lewis-acid catalyzed mechanism for activation of propargyl alcohols

    1.1.6 Earlier Work in the Dudley Lab

    Alkyne addition to carbonyl groups is relatively insensitive to sterics, whereas the resulting

    congested tertiary ethoxyalkynyl carbinols are sterically and electronically primed for rearrangement. In

    2006, we reported gold-catalyzed MeyerSchuster reactions of tertiary ethoxyalkynyl carbinols for the

    synthesis of ,-unsaturated ethyl esters (Fig. 12).11,12 In conjunction with ethoxyacetylide addition to

    ketones, this work provided the blueprint for general implementation of the two-stage olefination

    strategy outlined below (1b3b, R3=OEt, Fig. 12) for the synthesis of ,-unsaturated esters.

  • 10

    Figure 12. Gold-catalyzed MeyerSchuster reactions of tertiary ethoxyalkynyl carbinols

    The combination of the electron-rich ethoxyacetylenic -system and soft gold (III) chloride

    catalyst13 provided excellent reactivity in the MeyerSchuster reaction: consumption of the intermediate

    tertiary ethoxyalkynyl carbinols occurred within minutes of adding the catalyst. The MeyerSchuster

    reactions were conducted open to the air without external heating or cooling. Yields for both the

    acetylide addition and the formal rearrangement14, 14a, 14b and 14c were essentially quantitative in the

    majority of cases, but stereocontrol of the olefin geometry was non-existent. The second drawback of

    the reported conditions is the requirement for 5 mol % of the (expensive) gold catalyst. At 5 mol %

    catalyst loading, the reactions were complete within minutes, but at 1 mol %, the reaction failed to reach

    full conversion even after prolonged reaction times.11

  • 11

    1.1.7 Conclusion

    The Aldol, the Wittig and the HWE reactions are well known reactions for the conversion of

    aldehydes and ketones in to , - unsaturated esters. The aldol condensation is most attractive from the

    atom economy perspective, but it is the least general in terms of scope and efficiency. Although the

    Wittig and the HWE although more efficient, they produce toxic and/or undesirable phosphorus by-

    products. Moreover, the steric sensitivity of these classical methods impeded the olefination of hindered

    ketones, which led us to seek an alternative synthetic route for the preparation of the sterically congested

    ,-unsaturated esters. The use of electron-rich ethoxyacetylenic propargyl alcohols in combination with

    a gold(III)chloride catalyzed Meyer-Schuster rearrangement, provided an efficient alternative route to

    obtain the desired , -unsaturated esters.

  • 12

    CHAPTER II

    RESULTS AND DISCUSSION 2.2.1 Lewis-acid Catalyzed Rearrangement of Ethoxyalkynyl Carbinols

    The last few years have seen a surge of interest in the MeyerSchuster reaction.17a, 17b, 17c, 17d, 17e,

    17f, 17g, 17h, 17i Whereas our Laboratory has focused on electronically activated propargyl alcohols for the

    synthesis of ,-unsaturated esters,15,16 Zhang and co-workers reported a method for obtaining ,-

    unsaturated ketones through independent activation of Lewis basic sites of electronically neutral

    propargyl alcohols (Fig. 13).17d They and others17f have shown that pre-activation of the hydroxyl group

    as an acetate ester followed by a gold-catalyzed hydrolysis process of the propargyl acetate delivers

    MeyerSchuster products. 18,18a, 18b and 19 The Yamada Lab used high-pressure carbon dioxide, base, and

    a silver catalyst to merge this multi-step process into a single operation (Fig. 14).17i

    .

    Figure 13. Activation of Lewis basic sites of electronically neutral propargyl alcohols

    Figure 14. Single step formation of ,-unsaturated ketones

  • 13

    2.2.2 Gold-catalyzed Meyer-Schuster Reaction of Secondary Ethoxyalkynyl Carbinols

    Secondary ethoxyalkynyl carbinols could be converted into the corresponding ethyl trans-,-

    unsaturated esters with moderate to good stereocontrol using a mixed catalyst system of gold(I) chloride

    and silver(I) hexafluoroantimonate (Fig. 15)15.

    Figure 15. Gold (I) and silver (I) hexafluoroantimonate Meyer-Schuster rearrangement

    Inclusion of camphorsulfonic acid as a co-catalyst resulted in better selectivity for the trans

    isomer. In particular, our efforts focused on the rearrangement of secondary propargyl alcohols with

    simple alkyl substituents. These aliphatic substrates are less reactive towards the Meyer-Schuster

    reaction than tertiary propargyl alcohols, which ionize more easily. However, the dampened reactivity of

    secondary alcohols (and the steric distinction between the alkyl substituent and a hydrogen atom)

    provides greater control and the opportunity to enhance stereoselectivity in the formation of , -

    unsaturated ester products.

    This study11 focuses on using electron-rich alkoxyacetylenes to control selectivity so as to access

    the Meyer-Schuster rearrangement,1 a formal [1, 3]-hydroxy migration followed by tautomerization. We

    examined three main variables: gold catalyst, additive, and solvent.

  • 14

    As shown in Table 1, minor differences were observed among the various gold catalysts. Both

    gold (I) and gold (III) were effective. Silver (I) hexafluoroantimonate (AgSbF6) showed little activity on

    its own, but when employed in conjunction with the gold catalysts it exerted a positive effect on the E/Z-

    selectivity of the reaction.

    Figure 16. Condtions used for catayltic screenng

    Table 1. Catalyst screenings

    MeOH

    MeMe

    Me

    MeMe

    OEtOR

    O(10 mol %)

    5.0 equiv EtOHCH2Cl2

  • 15

    Solvent screenings were conducted (Table 2). Both dichloromethane and water were both

    suitable solvents, whereas THF was not. Interestingly, however, reactions conducted in a mixed system

    of THF and CH2Cl2 were most efficient (qualitatively) and selective for the E-alkene isomer

    (quantitatively).

    Figure 17. Conditions for solvent screenings

    Table 2. Solvent screenings

    Additives were employed to accelerate the rearrangement and increase the stereoselectivity

    (Table 2). Among the protic additives, which are envisioned to assist in the formal [1, 3]-hydroxy

    migration, ethanol was significantly more effective than other agents tested. Inclusion of

    camphorsulfonic acid (CSA) in the reaction mixture improved the stereoselectivity of most reactions;

    however, in this protocol the substrates must tolerate more acidic conditions.

