242
Florida State University Libraries Electronic Theses, Treatises and Dissertations The Graduate School 2010 Addition / C-C Bond Cleavage Reactions of Vinylogous Acyl Triflates and Their Application to Natural Product Synthesis David Mack Jones Follow this and additional works at the FSU Digital Library. For more information, please contact [email protected]

Florida State University Libraries - FLVC

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Florida State University Libraries - FLVC

Florida State University Libraries

Electronic Theses, Treatises and Dissertations The Graduate School

2010

Addition / C-C Bond Cleavage Reactionsof Vinylogous Acyl Triflates and TheirApplication to Natural Product SynthesisDavid Mack Jones

Follow this and additional works at the FSU Digital Library. For more information, please contact [email protected]

Page 2: Florida State University Libraries - FLVC

THE FLORIDA STATE UNIVERSITY

COLLEGE OF ARTS AND SCIENCES

ADDITION / C-C BOND CLEAVAGE REACTIONS OF VINYLOGOUS

ACYL TRIFLATES AND THEIR APPLICATION

TO NATURAL PRODUCT SYNTHESIS

By

DAVID MACK JONES

A Dissertation submitted to the Department of Chemistry and Biochemistry

in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

Degree Awarded: Spring Semester, 2010

Copyright © 2010 David M. Jones

All Rights Reserved

Page 3: Florida State University Libraries - FLVC

ii

The members of the committee approve the dissertation of David M. Jones

defended on December 3, 2009.

__________________________________ Gregory B. Dudley Professor Directing Dissertation

__________________________________ Kenneth Taylor University Representative

__________________________________ Jack Saltiel Committee Member

__________________________________

D. Tyler McQuade Committee Member

__________________________________

Kenneth Goldsby Committee Member

Approved: _____________________________________ Joseph B. Schlenoff, Chair, Department of Chemistry and Biochemistry The Graduate School has verified and approved the above-named committee

members.

Page 4: Florida State University Libraries - FLVC

iii

This manuscript is dedicated to my Mother and Father, without whom I would have been lost. Their constant and unwavering support has made all that I am,

and all that I will be, possible.

Page 5: Florida State University Libraries - FLVC

iv

ACKNOWLEDGEMENTS This body of work has been made possible not only through my hard work, but through the personal and academic support of many people. I would like acknowledge Professor Gregory Dudley. He was charged with the difficult task of not only providing challenging problems for me, his student, to explore, but also he had to provide an environment in which I could hone my own set of tools for future scientific endeavors. As a naïve 1st year graduate student I joined his research group and his constant guidance set me on the right path. As the years progressed he no longer provided answers, but only answered my questions with yet more questions. I remember being completely frustrated at the time with this tact. However now, in the waning moments of my graduate studies, I understand the role that a research advisor must play in the development of a Ph.D. student. I owe much to Dr. Dudley and I am very appreciative of his ability to change me from that naïve graduate student into the independent scientist that I have become today. I would also like to thank the members of the Dudley research group: Dr. Tim Briggs, who introduced me to lab techniques, and guided my early research; Dr. Shin Kamijo, who made my work possible through his early efforts; Dr. Doug Engel, who entered the lab at the same time as I and provided constant competition; Sami Tlais and Jingyue Yang, who often provided company late into the night in the lab; Marilda Lisboa, who provided several intermediates in my palmerolide research; and the rest of the members, past and present. I would like to acknowledge my family for providing constant support, financial and otherwise. Mom and Dad, you have truly been the foundation of my life. Although many times in grad school, you could not offer any advice to help me with my problems, you always made sure that I knew you would do anything in your power to help me. Amy, Laura, Ken, and your families, you have provided support to me in ways that you cannot even understand. I am thankful for your understanding of my inability to attend family gatherings, niece and nephew birthday parties, and other important milestones. Bamp, June, Grammy, and everyone else in the family, thank you. I would like to thank Kerri, a very big part of my life throughout graduate school; you have helped me through many difficult times. I would like to thank the Pritchard family for being like a second family to me. Thanks to Doug, Kerry, Phil, Chris, Antonio, Matt, Mike, Scott and all the other great friends in my life. I wish I had more space to mention all of those people that deserve recognition for supporting me in the generation of this manuscript, please forgive me for any omissions. Lastly, I would like to thank all of those who helped me edit this manuscript, without whom, this document would not have been possible: Kerry Gilmore, Sami Tlais, Marilda Lisboa, and Professor Dudley.

Page 6: Florida State University Libraries - FLVC

v

TABLE OF CONTENTS List of Tables .................................................................................. ….. vi List of Figures ................................................................................. ….. vii List of Abbreviations ....................................................................... ….. xii Abstract .................................................................................... ….. xviii

1. INTRODUCTION: C-C BOND CLEAVAGE AND FRAGMENTATION REACTIONS IN ORGANIC SYNTHESIS........ 1

2. SYNTHESIS OF (Z)-6-HENEICOSEN-11-ONE: THE SEX PHEROMONE OF THE DOUGLAS-FIR TUSSOCK MOTH ..... ….. 13

The Doulas-Fir Tussock Moth ........................................... ….. 13 Synthesis of (Z)-6-Heneicosen-11-one ............................. ….. 16 Experimental ..................................................................... ….. 21

3. A FRAGMENTATION / BENZANNULATION STRATEGY TO

PROVIDE ACCESS TO BENZO-FUSED INDANES ................. 37 Introduction ....................................................................... ….. 37 The Alcyopterosins ........................................................... ….. 37 Retrosynthetic Analysis of Alcyopterosin A ....................... ….. 53 Exploring Gold and Copper Catalyzed Benzannulations .. .. 59 Experimental ..................................................................... ….. 70

4. SYNTHESIS OF THE EASTERN HEMISPHERE (C1-C15)

OF PALMEROLIDE A ............................................................... ….. 115 Introduction ....................................................................... ….. 115 The Melanoma Problem .................................................... ….. 116 Palmerolide A ................................................................... ….. 121 Synthesis of the Eastern Hemisphere of Palmerolide A ... ….. 138 Experimental ..................................................................... ….. 145

5. RE-EXPLORING THE CLAISEN-TYPE CONDENSATIONS OF VINYLOGOUS ACYL TRIFLATES ...................................... ….. 170

New Insights into the Mechanism ..................................... ….. 170 Synthesis of -Ketophosphonates .................................... ….. 177 Experimental ..................................................................... ….. 184 REFERENCES .............................................................................. ….. 202 BIOGRAPHICAL SKETCH ............................................................. ….. 222

Page 7: Florida State University Libraries - FLVC

vi

LIST OF TABLES

Table 1: Scope of Original Fragmentation Reaction with Respect to Nucleophiles ............................................................................... 11 Table 2: Grignard Triggered Fragmentation of 2 ....................................... 20 Table 3: DNA Binding Assay Performed by Iglesias et al. ......................... 51 Table 4: Average Values (MG-MID) for In Vitro Antitumor Activity on the NCI 60-Cell Line Panel ................................................................ 52

Table 5: Preliminary Screening of Benzannulation Reactions of Substrates 84a-f.......................................................................... 65 Table 6: Selected Data from Nicolaou’s SAR Study (GI50 Values in M) .. 133 Table 7: Claisen-Type Condensation of Vinylogous Acyl Triflate 2 ........... 141 Table 8: Comparison of the Acidities of Several Acetophenone Phosphonate and Phosphine Oxide Derivatives in DMSO .......... 174 Table 9: Reactions of Vinylogous Acyl Triflates with 1.1 Equivalents of Dimethyl lithiomethylphosphonate (152b) ................................... 180 Table 10: Fragmentation of VAT 2 Using 1.1 Equivalents of Various Phosphonate Derived Nucleophiles .......................................... 182

Page 8: Florida State University Libraries - FLVC

vii

LIST OF FIGURES

Figure 1: Representative examples of the (1) Diels-Alder, (2) Michael Addition, (3) Evans Aldol, (4) and Sonogashira Cross Coupling Reactions in Synthesis ............................................................... 2 Figure 2: Examples of Tandem Bond Forming / Bond Breaking Strategies in Organic Synthesis .................................................................. 3 Figure 3: Examples of Transition Metal Catalyzed C-C Bond Cleavage Reactions in Synthesis ............................................................... 4

Figure 4: Possible Mechanistic Pathways of Grob Fragmentations .......... 6 Figure 5: General Representation of the Wharton Fragmentation ............ 6

Figure 6: Wood and Njardarson’s Wharton fragmentation approach to CP-263,114 ................................................................................ 7

Figure 7: Base Promoted Eschenmoser-Tanabe Fragmentation Process. ..................................................................................... 8

Figure 8: Mander’s Reduction-Epoxidation-Oxidation Solution for the Eschenmoser-Tanabe Fragmentation in the Synthesis of GB 13 ......................................................................................... 9

Figure 9: Comparison of the Eschenmoser-Tanabe Fragmentation and Enone Formation from Vinylogous Acid Esters .......................... 9

Figure 10: Mechanistic Hypothesis for the Fragmentation of Vinylogous Acyl Triflates .......................................................... 12

Figure 11: (a)64 A Male Specimen of the Douglas-Fir Tussock Moth; (b)64 Distribution of Host Type Where Douglas-Fir Tussock Moth Has Been Found; (c) the DFTM Sex Pheromone..................... 13

Figure 12: Smith’s Synthesis of the Sex Pheromone of the Douglas-Fir Tussock Moth ........................................................................... 17

Figure 13: Fetizon and Lazare’s Synthesis of Z6 ...................................... 17

Figure 14: Kocienski and Cernigliaro’s Synthesis of Moth Pheromone Z6 18

Page 9: Florida State University Libraries - FLVC

viii

Figure 15: Synthesis of (Z)-6-Heneicosen-11-one Using an ABC Strategy 21 Figure 16: Illudalane Skeleton and Alcyopterosin A .................................. 38

Figure 17: Proposed Biosynthetic Pathway to the Illudalanes. ................. 38 Figure 18: (eq. 1) Proposed Decomposition of Stearodelicone (21) Upon Absorption on Silica Gel, and (eq. 2) the Observed Reactions of Ptaquiloside (23) in the Presence of Acid and/or Base ........ 40 Figure 19: Representative Sample of Illudalane Structures Isolated from A. paessleri and A. grandis ...................................................... 41 Figure 20: Possible Traditional Reppe Reaction Involving Three Different Unsymmetrical Acetylenes ....................................................... 42 Figure 21: Typical Solutions for Chemo- and Regioselective Cyclotrimerization of Alkynes ................................................... 43 Figure 22: Representative Examples of Sato’s One-Pot Metalative Reppe Reactions ...................................................................... 44 Figure 23: Sato’s Synthesis of Alcyopterosin A ........................................ 44 Figure 24: Key Steps in the Syntheses of Alcyopterosin E (28) (eq. 1) and Alcyopterosin I (30) by Witulski and Snyder (eq. 2).. ......... 46 Figure 25: Synthesis of Iglesias’ Key Intermediate ................................... 47 Figure 26: Synthesis of Unnatural Alcyopterosin Analogs Performed by Iglesias et al ............................................................................. 48 Figure 27: Completion of Iglesias’ Synthesis of Alcyopterosin A. ............. 49 Figure 28: Compounds Known to Intercalate DNA. .................................. 50 Figure 29: Retrosynthetic Analysis of Alcyopterosin A Using a Fragmentation / Benzannulation Approach .............................. 54 Figure 30: AuCl3- and Cu(OTf)2-Catalyzed [4+2] Benzannulation Reactions Described by Asao and Yamamoto. ........................ 55 Figure 31: Proposed Mechanism of the [4+2] Cycloaddition Reactions of 56 and 57 Catalyzed by AuCl3 and Cu(OTf)2 / CF2HCO2H… 56

Page 10: Florida State University Libraries - FLVC

ix

Figure 32: Intramolecular Lewis Acid-Catalyzed [4+2] Benzannulation Reactions Studied by Asao and Yamamoto ............................. 57 Figure 33: Yamamoto’s Key Benzannulation in the Synthesis of (+)- Rubiginone B2 and (+)-Ochromycinone .................................... 57 Figure 34: Contrast Between Known Benzannulations and Desired Benzannulation ........................................................................ 58 Figure 35: Comparison of Known Benzannulations and Those of a New Methodology............................................................................. 60 Figure 36: Originally Considered Reactions to Access Fragmentation Pre-nucleophile 77 ................................................................... 61 Figure 37: Proposed Route to Benzannulation Substrates 84. ................. 62 Figure 38: Synthesis and Fragmentation Reaction of Aryltriazene 80. ..... 62 Figure 39: Synthesis of Benzannulation Substrates 84a-e. ...................... 63 Figure 40: Synthesis of Benzannulation Substrate 84f. ............................ 64 Figure 41: Direct Comparison of the Benzannulation Reactions of 84c and 84f ..................................................................................... 66 Figure 42: Alternative Synthesis of Benzannulation Substrate 89 ............ 67 Figure 43: Benzannulation Reactions of Compound 89 ............................ 68 Figure 44: Proposed Route to Vinyl Nucleophile 54 Using Negishi’s Z-Selective Bromoboration ....................................................... 69 Figure 45: Several Chemotherapeutic Agents Used in the Treatment of Melanoma ................................................................................ 120 Figure 46: The Report Issued to Baker from the National Cancer Institute’s 60-Cell Line Panel Toxicity Assay for Palmerolide A181 ....................................................................... 123 Figure 47: Palmerolide A and Other Members of the Palmerolide Natural Products with Major Distinctions Highlighted in Red Ovals ................................................................................ 124 Figure 48: Palmerolide A and Strategic Disconnections. .......................... 125

Page 11: Florida State University Libraries - FLVC

x

Figure 49: De Brabander’s Synthesis of the C16-C24 Fragment of . Palmerolide A ........................................................................... 126 Figure 50: De Brabander’s Synthesis of the C9-C15 Fragment of Palmerolide A ........................................................................... 126 Figure 51: De Brabander’s Synthesis of the C1-C8 Fragment of Palmerolide A ........................................................................... 127 Figure 52: Completion of De Brabander’s Synthesis of 109, a Diastereomer of Palmerolide A ................................................ 128 Figure 53: Synthesis of Nicolaou’s C16-C23 (112) and C8-C15 (116) Fragments ................................................................................ 129 Figure 54: Nicolaou’s Synthesis of C1-C8 Fragment of Palmerolide A ..... 130 Figure 55: Nicolaou’s End-Game Strategy for the Synthesis of 109 ......... 131 Figure 56: Key Analogs of Palmerolide A Developed by the Nicolaou Lab ........................................................................................... 132 Figure 57: Key Reactions in Maier’s Formal Synthesis of Palmerolide A .. 135 Figure 58: Hall’s Asymmetric Crotylboration en Route to Palmerolide A’s C16-C24 Fragment .................................................................. 136 Figure 59: Hall’s Unique Approach to Install the C7, C10, and C11 Stereocenters ........................................................................... 137 Figure 60: Brief Overview of the Addition / Bond Cleavage Reactions of Vinylogous Acyl Triflates .......................................................... 139 Figure 61: Synthesis and Nucleophile-Triggered Decompositions of DHP Triflates .................................................................................... 140 Figure 62: Retrosynthetic Analysis of Palmerolide A Using a Fragmentation Approach .......................................................... 140 Figure 63: Synthesis of the C1-C8 Olefination Reagent for the Synthesis of Palmerolide A ....................................................................... 142 Figure 64: Synthesis of Aldehyde 150, the C9-C15 Fragment of Palmerolide A ........................................................................... 143 Figure 65: Possible Michael Reaction of 155 ............................................ 144

Page 12: Florida State University Libraries - FLVC

xi

Figure 66: Completion of the Eastern Hemisphere (C1-C15) Fragment Synthesis ................................................................................. 144 Figure 67: Mechanism of the Classical Claisen Condensation of Ethyl Acetate ..................................................................................... 170 Figure 68: Proposed Mechanism for the Reaction Between 2 and 152a .. 171 Figure 69: Reported Claisen-Type Condensation Reactions of VAT 2 ..... 172 Figure 70: Observations Made During the Synthesis of the C1-C15 Fragment of Palmerolide A (Chapter 4) ................................... 173 Figure 71: Proposed Fragmentation Reaction Pathway Between 2 and 166 .................................................................................... 175 Figure 72: Proposed Mechanism of the Reaction Between VAT 2 and 152 ........................................................................................... 176 Figure 73: Common Methods for the Preparation of Phosphonates ......... 178 Figure 74: Synthesis of 180, an Analog of Phosphonate 178 ................... 181

Page 13: Florida State University Libraries - FLVC

xii

LIST OF ABBREVIATIONS

ABC addition / C-C bond cleavage Ac acetyl app apparent (spectral) Aq aqueous Ar aryl, argon BAIB bis(acetoxy)iodobenzene (phenyliodonium diacetate) BINAP 2,2’-bis(diphenylphosphino)-1,1’-binaphthyl Bn benzyl BOLD bleomycin, vincristine, lomustine, and dacarbazine Bt. Bacillus thuringiensis n-Bu normal butyl t-Bu tertiary butyl c centi oC degrees Celsius ca. circa (approximately) Calcd calculated (in mass spectrometry) CBS Corey-Bakshi-Shibata reagent CD circular dichroism cf. confer (compare) CI chemical ionization (in mass spectrometry) CNS central nervous system

Page 14: Florida State University Libraries - FLVC

xiii

cod 1,5-cyclooctadiene CSA camphor-10-sulfonic acid d doublet (spectral) heat, double bond location chemical shift, in parts per million relative to tetramethylsilane dba dibenzylideneacetone DBU 1,8-diazabicylco[5.4.0]undec-7-ene DCE 1,2-dichloroethane DCM dichloromethane DDQ 2,3-dichloro-5,6-dicyano-1,4-benzoquinone de diastereomeric excess DEAD diethyl azodicarboxylate DHP 5,6-dihydro-2-pyrone DIBAL diisobutylaluminum hydride DIPEA diisopropylethylamine DMAP N,N-4-dimethylaminopyridine DMP Dess-Martin periodinane DMSO dimethylsulfoxide DNA deoxyribonucleic acid dr diastereomeric ratio DTFM Douglas-fir tussock moth DTIC dacarbazine E- entgegen or opposite (alkene geometry)

Page 15: Florida State University Libraries - FLVC

xiv

EDC-Cl 1-ethyl-(3-dimethylaminopropyl)carbodiimide hydrochloride ee enantiomeric excess e.g. exempli gratia (for example) eq equation EI electron ionization (in mass spectrometry) EPA United States Environmental Protection Agency equiv equivalent(s) ESI electrospray ionization (in mass spectrometry) Et ethyl et al. et alii (and the others) EWG electron withdrawing group FAB fast-atom bombardment (in mass spectrometry) FT-IR Fourier-transformed infrared g gram(s) gem- geminal GI50 half maximal growth inhibitory concentration h hour(s) ha hectares Hex hexanes HIV human immuno-deficiency virus HPLC high-performance liquid chromatography HRMS high-resolution mass spectrometry HWE Horner-Wadsworth-Emmons

Page 16: Florida State University Libraries - FLVC

xv

Hz hertz IC50 half maximal inhibitory concentration i.e. id est (that is) Ipc isopinocamphenyl IR infrared J coupling constant reported in hertz (in NMR spectroscopy) wavelength L liter(s) LC50 median lethal dose LDA lithium diisopropylamide LiHMDS lithium bis(trimethylsilyl)amide micro m multiplet (spectral), meter(s), milli m- meta- M moles per liter, mega mCPBA m-chloroperbenzoic acid Me methyl MG-MID meangraph midpoint min minute(s) MOM methoxymethyl mp melting point Ms methanesulfonyl n nano

Page 17: Florida State University Libraries - FLVC

xvi

NCI United States National Cancer Institute NMR nuclear magnetic resonance Nuc nucleophile OPP pyrophosphate p- para- PCC pyridinium chlorochromate Ph phenyl Pin pinacolato PMB p-methoxybenzyl ppm parts per million ppt precipitate PPTS pyridinium p-toluenesulfonate i-Pr isopropyl q quartet (spectral) RCM ring-closing metathesis ref reference retro retrograde r.t. room temperature s singlet (spectral) SAR structure-activity relationship SN2 substitution nucleophilic bimolecular t triplet (spectral) TBAF tetrabutylammonium fluoride

Page 18: Florida State University Libraries - FLVC

xvii

TBDPS tert-butyldiphenylsilyl TBS tert-butyldimethylsilyl TEMPO 2,2,6,6-tetramethyl-1-piperidinyloxy free radical TES triethylsilyl Tf trifluoromethanesulfonyl TFA trifluoroacetic acid TGI total growth inhibitory concentration THF tetrahydrofuran TIPS triisopropylsilyl TMS trimethylsilyl TMZ temozolomide Tol tolyl Ts p-toluenesulfonyl UV ultraviolet VAT vinylogous acyl triflate V-ATPases vacuolar adenosine triphosphatases wt. weight Z- zasammen or together (alkene geometry)

Page 19: Florida State University Libraries - FLVC

xviii

ABSTRACT

This dissertation describes the synthetic utility of tandem addition / C-C bond

cleavage reactions of vinylogous acyl triflates. The first chapter provides background

into carbon-carbon bond breaking reactions that have been applied in organic synthesis

and the preliminary data that allowed for the original work presented here. Chapter 2

explains the significance as well as the prior syntheses of a commercially important

moth pheromone, (Z)-6-heneicosen-11-one. The second chapter culminates in the

synthesis of the sex attractant through a fragmentation reaction made possible by the

direct extension of the initial nucleophile-triggered fragmentation studies to include the

use of Grignard reagents. Chapter 3 describes the application of the fragmentation

method, coupled to a benzannulation reaction, to afford penta- and hexasubstituted

indanes. This two step sequence provides the basis for future work directed toward the

synthesis of alcyopterosin A, a known cytotoxic agent with possible biological

applications.

The current difficulties pertaining to the treatment of melanoma are discussed in

Chapter 4. Recently, an exciting natural product that provides promising activity against

this horrible cancer was discovered. Palmerolide A has the ability to kill melanoma cells

selectively at low concentrations. The fragmentation method developed in these

laboratories provides entry into a key fragment. The Claisen-type condensation reaction

of vinylogous acyl triflates was expanded to the synthesis of a novel -ketophosphine

oxide olefinating reagent, which allowed for the rapid synthesis of the eastern

hemisphere (C1-C15) of this exciting natural product. Optimization of the Claisen-type

condensation reaction to provide the -ketophosphine oxide reagent, led to the optimal

reduction of the number of equivalents of the nucleophile. Intrigued by this, these

reactions were explored in more detail. The results of this investigation are described in

Chapter 5. The reduction in the number of equivalents of nucleophile, a key feature in

these reactions, may be attributed to the ability of the phosphorus atom to form of an

oxaphosphetane-like intermediate. As a result, new, potentially useful, -

ketophosphonates were synthesized.

Page 20: Florida State University Libraries - FLVC

1

CHAPTER 1

INTRODUCTION: C-C BOND CLEAVAGE AND FRAGMENTATION REACTIONS IN ORGANIC SYNTHESIS

Synthetic organic chemistry has largely focused on the use of carbon-carbon

bond forming reactions to assemble complex molecules. The means to install such

bonds is of the utmost importance. There is a constant struggle to provide new carbon-

carbon bond forming reactions that are tolerant to a diverse number of functional

groups, as well as reactions that are both regio- and stereoselective. Discoveries of

such reactions constantly expand the frontiers of organic chemistry. Tolerant and

selective C-C bond forming reactions, such as the Diels-Alder,1-4 Michael addition,5-8

Evans aldol,9 and Sonogashira10-12 reactions were at the forefront of chemistry at the

time of their discovery. These reactions have since been applied in the synthesis of

numerous complex molecules (Figure 1). If not for the innovation of such reactions, the

synthesis of many natural products would have proven to be a much more daunting

challenge; they have changed the way chemists have approached natural product

synthesis and have allowed the development of synthetic strategies which would have

otherwise been impossible.

The need to build up complexity quickly in synthesis requires bond forming

reactions. Not surprisingly, C-C bond breaking reactions receive far less attention.

However, these reactions often provide access to compounds that can be difficult to

prepare through other methods. Some of the most useful C-C bond breaking reactions

applied in organic synthesis are simply the reverse processes of C-C bond forming

events similar to those mentioned above (the aldol,13,14 Diels-Alder,15,16 and Michael

reactions,17,18 among others).

Page 21: Florida State University Libraries - FLVC

2

N

OO

MeO OMeO

O

(2.5 equiv)

r.t., 45 min97%

N

OO

MeO OMe

O

OH

H

Diels-Alder Reaction From Boger's Synthesis of Rubrolone Aglycon19

(1)

Tandem Michael-Addition Reactions From Ihara's Synthesis of ( )-Longiborneol20±

CO2Me

O LiHMDS(2 equiv)

THF

-78 oC, 1h,

then

0 oC, 3h

94%

CO2Me

O O

CO2Me

(2)

NO

O

Ph Me

O

1. Bu2BOTf, Et3N,

-5 oC, DCM; then add

O

H

12

2. MeOH, 30% H2O263%, 2 steps

NO

O

Ph Me

O OH

12

(3)

Sonogashira Reaction in Paterson's Synthesis of Callipeltoside aglycon22

(4)

O

O

OTBS

Me

MeO

I

O

Me Cl

HMe

MeOH +

O

O

OTBS

Me

MeO O

Me

Me

MeOH

Cl

1. Pd(Ph3P)2Cl2, CuI, HN(i-Pr)2, EtOAc

2. TBAF, THF3. PPTS, CH3CN, H2O 54% over 3 steps

Callipeltoside aglycon

3

Figure 1: Representative examples of the (1) Diels-Alder, (2) Michael Addition, (3)

Evans Aldol, and (4) Sonogashira Cross Coupling Reactions in Synthesis.

Often, reverse reactions are used in tandem with their forward counterparts to

access complex molecules. Figure 2 provides some representative examples that

Page 22: Florida State University Libraries - FLVC

3

demonstrate the utility of retrograde reactions in organic synthesis. Jacobi and co-

workers have utilized a Diels-Alder / retrograde Diels-Alder sequence to access (±)-

Petasalbine (scheme 1).23 Jacobi took advantage of the reactivity of oxazoles as diene

partners; after the cycloaddition reaction with a tethered alkyne, the heterocyclic

intermediate underwent a retro-Diels-Alder to afford the required furan moiety. In 2005,

Iwabuchi and co-workers synthesized cannabinoid receptor agonist (-)-CP55,940 using

a modified-proline catalyzed aldol reaction to achieve stereocontrol, followed by a retro-

aldol to generate the chiral cyclohexane carboskeleton (scheme 2).24

MeMe

H

OH

Me

N

O N

O

Me

MeOHH

-HCN

O

MeMe OHMe

H

Jacobi's Key Diels-Alder/Retro-Diels-Alder Reaction in the Synthesis of ( )-Petasalbine±

(1)

Diels-Alder

Retro-Diels-Alder

Iwabuchi's Aldol/Retro-Aldol Strategy in the Synthesis of (-)-CP55,940

O

CHO

NH

TBDPSO

CO2H

MeCN, rt., 68%(>99% de, 94% ee)

N

PO

OO

H O H

O OH

( )-Petasalbine±

O OMOM

cat. TsOHethylene glycol

xylene,reflux

O O

OO

OMe OMe

n-C6H13

n-C6H13

OH

n-C6H13

HO

OH

(-)-CP55,940

(2)

84%

68%2 steps

88%

49%3 steps

Figure 2: Examples of Tandem Bond Forming / Bond Breaking Strategies in Organic Synthesis.

Page 23: Florida State University Libraries - FLVC

4

Reactions such as the Cope rearrangement,25,26 as well as oxidative cleavages

of olefins27 and diols,28 represent some traditional C-C bond cleavage reactions.

Several new C-C bond breaking reactions have been made available through the

advance of transition metal chemistry. Although transition metal-catalyzed C-C bond

cleavage chemistry has made some headway in synthetic chemistry, many of these

reactions are heavily dependent on the presence of either highly strained bonds (e.g.

cyclopropane or cyclobutane moieties) or functional groups located about the reaction

site capable of coordinating to the metal center (Figure 3, scheme 1).29-33 The evolution

of metathesis catalysts has allowed for the development of ring opening metathesis

reactions, yet another defining example of C-C bond cleavage reactions in synthetic

chemistry (Figure 3, scheme 2).34-36

RR' O

OH

[Rh(OH)(COD)]2

(R)-BINAPToluene

up to 92% and 95% ee

R

O ORh

R'

O

R

R'

Rh

-carbon elimination

O

1,4-Rh shift

Rh

O

R

R'

O

"H+"

O

R

R'

O

Murakami's Asymmetric Rhodium Catalyzed Synthesis of 3,4-Dihydrocoumarins Through Cleavage of a

Cyclobutyl Intermediate37

(1)

Tandem Ring opening/Ring Closing Metathesis Strategy in Phillips' Synthesis of ( )-trans-Kumausyne38

(2)O

O Ru

PPh3

PPh3

PhClCl

CH2Cl2, H2C=CH2, r.t.83% O

O

±

H

H

O

BrHH

AcO

( )-trans-Kumausyne±

Figure 3: Examples of Transition Metal Catalyzed C-C Bond Cleavage Reactions in Synthesis.

Page 24: Florida State University Libraries - FLVC

5

Throughout the 1950’s and 60’s, Grob and co-workers carried out investigations

into heterolytic bond cleavage reactions of molecules consisting of various combinations

of carbons and heteroatoms.39-43 These reactions produce three distinct fragments /

products, and are thus referred to as Grob fragmentations. The three ―products‖

generated from the fragmentation are all included in the starting molecule with the

general formula a—b—c—d—X (Figure 4). ―X‖ is referred to as the nucleofuge; leaving

with the electron pair with which it was originally attached to the starting molecule, thus

it becomes more negative. Prior to fragmentation, the nucleofugal fragment can be

neutral (e.g. halide, sulfonate, or carboxylate) or charged (diazonium, oxonium,

ammonium or sulfonium). The electrofuge, a—b, loses a bonding pair of electrons and

becomes more positive. The electrofugal fragment is typically a carbonyl containing

compound; however, carbon dioxide, olefins, dinitrogen, immonium-, carbonium-, and

acylium ions have been generated as electrofuges. The central portion of the starting

material, c—d, becomes the unsaturated fragment. The most commonly encountered

unsaturated fragments are olefins, acetylenes, nitriles and imines.

The most probable mechanistic pathway (Figure 4) of the Grob fragmentation is

substrate dependent. Both steric and electronic properties of the substrate influence the

nature by which the fragmentation takes place. Very narrow stereochemical

requirements must be met in order to achieve proper orbital overlap for the one-step

synchronous (concerted) mechanism to proceed. The transition state of the concerted

process involves all five atoms, and thus, this mechanism is invoked usually in Grob

fragmentations of conformationally rigid molecules. If necessary orbital overlap is

insufficient or absent, the concerted process is not possible; in this case, a two-step

process (usually cationic) must take place if the fragmentation is to occur. A two-step

fragmentation pathway typically provides the possibility for side reactions (e.g.

elimination), making fragmentations that proceed through stepwise mechanisms less

useful.

Page 25: Florida State University Libraries - FLVC

6

ab

cd

X

a b c d XElectrofugal

fragmentUnsaturated

fragmentNucleofugal

fragment

(A) One-step synchronous:

(B) Two-step cationic:

ab

cd

XX-

ab

cd a b

+ +

c d+

Electrofuge Nucleofuge

(C) Two-step anionic:

ab

cd

XX- a b c d +c

dX

Figure 4: Possible Mechanistic Pathways of Grob Fragmentations.

P. S. Wharton pioneered the base-induced heterolytic fragmentation reaction of

bicyclic-1,3-diol monosulfonate esters, now referred to as a Wharton fragmentation

(Figure 5).44-47 Although the Wharton fragmentation falls into the category of a Grob

fragmentation, it is a more specific term referring to the synthesis of alkenes from 1,3-

diols. The most common substrates for the Wharton fragmentation are bicyclic-1,3-

hydroxy monotosylates or monomesylates generated from unsymmetrical 1,3-diols.

n

OH

OSO2R

n

base

O

Figure 5: General Representation of the Wharton Fragmentation.

The Wharton fragmentation is often employed for the synthesis of medium sized

rings which are difficult to prepare. The rate of fragmentation depends both on the ring

Page 26: Florida State University Libraries - FLVC

7

strain of the bicycle and the concentration of the base. Typically strong, non-

nucleophilic, bases (t-BuOK, NaH, dimsylsodium, etc.) are best for promoting the

fragmentation. Alkenes from the Wharton fragmentation are generated

stereospecifically from the bicyclic precursor. Wood and Njardarson successfully

applied the Wharton fragmentation in their approach to the bicyclic core of CP-263,114

(Figure 6).48 The synthetic strategy outlined by Wood highlights the utility of the Wharton

fragmentation, as the originally envisioned oxy-Cope rearrangement failed.

AcO

OH

MsCl, pyrDMAP

AcO

OMs

K2CO3, MeOH

Me Mer.t.

95%2 steps

AcO

Me

Figure 6: Wood and Njardarson’s Wharton fragmentation approach to CP-263,114.

During the time Grob was describing the fragmentation reactions that now bear

his name, Eschenmoser49,50 and Tanabe51,52 were independently exploring the ring

opening reaction of ,-epoxyhydrazones. The Eschenmoser-Tanabe fragmentation

process (Figure 7) is classified as a 7-centered Grob-type fragmentation process,

yielding an electrofugal fragment (ketone) tethered to the unsaturated fragment (alkyne)

and two nucleofugal fragments (N2 and typically an arylsulfinate).

Page 27: Florida State University Libraries - FLVC

8

O

O

R

R'

TsNHNH2

-H2O

N

O

R

R'

N

H

TsBase

N

O

R

R'

NTs

N

R

R'

NTs

O

O

R'

R

+N NTs +

Nucleofugal fragments

Unsaturated fragment

Nucleofugalfragment

Figure 7: Base Promoted Eschenmoser-Tanabe Fragmentation Process.

The substrates of the Eschenmoser-Tanabe fragmentation are typically prepared

in a multistep sequence from ,-unsaturated ketones; first through the epoxidation of a

cyclic enone, followed by a condensation reaction with tosyl hydrazide. The

fragmentation is induced by treatment with acid or base in a protic medium induces

fragmentation. Although the multistep sequence from cyclic enone to tethered keto-

alkyne has found some application as a synthetic strategy through the years,53-

58,79,80,86,114,117 it remains largely pedagogical. The substrates required for the

Eschenmoser-Tanabe process, epoxy hydrazones, can be difficult to prepare, as

illustrated by Mander’s synthesis of the Galbulimima alkaloid GB 13.55 Direct

epoxidation of the enone of the pentacyclic late-stage intermediate was unsuccessful. In

an effort to obtain the necessary epoxy hydrazone, a reduction-epoxidation-oxidation

sequence was performed (Figure 8). The possible difficulty in the synthesis of epoxy

hydrazones and the protic medium (commonly ethanol or acetic acid) present potential

drawbacks to the method.

Page 28: Florida State University Libraries - FLVC

9

H

H

H

OMOM

H

O

MOMO

1. LiAlH4, THF2. mCPBA, DCM

3. DMP, NaHCO3;

4. p-NO2ArSO2NHNH2,pyridine, EtOH, THF

59% 4 steps

O

H

H

H

OMOM

H

MOMO

Figure 8: Mander’s Reduction-Epoxidation-Oxidation Solution for the Eschenmoser-

Tanabe Fragmentation in the Synthesis of GB 13.

Prior to the description of the Eschenmoser-Tanabe fragmentation process,

Woods and Tucker described the reaction of vinylogous acid esters with

phenylmagnesium bromide, providing cyclic enones.59 This method has been utilized in

cyclic enones that are difficult to prepare using other methods.60 There is a marked

similarity between the presumed intermediates of the Eschenmoser-Tanabe

fragmentation and the synthesis of enones from vinylogous acid esters (Figure 9).

Although there is a parallel between the intermediates A and B, they diverge in the

manner by which they decompose.

NNHTs

OR'

R

R' OH

NNTs

R

A

- N2

- TsH

R'

O R

The Eschenmoser-Tanabe Fragmentation

Enone Formation from Vinlogous Acid Esters

OR2

R1

R3 OM

OR2

R1

B

- R2OH

R3O

R3 M H3O+R1

O

Figure 9: Comparison of the Eschenmoser-Tanabe Fragmentation and Enone Formation from Vinylogous Acid Esters.

Page 29: Florida State University Libraries - FLVC

10

In 2005, our lab sought to introduce an intermediate similar to B which would

allow for a fragmentation similar to the Eschenmoser-Tanabe fragmentation under mild

conditions in an aprotic solvent. Such a reaction would have important mechanistic

implications and would provide a new tool in the synthesis of complex molecules. A

crossover in the mechanistic pathway was envisioned to occur if the nucleofugacity of

the –OR2 group in intermediate B were increased.

Kamijo and Dudley carried out a preliminary investigation into a tandem

carbanion addition / C-C bond cleavage reaction that provided tethered alkynyl ketones

that are similar, yet regioisomeric, to those obtained by the Eschenmoser-Tanabe

fragmentation.61 A change in the –OR2 group from alkoxy (enone formation, Figure 9)

to trifluoromethanesulfonyloxy allowed for the desired crossover mechanism to take

place both in an aprotic medium and under mild conditions (displayed in Table 1).

Kamijo and Dudley found that the synthesis of vinylogous acyl triflates (2) was

general and high yielding. Symmetric diketones such as 1 were converted into

vinylogous acyl triflates (VATs), similar to 2, in nearly quantitative yields using a

modified procedure.62 The fragmentation reaction was optimized for the addition of

phenylmagnesium bromide, and ethereal solvents were found to provide the most

suitable environment for the fragmentation. Table 1 summarizes the original scope of

the fragmentation reaction with respect to nucleophiles explored by Kamijo and Dudley.

Nucleophiles with electron donating groups had significant effect and accelerated the C-

C bond cleavage process (entries 1—4 vs. entries 5—6), suggesting a transition state

with significant carbonyl character. Aryl organolithium reagents were also found to

trigger fragmentation more readily, presumably due to an increase in ionic character of

the alkoxide intermediate (entries 7—9).

Page 30: Florida State University Libraries - FLVC

11

Table 1: Scope of Original Fragmentation Reaction with Respect to Nucleophiles.61,a

Me

O

O 1.2 equiv Tf2O2.0 equiv pyridine

CH2Cl2, 95-100%

O

OTf

Me

1 2

R1 M

THF

R1

O Me

3

entry R1__M conditions 3 yield (%)b

1 Ph—MgBr 0 oC to r.t. 3a 80c

2 p-MeO—C6H4—MgBr 0 oC to r.t. 3b 86

3 m-MeO—C6H4—MgBr 0 oC to r.t. 3c 57

4 o-MeO—C6H4—MgBr 0 oC to r.t. 3d 34

5 p-Cl—C6H4—MgBr 0 — 60 oC 3e 61

6 2-thienyl—MgBr 0 — 60 oC 3f 63

7 Ph—Li -78 oC to r.t. 3a 93c

8 m-MeO—C6H4—Li -78 oC to r.t. 3g 78

9 o-MeO—C6H4—Li -78 oC to r.t. 3h 57

10 Me—Li -78 oC to r.t. 3i 65 a

Typical procedure: enol triflate 2 (0.55 mmol) in 2 mL cold THF was treated with R1—M (0.50 mmol). All

reactions complete within 90 min. b Isolated yield.

c Average of two runs.

The mechanistic hypothesis (Figure 10) that guided Kamijo and Dudley’s original

studies has many interesting qualities as well as some guiding assumptions: (1)

nucleophilic addition is fast and proceeds in a 1,2- fashion; (2) decomposition of

intermediate C is the rate limiting step; (3) lithium triflate is extruded from C as a

dissociated ion pair that subsequently recombines;63 (4) an increase in the ionic

character of C promotes fragmentation; and (5) the stability of the resulting alkynyl

ketone and the dissociated ion pair are reflected in the transition-state (concerted);

however a two-step mechanism cannot be ruled out.

Page 31: Florida State University Libraries - FLVC

12

O

Me

OTf

R1 Li

THF

R1 OMe

OTf

Li THFn

C2

O

R1

Me

3fast

- LiOTf

slow

Figure 10: Mechanistic Hypothesis for the Fragmentation of Vinylogous Acyl Triflates.61

The fragmentation was found to be general with respect to the VAT, affording

alkynyl ketones of varying tether lengths and substitution patterns. Having established a

nucleophile-triggered fragmentation pathway of vinylogous acyl triflates under mild

reaction conditions, the Dudley lab directed further efforts towards expanding the scope

of the fragmentation reaction. This dissertation is focused on the development of this

method as well as its uses as a strategy for obtaining complex intermediates capable for

application in the synthesis of natural products. Since the discovery of this new reaction,

we have demonstrated the value of vinylogous acyl triflates as useful tools in complex

molecule synthesis.

Page 32: Florida State University Libraries - FLVC

13

CHAPTER 2

SYNTHESIS OF (Z)-6-HENEICOSEN-11-ONE: THE SEX PHEROMONE OF THE DOUGLAS-FIR TUSSOCK MOTH

The Douglas-Fir Tussock Moth

The Douglas-fir tussock moth (DFTM), Orgyia pseudotsugata seen in Figure 11a,

is a major contributor to the defoliation of fir trees in the Pacific Northwest (Figure 11b).

The populations of the DFTM typically remain stable; however they can explode,

leading to significant defoliation.64 For instance, in 1974 a DFTM outbreak gave rise to

the defoliation of 279,000 hectares (ha) of forest. The Environmental Protection Agency

(EPA) allowed the use of DDT, an otherwise banned substance, on 161,000 ha of forest

in order to contain the outbreak.65 The discovery of the sex pheromone (Figure 11c) of

the DFTM in 1975,66 (Z)-6-Heneicosen-11-one (Z6) has played an integral role in the

defense against such outbreaks.

Figure 11: (a)64 A Male Specimen of the Douglas-Fir Tussock Moth; (b) 64 Distribution of

Host Type Where Douglas-Fir Tussock Moth Has Been Found; (c) The DFTM sex pheromone.

O

8

Z6

(a) (b)

(c)

Page 33: Florida State University Libraries - FLVC

14

Outbreaks in the population of the DFTM are typically short in duration, one to

two years. The defoliation caused by outbreaks of the DFTM may result in complete

tree death or in the top-kill of trees, which retards vegetation growth and may induce

susceptibility of the tree to other pests. The defoliation of forestland caused by the

DFTM also increases the risk and severity of forest fires. The preferred food source for

the DFTM varies regionally, however the Douglas-fir is the dominant food source in

most areas where they are found.67 The caterpillar larvae of the DFTM are the source of

the defoliation. They are incapable of flight and are limited to the environment of the

host tree. Newly hatched larvae feed on the current year’s foliage, as the larvae

continue to grow, their demand for food increases and both new and old vegetation is

consumed.68

After consuming copious amounts of vegetation, the larvae build their cocoon

and pupation begins. Female moths emerge from their cocoon approximately 2 weeks

later and mate soon after. Being unable to fly, the female moth is limited to the use of

chemical communication in the form of pheromones to attract sexual partners. During

the daylight hours of the male flight season, usually in the months from July to

November, the females release their sex pheromone to signal potential mates. The

females lay their eggs soon after mating and subsequently die.69

There are many natural controls by which the population of the DFTM is

regulated. Eggs are preyed upon by small birds and parasitized by small wasp species.