    OH

    OEt

    CO2Et10 mol % AuClAgSbF6

    10 equiv EtOH

  • 16

    Figure 18. Conditions used for additive screenings

    Table 3. Additive screenings

    Addition of camphorsulfonic acid (CSA) accelerated the reaction, whereas an acid scavenger [2,

    6-di-(tert-butyl)-4-methylpyridine, DTBMP] inhibited the reaction. These results, along with earlier

    experiments, 18 indicate that exchangeable protons play an important supporting role in the gold-

    catalyzed rearrangement.

    MeOH

    MeMe

    Me

    MeMe

    OEtOR

    OAuCl or AuCl3(10 mol %)

    5.0 equiv CH2Cl2

  • 17

    2.2.3 Scope and limitations- Substrate Stereoselectivity

    Reactions were typically conducted under an inert atmosphere of argon using anhydrous THF

    and CH2Cl2, but similar results were obtained in open-flask' reactions using reagent-grade solvents. The

    small amount of water present in reagent-grade ethanol does not interfere with (and may facilitate) the

    reaction. Further experimentation indicated that a catalyst loading of 5 mol% was optimal.

    Figure 19. Optimized conditions for the gold (I) silver hexafluoroantimonate Meyer-Schuster rearrangement

    Table 4. Series of representative secondary alcohol substrates

    R

    OH 5 mol % AuClAgSbF610 equiv EtOH

    THFCH2Cl2 (1:1)rt, 3060 min

    ROEt

    OEt

    O

    1 2

  • 18

    We tested the rearrangement protocol on a series of representative secondary alcohol substrates

    (1a-f, Table 4]).19 Neopentyl alcohol (1a) gave rise to nonenolizable enoate 2a with nearly complete

    stereoselectivity (entry 1a). Alkyl-substituted alcohols 1b-d afforded enoates 2b-d (entries 2a-4a) to the

    complete exclusion of dehydration products (cf. path a of Fig. 9). Sequential addition of the silver and

    gold precatalysts in solution to the reaction mixture provided optimal stereoselectivity and

    reproducibility. In fact, simultaneous addition of the solutions of the gold and silver salts to the reaction

    mixture provided the enoate products with slightly better selectivity, but we consider the sequential

    addition protocol to be more easily duplicated and thus preferable. Premixing the gold and silver salts

    gave poorer results with respect to selectivity, as did addition of the precatalysts as solids.

    2.2.4 Conclusion

    In summary, , -unsaturated esters were prepared from ethoxyalkynyl carbinols using cationic

    gold catalysts. Substitution on the alcohol substrate, including aryl, alkyl, and vinyl groups, is well

    tolerated, with aliphatic substituents providing the highest stereoselectivity. Neither Rupe-type

    elimination products (from loss of water) nor -hydroxy ester products (from addition of water) were

    observed. The use of the secondary ethoxyalkynyl carbinols proved useful due their dampened

    reactivity, allowing investigation of the mechanistic hypothesis of the rearrangement reactions.

    The mild, efficient, and convenient reaction conditions should find use in chemical synthesis.

    This work illustrates the potential role of activated, electron-rich alkyne substrates in the rapidly

    emerging field of catalysis using soft, late-transition-metal cations. 20

  • 19

    2.2.5 Alternative Catalysts for the Meyer-Schuster Reaction of Secondary and Tertiary

    Ethoxyalkynyl Carbinols

    Terminal alkynes offer an alternative addition/rearrangement pathway for the homologation of

    aldehydes and ketones that can be executed in the two-stage process outlined in Figure 12: (1) alkyne

    addition to the carbonyl and (2) MeyerSchuster rearrangement.21 The strength of this latter approach

    stems from the use of acetylide nucleophiles to generate the initial carboncarbon bond; acetylide

    nucleophiles are suitable for addition to even the most hindered of carbonyl systems. Therefore, step (1)

    of the two-step process is quite general. In contrast, the MeyerSchuster rearrangement, on the other

    hand, has received little attention 22, 22a, 22b, 22c, 22d, 22d, 22e, 22f, 22g over the years due to the limited scope,

    harsh conditions, and the competing Rupe rearrangement pathway.23

    Efficient methods for promoting MeyerSchuster rearrangements thereby expand the olefination

    of aldehydes and ketones, including hindered ketones that may not be suitable substrates for any of the

    other olefination strategies listed above. The recent emergence of soft Lewis acids24a, 24b and 25often

    late transition metal salts with an affinity for -bonds over non-bonded electron pairsbrings attention

    to alternative Lewis basic sites (Fig. 20) and suggests the possibility of exploiting a previously

    unexplored mechanism for promoting the MeyerSchuster rearrangement: activation of the propargylic

    alcohol via the alkyne -bond rather than the hydroxyl group.26

    Figure 20. Lewis basic sites of propargyl alcohols

  • 20

    Data and observations reported herein include (1) alternative catalysts that are more economical

    and widely available than gold or silver salts, (2) lower catalyst loadings than our previously reported

    methods using gold and silver salts, (3) excellent stereoselectivity in the formation of the E-alkene

    isomer for most disubstituted alkenes, and (4) new mechanistic data suggesting that the higher

    stereoselectivity associated with the new catalysts may stem from a subtle alteration of the reaction

    mechanism.

    2.2.6 Screening of alternative catalysts

    Under the hypothesis that late transition metal-catalysis of the MeyerSchuster reaction of

    ethoxyalkynyl carbinols is derived from Lewis acid/base interactions, we became interested in

    identifying similar (or better) catalytic activity in other Lewis acids. Table 5 provides a summary of our

    catalyst screenings, which focused primarily (though not exclusively) on soft transition metal

    salts.27a,27b,27c

    From this general catalyst screening emerged three top choices: copper (II) triflate, indium (III)

    chloride, and scandium (III) triflate. Of these, indium (III) chloride is the least reactive; the copper and

    scandium catalysts are comparable in reactivity. All three are air-stable powders and are convenient to

    handle and use.

  • 21

    Table 5. Catalytic screenings of alternative Lewis-acids

  • 22

    2.2.7 Effects of additives

    Further information on these Lewis acid-catalyzed MeyerSchuster reactions was gleaned by

    observing the effect of additives on the reaction rate (qualitatively) and stereoselectivity (quantitatively).

    Table 6 recounts the outcome of a small grid of reactions in which the three top Lewis acid catalyst

    choices were each coupled with two acidic and two basic additives: 1 mol % CSA, 1.0 equiv acetic acid

    (AcOH), 1 mol % 2, 6-di-tert-butyl-4-methylpyridine (DTBMP), and 1.0 equiv magnesium oxide

    (MgO).

    Table 6. Effect of additives on top three Lewis-acid catalysts

  • 23

    Studying the effect of additives aids in the identification of optimal conditions, and it provides

    insight into the reaction mechanism. Lewis and protic acids catalyze the MeyerSchuster reaction, so

    one would expect acidic additives to accelerate the reaction and basic additives to quench or retard the

    reaction. This hypothesis is supported by the data presented in Table 6. However, the fact that basic

    additives retard but do not quench the reaction suggests that protic acid, though helpful, is not required

    for catalytic activity. Therefore, one can choose between a short reaction time (e.g., entries 9 or 14) and

    reaction conditions that are presumably free of protic acid (e.g., entry 5).

    2.2.8 Optimization of reaction conditions and stereoselectivity

    All of these experiments were conducted on an exploratory scale to gauge reactivity and

    selectivity. Because the scandium (III) and copper (II) catalysts in the absence of additives were

    significantly more reactive and slightly more selective than indium (III) chloride, the triflate salts were

    employed throughout the next stage of the methodology.

    Entries 14 in Table 7 document the comparison between including ethanol as an additive

    (5 equiv, as in our earlier studies)11, 15 and employing ethanol as a co-solvent, which provided superior

    results under the current conditions (entries 3 and 4). Aliphatic substituents on the propargyl alcohols

    were universally tolerated, whether the substituent was linear (2d), branched (2f), or even quaternary

    (2e). Some erosion of stereoselectivity was observed in the benzylic case (2g3g, entries 11 and 12).

    Entries 710 reveal that stereoselectivity was better for disubstituted alkenes than trisubstituted alkenes.

    Figure 21. Optimized conditions for Cu(II) and Sc(III) Meyer-Schuster rearrangement

  • 24

    Table 7. Ethanol as an additive vs. ethanol as co-solven

  • 25

    2.2.9 Two-stage Olefination of Aldehydes and Ketones

    When performed immediately following addition of ethoxyacetylene to a carbonyl compound,

    the MeyerSchuster reactions described above complete a two-stage olefination of aldehydes and

    ketones. Illustrative examples are presented in this section.

    Scandium (III) and copper (II)-catalyzed MeyerSchuster reactions of secondary and tertiary

    propargyl alcohols are shown in Table 7 (23). In all cases, both catalysts provided similar results, with

    scandium (III) triflate consistently (albeit perhaps insignificantly) out-performing copper (II) triflate.

    From an industrial perspective, the scarcity of scandium salts is off-set by the fact that scandium (III)

    triflate is water-soluble, recoverable after aqueous workup, and reusable without noticeable loss of

    activity.

    The experiment outlined in Figure 21 provides insight into the compatibility of the Meyer

    Schuster reaction conditions with common functionality. N-Boc-serine methyl ester (4) was converted

    into tert-butyldimethylsilyl (TBS) ether 5, which was then included in the reaction mixture during the

    conversion of 2f to 3f (75% yield; cf. Table 7, entry 6).

    Figure 21. Compatibility of the MeyerSchuster reaction conditions with common functionality

    Recovery of 5 from this control experiment in 99% yield indicates that the present Meyer

    Schuster reaction conditions will prove to be compatible with typical alkyl esters, amine carbamates, and

    silyl ethers.

  • 26

    Given the dearth of methods suitable for the homologation of hindered ketones into , -

    unsaturated esters,28 the two-stage acetylide addition/MeyerSchuster strategy as applied to hindered

    ketones is particularly valuable. We earlier investigated the utility of gold (III) chloride (5 mol %) as a

    catalyst for such processes. 11

    Table 8 illustrates that only 1 mol % of the less-expensive scandium (III) triflate provides

    similarly outstanding results: near-quantitative overall yield for the olefination of menthone (entry 1,

    1h3h, 98%), 28 verbenone (entry 2, 1c3c, 97%), benzophenone (entry 3, 1i3i, 99%), and

    adamantanone (entry 4, 1a3a, 96%). Verbenone gave rise to 3c as a 58:42 mixture of olefin isomers,

    whereas the isomeric mixture of esters 3h could not be reliably estimated by 1H NMR.

    Table 8. Homologation of hindered ketones

  • 27

    2.3.0 Mechanistic Hypothesis of the Lewis-acid Catalyzed Meyer-Schuster Reaction

    Earlier experiments in our Lab using gold and silver salts to catalyze the MeyerSchuster

    reaction of ethoxyalkynyl carbinols support a mechanism in which the alcoholic additive included in the

    reaction mixture (i.e., ethanol) becomes incorporated into 50% of the product via an intermediate 1,1-

    diethoxy-allene (7, Fig. 22).15 This gold-catalyzed reaction pathway is distinct from that of analogous

    reactions catalyzed by protic or hard Lewis acids27, 27a, 27b, and 27c, which are known27b and 27c to produce -

    hydroxy ester by-products (i.e., 6) from initial hydration of the alkyne. -Hydroxy esters (6) have not

    been observed in any of the MeyerSchuster reactions catalyzed by soft Lewis acids in our study.

    Figure 22. Hypothesized gold catalyzed reaction pathway

    R

    OH

    OEt

    5 mol% [Au+]5.0 equiv EtOH

    THFCH2Cl2 (1:1)R

    CO2Et

    R

    OEtOEt

    R OEt

    OOH

    H3O+EtOH H2O

    R OEt

    OHOEt

    +H2O

    (not observed)

    EtOH

    5 mol % AuClAgSbF65 equiv n-PrOH

    THFCH2Cl2 (1:1)

    CO2Et CO2nPr+

    ca. 1:15 mol% Au/AgSbF6, 5 equiv n-PrOH

    THFCH2Cl2 (1:1)

    transesterification:(does not occur)

    OH

    OEt

    7

    6

  • 28

    Isomerization of the Z-enoates to the E-enoates does not occur under the reaction conditions:

    extending the reaction time does not have a significant effect on the product ratio, and resubjecting the

    enoate mixtures to the rearrangement conditions does not change the ratio of stereoisomers. Therefore,

    we assume that the non-thermodynamic product distribution is purely the result of kinetic control.

    Perhaps the most compelling observation relevant to the mechanistic hypothesis laid out (Fig.

    22) is that when n-propanol was used in place of ethanol, the resulting product mixture comprised ethyl

    and propyl esters in a roughly 1:1 ratio.