After hatching, the caterpillars are eaten by various predators such as birds, spiders and

other insects. Carcelia yalensis, a parasitic fly species, is one of the primary foes of the

DFTM larvae, laying eggs inside of the caterpillar, which then hatch and eat the

caterpillar from within.70 When moth densities approach outbreak levels, there is a

nuclear polyhedrosis virus that frequently infects many colonies of the moth. Once

infected, a moth’s internal organs liquefy. The virus is spread throughout the colony

when a diseased body ruptures and is spread on the surface of the vegetation, and is

later ingested by other members of the species. Routinely, the virus is fatal and

commonly spreads rampantly throughout the colony, thus resulting in outbreak

suppression.71

Page 34: Florida State University Libraries - FLVC

15

When the natural means by which the DFTM populations are regulated become

insufficient, outbreaks, and subsequent tree damage, may result. A well integrated

management program must be maintained in order to handle population outbreaks and

minimize destructive defoliation. The early detection of increasing populations is the

foundation of any management program. Because the DFTM population produces only

one generation per year, it is possible for outbreaks to be detected one to two years

prior to any significant defoliation. Early detection of population outbreaks is made

possible, primarily, through the annual monitoring of male populations. The males can

be lured into traps baited with the sex attractant (Z6, Figure 11c) of their female

counterparts, allowing for sampling to be performed.72

When an outbreak is perceived to be eminent, measures to suppress moth

populations are determined through careful analysis of the potential threat to the forest.

Most recently, biological insecticides have emerged as the preferred method for

suppressing populations of the DFTM. Biological insecticides are regarded as

environmentally benign, making them preferred over persistent chemical based

alternatives. The two most common biological agents used to collapse populations of

the DFTM are: Bacillus thuringiensis (Bt), marketed under several trade names (e.g.

ThuricideTM from Bonide Products, Inc.), as well as the aforementioned tussock moth

nucleopolyhedrosis virus (TM-Biocontrol-1, produced by the U.S. Forestry Service).73

These agents are very successful in decreasing the population of feeding larvae,

however they are only used once outbreak population levels have been reached. As a

result, significant defoliation remains possible.

Pheromones have been used as species selective management agents.74 The

sex pheromone of the DFTM offers a potentially new means of controlling moth

populations at pre-outbreak levels through mating disruption.73,75,76 By spraying

synthetic Z6, impregnated in controlled-release capsules, male moths become confused

and unable to chemo-locate their female mating partners. By disrupting the mating

habits of the DFTM, reductions in the number of caterpillars in the following year are

likely. In 2005, the EPA has registered the use of Hercon® laminated plastic bio-flake

formulation of Z6 for the control of tussock moths and other lepidopteran insects.77

Page 35: Florida State University Libraries - FLVC

16

Continued research will assist in providing the necessary data and determine the

efficacy of Z6 as a mating disruption agent suitable for wide spread use.

Synthesis of (Z)-6-Heneiosen-11-one

Since its isolation and characterization by Smith, Daterman, and Davies in

1975,66 (Z)-6-heneicosen-11-one has arguably been the most important factor in the

fight against severe defoliation by the DFTM. The use of Z6 in baited traps, allowing for

population analysis and outbreak detection, and its potential for mating disruption lends

credence to its commercial importance. There have been considerable efforts directed

towards the synthesis of the DTFM sex pheromone.78-92 Most synthetic approaches to

Z6 rely on one of two strategies: (1) elaboration of the moth pheromone through a

series of steps that piece together the carbon back bone originating with the C11

carbonyl / protected-carbonyl through carbon-carbon bond forming reactions like the

SN2 reaction, or (2) beginning with a cyclic starting material and performing a ring-

opening event to install the necessary carbons.

The first synthesis78 of the sex attractant of the DFTM (Figure 12) exemplifies the

first synthetic strategy. Smith and co-workers began their synthesis with the protection

of aldehyde 4 as a dithiane. The dithiane (5) was then deprotonated with n-butyllithium,

and the resulting anion was alkylated with 1-chloro-5-decyne. Subsequent deprotection

and reduction of the ketone afforded alcohol 6. A syn-hydrogenation and oxidation

provided Z6 in 44% over 6 steps.

Page 36: Florida State University Libraries - FLVC

17

HS SH

BF3 OEt2

98%

4 5

1. n-BuLi;

Cl

2. CuO, CuClAcetone/H2O

3. LiAlH58% over 3 steps

O

8

Z6

O

8 H 8 H

SSOH

8

1. H2, P-2 NiEthylene Diamine

2. CrO3, pyridine

77% 2 steps

6

Figure 12: Smith’s Synthesis of the Sex Pheromone of the Douglas-Fir Tussock Moth.

Fetizon and Lazare’s synthesis of the DFTM sex pheromone (Figure 13),81 in

some ways, represents a hybrid of the strategies highlighted above. Their synthesis

began with 2-hydroxytetrahydropyran (7). Although 7 is a cyclic starting material,

hydroxy-aldehyde 8 is present in an equilibrium amount. Fetizon and Lazare took

advantage of this equilibrium and olefinated the aldehyde using Wittig reagent 9 to

install the Z-olefin of 10. Oxidation of alcohol 10, addition of n-decylmagnesium

bromide, and oxidation of the resulting alcohol provided Z6 in short order (four steps)

from a simple starting material in 51%.

O

H

OH

O

OH

Ph3P

7 8

9OH 1. CrO3, pyridine

2. n-C10H21MgBr

3. CrO3, pyridine85% 3 steps

O

8

Z6

10

60%

Figure 13: Fetizon and Lazare’s Synthesis of Z6.

In 1976, Kocienski and Cernigliaro published the synthesis of (Z)-6-heneicosen-

11-one (Figure 14);79 their synthesis exemplified the second strategy towards Z6, the

Page 37: Florida State University Libraries - FLVC

18

utilization of a ring-opening reaction. The ring opening reaction that Kocienski and

Cernigliaro envisioned as providing efficient access to the moth pheromone was the

Eschenmoser-Tanabe fragmentation49-52 (discussed in Chapter 1). Beginning with

vinylogous acid ester 11, they performed the enone synthesis first described by Woods

and Tucker.59 With all the necessary carbons installed, epoxidation of enone 12,

followed by condensation with p-tosylhydrazide in an acetic acid / methylene chloride

reaction medium provided tethered alkynyl-ketone 14. The alkyne was then

hydrogenated using palladium on barium sulfate in methanol and pyridine. Z6 was

synthesized in 4 steps from vinylogous acid ester 11 in 61% yield. Interestingly, there

have been 2 other syntheses that have also applied an Eschenmoser-Tanabe

fragmentation reaction similar to Kocienski and Cernigliaro to synthesize Z6.80,86

OMe 1. n-C10H21MgBr,Et2O;

H3O+

93%

O

O6

H2O2, NaOH,MeOH

O

6

O

O

8

O

8

Z6

11 12 13

14

p-TsNHNH2

AcOH/CH2Cl2

95%

71%

H2, Pd/BaSO4

MeOH/Pyridine97%

Figure 14: Kocienski and Cernigliaro’s Synthesis of Moth Pheromone Z6.

Having disclosed a preliminary study into the carbanion-triggered addition / C-C

bond cleavage (ABC) fragmentation methodology (Chapter 1),61 our lab envisioned the

synthesis of the sex pheromone of the Douglas-fir tussock moth to highlight our new

method.93 The impetus for this endeavor was derived from the fact that alkyl Grignard

nucleophiles were beyond the scope of our original report. Alkyl Grignards are often

more accessible and, in many cases, more reasonably priced than the corresponding

organolithiums; for instance: n-decylmagnesium chloride, needed for the synthesis of

Z6, is commercially available, n-decyllithium is not; ethylmagnesium chloride is far

Page 38: Florida State University Libraries - FLVC

19

cheaper than ethyllithium ($36.40 / 100 mL of 2.0 M in Et2O vs. $77.80 / 100mL of 0.5 M

in 9:1 benzene / cyclohexane, respectively).94 We therefore set out to optimize the

reaction between VAT 2 and Grignard nucleophiles (Table 2).

Aryl Grignard reagents (e.g., entry 1) were found to trigger fragmentation under

our original conditions;61 however, alkyl Grignards were not competent partners (entry

2). A quick screening of the reaction medium (entries 2-4) revealed that toluene was the

preferred solvent in our addition / fragmentation method using alkyl Grignards. The

reaction of 2 in toluene with an ethereal solution of n-butylmagnesium chloride afforded

the desired alkynyl ketone 3j (entry 4). Benzylmagnesium bromide provided 3k in 73%

(entry 6), however branched alkyl Grignards (e.g., i-propylmagnesium chloride, entry 5)

were significantly less efficient in the ABC process. The ability of n-decylmagnesium

bromide, relevant for the synthesis of Z6, in the fragmentation reaction was explored; it

was found to trigger the fragmentation of 2 (entry 7). In order to determine if toluene’s

effect on the reaction involving Grignard nucleophiles was general, we reexamined

phenylmagnesium bromide (entry 8), finding a modest and perhaps insignificant

decrease in the yield as compared to THF (entry 1).

Page 39: Florida State University Libraries - FLVC

20

Table 2: Grignard Triggered Fragmentation of 2.a

OTf

Me

O R1 M

-78 oC to 60 oCO Me

R1

2 3

Entry R1—M Solvent Product Yield (%)

1 PhMgBr THF 3a 80

2 n-BuMgCl THF 3j —c

3 n-BuMgCl Et2O 3j 24d

4 n-BuMgCl Toluene 3j 63-83

5 i-PrMgCl Toluene 3k —c

6 BnMgBr Toluene 3l 73

7 n-decylMgBr Toluene 3m 58

8 PhMgBr Toluene 3a 71 a

Solution of VAT 2 (1.1 equiv) treated with 1.0 equiv of R1—M (in Et2O) at -78

oC, warmed to r.t., and

then heated to 60 oC for 30 min.

b Note that Et2O is present in each case.

c Product was not isolated in

acceptable purity. d Reaction mixture was heated for 1 h at reflux; fragmentation was incomplete.

Having optimized the ABC reaction for alkyl Grignard reagents, we turned our

attention to the synthesis of the (Z)-6-heneicosen-11-one (Figure 15). Vinylogous acyl

triflate 15 was synthesized from 2-pentyl-1,3-cyclohexane dione95,96 using

trifluoromethanesulfonic anhydride and pyridine by analogy to the procedure published

by Kamijo and Dudley.61 Treatment of 15 with n-decylmagnesium bromide using our

optimized conditions afforded tethered alkynyl ketone 16 in 80% yield. Subsequent

hydrogenation of alkyne 16 provided Z6.79 Spectral data (1H NMR, 13C NMR. IR, and

HRMS) for our synthetic sample was in accordance with literature reports.78-92

Page 40: Florida State University Libraries - FLVC

21

-78 oC to 60 oC

O

OTf

n-decyl MgBr

toluene, 2.5 h80%

C10H21

C5H11O

5% Pd/BaSO4H2, pyridine

MeOH, 97%(ref. 77)

O

8

Z615 16

Figure 15: Synthesis of (Z)-6-Heneicosen-11-one Using an ABC Strategy.

Our synthesis is reminiscent of the Eschenmoser-Tanabe fragmentation

approach applied by Kocienski and Cernigliaro79 (discussed above). For example, in

their synthesis, vinylogous acid ester 11 was advanced to the moth pheromone in a four

step sequence that featured the Eschenmoser-Tanabe reaction. By enhancing the

nucleofugacity of the leaving group (methoxy of 11 vs. trifluoromethanesulfonyloxy of

15), we gained immediate access to the fragmentation product, streamlining the

synthetic sequence.

In summary, we extended the scope of our anion-triggered / C-C bond cleavage

reaction of vinylogous acyl triflates to include alkyl Grignard reagents. We applied the

ABC method to the synthesis of a commercially important natural product, (Z)-6-

heneicosen-11-one, the sex pheromone of the Douglas-fir tussock moth. Within the

context of our study, toluene proved to be a significantly more effective solvent than

THF for alkyl Grignard-triggered fragmentation reactions. The following chapters will

provide more insight into the development of our fragmentation method and its

extension to other synthetic applications.

Experimental

General Information:

1H NMR and 13C NMR spectra were recorded on a Varian 300 (300 MHz) spectrometer,

unless otherwise stated, using CDCl3 as the deuterated solvent. The chemical shifts ()

are reported in parts per million (ppm) relative to the residual chloroform peak (7.26

ppm for 1H NMR and 77.00 for 13C NMR). Coupling constants (J) are reported in Hertz

Page 41: Florida State University Libraries - FLVC

22

(Hz). IR spectra were recorded on a Perkin-Elmer FTIR paragon 1000 spectrometer

using NaCl discs. Mass Spectra were recorded on a JEOL JMS600H spectrometer. All

chemicals were used as received unless otherwise noted. Tetrahydrofuran (THF) and

toluene were dried through a solvent purification system packed with alumina and

molecular sieves under an Ar atmosphere. The Grignard solutions were titrated with a

known amount of iodine dissolved in ether. The purifications of the compounds were

performed by flash column chromatography using silica gel F-254 (230-499 mesh

particle size).

Representative procedure for the reaction of vinylogous acyl triflates (2) with

alkyl Grignard reagents (Table 2). To a toluene solution (2 mL) of n-BuMgCl (0.25 mL,

0.50 mmol; 2.0 M in Et2O) was added 2-methyl-3-(trifluoromethanesulfonyloxy)-2-

cyclohexenone (2) (142 mg, 0.55 mmol) at -78 oC under an Ar atmosphere. The mixture

was stirred at -78 oC for 10 min, at 0 oC for 10 min, at r.t. for 30 min, and then at 60 oC

for 30 min. Half-saturated aqueous solution of NH4Cl was added to quench the reaction

and the mixture was extracted with Et2O. The organic layer was washed with water,

dried over MgSO4, filtered, and concentrated. The residue was purified on silica gel

using 1% EtOAc/Hex to 5% EtOAc/Hex to give 9-undecyn-5-one (3j) in 63% yield (52

mg).

1-Phenyl-5-heptynone (3a): See reference 61 for analytical data.

9-Undecyn-5-one (3j): Pale yellow oil; 1H NMR (300 MHz, CDCl3) (t, J = 7.3 Hz,

2H), 2.40 (quintet, J = 7.3 Hz, 2H), 2.15 (tq, J = 7.3, 2.5 Hz, 2H), 1.77 (t, J = 2.5 Hz,

3H), 1.73 (quintet, J = 7.3 Hz, 2H), 1.55 (quintet, J = 7.3 Hz, 2H), 1.30 (sextet, J = 7.3

Hz, 2H), 0.90 (t, J = 7.3 Hz, 3H); 13C NMR (75 MHz, CDCl3) 210.8, 78.2, 76.2, 42.6,

41.3, 25.9,22.8, 22.3, 18.1, 13.8, 3.4; IR (neat) 1713, 1454, 1410, 1371, 1216, 1126

cm-1; HRMS (EI) Calcd for C11H18O (M+) 166.1357. Found 166.1357.

1-Phenyl-6-octyn-2-one (3l): Colorless oil; 1H NMR (300 MHz, CDCl3) 7.28-7.35 (m,

3H), 7.19-7.22 (s, 2H), 3.70 (s, 2H), 2.57 (t, J = 7.1Hz, 2H), 2.12 (tq, J = 7.1, 2.4 Hz,

Page 42: Florida State University Libraries - FLVC

23

2H), 1.74 (t, J = 2.4, 3H), 1.71 (quintet, J = 7.1, 2H); 13C NMR (75 MHz, CDCl3) 207.9,

134.2, 129.3, 128.6, 126.9, 78.1, 76.2, 50.1, 40.5, 22.8, 17.9, 3.3; IR (neat) 1713, 1602,

1495, 1453, 1367, 1093, 1031, 734, 700 cm-1; HRMS (CI) Calcd for C14H17O ([M+H]+)

201.1279. Found 201.1276.

2-Heptadecyn-7-one (3m): Colorless oil; 1H NMR (300 MHz, CDCl3) 2.52 (t, J = 7.4

Hz, 2H), 2.40 (t, J = 7.4 Hz, 2H), 2.16 (tq, J = 7.0, 2.5 Hz, 2H), 1.77 (t, J = 2.5 Hz, 3H),

1.73, quintet, 7.0 Hz), 1.26 (m, 16H), 0.88 (t, J = 6.6 Hz). 13C NMR (Bruker 600

spectrometer, 150 MHz, CDCl3) 210.98, 78.31, 76.25, 43.00, 41.41, 31.89, 29.70,

29.57, 29.48, 29.36, 29.30, 29.28, 23.91, 22.67, 18.18, 14.09, 3.49; IR (neat) 1701,

1470, 1418, 1374, 1091, 793 cm-1; HRMS (EI) Calcd for C17H30O (M+) 250.2297. Found

250.2296.

Synthesis of vinylogous acyl triflate 15: Prepared from 2-pentyl-1,3-

cyclohexanedione92,93 using triflic anhydride and pyridine by analogy to Kamijo and

Dudley’s published procedure, see reference 61. 1H NMR (300 MHz, CDCl3) 2.75 (t, J

= 6.2 Hz, 2H), 2.47 (t, J = 6.8, 2H), 2.32 (t, J = 7.6 Hz, 2H), 2.07 (app. quintet, J = 6.5

Hz, 2H), 1.22-1.42 (m, 6H), 0.88 (t, J = 6.8 Hz, 3H); 13C NMR (75 Hz, CDCl3) 197.52,

161.66, 132.32, 118.22 (q, J = 319.9 Hz), 36.84, 31.72, 28.63, 27.95, 23.66, 22.23,

20.58, 13.83; IR (neat) 1693, 1659, 1417, 1347, 1215, 1140, 1040 cm-1; HRMS (CI)

Calcd for C12H18OSF3 ([M+H]+) 315.0878. Found 315.0893.

Synthesis of alkynyl ketone 16: To a stirred solution of vinylogous acyl triflate 15 (100

mg, 0.32 mmol) in toluene (3 mL) at -78 oC was added n-decylmagnesium bromide

(0.31 mL, 0.93 M in Et2O, 0.29 mmol). The reaction mixture was warmed to r.t. for 1 h,

heated to 60 oC for 1.5 h, cooled to r.t., quenched with half-sat. NH4Cl solution (10 mL),

and extracted with Et2O. The combined extracts were washed with H2O, dried over

MgSO4, concentrated and purified on silica gel (elution with 1% EtOAc/Hexanes) to

afford alkynyl ketone 16 as an oil that solidified on standing; yielding 70 mg (80%). 1H

NMR (300 MHz, CDCl3) 2.52 (t, J = 7.3 Hz, 2H), 2.40 (t, J = 7.5 Hz, 2H), 2.08-2.23 (m,

4H), 1.74 (app. quintet, J = 7.0 Hz, 2H), 1.43-1.64 (m, 4H), 1.19-1.38 (m, 18H), 0.83-

Page 43: Florida State University Libraries - FLVC

24

0.95 (m, 6H); 13C NMR (75 Hz, CDCl3) 210.9, 81.1, 79.1, 43.0, 31.9, 29.6, 29.4, 29.3,

28.8, 23.9, 23.0, 22.6, 22.2, 18.7, 18.2, 14.1, 13.1; IR (neat) 1715, 1465, 1410, 1370,

1080, 720 cm-1; HRMS (CI) Calcd for C21H39O ([M+H]+) 307.3001. Found 307.2999.

Synthesis of (Z)-6-heneicosen-11-one (Z6): The reduction was performed in similar

manner to that presented in reference 77; note: the Pd/BaSO4 and pyridine must be

stirred in methanol for approximately 30 min before addition of alkyne for best results.

The following analytical data in accord with previous syntheses.76-90 1H NMR (500 MHz,

CDCl3) 5.37-5.41 (m, 2H), 2.35-2.41(m, 4H), 1.94-2.06 (m, 4H), 1.63 (app. quintet, J =

7.4 Hz, 2H), 1.21-1.37 (m, 22H), 0.85-0.90 (m, 6H); 13C NMR (Bruker 300 spectrometer,

75 Hz, CDCl3) 211.48, 130.96, 128.69, 42.90, 42.86, 42.82, 42.07; 31.97, 31.87,

29.55, 29.47, 29.40, 29.28, 27.19, 26.55, 23.88, 23.73, 22.66, 25.56, 14.09 (2 carbons);

IR (neat) 3020, 1715, 1465, 1420, 1380 cm-1; HRMS (CI) Calcd for C21H39O ([M+H]+)

309.3157. Found 309.3185.

Page 44: Florida State University Libraries - FLVC

25

1H NMR and 13C NMR spectra:

O

Bu

Me

3j

Page 45: Florida State University Libraries - FLVC

26

O

Bu

Me

3j

Page 46: Florida State University Libraries - FLVC

27

O Me

Ph

3l

Page 47: Florida State University Libraries - FLVC

28

O Me

Ph

3l

Page 48: Florida State University Libraries - FLVC

29

O

C10H21

Me

3m

Page 49: Florida State University Libraries - FLVC

30

O

C10H21

Me

3m

Page 50: Florida State University Libraries - FLVC

31

O

OTf

15

Page 51: Florida State University Libraries - FLVC

32

O

OTf

15

Page 52: Florida State University Libraries - FLVC

33

O

C10H21

C5H11

16

Page 53: Florida State University Libraries - FLVC

34

O

C10H21

C5H11

16

Page 54: Florida State University Libraries - FLVC

35

C10H21

O

C5H21Z6

Page 55: Florida State University Libraries - FLVC

36

C10H21

O

C5H21Z6

Page 56: Florida State University Libraries - FLVC

37

CHAPTER 3

A FRAGMENTATION / BENZANNULATION STRATEGY TO PROVIDE ACCESS TO BENZO-FUSED INDANES

Introduction

This chapter provides a detailed study into gold and copper catalyzed

benzannulation reactions of o-alkynyl aryl ketones bearing tethered acetylenes. The

primary motivation for the aforementioned study is derived from a desire to apply the

fragmentation reactions developed in the Dudley laboratory to an efficient synthesis of

the alcyopterosins, a rare subclass of natural products. However, before tackling the

synthesis of the alcyopterosins, a new methodology was required.