    Figure 23.Reaction of ,-unsaturated ester in Meyer-Schuster conditions

    Figure 24. Reaction of ester product in Meyer-Schuster conditions

    When this experiment was repeated on 2e using scandium (III) triflate as the catalyst (Fig. 23),

    the ratio of ethyl to propyl esters (3e:3e) was 25:75 (as estimated by 1H NMR). In other words, there

    was only about 25% retention of the ethyl ester in the product mixture. According to the mechanism

    outlined in Figure 22, however, the ethoxy group should be at least 50% retained, even at high levels of

    n-propanol. Transesterification does not occur under the reaction conditions (Fig. 24), so we conclude

    that the scandium (III) triflate-catalyzed reactions proceed by a slightly different mechanism than those

    catalyzed of cationic gold salts. One such potential mechanism is outlined in Figure 25.

  • 29

    Figure 25. Hypothesized scandium (III) triflate-catalyzed reaction

    Based on the consistent lack of -hydroxy ester by-products (i.e., 6), the scandium(III) triflate-

    catalyzed MeyerSchuster reactions of secondary alcohols 228 likely also proceed via intermediate 1,1-

    diethoxy-allene 7 (Fig. 25). Addition of a second equivalent of ethanol to allene 7 would give rise to

    ortho-ester 9, which can then hydrolyze via 8 to reach the , -unsaturated ester (3). Ortho-ester

    intermediate 9 thus easily accounts for up to 67% incorporation of the alcohol additive, but we observed

    75% (nearly statistical) incorporation of n-propanol in the experiment recounted in Figure 23. This high

    level of incorporation can be explained by dynamic alcohol exchange reactions of ortho-ester 9.

    A series of experiments were conducted in which the incorporation of the alcohol additive

    (propanol) was tracked with respect to the amount of alcohol added (Table 8). 29 In each case, the ratio

    of propyl and ethyl esters (3e:3e) was less than but close to statistical incorporation of propanol. These

    data are consistent with a hemi-labile intermediate (e.g., 9) that can undergo partial equilibration before

    giving way irreversibly to the observed , -unsaturated ester (3).

    R

    OH

    OEtR

    CO2Et

    R OEt

    OHOEt

    EtOH

    R

    OEtOEt

    EtOHH2O

    R OEt

    OEtOEtEtOH H2O

    EtOH

    1 mol% Sc(OTf)3

    CH2Cl2/EtOH (4:1)2

    79

    3

  • 30

    Figure 26. Conditions for Meyer-Schuster rearrangement with n-propanol

    Table 8. Statistical incorporation of n-propanol

    An attractive feature of this mechanistic hypothesis is that it can account for the high

    stereoselectivity observed for the E-olefin isomer in the scandium (III) triflate-catalyzed Meyer

    Schuster reactions of secondary alcohols. 28 Direct hydrolysis of allene 7 would most likely occur under

    kinetic control, whereas vinyl ortho-ester 9 provides the opportunity for thermodynamic establishment

    of olefin geometry using the exaggerated steric profile of ortho-ester 9.

    OH

    OEt

    1 mol% Sc(OTf)3

    CH2Cl2

    CO2R

    R = Et: 2R = nPr: 2'1

  • 31

    2.3.1 Conclusion

    Acetylide addition followed by the Lewis acid catalyzed MeyerSchuster reaction of

    ethoxyalkynyl carbinols provides a strategy for the olefination of aldehydes and ketones. Many different

    Lewis and protic acids catalyze MeyerSchuster reactions of ethoxyacetylenes; Lewis acids that

    demonstrate an affinity for -bonds were most effective in our methodology. After a detailed screening

    of many catalysts, we recommend scandium (III) triflate for the excellent reactivity and optimal

    stereoselectivity that it provides in the MeyerSchuster reactions, even at low catalyst loading. The

    method would appear to be limited only by the ability to access the requisite propargyl alcohols via

    ethoxyacetylide addition to carbonyls, and such reactions are known to be quite general.

    Stereoselectivities in the two-stage olefination of aldehydes range from good to excellent, whereas , -

    unsaturated esters derived from ketones are obtained with little to no stereocontrol. This method is likely

    to find widespread application in organic synthesis, particularly for its unique ability to complete the

    olefination of hindered ketones in excellent yield.

  • 32

    CHAPTER III

    EXPERIMENTAL

    3.3.1 General Information

    1H NMR and 13C NMR spectra were recorded on 300 MHz spectrometer using CDCl3 as the

    deuterated solvent. The chemical shifts () are reported in parts per million (ppm) relative to the residual

    CHCl3 peak (7.26 ppm for 1H NMR, 77.0 ppm for 13C NMR). The coupling constants (J) were reported

    in hertz (Hz). IR spectra were recorded on an FTIR spectrometer on NaCl discs. Mass spectra were

    recorded using chemical ionization (CI) or electron ionization (EI) technique. Yields refer to isolated

    material judged to be 95% pure by 1H NMR spectroscopy following silica gel chromatography. All

    chemicals were used as received unless otherwise stated. Tetrahydrofuran (THF) and methylene chloride

    (CH2Cl2) were purified by passing through a column of activated alumina. The n-BuLi solutions were

    titrated with menthol dissolved in tetrahydrofuran using 1,10-phenanthroline as the indicator. The

    purifications were performed by flash chromatography using silica gel F-254 (230499 mesh particle

    size)

  • 33

    3.3.2 Synthesis of Substrates

    General procedure for the preparation of ethoxyalkynyl carbinols (12)

    To a THF solution (7 mL) of ethyl ethynyl ether (0.7 g, ca. 40% by weight in hexanes, ca.

    9 mmol) was added n-BuLi (1.5 mL, 3.4 mmol, 2.3 M) dropwise over 5 min at 78 C under argon

    atmosphere. The solution was allowed to warm to 0 C over 1 h and held at 0 C for an additional

    30 min. The solution was then recooled to 78 C and pinacolone (1b, 0.30 mL, 2.4 mmol) was added in

    one portion. The solution was allowed to warm to room temperature over 1 h and held at room

    temperature for an additional 3 h. Saturated aqueous NH4Cl solution was added to quench the reaction,

    and the mixture was extracted with ethyl acetate. The organic layer was washed sequentially with water,

    saturated aqueous sodium bicarbonate, and brine. The organic layer was dried over MgSO4, filtered, and

    concentrated under reduced pressure. The residue was purified using silica gel column chromatography

    (gradient elution with 20:1 to 7:1 hexanes/ethyl acetate) to give 1-ethoxy-3-methyl-3-tert-butyl-1-

    propyn-3-ol (2b) in 83% yield (0.34 g).

    1-Ethoxy-3-methyl-3-tert-butyl-1-propyn-3-ol (2b) 1H NMR (300 MHz, CDCl3) 1.03 (s, 9H), 1.37 (t, J=7.1 Hz, 3H), 1.41 (s, 3H), 1.71 (s, 1H), 4.08 (q,

    J=7.1 Hz, 2H); 13C NMR (75 MHz, CDCl3) 14.3, 25.2, 25.6, 38.4, 41.9, 73.9, 74.2, 92.8; IR (neat)

    3479, 2971, 2873, 2261, 1481, 1392, 1369, 1219, 1094, 1007, 908, 878 cm1; HRMS (CI) calcd for

    C10H19O2 ([M+H]+) 171.1385. Found 171.1390.

    Ethoxy-dec-1-yn-3-ol (2d)

    The title compound was prepared in a similar manner as described above (>99% yield); 1H NMR

    (300 MHz, CDCl3) 0.860.90 (m, 3H), 1.211.46 (m, 10H), 1.37 (t, J=7.1 Hz, 3H), 1.561.70 (m, 3H),

    4.09 (q, J=7.1 Hz, 2H), 4.39 (q, J=6.3 Hz, 1H); 13C NMR (75 MHz, CDCl3) 14.0, 14.2, 22.6, 25.3,

    29.2, 29.2, 31.7, 38.7, 39.7, 62.4, 74.4, 93.6; IR (neat) 3381, 2927, 2263, 1722, 1467 cm1; HRMS (CI)

    calcd for C12H22O2 (M+H+) 199.1698. Found 199.1692.

  • 34

    1-Ethoxy-4, 4-dimethyl-pent-1-yn-3-ol (2e)

    The title compound was prepared in a similar manner as described above (97% yield); 1H NMR

    (300 MHz, CDCl3) 0.97 (s, 9H), 1.38 (t, J=7.1 Hz, 3H), 4.03 (d, J=6.0 Hz, 1H), 4.10 (q, J=7.1 Hz,

    2H); 13C NMR (75 MHz, CDCl3) 14.2, 25.2, 35.8, 38.0, 71.0, 74.3, 94.0; IR (neat) 3431, 2956, 2714,

    2264, 1629 cm1; HRMS (EI) calcd for C9H16O2 (M+) 156.1150. Found 156.1103.

    1-Cyclohexyl-3-ethoxy-prop-2-yn-1-ol (2f)

    The title compound was prepared in a similar manner as described above (76% yield); 1H NMR

    (300 MHz, CDCl3) 0.831.3 (m, 6H), 1.38 (t, J=7.1 Hz, 3H), 1.571.84 (m, 6H), 4.10 (q, J=7.1 Hz,

    2H), 4.18 (t, J=5.7 Hz, 1H); 13C NMR (75 MHz, CDCl3) 14.3, 25.9, 25.9, 26.4, 28.1, 28.6, 38.3, 44.6,

    67.0, 74.5, 94.3; IR (neat) 3411, 2980, 2460, 1719, 1450 cm1; HRMS (CI) calcd for C11H18O2 (M+H+)

    183.1385. Found 183.1390.

    3-Ethoxy-1-phenyl-prop-2-yn-1-ol (2g)

    The title compound was prepared in a similar manner as described above (92% yield); 1H NMR

    (300 MHz, CDCl3) 1.39 (t, J=7.1 Hz, 3H), 2.02 (d, J=6.0 Hz, 1H), 4.15 (q, J=7.1 Hz, 2H), 5.51 (d,

    J=6.0 Hz, 1H), 7.317.40 (m, 3H), 7.527.56 (m, 2H); 13C NMR (75 MHz, CDCl3) 14.4, 38.8, 64.6,

    74.8, 95.4, 126.5, 128.0, 128.5, 129.2; IR (neat) 3401, 2981, 2226, 1718, 1450 cm1; HRMS (EI) calcd

    for C11H12O2 (M+) 176.0834. Found 176.0837.

    General procedure for the preparation of ,-unsaturated esters (23)

    To a 4:1 v/v CH2Cl2/ethanol solution (10 mL) of 1-ethoxy-dec-1-yn-3-ol (2d, 0.10 g, 0.51 mmol) in an

    open flask was added Sc(OTf)3 (2.5 mg, 0.005 mmol). Progress of the reaction was monitored by TLC

    analysis. After 1 h, the reaction mixture was concentrated under reduced pressure and purified using

    silica gel column chromatography (hexanes/ethyl acetate, 50:1) to give ethyl (E)-dec-2-enoate (3d) in

    70% yield (70 mg).

  • 35

    (E/Z)-3,4,4-Trimethyl-1-pent-2-enoic acid ethyl ester (3b)

    The title compound was prepared in a similar manner as described above (89% yield, E/Z ratio, 58:42); 1H NMR (300 MHz, CDCl3, E isomer) 1.10 (s, 9H), 1.28 (t, J=7.1 Hz, 3H), 2.16 (br d, J=1.1 Hz, 3H),

    4.14 (q, J=7.1 Hz, 2H), 5.74 (q, J=1.1 Hz, 1H); 1H NMR (300 MHz, CDCl3, Z isomer) 1.20 (s, 9H),

    1.28 (t, J=7.1 Hz, 3H), 1.84 (br d, J=1.3 Hz, 3H), 4.14 (q, J=7.1 Hz, 2H), 5.63 (q, J=1.3 Hz, 1H); 13C

    NMR (75 MHz, CDCl3, E/Z mixture) 14.1, 14.3, 15.1, 23.9, 28.5, 29.0, 36.4, 37.9, 59.4, 60.0, 112.9,

    116.6, 158.5, 167.2, 167.5, 167.9; IR (neat, E/Z mixture) 2970, 2873, 1719, 1634, 1466, 1372, 1262,

    1182, 1123, 1054, 868 cm1; HRMS (EI) calcd for C10H18O2 (M+) 170.1307. Found 170.1306.