A detailed background of the alcyopterosins, including previous synthetic

strategies and biological importance, will provide the necessary context for the

development of a new convergent synthetic strategy towards these natural products.

Furthermore, a critical evaluation of benzannulation reactions similar to those

envisioned necessary in our focused retrosynthesis will set the stage for the original

work presented here.

The goal of this work is to determine the optimal conditions governing

intramolecular benzannulation reactions, while at the same time providing a method to

prepare benzo-fused indanes. Our research has been designed to bridge the gap that

exists between known benzannulation reactions and those which are required for our

proposed synthesis. The results of this study will play a vital role in future synthesis of

these natural products, new analogs, and other substituted indanes.

The Alcyopterosins

The illudalane sesquiterpenes,97 which include the alcyopterosins, represent a

class of rarely encountered natural products. These secondary metabolites consist of

Page 57: Florida State University Libraries - FLVC

38

bicyclo[4.3.0]nonane carboskeleton as seen in Figure 16. In most cases the 6-

membered ring is aromatic.

Cl

Alcyopterosin AIlludalane Skeleton

Figure 16: Illudalane Skeleton and Alcyopterosin A.

The biosynthesis of the illudalanes (Figure 17) originates from farnesyl

pyrophosphate (17) via a humulene intermediate 18.98 The humulene intermediate is

theorized to undergo cyclization to provide a protoilludane 19; a subsequent

rearrangement could give rise to an illudane (20). From illudane intermediate 20, a bond

cleavage reaction and aromatization would afford a molecule with the illudalane

carboskeleton.

O P

O

O

O P

O

O

O

OPP17

18

H

19 20

aromatization

Aromatic IlludalaneSkeleton

bond cleavageand

Protoilludane Illudane

Figure 17: Proposed Biosynthetic Pathway to the Illudalanes.

The chemistry of protoilludanes and illudanes has been studied by several

researchers.99-105 Some members of these natural products have been found to be

unstable under acidic or basic conditions, leading to the formation of aromatic illudalane

Page 58: Florida State University Libraries - FLVC

39

sesquiterpenes. Sterner and co-workers reported that the protoilludane stearodelicone

(21) decomposes to illudalane 22 upon absorption onto silica gel (Figure 18, equation

1). The decomposition is presumably due to traces of acid in the silica gel resulting in

the protonation of the enone, cleavage of the cyclobutyl moiety, and aromatization of

the cyclohexyldienone.100

The degradation of ptaquiloside 23 (Figure 18, equation 2), the major illudane

toxin isolated from the bracken fern, was examined by Saito and co-workers.101 The

glycosidic bond of ptaquiloside is easily cleaved in the presence of acid or base. Upon

cleavage of the glycosidic bond, the resulting alcohol is eliminated to produce bracken

dienone (24). If acid is present the tertiary alcohol of 24 ionizes and the cyclopropyl ring

undergoes heterolytic cleavage, resulting in the aromatization of the cyclohexadienyl

moiety; the cation is thus trapped by water to produce 25. The fact that ptaquiloside and

stearodelicone decompose to form illudalane-type products seems to support the

likelihood of their biosynthesis from the protoilludanes and illudanes.

Page 59: Florida State University Libraries - FLVC

40

O

OR

R = stearoyl

21

silicagel

O

OR

O

OR

H

OH2

(1)

O

O

HOH

OHO

OH

OH

OH

22

H+/H2O

D-glucoseO

HO

D-glucose

-OH/H2O

(pH 8-11)

OOH23

24

25

(2)

H+/H2O

Figure 18: (eq. 1) Proposed Decomposition of Stearodelicone (21) Upon Absorption on Silica Gel, and (eq. 2) the Observed Reactions of Ptaquiloside (23) in the

Presence of Acid and/or Base.

The illudalanes are typically isolated from both fungi of the Basidomycotina

subdivision105 and ferns of the Pteridaceae family.107 As rare as the illudalanes isolated

from terrestrial sources are, the alcyopterosins are even more rare. This subclass of

natural products represents the first illudalanes isolated from marine sources. The

alcyopterosins were first isolated from a deep water soft coral species, Alcyonium

paessleri, in sub-Antarctic waters by Palmero and co-workers in 2000.108 In 2009,

Gavagnin and co-workers isolated several new members of the alcyopterosins from a

different soft coral species, Alcyonium grandis.109

The alcyopterosins (Figure 19) have an aromatized six-membered ring, and

almost all members have either a chlorine atom or a nitrate ester present on the

ethylene side chain. Prior to the discovery of the alcyopterosins, there had never been a

Page 60: Florida State University Libraries - FLVC

41

natural nitrate ester secondary metabolite isolated from a marine source, despite the

fact that nitrates are common solutes in seawater.108 Sulfates and phosphates, which

are other common marine nutrients, are frequently observed in natural products isolated

from marine organisms. The fact that the alcyopterosins have been isolated as nitrate

esters makes them even more remarkable.

Cl O2NO

O2NO

O2NO

OH

O

O2NO

O

O

Cl

HO

OH O2NO

HOOH

Cl

Cl

O

O

O

AcO

AcO

O

25 26 28

HO

O

29

27

30

OH

Figure 19: Representative Sample of Illudalane Structures Isolated from A. paessleri and A. grandis.

Several members of the illudalane sesquiterpenes possess some interesting

biological activities; antimicrobial,99,110 cytotoxic,103,111 and antispasmodic activities112

being among them. Extracts containing members of the alcyopterosins have also been

found to possess feeding-deterrent activity against a generalist Antarctic sea-star

predator (Odontaster validus), implicating their chemical evolution as a defensive

mechanism (further discussed in Chapter 4).109 Alcyopterosins A (25), C (26), and H

(27), are cytotoxic towards the HT-29 (human colon carcinoma) cell line at 10 g/mL in

Page 61: Florida State University Libraries - FLVC

42

a preliminary in vitro test; and alcyopterosin E (28) has mild cytotoxicity (IC50 = 13.5 M)

towards the Hep-2 (human larynx carcinoma) cell line.108 In addition, several synthetic

analogs of the alcyopterosins show interesting DNA-binding properties (vide infra).113

The fact that the alcyopterosins are rarely observed as secondary metabolites,

their unusual structure, and their potential biological applications, provides motivation to

select them as synthetic targets. Since the initial report of their isolation and structural

elucidation,108 there have been several synthetic efforts directed towards members of

this sub-class of the illudalane sesquiterpenoids and several analogs.113-117

Most synthetic approaches to the alcyopterosin natural products include a

convergent transition metal promoted cycloaddition reaction.114-117 Unsymmetrical

polysubstituted aromatic rings are often difficult to prepare via sequential electrophilic

aromatic substitution reactions. Such reactions often result in regioisomeric products

that have to be separated. Therefore, several convergent aromatic annulations methods

have been developed to solve this challenging problem. The next section will address

the cyclotrimerization of alkynes and other aromatic annulation methods for assembly of

the core arenes of the illudalane sesquiterpenes.

The cyclotrimerization of acetylenes was first developed by Reppe in 1948.118

This method would be of particular value if selectivity could be obtained when

performed on substituted acetylenes; for instance, when this method is applied for the

synthesis of substituted aromatic compounds from three unsymmetrical acetylenes, 38

homo- and cross-coupled products are possible (Figure 20).

U

V X

W Y

Z

+ +

V

U

V

U

V

U

V

U

V

Z

Y

U

X

X

W W

X

W

Y

Z Z

Y

Z

Y

W

V

U

X

W

X X

W W

Y

Z

X

V

U

W

X

V

U

Plus 31 Other Isomers!

Metal

Catalyst

Figure 20: Traditional Reppe Reaction Involving Three Different Unsymmetrical Acetylenes.

Page 62: Florida State University Libraries - FLVC

43

Most of the recent solutions to the aforementioned issues associated with the

cyclotrimerization of acetylenes rely on a limited number of strategies (Figure 21): (a)

homo-coupling of acetylenes;119-125 (b) cross-coupling involving at least one symmetrical

acetylene;126-129 or (c) cross-coupling of tethered alkynes.130-135

X

YMetal Catalyst

X

Y

X

Y

X

Y

(a)

(b)

Y

X

X

Y

Z

Y

X

X

X

X

X

Y

Z

X

X

XMetal Catalyst

n

Y

Z

W

X

(c)

W

Y

Z

X

n

Figure 21: Typical Solutions for Chemo- and Regioselective Cyclotrimerization of Alkynes.

In 2001, Fumie Sato published a preliminary investigation into a metalative

Reppe reaction that allowed the use of three different unsymmetrical alkynes, one of

which being ethynyl-p-tolylsulfone, to provide a functionalized metalated arene as a

single isomer (Figure 22, equation 1).136 In the following year, Sato and co-workers

expanded their metalative Reppe process to the synthesis of arenes metalated at the

benzylic position. This extension was made possible by the replacement of the

ethynylsulfone with propargyl bromide (Figure 22, equation 2).114 In either case, the

metalated species could be trapped with a variety of electrophiles (e.g., H+, D+, I2),

leading to the synthesis of some potentially valuable compounds.

Page 63: Florida State University Libraries - FLVC

44

+

CO2t-Bu

C6H13 C6H13

Ti(O-i-Pr)4 /

2 i-PrMgClTi(O-i-Pr)2

CO2Bu-t

C6H13

H

C6H13 SO2Tol

-50 oC -50 oC to r.t.

CO2Bu-t

TiX3

C6H13

C6H13

(1)

+

CO2t-Bu

SiMe3

C6H13

Ti(O-i-Pr)4 /

2 i-PrMgClTi(O-i-Pr)2

SiMe3

C6H13

H

t-BuO2C CH2Br

-50 oC -50 oC to r.t.

MeSi3

C6H13

t-BuO2C

TiX3

(2)

X = (O-i-Pr)2Br

X = (O-i-Pr)2(O2STol)

Figure 22: Representative Examples of Sato’s One-Pot Metalative Reppe Reactions.

The metalative Reppe reaction developed by Sato was also demonstrated to

transform tethered alkynes, along with an external acetylene, to provide access to

bicyclic arenes. The Sato laboratory utilized its new method to accomplish the first

synthesis of alcyopterosin A (25) (Figure 23). The synthesis began with the reaction

between acetylenic ester 31 and tethered diyne 32, to provide the substituted indane 33

in 73% yield after hydrolysis. Diyne 32 was synthesized in 6 steps from isophorone,

featuring an Eschenmoser-Tanabe fragmentation (discussed in Chapter 1). The ethyl

ester of 33 was manipulated through a reduction, oxidation, and olefination sequence to

provide the ethylene side chain of 34. The olefin was then subjected to hydroboration-

oxidation, followed by conversion of the resulting alcohol to a chloride using standard

reaction conditions. The reaction sequence provided alcyopterosin A in 6 steps and

26% yield from diyne 32.

Page 64: Florida State University Libraries - FLVC

45

CO2Et

Me

+

Br

O

1. H2O2, NaOH

2. TsNHNH2, AcOH

O

4-steps

Isophorone

Ti(O-i-Pr)4 /i-PrMgCl;

then H+

73%

CO2Et

1. LiAlH4, 91%

2. PCC, 96%3. Ph3P=CH2, 86%

1. BH3 THF;H2O2, NaOH, 67% Cl

2. SOCl2, Pyridine

70%25

31 32 33

34

Figure 23: Sato’s Synthesis of Alcyopterosin A.

Since Sato’s synthesis of alcyopterosin A, two other members of this subclass of

natural products, alcyopterosins E and I (28 and 30, respectively) were synthesized

using a transition metal-catalyzed [2+2+2] cycloaddition strategy.115,117 Witulski and co-

workers completed the synthesis of alcyopterosin E115 (28) and confirmed the absolute

configuration originally assigned by Palmero et al.108 From tethered triyne 35, they

installed the tricyclic core (36) of alcyopterosin E in one synthetic operation using

Wilkinson’s rhodium(I) catalyst (Figure 24, equation 1). Much like Sato’s synthesis,

Witulski’s synthesis relied on the Eschenmoser-Tanabe fragmentation of isophorone to

provide access to the gem-dimethyl moiety.

Snyder and Jones provided the first synthesis of alcyopterosin I (30) in 2009 to

highlighting their newly discovered intramolecular rhodium-catalyzed [2+2+2]

cycloaddition reactions of diynes and enones.117 Cyclization precursor 37 was prepared

through a sequential double bromide displacement of 1,4-dibromo-2-butyne, first with

the enolate of ethyl isobutyrate, then with 3-pentynol. Conversion of the ethyl ester to

the terminal enone of 37 was carried out through common organic transformations. The

Page 65: Florida State University Libraries - FLVC

46

cycloaddition reaction was carried out using Wilkinson’s catalyst, and a DDQ work-up

produced the tricyclic core of alcyopterosin I in 71% yield (Figure 24, equation 2).

H

O

H3C

O

Isophorone

10 mol % RhCl(PPh3)3,

H

OTs

O

H

OTs O

CH2Cl2, 40 oC

72%

(1)

(2)

O

EtO

Ethyl Isobutyrate

O1. RhCl(PPh3)3,

PhCl, mW, 150 oC

2. DDQ, r.t.71% 2 steps

O

O

O

35 36

37 38

Figure 24: Key Steps in the Syntheses of Alcyopterosin E (28) (eq. 1), and Alcyopterosin I (30) by Witulski and Snyder (eq. 2).

In contrast to the more academically attractive methods used to prepare the

members of the alcyopterosins described above, Iglesias and co-workers presented a

more conventional approach to alcyopterosin A and several unnatural analogs.113 In the

course of Iglesias’ synthetic pathway, several compounds possessing the illudalane

skeleton were obtained, allowing for structure-activity relationship (SAR) studies to be

conducted. The Iglesias synthesis began with the construction of key intermediate 40

(Figure 25). Friedel-Crafts acylation of 4-bromo-m-xylene (38)—itself prepared through

a bromination of m-xylene and purification—provided -chloroketone 39. Intermediate

indanone 40 was obtained upon a subsequent acid promoted Nazarov reaction.

Page 66: Florida State University Libraries - FLVC

47

Br Br

O

Br

O

38

Cl

39 40

O

Cl Cl

AlCl3, CS299%

conc. H2SO4

67%

Figure 25: Synthesis of Iglesias’ Key Intermediate.

Iglesias and co-workers employed intermediate 40 to synthesize a variety of

analogs of the alcyopterosins (Figure 26). A reduction of the benzylic ketone of 40

provided bromoindane 41; another Friedel-Crafts acylation installed the necessary

carbons for the ethylene side chain of the illudalane skeleton. With the -chloroketone

42 in hand, the synthesis of various side chain functionalities (compounds 43-47) was

made possible through the use of several reduction methods. Compounds 45, 46, and

47 demonstrate an interesting divergence in reactivity; the reaction of -chloroketone 42

with excess sodium borohydride in refluxing ethanol provided three different analogs

simply by increasing the reaction time. Compound 47, most similar to alcyopterosin A,

was treated with lithium aluminum hydride to afford compound 48.

Page 67: Florida State University Libraries - FLVC

48

40

NaCNBH3, ZnI2, DCE

78%

Br Br

O

ClCl

OCl

AlCl3, CS279%

41 42

"conditions"

Br

R

Conditions:

(a) NaCNBH3, ZnI2, DCE, reflux (b) CuCl, NaBH4, EtOH, reflux (c) NaBH4, EtOH reflux, 15 min (d) NaBH4, EtOH, reflux, 4 h (e) NaBH4, EtOH, reflux, 16 h

43: R = CH2Cl2I 28%

44: R = 24%

45: R = 43%

46: R = 42%

47: R = CH2CH2OH 41%

O

CCH3

CHCH2Cl

OH

HC CH2

O

47LiAlH4

67%

HO

48

Figure 26: Synthesis of Unnatural Alcyopterosin Analogs Performed by Iglesias et al.

Having synthesized several analogs lacking the gem-dimethyl substituents on the

indane skeleton, Iglesias and co-workers turned their attention to the synthesis of

alcyopterosin A (Figure 27). Double methylation of intermediate 40, followed by

reduction of the ketone, generated compound 49. Friedel-Crafts acylation using

chloroacetyl chloride and subsequent reduction provided analog 50. Alcyopterosin A

(25) was obtained through the reduction of the arylbromide (providing 51) and

conversion of the side-chain alcohol to the necessary chloride.

Page 68: Florida State University Libraries - FLVC

49

40

Br

1. MeI, NaH,Toluene, 66%

2. NaCNBH3,ZnI2, DCE

88%

Cl

OCl

1. AlCl3, CS2

Br

HO

O

Br

49 50

69%

LiAlH4HO

, 79%

2. NaBH4, EtOH, reflux, 16 h

42%

SOCl2, pyridine

51

Cl

25

78%

Figure 27: Completion of Iglesias’ Synthesis of Alcyopterosin A.

The Iglesias laboratory, with numerous alcyopterosin analogs in hand, turned

their attention to performing DNA binding experiments. The ability of the alcyopterosin

analogs to bind to DNA was evaluated by measuring their hypochromic (decreased

absorbance at 260 nm) and bathochromic (red-shift) effects on the UV absorbance

spectrum of DNA.137 They validated their experiment through comparison of their test

assays and known intercalating agents (m-AMSA, mitoxantrone, and bis-benzamide;

Figure 28).

Page 69: Florida State University Libraries - FLVC

50

OH

OH

O

O

HN

HN

HN

OH

NH

OH

Mitoxantrone

N

HN

MeOHN

SO2Me

m-AMSA

N

N

NH

NNH

N

OH

H Cl3

bis-benzamide, Hoechst No. 33258

Figure 28: Compounds Known to Intercalate DNA.

The degree of interaction was expressed as a ratio between the final absorbance

area after stirring the compound for 24 h with DNA (a24) and the initial absorbance area

at max (a0). Values of 1 or higher indicate lack of affinity and values of 0 indicate

complete binding. The results of the DNA binding affinity assay demonstrate that

alcyopterosin A and various alcyopterosin analogs are potent DNA ligands (Table 3).

The gem-dimethyl substitution modifies, only slightly, the DNA binding affinity of the

compounds tested (47 vs. 50, and 48 vs. 51); whereas the ethylene side chain was of

the utmost importance for DNA ligation (compounds 41 and 49 had very poor affinity for

DNA). Perhaps most interesting was the fact that the presence of the bromine increased

the degree of binding of the analogs containing the hydroxy-functionalized ethylene side

chain (compounds 47, 48, 50, and 51).

Page 70: Florida State University Libraries - FLVC

51

Table 3: DNA Binding Assay Performed By Iglesias et al.113

Compound a24/a0 Compound a24/a0

41 0.90 49 0.87

42 0.12 50 0.40

43 0.16 51 0.71

44 0.26 25 0.38

45 0.69 Mxa 0.00

46 0.47 m-Ab 0.54

47 0.59 B-bc 0.57

48 0.89

a mitoxantrone; b m-AMSA; c bis-benzamide.

Br

OCl

42

Cl

25

HO

Br

HO

Br

47 50

A preliminary test was then carried out by The National Cancer Institute (NCI).

Compounds 44, 47, and 50 were evaluated in a three cell-line one dose pre-screen to

determine if they possess any ability to inhibit the growth of tumor cells in vitro. The cell

lines were MCF-7 (breast), NCI-H460 (lung), and SF-268 (CNS). Compounds found to

reduce the growth of any of the three cell lines to 32% or less, when compared to

untreated cells, were considered a positive in vitro lead. Compound 44 was found not to

inhibit growth to any significant extent. Compound 50 was found to produce a 0%

relative growth rate on all three cell lines, and compound 47 had the same effect on two

of the cell lines (breast and lung).

Analogs 47 and 50, having passed the first criterion for activity, were then

subjected to further testing against a 60-cell line panel at varying concentrations (10-4 to

Page 71: Florida State University Libraries - FLVC

52

10-8 M). The cell lines consisted of subpanels representing melanoma, leukemia, and

cancers of the breast, prostate, lung, colon, ovary, kidney, and brain. Dose-dependent

responses were found for three different activity parameters: the molar concentration

required to cause 50% growth inhibition (GI50), the concentration required to completely

inhibit growth (TGI), and the concentration that leads to 50% cell death (LC50). The

meangraph midpoints (MG-MID) correspond to the average sensitivity exhibited by the

entire panel of cell lines to a specific compound. The comparison of the MG-MID and

the activity against specific cell lines is often used to determine a compound’s selective

activity.

Compounds 47 and 50 demonstrated promising activities in the in vitro antitumor

screening (Table 4). The concentrations that promoted cytostatic (MG-MID GI50) and

cytotoxic (MG-MID LC50) effects for compounds 47 and 50 were found to have a marked

difference (ca. 5-fold). The ability to selectively control cancer cell growth or induce cell

death is an interesting trait observed for these natural product analogs.