    (E)-Dec-2-enoic acid ethyl ester (3d)

    The title compound was prepared as described above (70% yield); 1H NMR (300 MHz, CDCl3) 0.86

    0.90 (m, 3H), 1.261.31 (m, 8H), 1.28 (t, J=7.1 Hz, 3H), 1.421.47 (m, 2H), 2.19 (ddd, J=14.6, 7.1,

    1.2 Hz, 2H), 4.18 (q, J=7.1 Hz, 2H), 5.80 (br d, J=15.6 Hz, 1H), 6.96 (dt, J=15.6, 7.0 Hz, 1H).

    (E)-4,4-Dimethyl-pent-2-enoic acid ethyl ester (3e)

    The title compound was prepared in a similar manner as described above (97% yield); 1H NMR

    (300 MHz, CDCl3) 1.08 (s, 9H), 1.29 (t, J=7.1 Hz, 3H), 4.19 (q, J=7.1 Hz, 2H), 5.73 (d, J=15.9 Hz,

    1H), 6.97 (d, J=15.9 Hz, 1H).

    (E)-4,4-Dimethyl-pent-2-enoic acid ethyl ester (3e)

    The title compound was prepared in a similar manner as described above (97% yield); 1H NMR

    (300 MHz, CDCl3) 1.08 (s, 9H), 1.29 (t, J=7.1 Hz, 3H), 4.19 (q, J=7.1 Hz, 2H), 5.73 (d, J=15.9 Hz,

    1H), 6.97 (d, J=15.9 Hz, 1H).

    (E)-3-Cyclohexyl-acrylic acid ethyl ester (3f)

    The title compound was prepared in a similar manner as described above (75% yield); 1H NMR

    (300 MHz, CDCl3) 1.121.31 (m, 5H), 1.29 (t, J=7.1 Hz, 3H), 1.641.77 (m, 5H), 2.042.17 (m, 1H),

    4.18 (q, J=7.1 Hz, 2H), 5.75 (dd, J=15.8, 1.4 Hz, 1H), 6.91 (dd, J=15.8, 6.7 Hz).

  • 36

    (E/Z)-3-Phenyl-2-propenoic acid ethyl ester (3g)

    The title compound was prepared in a similar manner as described above (93% yield, E/Z ratio, 77:23); 1H NMR (300 MHz, CDCl3, E isomer) 1.34 (t, J=7.1 Hz, 3H), 4.27 (q, J=7.1 Hz, 2H), 6.44 (d,

    J=16.0 Hz, 1H), 7.377.40 (m, 3H), 7.517.54 (m, 2H), 7.69 (d, J=16.0 Hz, 1H); 1H NMR (300 MHz,

    CDCl3, Z isomer) 1.24 (t, J=7.1 Hz, 3H), 4.17 (q, J=7.1 Hz, 2H), 5.95 (d, J=12.6 Hz, 1H), 6.95 (d,

    J=12.6 Hz, 1H), 7.337.38 (m, 3H), 7.567.59 (m, 2H).

    General two-step procedure for the preparation of ,-unsaturated esters (13)

    To a THF solution (2.6 mL) of ethyl ethynyl ether (0.13 g, ca. 40% by weight in hexanes, ca.

    2 mmol) was added n-BuLi (0.40 mL, 0.75 mmol, 2.0 M) dropwise over 5 min at 78 C under argon

    atmosphere. The solution was allowed to warm to 0 C over 1 h and held at 0 C for an additional

    30 min. The solution was then recooled to 78 C and 2-adamantanone (1a, 75 mg, 0.50 mmol) was

    added in one portion. The solution was allowed to warm to room temperature over 1 h and held at room

    temperature for an additional 3 h. Saturated aqueous NH4Cl solution was added to quench the reaction

    and the mixture was extracted with ethyl acetate. The organic layer was washed sequentially with water,

    saturated aqueous sodium bicarbonate, and brine. The organic layer was dried over MgSO4, filtered, and

    concentrated under reduced pressure. To the concentrated mixture in an open flask were added CH2Cl2

    (8 mL), absolute ethanol (2 mL), and Sc(OTf)3 (2.5 mg, 0.005 mmol). After 6 h, the reaction mixture

    was concentrated under reduced pressure and purified using silica gel column chromatography

    (hexanes/ethyl acetate, 50:1) to give adamantan-2-ylidene-acetic acid ethyl ester (3a) in 96% yield over

    two steps (106 mg).

    Adamantan-2-ylidene-acetic acid ethyl ester (3a) 1H NMR (300 MHz, CDCl3) 1.27 (t, J=7.1 Hz, 3H), 1.86 (br s, 6H), 1.931.96 (m, 6H), 2.43 (br s,

    1H), 4.07 (br s, 1H), 4.13 (q, J=7.1 Hz, 2H), 5.58 (s, 1H).

  • 37

    (4,6,6-Trimethyl-bicyclo[3.1.1]hept-3-en-(2E/Z)-ylidene)-acetic acid ester (3c)

    Title compound was prepared in a similar manner as described above (97% yield); 1H NMR (300 MHz,

    CDCl3, minor isomer) 0.86 (s, 3H), 1.28 (t, J=7.1 Hz, 3H), 1.40 (s, 3H), 1.68 (d, J=7.9 Hz, 1H), 1.90

    (d, J=1.5 Hz, 3H), 2.202.45 (m, 1H), 2.532.63 (m, 2H), 4.084.20 (m, 2H), 5.32 (s, 1H), 7.13 (s, 1H); 1H NMR (300 MHz, CDCl3, major isomer) 0.84 (s, 3H), 1.26 (t, J=7.1 Hz, 3H), 1.44 (s, 3H), 1.58 (d,

    J=8.8 Hz, 1H), 1.86 (d, J=1.4 Hz, 3H), 2.202.45 (m, 1H), 2.532.63 (m, 2H), 4.084.20 (m, 2H), 5.46

    (s, 1H), 5.77 (s, 1H); 13C NMR (75 MHz, CDCl3, E/Z mixture) 14.3, 14.4, 21.7, 21.8, 23.2, 23.6, 26.4,

    26.5, 37.5, 38.1, 45.3, 47.8, 48.2, 49.0, 49.1, 53.1, 59.2, 59.3, 107.6, 110.0, 117.6, 121.6, 156.9, 158.0,

    159.6, 161.2, 166.8, 167.4; IR (neat) 2979, 2930, 2870, 1708, 1622, 1466, 1443, 1380, 1370, 1226,

    1164, 1040, 874, 705 cm1; HRMS (EI) calcd for C14H20O2 (M+) 220.1463. Found 220.1462.

    (2-Isopropyl-5-methyl-cyclohexylidiene)-acetic acid ethyl ester (3h)

    The title compound was prepared in a similar manner as described above (98% yield); 1H NMR

    (300 MHz, CDCl3, E/Z mixture, diagnostic peaks) 2.55 (ddd, J=12.9, 5.5, 1.5 Hz), 3.14 (dd, J=12.9,

    4.3 Hz), 3.483.52 (m), 4.13 (q, J=7.1 Hz), 4.14 (q, J=7.1 Hz), 5.63 (br s); 13C NMR (75 MHz, CDCl3,

    E/Z mixture) 14.3, 18.1, 19.5, 20.5, 20.8, 21.8, 23.4, 26.1, 26.8, 27.0, 27.6, 30.4, 31.6, 33.6, 33.9, 35.4,

    36.1, 40.0, 43.5, 50.8, 52.6, 55.9, 59.3, 59.4, 113.3, 116.3, 164.8, 167.1.

    Ethyl 3,3-diphenylpropenoate (3i)

    The title compound was prepared in a similar manner as described above (99% yield); 1H NMR

    (300 MHz, CDCl3) 1.11 (t, J=7.1 Hz, 3H), 4.05 (q, J=7.1 Hz, 2H), 6.37 (s, 1H), 7.207.23 (m, 2H),

    7.307.39 (m, 8H).

  • 38

    LIST OF REFERENCES

    1. L. Krti and B. Czak, Strategic Applications of Named Reactions in Organic Synthesis, Elsevier, New York, NY (2003)

    2. (a) B.E. Maryanoff and A.B. Reitz, Chem. Rev. 89 (1989), pp. 863927.

    3. S.E. Kelly In: B.M. Trost and I. Fleming, Editors, Alkene Synthesis, Comprehensive Organic Synthesis Vol. 1, Pergamon, Oxford (1991), pp. 729817.

    4. B.M. Trost, Science 254 (1991), pp. 14711477.

    5. Brandsma L, Acetylenes, Allenes and Cumulenes Elsevier; Amsterdam: 2003.

    6. Modern Acetylene Chemistry Stang PJ, Diederich F,VCH; Weinheim: 1995.

    7. S. Swaminathan and K.V. Narayanan, Chem. Rev. 71 (1971), pp. 429438.

    8. Lewis Acids in Organic Synthesis Yamamoto H,Wiley-VCH; New York: 2000.

    9. (a)Dyker G,Angew. Chem. Int. Ed. 2000, 39: 4237. (b) Hashmi AS,Gold Bull. 2003, 36: 3. (c) Ma S, Yu S, Gu Z,Angew. Chem. Int. Ed. 2006, 45: 200. (d) Asao N,Synlett 2006, 1645.

    10. Asao N, Sato K,Org. Lett. 2006, 8: 5361. (b )Staben ST, Kennedy-Smith JJ, Huang D, Corkey BK,

    LaLonde RL, Toste FD,Angew. Chem. Int. Ed. 2006, 45: 5991. (c) Belting V, Krause N,Org. Lett. 2006, 8: 4489. (d)Wang S, Zhang L,Org. Lett. 2006, 8: 4585. (e) Huang B, Yao X, Li C.-J,Adv. Synth. Catal. 2006, 348: 1528. (f) Seregin IV, Gevorgyan V,J. Am. Chem. Soc. 2006, 128: 12050. (g)Carretin S, Blanco MC, Corma A, Hashmi AS,Adv. Synth. Catal. 2006, 348: 1283. (h)Park S, Lee D,J. Am. Chem. Soc. 2006, 128: 10664. (i) Nakamura I, Sato T, Yamamoto Y,Angew. Chem. Int. Ed. 2006, 45: 4473. (j)Kang J.-E, Kim H.-B, Lee J.-W, Shin S,Org. Lett. 2006, 8: 3537. (k) Sun J, Conley MP, Zhang L, Kozmin SA,J. Am. Chem. Soc. 2006, 128: 9705. (l) Liu Y, Liu M, Guo S, Tu H, Zhou Y, Gao H,Org. Lett. 2006, 8: 3445. (m) Robles-Machin R, Adrio J, Carretero JC,J. Org. Chem. 2006, 71: 5023. (n) Harrison TJ, Kozak JA, Corbella-Pane M, Dake GR,J. Org. Chem. 2006, 71: 4525. (o) Frstner A, Hannen P,Chem. Eur. J. 2006, 12: 3006. (p) Georgy M, Boucard V, Campagne J.-M,J. Am. Chem. Soc. 2005, 127: 14180. (q) Fukuda Y, Utimoto K,Bull. Chem. Soc. Jpn. 1991, 64: 2013.

    11. D.A. Engel and G.B. Dudley, Org. Lett. 8 (2006), pp. 40274029.

  • 39

    12. Examples of MeyerSchuster reactions of ethoxyalkynyl carbinols using hard Lewis or protic acids: (a) M. Duraisamy and H.M. Walborsky, J. Am. Chem. Soc. 105 (1983), pp. 32523264. (b) S.C. Welch, C.P. Hagan, D.H. White, W.P. Fleming and J.W. Trotter, J. Am. Chem. Soc. 99 (1977), pp. 549556. (c) D. Crich, S. Natarajan and J.Z. Crich, Tetrahedron 53 (1997), pp. 71397158.

    13. The combination of an electron-rich -system and soft Lewis acid catalyst has also been used in

    cycloisomerization reactions: J. Sun, M.P. Conley, L. Zhang and S.A. Kozmin, J. Am. Chem. Soc. 128 (2006), pp. 97059710.

    14. Because the 1,3-hydroxy shift is not concerted and there is ample opportunity for the hydroxy to

    exchange with water in the reaction medium, the MeyerSchuster reaction is not a true rearrangement. For mechanistic investigations of the MeyerSchuster reaction, see:.(a)M. Edens, D. Boerner, C.R. Chase, D. Nass and M.D. Sciavelli, J. Org. Chem. 42 (1977), pp. 34033408. (b)J. Andres, R. Cardenas, E. Silla and O. Tapia, J. Am. Chem. Soc. 110 (1988), pp. 666674. (c)S. Yamabe, Tsuchida and S. Yamazaki, J. Chem. Theory Comput. 2 (2006), pp. 13791387.

    15. S.S. Lpez, D.A. Engel and G.B. Dudley, Synlett (2007), pp. 949953.

    16. The question as to what reaction pathway(s) leading from the propargyl acetate to the ,-unsaturated ketone is catalyzed by cationic gold salts remains open. For further discussion, see Ref.17f.