Table 4: Average values (MG-MID) for in vitro antitumor activity on the NCI 60-Cell Line

Panel

Compound MG-MIDa

Log10GI50b (GI50) Log10TGIc (TGI) Log10LC50

d (LC50)

47 -4.77 (17 M) -4.40 (40 M) -4.12 (76 M)

50 -4.71 (19 M) -4.41 (39 M) -4.14 (72 M) a MG-MID = meangraph midpoint, average across all cell lines tested.

b GI50 = concentration required to

inhibit cell growth by 50%. c TGI = concentration required to completely inhibit cell growth.

d LC50 =

concentration required to kill 50% of tumor cells.

HO

Br

HO

Br

47 50

Page 72: Florida State University Libraries - FLVC

53

Nearly all members of the 60-cell line panel were found to be responsive to compound

50, whereas compound 47 was found to be more selective towards leukemia and

cancers of lung, colon, and breast (GI50 < 15 M). The antitumor activity of compounds

47 and 50 observed by Iglesias support the findings that the gem-dimethyl substituents

have little effect on DNA binding affinity as discussed above. The lack of the gem-

dimethyl, on the contrary, produced an increase in the selectivity of compound 47’s

ability to inhibit tumor growth.

The studies performed by Iglesias and co-workers identified some new

interesting anticancer leads as well as a straightforward approach to the alcyopterosins.

Their research, and the studies conducted by the other researchers referenced above,

have provided insight into the synthesis of compounds from this interesting subclass of

natural products. As part of our lab’s research goals, ―to devise, develop, and apply new

ideas in organic chemistry to the efficient synthesis of interesting molecules,‖138 we

identified the alcyopterosin natural products as potential targets that could benefit from

our fragmentation methodology. The remainder of this chapter will demonstrate the

synthetic approach we devised to access these natural products and to provide the

foundation for future synthetic efforts.

Retrosynthetic Analysis of Alcyopterosin A

In an effort to apply the carbanion-triggered fragmentation reaction of vinylogous

acyl triflates (VATs) (discussed in the previous chapters) to the synthesis of additional

natural products, we identified the alcyopterosins, specifically, alcyopterosin A, as

potential targets. Our retrosynthetic analysis (Figure 29) began with bicyclic arene 52,

which we envisioned gaining access to via an unprecedented benzannulation reaction

of acyclic enediyne intermediate 53. Based on our previous work, we believed that a

reaction between the metalated vinyl pre-nucleophile 54 and VAT 55 (derived from

dimedone) would provide our key acyclic intermediate (53).

Page 73: Florida State University Libraries - FLVC

54

Cl

Alcyopterosin A

functional group

manipulation

Z

Z = COR, H

52

benzannulation

R

O

fragmentationX

R

+

OTf

O

53(25)

54 55

Figure 29: Retrosynthetic Analysis to Alcyopterosin A Using a Fragmentation /

Benzannulation Approach.

Our strategy to synthesize alcyopterosin A hinges upon two key synthetic

transformations: (a) the fragmentation of vinylogous acyl triflate 55, and (b) the

benzannulation of enediyne 53. The basis of the desired benzannulation reaction stems

from the work of Yoshinori Yamamoto, Naoki Asao, and other members of the

Yamamoto laboratory.139-142 The fragmentation reactions of vinylogous acyl triflates has

been addressed in previous chapters. The following section will provide the relevant

background of the Yamamoto / Asao methodology for benzannulation and significant

questions that must first be addressed in order for the successful implementation of our

strategy.

In 2002, Yamamoto and co-workers published a preliminary communication

regarding a regioselective AuCl3-catalyzed formal [4+2] cycloaddition reaction between

o-alkynylbenzaldehydes (56) and alkynes (57) to produce naphthyl ketones (58 and 59)

(Figure 30, equation 1).139 A more thorough full paper ensued the following year.140 The

detailed study chronicled this benzannulation and also provided insight into a similar

[4+2] benzannulation of o-alkynylbenzaldehydes (or enals) (60) and alkynes (57) using

a copper catalyst. The copper catalyst system, in contrast to gold, produced

debenzoylated arenes (61 and 62) (Figure 30, equation 2). Naphthalenes were

Page 74: Florida State University Libraries - FLVC

55

generated in most cases, but a few examples of simple benzene derivatives, derived

from enals (as would be required for the synthesis of the alcyopterosins), were included.

Similar benzannulation reactions have also been explored through the use of

electrophilic iodine sources as stoichiometric reagents, however they fall outside the

scope of this discussion.143,144

H

O

R1

+

R3

R2

3 mol % AuCl3

DCE, 80 oCR2

R3

O R1

+ R3

R2

O R1

H

O

Ph

+

R3

R2

5 mol % Cu(OTf)2

1 equiv CF2HCO2H

DCE, 80 to 100 oC

R2

R3

+ R3

R2

H H

58

R2 = EDG

major

59

R2 = EWG

major

56 57

61

R2 = EDG

major

62

R2 = EWG

major

60 57

(1)

(2)

Figure 30: AuCl3- and Cu(OTf)2-Catalyzed [4+2] Benzannulation Reactions Described By Asao and Yamamoto.

The proposed mechanisms of these benzannulation reactions are presented in

Figure 31. Upon treatment with the Lewis acid, the soft -system of the alkyne 56

undergoes coordination to the Lewis acid (MLn: AuCl3 or Cu(OTf)2), enhancing the

electrophilicity of the alkyne. Subsequent nucleophilic 6-endo-dig cyclization of the

carbonyl oxygen onto the electron-deficient alkyne (as seen in 65) would form ate-

complex 66. The [4+2] cycloaddition of 66 with alkyne 57 would form intermediate 68 via

67. In the case of AuCl3-catalysis, subsequent bond rearrangement (as shown in 69)

would afford ketones 58 and 59 and regenerate the AuCl3. However, in the case of the

Cu(OTf)2 / CF2HCO2H system, protonolysis of the copper-carbon bond of 68, followed

by the attack of the conjugate base on the oxocarbenium ion, would produce

Page 75: Florida State University Libraries - FLVC

56

intermediate 70. A retro-Diels-Alder reaction would then release a mixed anhydride and

lead to the formation of products 61 and 62.

O

H

R1

O

H

R1LnM

O

H

MLn

MLn

RYRX

O

MLn

RX

RY

OR1

LnM

RX

RY

OR1

LnM

RX

RY

O R1

RY

RX

H

O

H

R1 A

A = CF2HCO2H

RX

RY

RX

RY

H

R1

O

A

_56

65

R1

6667

68

57

69

70

58, 59

61, 62

Figure 31: Proposed Mechanism of [4+2] Cycloaddition Reactions of 56 and 57 Catalyzed by AuCl3 and Cu(OTf)2 / CF2HCO2H.

Having successfully carried out intermolecular [4+2] benzannulation reactions,

Yamamoto and co-workers turned their attention towards the synthesis of polycyclic

naphthalene derivatives through the use of tethered alkyne dienophiles.142 The ―top-

down approach‖ (Figure 32, equation 1), in which the tethered alkyne is linked through

the carbonyl group, was found to convert compound 71 into naphthyl ketone 72 in high

yields. Interestingly, the reaction was found to occur even in the absence of a Lewis

acid at high temperatures, albeit in low yield (34% at 80 oC for 10 days). A related ―top-

down‖ benzannulation (without the prepositioned benzene ring) is envisioned for our

synthesis of alcyopterosin A. These examples, although limited, are therefore highly

Page 76: Florida State University Libraries - FLVC

57

relevant to our studies. The ―bottom-up approach‖ (Figure 32, equation 2), in which the

tethered alkyne is linked through the aryl-alkyne group (73), provided corresponding

polycyclic ketone 74 in yields ranging from 66 to 91%.

O R

Ph

n

n = 3 and 4R = Ph, Bu, H, TMS

O R

n-3 "Top-Down"40 to 92%

71 72

R

O

H

3O

R

R = Ph, p-Tolyl, p-CF3C6H4, n-Bu, H, TIPS, (CH2)2OTIPS, I

"Bottom-Up"66 to 91%

(2)

(1)AuX3

AuX3

73 74

Figure 32: Intramolecular Lewis Acid-Catalyzed [4+2] Benzannulation Reactions

Studied by Asao and Yamamoto.

Yamamoto and co-workers applied the ―bottom-up‖ approach to the synthesis of (+)-

ochromycinone and (+)-rubiginone B2 (Figure 33), demonstrating the power of these

reactions in synthesis.145

CHO

OMe

OMeOMe

cat. AuX3

O

OMe

MeO

OMe

O

OR

O

O

R = OMe: (+)-rubiginone B2

R = OH: (+)-ochromycinone

Figure 33: Yamamoto’s Key Benzannulation in the Synthesis of (+)-Rubiginone B2 and (+)-Ochromycinone.

Page 77: Florida State University Libraries - FLVC

58

In their studies, Asao and Yamamoto provided few examples of intermolecular

benzannulation reactions between alkynyl-enal substrates and alkynes (i.e. lacking the

prepositioned benzene backbone).139-142 Of these reactions, only the Cu(OTf)2 /

CF2HCO2H catalytic system were reported (Figure 34). Moreover, there were no reports

of the intramolecular benzannulation reaction taking place when Cu(OTf)2 was used as

the Lewis acid. The lack of such results prompts the questions: (1) Is AuCl3 capable of

effectively inducing the benzannulation of dialkynyl-enones similar to 53, and (2) is the

Cu(OTf)2 / CF2HCO2H a competent catalyst system in inducing the intramolecular

benzannulation reaction?

R1

R2

O

H

R3

R5

R6

Cu / H+ R1

R2

R5

R6

Few Examples(Only Intermolecular)

O

R

Ph

Au

R

PhO

Few Examples(Only Benzo-Fused,

Only Phenyl Ketones)

Yamamotoand Asao

O

R

?

Z

Z = COR, H52

Needed for Synthesis ofAlcyopterosin A

53

Figure 34: Contrast Between Known Benzannulations and Desired Benzannulation.

Using our fragmentation chemistry in conjunction with a new focused

methodology, we could make considerable contributions to the Lewis acid-catalyzed

intramolecular benzannulation reaction. We envisioned the ―top-down‖ approach, as

outlined by Yamamoto and co-workers, as being well suited for the synthesis of

Page 78: Florida State University Libraries - FLVC

59

alcyopterosin A. Ultimately we would require entry into an indane system, as opposed to

the benzo-fused indanes, potentially available via the Yamamoto / Asao methodology.

The following section describes this new methodology, highlighting the use of

fragmentation reactions to provide the needed monocyclic benzannulation precursors.

Exploring Gold and Copper Catalyzed Benzannulations

Prior to launching into the synthesis of alcyopterosin A, we sought to explore the

―top-down‖ intramolecular benzannulation in more detail. The substrates included in the

previous study by Asao and Yamamoto only varied the substituent at the terminus of the

tethered alkyne.142 An investigation into the effect of the substituents on the alkyne to

which the Lewis acid coordinates is envisioned to provide valuable knowledge of the

electronic requirements for benzannulation and catalyst selection, which may prove

useful in the synthesis of alcyopterosin A (Figure 35). We chose to perform our study on

benzo-fused systems for two reasons: (a) they would be most similar to those studied

by Yamamoto and Asao, and (b) the substrates would be easier to prepare due to their

inability to isomerize (e.g. E-, Z-isomerization). We believed that through the use of our

fragmentation methodology we could provide access to the benzo fused substrates in

short order.

Page 79: Florida State University Libraries - FLVC

60

O

R

Ph

R

O Ph

AuX3Asao and Yamamoto

J. Org. Chem. 2005, 70, 3682-3685.

O

Me

R

Me

O R

R = Ph, Bu, H, TMS

AuCl3 or Cu(OTf)2

R = p-MeOPh, Ph, t-Bu, n-Bu, TMS, p-CF3Ph

Required for newfocused methodology

Figure 35: Comparison of Known benzannulations and Those of a New Methodology.

This investigation would provide new knowledge into the steric and electronic

requirements of the intramolecular gold and copper catalyzed benzannulation reactions,

and allow access to new substituted benzo-fused indanes (polysubstituted

naphthalenes). The results would thereby further the current understanding of these

reactions as well as establish the ground work for future applications to alcyopterosin

synthesis.

We began our study by preparing the necessary substrates for the new

benzannulation study. Initially we considered two different starting materials for the

generation of o-alkynyl-haloarenes (77), which would serve as pre-nucleophiles for our

fragmentation reaction: (a) 1,2-dibromobenzene (75); and (b) 2-bromoiodobenzene (76)

(Figure 36). Upon further analysis, we identified some potential drawbacks in our initial

strategy. Attempting a Sonogashira reaction between 75 and 1-hexyne using standard

reaction conditions, we obtained an inseparable mixture of compounds 77 and 78 (ca.

35% yield, 1:1); similar results are not an uncommon occurance.146 Performing the

Sonogashira reaction on dihaloarene 76 would provide a selective reaction because of

Page 80: Florida State University Libraries - FLVC

61

the increased reactivity of the iodide, however the cost of 76 makes it less attractive for

use in a model study.

Br

Br

H R

Sonogashiracoupling

R

Br

R

R

possible side product78

Br

I

H R

Sonogashiracoupling

R

Br

75

76

77

(a)

(b)

Figure 36: Originally Considered Reactions to Access Fragmentation Pre-nucleophile 77.

In an effort to circumvent the problems associated with the strategy outlined

above, we identified 2-iodoaniline (79) as a potential alternative to the synthesis of

benzannulation test substrates. The advantages to the use of 79 as a starting material

would be three-fold: (a) 79 is intermediately priced (25 g/ $99.00) compared to 75 (25 g/

$74.10) and 76 (25 g/ $121.50);94 (b) the synthesis of iodotriazene 80147 would allow for

a directed metalation reaction of an aryl iodide, rather than an aryl bromide (cf. 77), to

provide nucleophile 81; and (c) triazene 82 could be converted into iodide 83 for the

selective synthesis of benzannulation substrates 84 through a Sonogashira reaction

(Figure 37). In effect, triazene 82 serves as a masked iodide that is also capable of

directing metalation chemistry.

Because of the fact that aryl iodides are more reactive than aryl bromides in both

Sonogashira reactions and halogen-metal exchange reactions, coupled to the fact that

our proposed halogen-metal exchange is envisioned to proceed through a directed

metalation, we believed this strategy would provide a general and efficient approach to

the synthesis of compounds similar to 84.

Page 81: Florida State University Libraries - FLVC

62

I

NH2

HCl, NaNO2;

then R2NH

I

N

NN

R

R

R LiLi

N3R2directed

metalation

+

OTf

O

O

N3R2

fragmentation "Conditions"O

ISonogashira

coupling

RH

R

O

79 80 81

82 83 84

2

Figure 37: Proposed Route to Benzannulation Substrates 84.

Our strategy proved very effective towards the synthesis of our model

benzannulation substrates. Conversion of 79 to diethyl iodotriazene 80147 was carried

out using standard conditions; first conversion of the arylamine to the diazonium salt,

and then an in situ trapping of the diazonium with diethylamine. Halogen-metal

exchange and subsequent fragmentation of vinylogous acyl triflate 2 provided triazene

82 in 82% over 2 steps (Figure 38).

I

NH2

HCl, NaNO2;

then Et2NH

97%

I

N

NN

Et

Et79 80

n-BuLi, Et2O

-78 oC;

then 2,

-78 oC to r.t.

85%

O

N3Et2

82

Figure 38: Synthesis and Fragmentation Reaction of Aryltriazene 80.

Aryltriazenes, similar to 80 and 82, are bench stable and chromatographable;

they have been used extensively in the synthesis of a large variety of phenylacetylene-

based systems.148 Typically these aryltriazenes are converted to the corresponding

iodoarene in high yields by heating in iodomethane at temperatures in excess of 100

Page 82: Florida State University Libraries - FLVC

63

oC.149 In the case of electron deficient aryltriazenes, decomposition of the triazene in

iodomethane requires higher temperatures. The toxicity of iodomethane and the high

temperatures and pressures required for the decomposition of aryltriazenes to

iodoarenes prompted us to search out other methods for this transformation. We found

reports in the literature that electron-deficient aryltriazenes undergo decomposition to

afford iodoarenes in high yields upon treatment with sodium iodide and sulfonic acid

cation exchange resins (H+ form) in dry acetonitrile at 75 oC; methanesulfonic acid and

trifluoroacetic acid also provided the product in acceptable yields.150

Armed with this knowledge, we completed the synthesis of our model

benzannulation substrates (Figure 39). Using slightly modified conditions, camphor-10-

sulfonic acid (CSA) in place of the sulfonic acid exchange resin, aryltriazene 82 was

converted to an aryl iodide 83. Sonogashira coupling reactions between various

terminal acetylenes and aryl iodide 83 provided benzannulation precursors 84a-e.

82

10 equiv CSA2 equiv NaI,

CH3CN, 75 oC

(ca. 75%) I

O

83

5 mol % PdCl2(PPh3)2,

10 mol % CuI, Et3N, 50 oC

H R

O

R

84a: R = Ph84b: R = n-Bu84c: R = t-Bu84d: R = p-MeO-C6H484e: R = TMS

68%85%52%60%80%

Figure 39: Synthesis of Benzannulation Substrates 84a-e.

The coupling reaction between 83 and an electron-deficient acetylene (R = p-

CF3-C6H4) did not proceed to any significant extent. In an effort to synthesize a

substrate with an acetylene having an electronic deficiency, the trimethylsilyl (TMS)

substituent of 84e was cleaved using a methanolic solution of potassium carbonate

affording 85; a Sonogashira reaction was performed between 85 and 4-

iodobenzotrifluoride to provide 84f in 85% yield (Figure 40).

Page 83: Florida State University Libraries - FLVC

64

O

TMS

K2CO3, MeOH

r.t., 92%O

H8584e

5 mol % PdCl2(PPh3)2,

10 mol % CuI, Et3N, 50 oC

CF3

I85%

O

84f CF3

Figure 40: Synthesis of Benzannulation Substrate 84f.

With a series of benzannulation substrates in hand similar to those prepared by

Yamamoto, ranging from electron-rich (R = p-MeO-C6H4, 84d) to electron-poor (R = p-

CF3-C6H4, 84f), we began to examine the benzannulation reaction using the AuCl3 and

Cu(OTf)2 / CF2HCO2H catalyst systems. The electron-neutral substrate included in our

study (R = Ph, 84a) most resembles those examined by Yamamoto and Asao.142

However, only the gold catalyzed reaction was reported from their related studies.

Table 5 summarizes the results obtained in the preliminary screening of

benzannulation reactions. Substrates 84a (as suggested by the results of Yamamoto

and Asao) and 84b provided promising reactivity when AuCl3-catalysis was employed.

However, they provided a mixture of the decarbonylated / reduced product 87 and

ketone products (86a and 86b, respectively) in the presence of the Cu(OTf)2 /

CF2HCO2H catalyst system (entries 1 and 2). Entries 4 and 5 demonstrate that the

electronically rich alkynes (84d) and silylacetylenes (84e) are not competent

benzannulation substrates in the presence of either catalytic system. Most interesting to

our future synthetic efforts was the divergence in reactivity between substrates 84c and

84f; both substrates provided ketone products 86c and 86f in the presence of AuCl3, but

when the Cu(OTf)2 / CF2HCO2H catalyst system was applied, the t-butylacetylene

containing substrate (84c) provided reduced product 87 as the sole product and the

electronically deficient acetylenic substrate (84f) provided only the ketone product (86f)

(entries 3 and 6, respectively). Thus, employing the Cu(OTf)2 / CF2HCO2H catalyst

system, one can switch between the two reaction pathways by changing the acetylene

substituent from t-butyl to p-CF3-C6H4. Likewise, in 84c (R = t-Bu) one can select

Page 84: Florida State University Libraries - FLVC

65

between the two products by simply changing the catalyst system from AuCl3 to

Cu(OTf)2 / CF2HCO2H.

Table 5: Preliminary Screening of Benzannulation Reactions of Substrates 84a-f.a

O

R

Catalyst

DCE, 80 oC

1 to 1.5 hO R

+

H

84 86 87

Entry R Substrate AuCl3 Catalystb Cu(OTf)2 / CF2HCO2H

Systemc

86 Yield, %b 86 Yield, %d 87 Yield, %d

1 Ph 84a 86a 75 86a 65 87 23

2 n-Bu 84b 86c 70 86c 10 87 71

3 t-Bu 84c 86c 50 86c 0 87 71

4 p-MeO-C6H4 84d 86d 0e 86d 0e 87 0e

5 Me3Si 84e 86e 0e 86e 0e 87 0e

6 p-CF3-C6H4 84f 86f 76 86f 80 87 0 a

Reactions performed on 10 mg scale for screening purposes. b 5 mol % AuCl3.

c 5 mol % Cu(OTf)2, 1.0

equiv. CF2HCO2H. dIsolated yields.

e No reaction was detected by TLC after 15 h, substrates were

recovered.

After performing our preliminary study into the ―top-down‖ benzannulation

reaction of Yamamoto on our series of test substrates, we chose to carry out the

benzannulation of substrates 84c and 84f on a larger scale under the copper(II)

catalysis conditions to confirm our preliminary results and provide more accurate yields.

Indeed our results were confirmed: compound 84c provided the reduced product (87) in

88% yield (Figure 41, equation 1); whereas compound 84f lead to naphthyl ketone 86f

in 89% yield (Figure 41, equation 2).

Page 85: Florida State University Libraries - FLVC

66

O

84c

5 mol % Cu(OTf)2,1.0 equiv. CF2HCO2H

DCE, 80 oC

88% H

87

O

84f

5 mol % Cu(OTf)2,1.0 equiv. CF2HCO2H

DCE, 80 oC

89%

CF3

O

CF3

(1)

(2)

86f

Figure 41: Direct Comparison of the Benzannulation Reactions of 84c and 84f.

In an effort to demonstrate the ability of vinylogous acyl triflate 55 to undergo the

desired fragmentation chemistry, as well as the ability of a substrate containing the

gem-dimethyl on the alkyne tether to participate in the benzannulation reaction, we set

out to synthesize a benzo-fused compound similar to 53 in our retrosynthetic analysis of

alcyopterosin A (Figure 29). Due to the fact that vinylogous acyl triflate 55 was

considered precious material,i we modified our synthetic approach (Figure 42). We

began with a Sonogashira reaction between 3,3-dimethylbutyne and 2-iodo-aryltriazene

80, which provided an inseparable mixture of the desired compound 88 and

unidentifiable byproducts. Conversion of the resulting o-alkynyl-aryltriazene into the

corresponding iodoarene 88 was performed using our previously described conditions

(see page 63). This reaction also provided an inseparable mixture that contained our

desired product as the major component by 1H NMR. Performing halogen metal-

exchange on the mixture containing iodoarene 88, followed by treatment with 1.0

equivalent of vinylogous acyl triflate 55 provided the desired product (89) in 61% yield

(90% based on recovered triflate).

i Synthesized in 2 steps: (1) methylation of dimedone;

151 and (2) subsequent triflation using standard

conditions from Kamijo and Dudley’s initial report.61

Page 86: Florida State University Libraries - FLVC

67

N3Et2

It-BuH

1. PdCl2(PPh3)2,

CuI, Et3N, 50 oC

2. CSA, NaI, CH3CN

(ca. 82% over 2 steps)

In-BuLi, Et2O, -78 oC;

then

OTf

O

5580 88

61% (>90% brsm)

O

89

Figure 42: Alternative Synthesis of Benzannulation Substrate 89.

Compound 89 underwent the desired intramolecular [4+2] benzannulation

reaction under either the gold or copper catalytic systems in 83 and 75%, respectively

(Figure 43). Much like our previous experiments, in the presence of the Cu(OTf)2 /

CF2HCO2H catalyst system compound 89 underwent benzannulation and

decarbonylation to provide the substituted naphthalene derivative 90 (equation 1). The

reaction of compound 89 in the presence of AuCl3 led to the formation of the naphthyl

ketone product (91) (equation 2). Interestingly, the 1H NMR spectrum of ketone 91

suggests it exists as a racemic mixture of atropisomers.ii The gem-dimethyl groups are

diastereotopic; each methyl group of the carbon bearing the gem-dimethyl can be

distinguished and one of the adjacent methylene units of the partially saturated ring

appears as an AB quartet (see page 113). We believe that the divergence in reactivity

of compound 89 upon treatment with either the gold or copper catalyzed benzannulation

conditions will prove useful, if it is observed when performed on acyclic intermediate 53,

as in the synthesis of alcyopterosin A.

ii However a slow rotation about the arene-ketone bond on the NMR time scale cannot be ruled out.

Page 87: Florida State University Libraries - FLVC

68

O

89

5 mol % Cu(OTf)2,1.0 equiv. CF2HCO2H

DCE, 80 oC

83% H

90

O

89

5 mol % AuCl3,

DCE, 80 oC

75%

91

O

+

O

(1)

(2)

Figure 43: Benzannulation Reactions of Compound 89.

The newfound knowledge in the gold and copper benzannulation chemistry

enables a new strategy for the synthesis of benzo-fused indanes. These benzannulation

reactions contribute to the observations made by Yamamoto and Asao. In addition, the

ability to obtain either the ketone or decarbonylated benzannulated products selectively,

either through choice of catalyst or by altering the substrate, was previously unreported.

This provides synthetic versatility in the synthesis of substituted indanes. For the Dudley

lab, it is this flexibility that may be the key toward the future synthesis of alcyopterosin

A. Our approach to benzo-fused indanes has incorporated the use of aryltriazenes for

the synthesis of useful intermediates and the fragmentation of vinylogous acyl triflates.

We have demonstrated the ability of vinylogous acyl triflates 2 and 55 to undergo

fragmentation reactions to provide synthetically useful compounds.

In regards to the synthesis of alcyopterosin A, a recently published study has

provided the necessary method for the synthesis of vinyl nucleophile 54. Negishi and

co-workers published the synthesis of various (Z)-2-alkynyl-vinyl iodides in a highly

stereoselective fashion (≥98% Z) (Figure 44).152 Bromoboration of propyne and trapping

Page 88: Florida State University Libraries - FLVC

69

of the resulting vinyldibromoborane with pinacol diminishes the stereoisomerization of

92 to provide cyclic boronate 93. Compound 93 is stable to air for several days at room

temperature without any change in the NMR spectrum. Negishi coupling of vinyl

bromide 93 and the appropriate terminal acetylene, followed by subsequent exchange

of the boronate for an iodide should provide access to 54 in high yield and selectivity.

Me

H 1.1 equiv BBr3,CH2Cl2

-78 oC to r.t., 2hMe

BBr2H

Br

1.2 equiv pinacol

-78 oC to r.t., 1h

Me Br

H BO

O

(85%, > 98% Z)

NegishiCoupling

H t-Bu Me

H BO

O2 equiv I2,

3 equiv NaOH

THF-H2O, r.t. Me

IH

54

92 93

Figure 44: Proposed Route to Vinyl Nucleophile 54 Using Negishi’s Z-Selective

Bromoboration.

The detailed benzannulation studies presented above provide a firm foundation

for future synthetic efforts. We have confirmed the ability of the Cu(OTf)2 / CF2HCO2H

catalyst system to promote intramolecular benzannulation reaction. When coupled with

the chemistry developed by Negishi, the results of this study pave the way for a

convergent approach to access the alcyopterosins and various analogs thereof. The

application of the fragmentation / benzannulation strategy to the synthesis of

alcyopterosin A and analogs thereof is currently underway in the Dudley laboratory.

Page 89: Florida State University Libraries - FLVC

70

Experimental

General Information:

1H NMR and 13C NMR spectra were recorded on a Varian 300 (300 MHz), Bruker 400

(400 MHz), or Bruker 600 (600 MHz) spectrometer, using CDCl3 as the deuterated

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

residual chloroform peak (7.26 ppm for 1H NMR and 77.00 for 13C NMR). Coupling

constants (J) are reported in Hertz (Hz). IR spectra were recorded on a Perkin-Elmer

FT-IR spectrometer with diamond ATR accessory as thin film. Mass Spectra were

recorded on a JEOL JMS600H spectrometer. Yields refer to isolated material judged to

be > 95% pure by 1H NMR spectroscopy following silica gel chromatography, F-254

(230-499 mesh particle size). All chemicals were used as received unless otherwise

noted. Acetonitrile (CH3CN) was distilled from calcium hydride (CaH2) and stored over

molecular sieves. Diethyl ether (Et2O) was dried through a solvent purification system

packed with alumina and molecular sieves under an Ar atmosphere. 1,2-Dichloroethane

(DCE) was used as received with no further purification. Triethylamine and diethylamine

were distilled from CaH2 and stored over KOH pellets. The n-butyllithium (n-BuLi)

solutions were titrated with a known amount of menthol, using 1,10-phenanthroline as

an indicator, in a solution of ether. All reactions were carried out under an inert argon

atmosphere unless otherwise stated.

Synthesis of 3-trifluoromethanesulfonyloxy-2,5,5-trimethyl-2-cyclohexenone (55):

Dimedone was methylated using iodomethane in a 5M aqueous KOH solution by

analogy to a published procedure, see reference 150; the resulting 2,5,5-trimethyl-1,3-

cyclohexanedione was converted to the corresponding triflate using triflic anhydride and

pyridine by analogy to our published procedure, see reference 61. Clear oil; 1H NMR

(300 MHz, CDCl3) 2.57 (app. q, J = 2.0 Hz, 2H), 2.33 (s, 2H), 1.85 (t, J = 2.0 Hz, 3H),

1.09 (s, 6H); 13C NMR (75 MHz, CDCl3) 197.52, 160.47, 126.96, 118.24 (q, J = 319.7

Hz), 50.57, 42.48, 32.73, 27.86, 8.89; IR (neat) 1689, 1671, 1418, 1207, 1136, 1029,

823 cm-1; HRMS (EI+) Calcd for C10H13OSF3 (M+) 286.0487. Found 286.0490.

Page 90: Florida State University Libraries - FLVC

71

Synthesis of 3,3-diethyl-1-(2-iodophenyl)-triazene (80): To a solution of 2-iodoaniline

(3 g, 13.7 mmol) in a minimal amount of acetonitrile (2 mL) was added ~6 g of ice,

followed by dropwise addition of concentrated HCl (9.1 mL, 109.6 mmol). The solution

was cooled to -10 oC and a solution of sodium nitrite (2.08 g, 30.14 mmol), in 33.3 mL of

water—acetonitrile (3:1), was added slowly. The reaction mixture was stirred at -10 oC

for 45 min. The solution of the generated diazonium salt was then transferred by

cannula to a solution (1.4 L) of acetonitrile—water (3:1) containing freshly distilled

diethylamine (14.2 mL, 137 mmol) and potassium carbonate (9.47 g, 68.5 mmol) at 0

oC. The resulting solution was allowed to warm upon stirring overnight. To the reaction

mixture was added 500 mL of water, and the products extracted three times with 250

mL of ether. The combined extracts were washed with brine, dried with magnesium

sulfate, and concentrated. The crude oil was purified by flash column chromatography

using 10% EtOAc/Hex, providing iodotriazene 80 in 97% as a yellow oil (4.025 g). 1H

NMR (300 MHz, CDCl3) 7.84 (dd, J = 7.9, 1.2 Hz, 1H), 7.35 (dd, J = 8.0, 1.7 Hz, 1H),

7.29 (ddd, J = 8.0, 7.3, 1.2 Hz, 1H), 6.83 (ddd, J = 7.9, 7.3, 1.7 Hz, 1H), 3.80 (q, J = 7.2

Hz, 4H), 1.33 (t, J = 7.2 Hz, 6H); 13C NMR (100 MHz, CDCl3) 150.45, 139.06, 128.67,

126.52, 117.56, 96.61, 49.20, 42.16, 14.57, 11.14; IR (neat) 1577, 1561, 1457, 1399,

1327, 1100, 749 cm-1; HRMS (ESI+) Calcd for C10H15IN3 ([M+H]+) 304.0311. Found

304.0309.

Synthesis of 1-(2-(3,3-diethyl-1-triazo)phenyl)-1-oxo-5-heptyne (82): To a solution of

iodotriazene 80 (2.0 g, 6.6 mmol) in diethyl ether (180 mL) at -78 oC was added n-BuLi

(4.13 mL, 6.6 mmol, 1.6M solution in hexane) dropwise. The mixture was stirred at -78

oC for 30 min. To the solution was added triflate 2 (1.87 g, 7.26 mmol), as an ethereal

solution (50 mL), dropwise. The solution was stirred at -78 oC for 15 min, 0 oC for 15

min, and at r.t. for 30 min. The reaction was then quenched with ½ sat. NH4Cl, extracted

2 times with Et2O, washed with H2O and brine, and dried with MgSO4. The concentrated

solution provided a crude oil, which was purified by flash column chromatography using

1% EtOAc/Hex. The product (82) was isolated as a yellowish-brown oil in 85% yield

(1.59 g). 1H NMR (300 MHz, CDCl3) 7.48 (dd, J = 8.2, 1.0 Hz, 1H), 7.43 (dd, J = 7.6,

1.3 Hz, 1H), 7.37 (ddd, J = 8.2, 7.6, 1.5, 1H), 7.14 (dt, J = 7.6, 1.0 Hz, 1H), 3.77 (q, J =

Page 91: Florida State University Libraries - FLVC

72

7.0 Hz, 4H), 3.06 (t, J = 7.4 Hz, 2H), 2.19 (tq, J = 7.0, 2.5 Hz, 2H), 1.93-1.77 (app.

quintet, J = 7.2 Hz, 2H), 1.74 (t, J = 2.5 Hz, 3H), 1.40-1.14 (broad multiplet, 6H); 13C

NMR (100 MHz, CDCl3) 206.36, 148.91, 135.12, 131.09, 127.96, 124.88, 118.14,

78.60, 75.95, 49.02, 43.48, 41.52, 23.84, 18.45, 14.46, 11.27, 3.47; IR (neat) 1674,

1592, 1403, 1328, 1092, 757 cm-1; HRMS (ESI+) Calcd for C17H24N3O ([M+H]+)

286.1919. Found 286.1915.

Representative procedure for the decomposition of aryl triazenes to provide aryl

iodides and Synthesis of 1-(2-iodophenyl)-1-oxo-5-heptyne (83): To a solution of

camphor-10-sulfonic acid (2.44 g, 10.5 mmol) and NaI (0.315 g, 2.1 mmol) in

acetonitrile (25 mL) at 75 oC was added a solution of triazene 82 (300 mg, 1.05 mmol in

5 mL of acetonitrile) dropwise. The evolution of nitrogen was complete after 5 minutes

of stirring at 75 oC. The mixture was cooled to r.t. and diluted with 25 mL of hexane. The

product was extracted 5 times with hexane. The combined hexane layers were then

dried with Na2SO4 and concentrated to provide a reddish oil. The crude material was

then purified by flash column chromatography using hexane. The resulting red—brown

oil (83) was not completely pure by 1H NMR, and was used in the next step without

further purification (ca. 75% yield, >85% pure).

Representative procedure for the Sonogashira coupling reaction for the synthesis

of compounds 84a-f and 88: To a solution of aryl iodide 83 (32 mg, 0.1 mmol) in

triethylamine (1 mL) was added dichlorobis(triphenylphosphine)palladium (4 mg, 5

mol) and copper(I) iodide (2 mg, 10 mol). The heterogeneous solution was degassed

using the freeze—pump—thaw method (5 times) and then warmed to room

temperature. Hexyne (30 L, 0.22 mmol) was then added to the reaction mixture in one

shot. The solution was then warmed to 50 oC and stirred for 3 h. The mixture was

cooled to r.t., diluted with ether, and filtered through Celite™. The filter cake was

washed three times with ether and the combined filtrates were concentrated. The crude

product was then purified by flash column chromatography using pure hexanes up to

1% EtOAc/Hex, providing 84b as a pale yellow oil in 85% yield (22 mg).

Page 92: Florida State University Libraries - FLVC

73

1-(2-(2-phenylethynyl)phenyl)-5-heptyn-1-one (84a): Pale yellow oil; 1H NMR (300

MHz, CDCl3) 7.70 (dd, J = 7.5, 1.6 Hz, 1H), 7.63 (dd, J = 7.5, 1.5 Hz, 1H), 7.59-7.53

(m, 2H), 7.47 (dt, J = 7.5, 1.6 Hz, 1H), 7.40 (dt, J = 7.5, 1.5 Hz, 1H), 7.37-7.34 (m, 3H),

3.30 (t, J = 7.0 Hz), 2.26 (tq, J = 7.0, 2.5 Hz, 2H), 1.94 (quintet, J = 7.0, 2H), 1.69 (t, J =

2.5 Hz, 3H); 13C NMR (75 MHz, CDCl3) 202.64, 141.02, 133.74, 131.48, 130.86,

128.62, 128.36, 128.20, 122.83, 121.20, 94.57, 88.25, 78.30, 76.35, 40.99, 23.61,

18.29, 3.32; IR (neat) 2215, 1678, 1493, 1217, 753, 689 cm-1; HRMS (EI+) Calcd for

C21H18O (M+) 286.1358. Found 286.1353.

1-(2-(2-n-butylethynyl)phenyl)-5-heptyn-1-one (84b): Clear oil; 1H NMR (300 MHz,

CDCl3) 7.59 (dd, J = 7.6, 1.6 Hz, 1H), 7.47 (dd, J = 7.6, 1.5 Hz, 1H), 7.38, (dt, J = 7.6,

1.6 Hz, 1H), 7.32 (dt, J = 7.6, 1.5 Hz, 1H), 3.20 (t, J = 7.3 Hz, 2H), 2.46 (t, J = 7.0, 2H),

2.31-2.16 (m, 2H), 1.89 (quintet, J = 7.3 Hz, 2H), 1.76 (t, J = 2.5 Hz, 3H), 1.67-1.55 (m,

2H), 1.54 (m, 2H), 0.95 (t, J = 7.2 Hz, 3H); 13C NMR (75 MHz, CDCl3) 203.50, 141.45,

133.77, 127.82, 127.47, 121.93, 96.28, 79.39, 78.41, 76.13, 41.10, 30.52, 22.07, 18.32,

13.57, 3.43; IR (neat) 2228, 1679, 1440, 758 cm-1; HRMS (EI+) Calcd for C19H22O (M+)

266.1671. Found 266.1669.

1-(2-(2-t-butylethynyl)phenyl)-5-heptyn-1-one (84c): Clear oil; 1H NMR (300 MHz,

CDCl3) 7.60 (dd, J = 7.5, 1.5 Hz, 1 H), 7.46 (dd, J = 7.5, 1.5 Hz), 7.38 (dt, J = 7.5, 1.5

Hz, 1H), 7.31 (dt, J = 7.5, 1.5 Hz, 1H), 3.24 (t, J = 7.2 Hz, 2H), 2.24 (tq, J = 6.0, 2.5 Hz,

2H), 1.89 (quintet, J = 7.2 Hz, 2H), 1.76 (t, J = 2.5 Hz, 3H), 1.33 (s, 9H); 13C NMR (75

MHz, CDCl3) 203.52, 141.30, 133.62, 127.82, 127.48, 104.09, 78.39, 78.18, 76.14,

41.37, 30.58, 28.18, 23.63, 18.31, 3.45; IR (neat) 2235, 1679, 1362, 1273, 757 cm-1;

HRMS (EI+) Calcd for C19H22O (M+) 266.1671. Found 266.1669.

1-(2-(2-(p-methoxyphenyl)ethynyl)phenyl)-5-heptyn-1-one (84d): Pale yellow oil; 1H

NMR (300 MHz, CDCl3) 7.69 (dd, J = 7.5, 1.2 Hz, 1H), 7.59 (dd, J = 7.5, 1.1 Hz, 1H),

7.50 (d, J = 8.7 Hz, 2H), 7.44 (dt, J = 7.5, 1.2 Hz, 1H), 7.36 (dt, J = 7.5, 1.1 Hz, 1H),

6.88 (d, J = 8.7 Hz, 2H), 3.82 (s, 3H), 3.29 (t, J = 7.3 Hz, 2H), 2.32-2.18 (m, 2H), 1.93

(app. quintet, J = 7.0 Hz, 2H), 1.69 (t, J = 2.4 Hz, 3H); 13C NMR (75 MHz, CDCl3)

Page 93: Florida State University Libraries - FLVC

74

202.96, 159.94, 140.82, 133.57, 133.02, 130.87, 128.21, 127.85, 121.63, 114.96,

114.05, 94.84, 87.14, 78.37, 76.36, 55.28, 41.04, 23.66, 18.33, 3.37; IR (neat) 2212,

1678, 1605, 151, 1247, 1028, 831, 758 cm-1; HRMS (EI+) Calcd for C22H20O2 (M+)

316.1463. Found 316.1462.

1-(2-(2-trimethylsilylethynyl)phenyl)-5-heptyn-1-one (84e): Yellow oil; 1H NMR (300

MHz, CDCl3) 7.64-7.59 (m, 1H), 7.57-7.52 (m, 1H), 7.41 (dt, J = 7.4, 1.9 Hz, 1H), 7.37

(dt, J = 7.4, 1.7 Hz, 1H), 3.23 (t, J = 7.3 Hz, 2H), 2.42 (tq, J = 7.1, 2.5 Hz, 2H), 1.90

(quintet, J = 7.1 Hz, 2H), 1.76 (t, J = 2.5 Hz, 3H), 0.26 (s, 9H); 13C NMR (75 MHz,

CDCl3) 203.03, 141.81, 133.99, 130.61, 128.48, 127.88, 120.80, 103.53, 100.44,

78.30, 76.13, 41.17, 23.66, 18.27, 3.41, 0.39; IR (neat) 2157, 1682, 1249, 863, 840, 758

cm-1; HRMS (ESI+) Calcd for C18H22OSiNa ([M+Na]+) 305.1338. Found 305.1333.

1-(2-(2-(p-Trifluoromethylphenyl)ethynyl)phenyl)-5-heptyn-1-one (84f): Yellow oil;

1H NMR (300 MHz, CDCl3) 7.74 (dd, J = 7.4, 1.5 Hz, 1H), 7.67 (d, J = 8.3 Hz, 2H),

7.65-7.58 (m, 3H), 7.50 (dt, J = 7.4, 1.6 Hz, 1H), 7.44 (dt, J = 7.4, 1.5 Hz, 1H), 3.24 (t, J

= 7.3 Hz, 2H), 2.26 (tq, J = 7.0, 2.5 Hz, 2H), 1.94 (quintet, J = 7.0 Hz, 2H), 1.70 (t, J =

2.5 Hz, 3H); 13C NMR (150 MHz, CDCl3) 202.15, 141.14, 134.82, 131.05, 130.31 (q,

32.7 Hz), 128.78, 125.34 (q, 3.9 Hz), 123.88 (q, 270.9 Hz), 120.61, 92.74, 90.61, 78.28,

76.48, 40.77, 23.59, 18.31, 3.53; IR (neat) 1683, 1614, 1320, 1126, 1065, 824 cm-1;

HRMS (EI+) Calcd for C22H17OF3 (M+) 354.1232. Found 354.1230.