    17. Recent examples: Refs. (a) I. Imagawa, Y. Asai, H. Takano, H. Hamagaki and M. Nishizawa, Org.

    Lett. 8 (2006), pp. 447450.(b)V. Cadierno, J. Dez, S.E. Garca-Garrido, J. Gimeno and N. Nebra, Adv. Synth. Catal. 348 (2006), pp. 21252132. (c) C. Sun, X. Lin and S.M. Weinreb, J. Org. Chem. 71 (2006), pp. 31593166. (d) M. Yu, G. Li, S. Wang and L. Zhang, Adv. Synth. Catal. 349 (2007), pp. 871875. (e) E. Bustelo and P.H. Dixneuf, Adv. Synth. Catal. 349 (2007), pp. 933942. (f)N. Marion, P. Carlqvist, R. Gealageas, P. de Frmont, F. Maseras and S.P. Nolan, Chem.Eur. J. 13 (2007), pp. 64376451. (g)S.I. Lee, J.Y. Baek, S.H. Sim and Y.K. Chung, Synthesis (2007), pp. 21072114. (h)M.J. Sandelier and P. DeShong, Org. Lett. 9 (2007), pp. 32093212. (i)Y. Sugawara, W. Yamada, S. Yoshida, T. Ikeno and T. Yamada, J. Am. Chem. Soc. 129 (2007), pp. 1290212903.

    18. Protic acids in the absence of gold are significantly less effective (ref. 2). Examples of the protic-acid-catalyzed Meyer-Schuster rearrangement of ethoxyalkynyl carbinols: (a) Welch SC, Hagan CP, White DH, Fleming WP, Trotter JW,J. Am. Chem. Soc. 1977, 99: 549. (b) Duraisamy M, Walborsky HM,J. Am. Chem. Soc. 1983, 105: 3252. (c) Crich D, Natarajan S, Crich JZ,Tetrahedron 1997, 53: 7139.

    19. Satisfactory characterization data (1H NMR, 13C NMR, IR, HRMS) were obtained for all compounds. Yields refer to at least 50 mg of material isolated in >95% purity.

    20. For a previous report on the combined use of oxygen-activated alkynes and cationic gold catalysts, see: Zhang L, Kozmin SA,J. Am. Chem. Soc. 2004, 126: 11806

  • 40

    21. K.H. Meyer and K. Schuster, Chem. Ber. 55 (1922), pp. 819822

    22. Examples:.(a)M.B. Erman, I.S. Aul'chenko, L.A. Kheifits, V.G. Dulova, J.N. Novikov and M.E. Vol'pin, Tetrahedron Lett. (1976), pp. 29812984. (b)P. Chabardes, Tetrahedron Lett. 29 (1988), pp. 62536256. (c)B.M. Choudary, A. Durga Prasad and V.L.K. Valli, Tetrahedron Lett. 31 (1990), pp. 75217522. (d)K. Narasaka, H. Kusama and Y. Hayashi, Chem. Lett. (1991), pp. 14131416. (e)M. Yoshimatsu, M. Naito, M. Kawahigashi, H. Shimizu and T. Kataoka, J. Org. Chem. 60 (1995), pp. 47984802. (f)C.Y. Lorber and J.A. Osborn, Tetrahedron Lett. 37 (1996), pp. 853856. (g)T. Suzuki, M. Tokunaga and Y. Wakatsuki, Tetrahedron Lett. 43 (2002), pp. 75317533.

    23. The Rupe rearrangement ( H. Rupe and E. Kambli, Helv. Chim. Acta 9 (1926), p. 672 ).

    24. (a)T.-L. Ho, Hard and Soft Acids and Bases Principle in Organic Chemistry, Academic, New York, NY (1977).(b)P.K. Chattaraj, H. Lee and R.G. Parr, J. Am. Chem. Soc. 113 (1991), pp. 18551856.

    25. In: H. Yamamoto, Editor, Lewis Acids in Organic Synthesis, Wiley-VCH, New York, NY (2000).

    26. For seminal examples illustrating this concept, see: M. Georgy, V. Boucard and J.-M. Campagne, J. Am. Chem. Soc. 127 (2005), pp. 1418014181.

    27. Examples of MeyerSchuster reactions of ethoxyalkynyl carbinols using hard Lewis or protic

    acids:(a)M. Duraisamy and H.M. Walborsky, J. Am. Chem. Soc. 105 (1983), pp. 32523264. (b)S.C. Welch, C.P. Hagan, D.H. White, W.P. Fleming and J.W. Trotter, J. Am. Chem. Soc. 99 (1977), pp. 549556. (c)D. Crich, S. Natarajan and J.Z. Crich, Tetrahedron 53 (1997), pp. 71397158.

    28. The MeyerSchuster reactions of tertiary alcohols may take a different course. Further investigations

    are planned and will be communicated in due course. 29. Reactions conducted with less than a full equivalent of propanol were slow and inefficient, and are

    omitted from Table 8.

  • 41

    BIOGRAPHICAL SKETCH

    Susana Sorina Lpez was born on September 23rd 1980 in Miami Beach, Florida. She grew up in

    North Miami, Florida moving to Hollywood, Florida during her freshman year of high school where her

    parents, Oscar and Susana Mercedes Lpez, still reside. Susana was classically trained in voice and the

    flute as well as in various forms of dance, including ballet, tap jazz and modern from an early age.

    During her high school years, she figure skated competitively winning several competitions at the

    sectional, regional and national level. Upon graduating high school in 1999, she received a theatre and

    dance scholarship to attend Lees-McRae College in Banner Elk, North Carolina but decided to return to

    South Florida after her freshman year to pursue pre-medical studies. Susana received her Associates of

    Science in Biology from Broward Community College in the spring of 2003. She began her

    undergraduate studies in the fall of 2003 at Barry University and discovered a passion for organic

    chemistry while taking the course for her pre-medical major requirements. She changed her major in the

    fall of 2004 to chemistry and did active research under the direction of Dr. George Fisher and Dr. Paul I.

    Higgs. She also worked under the guidance of Dr. Anthony J. Pearson of Case Western Reserve

    University in Cleveland, Ohio during the summer of 2005. In the fall of 2005 Susana graduated with a

    Bachelors of Science degree from Barry University and continued to do research at Barry until moving

    to Tallahassee, Florida in the summer of 2006 to pursue her graduate studies at Florida State University.

    The Florida State UniversityDigiNole Commons4-2-2009

    Methodology for the Olefination Of Aldehydes and Ketones Via the Meyer-Schuster ReactionSusana Sorina LpezRecommended Citation