Synthesis of 1-(2-ethynylphenyl)-5-heptyn-1-one (85): To a methanolic solution (2

mL) of trimethylsilylacetylene (84e) (136 mg, 0.48 mmol) was added potassium

carbonate (100 mg, 0.72 mmol) at room temperature. The reaction mixture was stirred

at room temperature until the starting material was no longer detected by TLC (ca. 30

min). The mixture was diluted with ether and water. The reaction was quenched with 1N

HCl until CO2 evolution was no longer observed. The product was extracted twice with

EtOAc. The combine organics were washed with water and brine, dried with Na2SO4,

filtered and concentrated. The resulting crude oil was purified by flash column

chromatography using 100% hexane up to 1% EtOAc/Hex, providing 85 as a clear oil in

Page 94: Florida State University Libraries - FLVC

75

94% yield (94 mg). 1H NMR (300 MHz, CDCl3) 7.69-7.63 (m, 1H), 7.63-7.56 (m, 1H),

7.48-7.38 (m, 2H), 3.35 (s, 1H), 3.18 (t, J = 7.1 Hz, 2H), 2.25 (tq, J = 7.1, 2.5 Hz), 1.91

(quintet, J = 7.1 Hz, 2H), 1.77 (t, J = 2.5 Hz, 3H); 13C NMR (150 MHz, CDCl3) 202.52,

141.92, 134.64, 130.85, 128.78, 120.04, 82.42, 82.24, 78.36, 76.32, 40.58, 23.51,

18.23, 3.44; IR (neat) 1687, 1440, 1225, 757 cm-1; HRMS (EI+) Calcd for C15H13O ([M-

H]+) 209.0967. Found 209.0964.

Representative procedure for the AuCl3—catalyzed benzannulation reaction: To a

solution of AuCl3 (0.6 mg, 1.8 mol) in 100 L dichloroethane (DCE), obtained from a

stock solution (6 mg/mL), was added an additional 200 L of DCE and diyne 84b (10

mg, 37 mol, in 200 L of DCE). The solution was then heated to 80 oC for 1.5 h. The

mixture was cooled to r.t. and filtered through a plug of silica gel. The filtrate was

concentrated and purified by flash column chromatography using 1% EtOAc/Hex ,

providing naphthyl ketone 86b in 70% yield (7 mg) as a clear oil.

(4-Methyl-2,3-dihydro-1H-cyclopenta[a]naphthalene-5-yl)-phenyl-methanone (86a):

1H NMR (300 MHz, CDCl3) 7.85-7.80 (m, 2H), 7.61-7.53 (m, 1H), 7.50-7.38 (m, 3H),

7.34-7.26 (m, 1H), 3.35 (t, J = 7.5 Hz, 2H), 3.08 (t, J = 7.5 Hz, 2H), 2.30 (quintet, J = 7.5

Hz, 2H), 2.22 (s, 3H); 13C NMR (75 MHz, CDCl3) 200.88, 140.99, 140.41, 137.95,

134.73, 133.58, 130.30, 129.76, 129.55, 128.73, 125,50, 125.45, 125.26, 124.54; 32.63,

31.65, 23.89, 17.23; IR (neat) 1664, 1448, 1219, 884, 756, 717 cm-1; HRMS (EI+) Calcd

for C21H18O (M+) 286.1358. Found 286.1353.

n-Butyl-(4-methyl-2,3-dihydro-1H-cyclopenta[a]naphthalene-5-yl)-methanone

(86b): 1H NMR (300 MHz, CDCl3) 7.78 (dd, J = 7.8, 1.8 Hz, 1H), 7.55 (dd, J = 7.2, 1.8

Hz, 1H), 7.43-7.37 (m, 2H), 3.29 (t, J = 7.2 Hz, 2H), 3.04 (t, J = 7.8 Hz, 2H), 2.87 (t, J =

7.8 Hz, 2H), 2.31 (s, 3H), 2.26 (quintet, J = 7.8 Hz, 2H), 1.78 (quintet, J = 7.8 Hz), 1.44

(app. sextet, J = 7.6 Hz, 2H), 0.953 (t, J = 7.2 Hz, 3H); 13C NMR (150 MHz, CDCl3)

211.21, 141.03, 140.05, 137.75, 128.91, 128.87, 127.53, 125.57, 125.30, 124.74,

124.55, 45.65, 32.64, 31.59, 25.75, 23.87, 22.47, 16.91, 13.94; IR (neat) 1697, 1130,

751 cm-1; HRMS (EI+) Calcd for C19H22O (M+) 266.1671. Found 266.1670.

Page 95: Florida State University Libraries - FLVC

76

t-Butyl-(4-methyl-2,3-dihydro-1H-cyclopenta[a]naphthalene-5-yl)-methanone (86c):

1H NMR (300 MHz, CDCl3) 7.77 (dd, J = 7.6, 1.1 Hz, 1H), 7.53-7.33 (m, 3H), 3.29

(app. triplet, J = 7.8 Hz, 2H), 3.13-2.94 (m, 2H), 2.29 (s, 3H), 2.26 (app. quintet, J =

7.8Hz, 2H), 1.27 (s, 9H); 13C NMR (150 MHz, CDCl3) 219.07, 141.05, 139.42, 137.12,

129.31, 128.83, 127.46, 125.68, 129.19, 125.06, 124.66, 45.69, 32.71, 31.51, 28.09,

23.83, 18.34; IR (neat) 1688, 1276, 1261, 764, 749 cm-1; HRMS (EI+) Calcd for

C19H22O (M+) 266.1671. Found 266.1668.

(4-Methyl-2,3-dihydro-1H-cyclopenta[a]naphthalene-5-yl)-p-trifluoromethylphenyl-

methanone (86f): 1H NMR (300 MHz, CDCl3) 7.95 (d, J = 8.1 Hz, 2H), 7.84 (d, J = 8.1

Hz, 1H), 7.69 (d, J = 8.1 Hz, 2H), 7.52-7.39 (m, 2H), 7.36-7.28 (m, 1H), 3.37 (t, J = 7.5

Hz, 2H), 3.10 (t, J = 7.5 Hz, 2H), 2.32 (quintet, J = 7.5 Hz, 2H), 2.22 (s, 3H); 13C NMR

(75 MHz, CDCl3) 199.80, 141.03, 140.47, 134.73 (q, J = 32.5 Hz), 133.78, 130.12,

129.99, 129.83, 128.82, 125.83 (q, J = 3.7 Hz), 125.72, 125.15, 124.71, 123.55 (q, J =

272.9 Hz), 32.60, 31.67, 23.84, 17.26; IR (neat) 1673, 1409, 1322, 1168, 1128, 1069

cm-1; HRMS (ESI+) Calcd for C22H17F3ONa ([M+Na]+) 377.1129. Found 377.1135.

Representative procedure for reactions performed with the Cu(OTf)2 / CF2HCO2H

catalyst system: To a solution of Cu(OTf)2 (4 mg, 11 mol) and difluoroacetic acid (14

L, 0.22 mmol) in DCE, was added a solution of diyne 84c (58 mg, 0.22 mmol in 1 mL

of DCE). The solution was heated to 80 oC and stirred for 40 min. The reaction mixture

was cooled to r.t. and filtered through silica gel. The filtrate was concentrated; the

resulting oil was purified by flash column chromatography using hexanes, providing

naphthalene derivative 87 in 88% yield (35 mg).

4-Methyl-2,3-dihydro-1H-cyclopenta[a]naphthalene (87): 1H NMR (300 MHz, CDCl3)

7.78 (d, J = 7.3 Hz, 1H), 7.75 (d, J = 7.3 Hz, 1H), 7.47 (s, 1H), 7.46-7.35 (m, 2H), 3.28

(t, J = 7.5 Hz, 2H), 3.04 (t, J = 7.5 Hz, 2H), 2.43 (s, 3H), 2.26 (quintet, J = 7.5 Hz, 2H);

13C NMR (150 MHz, CDCl3) 199.80, 141.03, 140.47, 134.73 (q, J = 32.5 Hz), 133.78,

130.12, 129.99, 129.83, 128.82, 125.83 (q, J = 3.7 Hz), 125.72, 125.15, 124.71, 123.55

(q, J = 272.9 Hz), 32.60, 31.67, 23.84, 17.26. 141.26, 139.08, 133.20, 132.86, 129.03,

Page 96: Florida State University Libraries - FLVC

77

127.62, 125.65, 124.87, 124.72, 124.24, 32.39, 31.32, 24.14, 19.76; IR (neat) 1595,

1382, 1020, 872, 766, 743 cm-1; HRMS (ESI+) Calcd for C14H14 (M+) 182.1096. Found

182.1094.

Synthesis of 1-(2-(2-t-butylethynyl)phenyl)-3,3-dimethyl-5-heptyn-1-one (89):

Iodotriazene 80 was coupled to 3,3-dimethylbutyne using the same method as outlined

above. The resulting alkynyl triazene was converted to the corresponding iodide under

our modified conditions (see above). The product was purified by flash column

chromatography providing 88 as the major component in an inseparable mixture of

compounds. To a solution of the this mixture (165 mg, ~0.58 mmol) in diethyl ether (20

mL) was added n-BuLi (0.39 mL, 0.52 mmol, as a 1.31M solution in hexane) dropwise

at -78 oC. The mixture was stirred for 30 min, at which time, an ethereal solution (5 mL)

of triflate 55 (150 mg, 0.52 mmol) was added dropwise at -78 oC. The solution was

stirred for 15 min at -78 oC, then at 0 oC, and finally at r.t. for 30 min. The reaction was

quenched with ½ sat. ammonium chloride. The products were extracted with ether (2

times). The combined organic layers were washed with water and brine, dried with

MgSO4, filtered and concentrated. The resulting crude oil was purified by flash column

chromatography using 1% EtOAc/Hex up to 5% EtOAc/Hex to provide diyne 89 in 61%

yield (94 mg); (>90% brsm). Clear oil; 1H NMR (300 MHz, CDCl3) 7.47 (dd, J = 7.4,

1.5 Hz, 1H), 7.43 (dd, J = 7.4, 1.5 Hz, 1H), 7.34 (dt, J = 7.4, 1.5 Hz, 1H), 7.29 (dt, J =

7.4, 1.5 Hz, 1H), 3.13 (s, 2H), 2.21 (q, J = 2.5 Hz, 2H), 1.75 (t, J = 2.5 Hz, 3H), 1.32 (s,

9H), 1.08 (s, 6H); 13C NMR (75 MHz, CDCl3) 204.27, 143.58, 133.39, 130.15, 127.41,

121.40, 103.72, 78,99, 77.50, 76.75, 76.49, 51.44, 34.58, 32.39, 30.65, 28.17, 27.14,

3.48; IR (neat) 2359, 1684, 1472, 1362, 758 cm-1; HRMS (EI+) Calcd for C21H26O (M+)

294.1984. Found 294.1984.

3,3,4-Trimethyl-2,3-dihydro-1H-cyclopenta[a]naphthalene (90): 1H NMR (300 MHz,

CDCl3) 7.76 (dd, J = 6.8, 2.5 Hz,1H), 7.70 (dd, J = 6.6, 2.5 Hz, 1H), 7.46 (s, 1H), 7.44-

7.34 (m, 2H), 3.08 (s, 2H), 2.86 (s, 2H), 1.26 (s, 6H); 13C NMR (75 MHz, CDCl3)

144.60, 142.46, 137.53, 137.41, 133.61, 131.99, 129.97, 129.09, 128.94, 128.41, 51.77,

Page 97: Florida State University Libraries - FLVC

78

50.61, 43.64, 34.16, 23.98; IR (neat) 1364, 872, 843, 743 cm-1; HRMS (EI+) Calcd for

C16H18 (M+) 210.1408. Found 210.1404.

t-Butyl-(3,3,4-trimethyl-2,3-dihydro-1H-cyclopenta[a]naphthalene-5-yl)-methanone

(91): 1H NMR (300 MHz, CDCl3) 7.71 (dd, J = 7.8, 1.4 Hz, 1H), 7.47 (dd, J = 7.8, 1.3

Hz, 1H), 7.42 (dt, J = 7.8, 1.4 Hz, 1H), 7.36 (dt, J = 7.8, 1.4 Hz), 3.09 (s, 2H), 2.89 (d, J

= 16.0 Hz, 1H), 2.80 (d, J = 16 Hz, 1H), 2.25 (s, 2H), 1.29 (s, 3H), 1.27 (s, 9H), 1.23 (s,

3H); 13C NMR (150 MHz, CDCl3) 219.15, 140.10, 128.55, 137.12, 129.29, 129.10,

127.68, 125.71, 125.13, 124.99, 124.54, 47.71, 46.40, 45.69, 39.25, 29.82, 29.79,

28.11, 18.25; IR (neat) 1685, 1463, 1102, 903, 738 cm-1; HRMS (EI+) Calcd for C21H26O

(M+) 294.1984. Found 294.1984.

Page 98: Florida State University Libraries - FLVC

79

1H NMR and 13C NMR Spectra:

OTf

O

55

Page 99: Florida State University Libraries - FLVC

80

OTf

O

55

Page 100: Florida State University Libraries - FLVC

81

I

N

NN

80

Page 101: Florida State University Libraries - FLVC

82

I

N

NN

80

Page 102: Florida State University Libraries - FLVC

83

O

N

NN

82

Page 103: Florida State University Libraries - FLVC

84

O

N

NN

82

Page 104: Florida State University Libraries - FLVC

85

O

84a

Page 105: Florida State University Libraries - FLVC

86

O

84a

Page 106: Florida State University Libraries - FLVC

87

O

84b

Page 107: Florida State University Libraries - FLVC

88

O

84b

Page 108: Florida State University Libraries - FLVC

89

O

84c

Page 109: Florida State University Libraries - FLVC

90

O

84c

Page 110: Florida State University Libraries - FLVC

91

O

84dOMe

Page 111: Florida State University Libraries - FLVC

92

O

84dOMe

Page 112: Florida State University Libraries - FLVC

93

O

Si

84e

Page 113: Florida State University Libraries - FLVC

94

O

Si

84e

Page 114: Florida State University Libraries - FLVC

95

O

84fCF3

Page 115: Florida State University Libraries - FLVC

96

O

84fCF3

Page 116: Florida State University Libraries - FLVC

97

O

85

H

Page 117: Florida State University Libraries - FLVC

98

O

85

H

Page 118: Florida State University Libraries - FLVC

99

O

86a

Page 119: Florida State University Libraries - FLVC

100

O

86a

Page 120: Florida State University Libraries - FLVC

101

O

86b

Page 121: Florida State University Libraries - FLVC

102

O

86b

Page 122: Florida State University Libraries - FLVC

103

O

86c

Page 123: Florida State University Libraries - FLVC

104

O

86c

Page 124: Florida State University Libraries - FLVC

105

O

86f CF3

Page 125: Florida State University Libraries - FLVC

106

O

86f CF3

Page 126: Florida State University Libraries - FLVC

107

87

Page 127: Florida State University Libraries - FLVC

108

87

Page 128: Florida State University Libraries - FLVC

109

O

89

Page 129: Florida State University Libraries - FLVC

110

O

89

Page 130: Florida State University Libraries - FLVC

111

91

Page 131: Florida State University Libraries - FLVC

112

91

Page 132: Florida State University Libraries - FLVC

113

O

90

Page 133: Florida State University Libraries - FLVC

114

O

90

Page 134: Florida State University Libraries - FLVC

115

CHAPTER 4

SYNTHESIS OF THE EASTERN HEMISPHERE (C1-C15) OF PALMEROLIDE A

Introduction

The goal of this work was to address the shortcomings of a previously reported

fragmentation reaction in order to provide an efficient synthesis of the eastern

hemisphere of palmerolide A. Palmerolide is an exciting natural product that possesses

anti-cancer properties and selectively targets melanoma. This work will provide the

basis for future synthetic efforts applied to the large scale synthesis of this natural

product by the Dudley laboratory, including the expedient synthesis of the eastern

hemisphere (C1-C15).

The following chapter will highlight some of the deficiencies related to the

treatment of the growing melanoma epidemic in western countries. There exists a need

for new drugs that can selectively lead to cell death in melanoma tumors. The discovery

of palmerolide A has provided a possible target that may ultimately lead to a better

prognosis for patients that suffer from melanoma. Unfortunately, synthesis is the only

means at present to obtain enough quantities of this natural product to perform further

clinical studies.

The examination of previous synthetic methods applied to the generation of

palmerolide A draws attention to the fact that all approaches to this natural product have

focused on a convergent process with the production of relatively few strategically

generated C-C bonds in an effort to access the core structure. Thus, fragment synthesis

is paramount to provide an efficient synthesis capable of producing quantities of

palmerolide to support future biological studies.

The eastern hemisphere contains three of the five stereocenters which are

isolated by hydrocarbon regions. We envisioned our fragmentation strategy as being

well suited for the synthesis of this C1-C15 fragment. However, an optimization of a

known fragmentation reaction had to be improved prior to beginning this endeavor.

Page 135: Florida State University Libraries - FLVC

116

The Melanoma Problem

The skin is the largest organ of the human body. It is responsible for providing a

protective barrier against infection and injury, and serves a key role in thermoregulation.

The skin is composed of three distinct layers: the epidermis, the dermis, and the

subcutis. For this discussion, the focus is on the epidermis, the most superficial layer of

the skin.

The dead cells at the surface are composed of squamous cells that have been

flattened and keratinized; they provide the primary protective barrier for the body.

Several types of cells exist below this outer most layer of the epidermis, including:

Merkel cells (tactile receptors), Langerhans cells (antigen processing cells),

keratinocytes, melanocytes and basal cells, among others.153

Melanocytes are responsible for the production of a pigment called melanin,

which provides the skin with its color and protects the deeper layers of the skin from

ultraviolet radiation. The sun stimulates the melanocytes to produce more melanin

resulting in tanning of the skin. As do most cells, melanocytes grow, divide, and die.

When these cells begin to divide and grow in an unregulated fashion, a melanoma

tumor results. These tumors, as with all tumors, can either be benign (non-cancerous)

or malignant (cancerous) in nature. Most melanocyte-derived tumors commonly develop

in the skin, however, melanoma can develop anywhere melanocytes are found (e.g. the

eye, meninges, digestive tract, and lymph nodes).

Melanoma is one of the most common types of cancer. It is estimated that in

2009 alone, 68,720 adults in the United States will have been diagnosed with

melanoma, resulting in 8,650 deaths.154 The prevalence of melanoma in Western

countries increases every year. In the United States, Australia, and Europe, melanoma

has been considered an epidemic cancer.155 In fact, the percentage of people

developing melanoma in the United States has more than doubled in the past 30

years.156

People with fair skin are more susceptible towards developing melanoma, and

white people develop melanoma at more than 10 times a higher rate than black

people.154 Although the occurrence of this disease is more likely as individuals age,

Page 136: Florida State University Libraries - FLVC

117

melanoma has been detected at all age groups. People with personal and family

histories of melanoma are at an increased risk, as are individuals with increased

numbers of ordinary moles (benign clusters of melanocytes). Weakened immune

systems, resulting from a number of conditions (e.g. HIV, different forms of cancer, or

drugs prescribed following organ transplantation), increased exposure to ultraviolet

radiation, and sunburns resulting in blistering are all thought to increase the likelihood of

developing melanoma. However, it is not known why a person develops this type of

cancer while others do not, and multiple factors probably give rise to melanoma

tumors.156

The beginning signs and symptoms associated with melanoma often include, but

are not limited to: changes in size, shape, color, or texture of an existing mole. Often

times, these lesions have black or bluish-black areas, they are often referred to as ―ugly

looking‖ moles. Leading cancer research advises self-examinations involving the so-

called, ―ABCDE’s‖ of melanoma:153,154,156 Asymmetry, the shape of part of a mole does

not match the other; Border irregularity, the edges are ragged, notched or blurred;

Color, the color of the mole is not consistent, often shades of tan, brown, blue, pink, red,

black, or white; and Diameter, the mole is larger than ¼ of an inch in diameter, larger

than the size of a pencil eraser; and Evolution, mole has changed in size, shape, color,

or has risen. Although these guidelines for self-examination are general signs of

melanoma, some melanomas do not fit these rules. Only medical professionals can

confirm the presence of melanoma.

The onset of melanoma progresses through various stages of increasing

severity. At stage 0, cells determined to be cancerous melanoma are found only in the

most superficial layers of skin, and have not invaded any of the deeper tissues.

Melanoma is considered to be stage I when the tumor is either no more than 1 mm in

thickness and appears to be ulcerated, or between 1 to 2 mm thick and has no

ulceration. Stage II melanoma is established when the tumor is between 1 to 2 mm thick

and appears with ulceration. In stage I and stage II, the tumor may or may not have

begun to penetrate into the deeper tissues of the skin. However, the cancerous cells

cannot have spread to nearby lymph nodes. At stage III, the signs and symptoms of

melanoma are progressively worse, the cancerous cells having spread (metastasized)

Page 137: Florida State University Libraries - FLVC

118

to nearby lymph nodes or other tissues just outside the original tumor. Finally stage IV

melanoma refers to the condition resulting in the metastasis of cancerous melanoma

cells to lymph nodes and/or tissues distant from the original tumor.156 Once

metastasized, the development of a new tumor in the distant tissues ensues. If this

occurs, the new tumor is still comprised of cancer cells originating from melanocytes,

and the patient is said to have metastatic melanoma.

The treatment of melanoma is usually carried out as a prescribed plan involving

combinations of surgery, chemotherapy, biotherapy, and / or radiation therapy; these

treatments are also integrated with a symptom management program (supportive care,

or palliative care) due to the fact that many of the primary treatments are associated

with negative side effects. Treatment plans are often case dependent and are based on

the age and general health of the patient, as well as the severity of the cancer being

treated.156

The surgical removal of the tumor is the most common practice for melanoma

that is found in the superficial tissues, and is often accompanied by necessary skin

grafts for larger tumors. Typically when surgery is performed the tumor and some of the

surrounding normal tissue is removed for analysis to ensure removal of the cancer in its

entirety. Surgery is also used for the removal of cancerous lymph nodes in the

surrounding area of the original tumor. Although surgery is the most common treatment

for melanoma, it is typically not effective in controlling melanoma that has spread to

other areas of the body.153

Biotherapy (also referred to as immunotherapy) provides assistance to a

person’s immune system, allowing for the body’s natural defenses to aid in fighting

cancer. This type of therapy involves the use of proteins, small molecules, harmless

bacterial microbes, and sometimes even weakened melanoma cells to initiate an

immune response. Commonly used biotherapies are: injections of cytokines (proteins

that activate the immune system) and injections of Interferon-alpha. Both cytokine

proteins and Interferon-alpha are associated with flu-like side-effects that can be severe

in some cases. This type of therapy is often utilized as an adjuvant therapy to limit the

growth and metastasis of any remaining cancerous cells after surgery.153

Page 138: Florida State University Libraries - FLVC

119

Radiation therapy requires the use of high frequency electro-magnetic radiation

to kill cancer cells and reduce the size of tumors. This type of therapy is typically not

used to eradicate cancer, but prevent growth and metastasis. Radiation often results in

general malaise of the patient, diarrhea, upset stomach, and skin irritation among other

side effects. The use of radiation therapy is often reserved for those that have

metastatic melanoma and recurrent melanoma.154

Chemotherapy is the use of chemicals to kill the cancerous cells selectively as

opposed to healthy cells, and is often prescribed for more advanced cases of

melanoma. Although there are many types of pharmacological agents used for the

treatment of patients with stage III and IV melanoma, the prognosis for patients with

metastatic cancer remains very poor; once the cancer has spread to organs and tissues

distant from the originating tumor, the median overall survival rate is approximately 6

months.157 Despite vast research, the use of prescribed chemotherapeutic agents for

the treatment of metastatic melanoma remains marginally beneficial. This form of

cancer is one of the most chemo-resistant.158

Chemotherapy is often administered as single agents (Figure 45) such as:

dacarbazine (DTIC), temozolomide (TMZ), cisplatin, carboplatin, carmustine, lomustine,

docetaxel, and paclitaxel; DTIC and TMZ being the most common.158 In some cases the

chemotherapy is administered as a combination of drugs / therapeutic agents: the

Dartmouth Regimen (DTIC, carmustine, cisplatin, and tamoxifen), CVD (cisplatin,

vinblastine, and DTIC), and BOLD (bleomycin, vincristine, lomustine, and DTIC) being

representative examples.157 The fact that so many chemotherapeutic options are being

used for the treatment of late stage melanoma, substantiates the claim that melanoma

is a chemo-resistant type of cancer. Furthermore, 90 to 95% of patients with advanced

melanoma do not survive more than three years, regardless of treatment modality.159

Page 139: Florida State University Libraries - FLVC

120

HN

N

O

NH2

N

Dacarbazine (DTIC)

NN

CH3

CH3

N

NN

NN

O

OH2N

H3C

Temozolomide (TMZ)

PtCl NH3

Cl NH3

Cisplatin

NH3

NH3

OPt

OO

O

Carboplatin

NH

Cl

O

NCl

NO

Carmustine

NH

O

NCl

NO

Lomustine

OO

O

O

H

HO

HO

OH

ONH

OH

O

O

O

O

OH

OO

O

O

H

HO

O

OH

ONH

OH

O

O

O

O

OHO

Docetaxel Paclitaxel

Figure 45: Several Chemotherapeutic Agents Used in the Treatment of Melanoma.

The limited efficacy of treatment programs for metastatic cancer patients and the

harsh side effects that accompany current cancer treatments provide researchers with

the daunting task of finding effective alternatives to the drugs referenced above that will

selectively target melanoma. This chapter highlights synthetic efforts towards an

interesting new lead compound, palmerolide A, which possesses selective

pharmacological activity towards melanoma cell lines. It describes a highly efficient and

convergent method to generate the eastern hemisphere (C1-C15) of palmerolide A.

Page 140: Florida State University Libraries - FLVC

121

Palmerolide A

Chemical defense mechanisms are of the utmost importance in marine

ecosystems.160-164 Sessile marine invertebrates, much like marine plants, are

particularly susceptible to predation, fouling by settling larvae, diatoms, algae, and

overgrowth by competing species for space and resources. Most research on the

chemical ecology of marine invertebrate communities has focused on tropical regions

because of their high levels of species diversity and density. Not surprisingly, members

of these communities have evolved interesting chemical defenses.160 Many of the initial

studies focusing on geographic patterns of chemical defense in marine invertebrates

found that there was an inverse relationship between species chemically defended

against predatory fish and latitude.165-167 This may be, in part, responsible for the focus

of chemical ecology remaining on tropical locations, leaving benthic communities in the

Antarctic to be relatively understudied.

Another possible reason for the lack of research in the marine habitats of

Antarctica is that it remains one of the least accessible marine environments. However,

outside of the shallow waters, where anchor ice and ice scour dominate the landscape

(< 33 m depth),168 exist diverse and stable communities of invertebrates.169 The

Antarctic, cold adapted, marine ecosystem has been largely isolated for approximately

20 million years. Many of the marine organisms that are found in these communities

emerged prior to the breakup of Gondwanaland and the movement of the continent to

the pole; the biological isolation of the Antarctic marine ecosystem is perpetuated by the

Antarctic Polar Front, an oceanic water current encircling the continent.170

The predation of fish on the sessile invertebrates is rather rare at higher latitudes,

limiting the need for chemical defenses against vertebrate predators.171 However,

predation in the Antarctic benthos is dominated by mobile invertebrates like sea stars.172

These waters harbor of some of the oldest, most environmentally and biologically stable

marine environments, making them well suited for the evolution of chemical defense

mechanisms against such predators.164 Indeed, the fact that the marine ecosystem

around Antarctica is largely isolated from subtropical and temperate waters, in addition

to the fact that mobile invertebrates control the predatory landscape of sessile

Page 141: Florida State University Libraries - FLVC

122

invertebrates, has led to an environment where interspecies chemical warfare plays a

pivotal role in survival.173 The predation suppression exhibited by members of the

alcyopterosins (discussed in Chapter 3) may be a representative example.109

For these reasons, natural products research in Antarctica has the potential to

produce tantalizing leads for drug discovery and development. Although it is up to

isolation chemists and marine ecologists to ascertain and characterize these

compounds, it falls upon synthetic chemists to translate promising leads into viable drug

development candidates. This responsibility is solely reserved for the synthetic

community because of an international treaty that prohibits the exploitation of Antarctic

resources for commercial development.174

Palmerolide A (94)175 was isolated by Bill Baker and co-workers from the

Antarctic tunicate Synoicum adareanum. It represents one of the most exciting synthetic

challenges in natural products organic chemistry today. Palmerolide A is a potent

inhibitor of vacuolar ATPase proton pumps (IC50 = 2 nM),175 which are highly expressed

in metastatic cancer cells176 where they modulate pH. V-ATPases are also the target of

several other interesting cytotoxic natural products including salicylihalamide A,177,178

bafilomycin A1,179 and oximidines.180 What’s more, palmerolide A was found to exhibit

cytotoxicity three orders of magnitude greater towards the melanoma UACC-62 cell line

(LC50 = 18 nm)175 when compared to the rest of the NCI’s 60-cell line panel. Figure 46

depicts data from the report issued to Baker and co-workers from the NCI,181 the larger

region highlighted in red shows the promising cytostatic activity demonstrated by

palmerolide A (most notably against leukemia, colon, melanoma, renal and breast

cancer). The smaller region highlighted in green provides the evidence of the natural

product’s ability to kill melanoma cells selectively.

Page 142: Florida State University Libraries - FLVC

123

Figure 46: The Report Issued to Baker From the National Cancer Institute’s 60-Cell Line Panel Toxicity Assay for Palmerolide A.181

Baker and co-workers have since isolated several other members of the family of

palmerolide natural products, some of which inhibits V-ATPase activity (Figure 47).181

Page 143: Florida State University Libraries - FLVC

124

Most notably, these compounds vary from the 94 in the position of the C8, C9 olefin

(palmerolide A 94 vs. B, C, and H), and their enamide side chain (palmerolide A vs. D,

E, F, and H); palmerolide E lacks the enamide moiety altogether.

O

O

OH

N

O

HO

O

O

NH2

Palmerolide A (94)

V-ATPase (IC50 = 2 nM)

O

O

O

N

O

OSO3

Palmerolide B

O

NH2HO

O

O

OH

N

O

HO

O

O

NH2

Palmerolide D

O

O

N

O

Palmerolide C

V-ATPase (IC50 = 150 nM)OH

O

OH

O

NH2

O

O

OH

HO

O

O

NH2

Palmerolide E

V-ATPase (IC50 = 6.5 M)

O

H

O

O

OH

N

O

HO

O

O

NH2

Palmerolide F

V-ATPase (IC50 = 62.5 nM)

O

O

O

N

O

OSO3Palmerolide H

O

NH2HO

O

O

OH

HO

O

O

NH2

NHO

1

7

11

15

1924

Palmerolide E

V-ATPase (IC50 = 10M)

Figure 47: Palmerolide A and Other Members of the Palmerolide Natural Products With

Major Distinctions Highlighted in Red Ovals.181

Page 144: Florida State University Libraries - FLVC

125

This natural product has attracted several synthetic efforts182-188 due to its

exciting biological activity and interesting structure. The chemical synthesis of

palmerolide A requires one to address several independent challenges. Hydrocarbon

regions isolate several stereocenters within the macrocyclic core, making it ideal for

convergent fragment assembly strategies. Most synthetic approaches are focused on a

few strategic bond disconnections: generation of the C15-C16 bond using a transition

metal catalyzed coupling reaction, an esterification reaction to provide the C1-Oxygen

bond, closure of the macrolide at the C8-C9 bond, and late stage enamide installation

(Figure 48).

O

O

OH

N

O

HO

O

O

NH2

Palmerolide A (94)

1

7

1115

1924

Esterification

EnamideInstallation

HWE or RCMCouplingReaction

Figure 48: Palmerolide A and Strategic Disconnections.

In 2007, the labs of Jef De Brabander181 and K. C. Nicolaou182 independently

reported the total synthesis (and structural reassignment) of palmerolide A. The De

Brabander synthesis began with a chiral vinylogous Mukaiyama aldol reaction between

known vinylketene silyl N,O-acetal 95189 and aldehyde 96190 to provide alcohol 97 in a

13:1 diastereomeric ratio and 80% yield. A Mitsunobu inversion, followed by

simultaneous reduction of the resulting benzoyl ester and chiral auxiliary, provided an

aldehyde that was homologated using Ph3PCHCO2Me. This sequence provided the

C16-C24 fragment of the palmerolide A (98), and contained the necessary

stereochemistry at C19 and C20 as well as the 16,17 Z-olefin (Figure 49).

Page 145: Florida State University Libraries - FLVC

126

N

TBSOMe

Me

O

O+

OHC

I

Me

TiCl4, CH2Cl2-78 oC

dr = 13:180%

N

O

Me Me

OH

I

Me

O

O

i-Pr

1. PhCO2H, DEAD, Ph3P2. DIBAL, CH2Cl2

3. Ph3PCHCO2Me 61% over 3-steps

O

Me

Me Me

OH

I

Me

95 96 97

98

Figure 49: De Brabander’s Synthesis of the C16-C24 Fragment of Palmerolide A.

De Brabander and co-workers began their synthesis of the C9-C15 fragment of

palmerolide A with D-arabitol (99). The formation of a 1,3-benzylidene acetal and

oxidative cleavage of the resulting 1,2-diol provided aldehyde 100.191 Conversion of

aldehyde 100 to aldehyde 101 was carried out using a series of standard reactions. The

condensation of aldehyde 101 with pinacol dichloromethylborane completed the

synthesis of the C9-C15 fragment (Figure 50).192

HO

HO

HO

OH

OHref. 190 O

O

OH

CHOPh

99 100

1. Ph3PCHCO2Me2. TIPSOTf, 2,6-Lutidine3. Pd/C, H2, EtOAc

4. TESCl, imidizole

5. DIBAL, -78 oC

81% over 5 steps

TESO

TESO

OHC

OTIPS

101

O BO

CHCl2

CrCl2, LiI, r.t.84%

102

OTES

RO

OTES

R = TIPS

PinB

Figure 50: De Brabander’s Synthesis of the C9-C15 Fragment of Palmerolide A.

Page 146: Florida State University Libraries - FLVC

127

The synthesis of the C1-C8 fragment began with -valerolactone (103). Upon

methanolysis and Swern oxidation, 103 was converted to aldehyde 104, and

subsequently olefinated. The resulting t-butyl ester was hydrolyzed under acidic

conditions and a Claisen-type condensation, using dimethyl methylphosphonate was

carried out to afford the C1-C8 fragment (105) (Figure 51).

O

O1. MeOH, H2SO4,

reflux

2. Swern OxidationRef. 191

72%

OHC

OMeO

103 104

1. Ph3PCHCO2t-Bu2. TFA, CH2Cl2

3. n-BuLi, (MeO)2P(O)Me,

THF, -78 to 0 oC

82% over 3 stepsO

(MeO)2P

O

O

HO

105

Figure 51: De Brabander’s Synthesis of the C1-C8 Fragment of Palmerolide A.

Having synthesized the necessary fragments, vinyl iodide 98 and vinylboronate

102 were coupled through a Suzuki coupling reaction to provide 106. Subsequent

Yamaguchi esterification194 of fragment 105 and 106, followed by cleavage of the

triethysilyl ethers provided compound 107. A selective oxidation195 of the primary

alcohol and an intramolecular Horner-Wadsworth-Emmons reaction (HWE)196

established the macrocyclic core of palmerolide A (108). The macrocycle was then

transformed into 109 through a series of several steps: a CBS-reduction197 that provided

the C7 stereochemistry (dr = 4:1); installation of the enamide via a Curtius

rearrangement at C24, followed by addition of 2-methyl-propenylmagnesium bromide to

the resulting isocyanate; carbamate synthesis at the C11 oxygen, and finally, global

deprotection (Figure 52). The spectroscopic data of compound 109 proved to be

inconsistent with the natural isolate. The De Brabander lab then carried out the same

sequence of reactions utilizing the enantiomer of vinylboronate fragment 102 which

provided a compound with the identical NMR spectrum to that of palmerolide A.

However, the circular dichroism (CD) spectrum proved to be the mirror image to that of

the natural product. De Brabander and co-workers had successfully completed the

synthesis of the originally proposed compound as well as the unnatural enantiomer.

Page 147: Florida State University Libraries - FLVC

128

Their synthesis employed a relatively straightforward approach; notably a HWE

macrocyclization, a selective vinylogous Mukaiyama aldol reaction, and a Curtius

rearrangement reaction were utilized to provide the first synthesis of (-)-palmerolide A.

98 + 102

cat. Pd(PPh3)4, Tl2CO3, THF, H2O

79% RO

OTES

OH

MeO2C

OTESR = TIPS

106

1. Yamaguchi conditions,105, 69%

2. PPTS, MeOH, 0 oC, 95%

RO

OH

O

MeO2C

OHR = TIPS

O

OP(OMe)2

O1. PhI(OAc)2, TEMPO

2. K2CO3, 18-Crown-6,

PhMe, 60 oC

70% over 2 steps

RO

OH

O

MeO2C

R = TIPS

O

O

107 108

HO

O

O

O

OH

Diastereomer of Palmerolide A109

HN

O

O

NH2

18% over 9 steps

Figure 52: Completion of De Brabander’s Synthesis of 109, a Diastereomer of Palmerolide A.

K. C. Nicolaou’s lab, having made the same disconnections as De Brabander,

initiated their synthesis through the preparation of the C16-C23 (112) and C15-C8 (116)

fragments (Figure 53).183 Nicolaou and co-workers, like De Brabander’s group, also

utilized vinyl iodide 96 in their synthesis, however they chose to perform an Evans’ aldol

reaction using imide 110 to set the C19 and C20 chiral centers (95% de), providing

111.197 The aldol reaction was followed by several standard reactions to reach the C16-

Page 148: Florida State University Libraries - FLVC

129

C23 fragment (112). The chiral centers at C10 and C11 were established through the

reaction of aldehyde 113 (2 steps from 4-pentyn-1-ol)198 and [(Z)--(methoxy-

methoxy)allyl]-(-)-diisopinocampheylborane (114),199 which afforded 115 upon

desilylation in 74% yield (>95% de, >90% ee). Hydroxy acetylene 115 was converted to

vinylstannane 116, first through carbamate installation,200 followed by a standard

manipulation of the acetylene moiety.201

NO

O O

Bn

96, n-Bu2BOTf,Et3N

46%( 95% de)

NO

110

O O OH

IBn

111

OTBS

I

OH

112

CHO

TBS MOMO

B[(-)-Ipc]2

113

1141.

2. K2CO3, MeOH74%

(>95% de, >90% ee)

20% over 7 steps

O

OH

MOM

115

Bu3Sn

O

O

MOM

O NH2116

62% over 3 steps

Figure 53: The Synthesis of Nicolaou’s C16-C23 (112) and C8-C15 (116) Fragments.

TBS protected 5-hexene-1-ol (117) served as the starting material for the

synthesis of Nicolaou’s C1-C8 fragment (120). Upon epoxidation and chiral resolution,

using the Jacobsen method,202 117 was converted to 118 in 42% yield (>99 ee). Ring-

opening of the epoxide using a sulfur ylide provided allylic alcohol 119, and a series of

reactions (protection, deprotection, oxidation, olefination, and saponification) provided

the acid fragment 120 in 59% from 119 (Figure 54).

Page 149: Florida State University Libraries - FLVC

130

TBSO1. m-CPBA

2. (R,R)-Jacobsen, Co(II), cat. AcOH, H2O 42% (>99% ee) over 2 steps

TBSO

O

117 118

I

n-BuLi

TBSO

OH

119

1. MOMCl, DIPEA, 85%

2. TBAF, 95%

3. DMP, NaHCO3, 95%

4. Ph3PCHCO2Me, 90%

5. KOH, 85%

HO2C

OMOM

120

Me3S

Figure 54: Nicolaou’s Synthesis of the C1-C8 Fragment of Palmerolide A.

The Nicolaou lab, having each of the key fragments in hand, turned their

attention towards the assemblage of the fragments and their elaboration into the

originally proposed structure of palmerolide A (Figure 55). A Stille reaction203 between

vinyl iodide 112 and vinylstannane 116, followed by a Yamaguchi esterification194 of the

resulting alcohol and acid 120 provided cyclization precursor 121. Conversion of the

allylic silyl ether into vinyl iodide 122 was carried out in a three step sequence:

deprotection, oxidation, and olefination.204 Removal of the MOM protecting groups and

a ring closing metathesis reaction35 led to the formation of the C8-C11 olefin and the

macrocyclic core of palmerolide A (123). From vinyl iodide 123, enamide installation205

completed the synthesis of the proposed structure of palmerolide A (109).

Having reached 109, Nicolaou came to the same conclusion as the De

Brabander group: the originally proposed structure had been assigned incorrectly. Their

synthetic scheme allowed for the synthesis of the naturally occurring enantiomer simply

by inserting ent-116 and ent-120 into their already established route. They completed

the synthesis of natural palmerolide A (94) with similar yields. Their product exhibited

identical analytical data to those of the natural isolate, including the CD spectrum.

Page 150: Florida State University Libraries - FLVC

131

112 + 116

1. cat. [Pd(dba)2],AsPh3, LiCl, 67%

2. Yamaguchi conditions, 120, 61%

O

OR

RO

O

OTBS

O

O NH2121

1. TBAF2. DMP, NaHCO3

3. CrCl2, CHI3

63% (>95:5 E/Z)

O

OR

RO

O

O

O NH2122

2. Grubbs II,35 76%

O

OH

HO

O

O

123

I I

1. BF3 OEt2, 46%

O

NH2

CuI, Cs2CO3

O

NH2

109

R = MOM

R = MOM

Figure 55: Nicolaou’s End-Game strategy for the Synthesis of 109.

Nicolaou’s synthesis of palmerolide A provided a very flexible route to the natural

product.183 It allowed the Nicolaou lab to synthesize various isomers as well as some

interesting analogs thereof.206,207 In the course of their research they performed some

structure-activity relationship studies (SAR) studies, identifying some structural

characteristics that influence cytotoxic activity. Figure 56 provides some representative

examples of the analogs developed by Nicolaou.207

Page 151: Florida State University Libraries - FLVC

132

R

O

O

HO

O

OH

O

NH2

HN

O

O

OH

OH

HN

O

O

HO

O

OH

O

NH2

OO

Palmerolide A (94)

HO

124

127: R =Me

O

NH

128: R =

O

NH

N

129 R =

O

NH

N

130: R =NH

N

S

Me

O

131: R =

O

NH

132: R =

O

NH

HN

O

O

O

OH

O

NH2

O

HN

O

O

HO

O

O

NH2

O

126125

Figure 56: Key Analogs of Palmerolide A Developed by the Nicolaou Lab.

The synthetic analogs were tested for cytostatic activity against a panel of seven

different cancer cell lines, including breast (MCF-7), melanoma (UACC-62), CNS

(SF268), lung (NCI-H460), ovarian (1A9), Taxol-resistant ovarian (PTX22), and

epothilone-resistant ovarian (A8) cells. The synthetic compounds tested were compared

to Taxol, doxorubicin, and natural palmerolide A (94). Table 6 summarizes selected

data (the less informative examples have been omitted).

Page 152: Florida State University Libraries - FLVC

133

Table 6: Selected Data from Nicolaou’s SAR Study (GI50 Values in M).

Entry Compound Cell Line

UACC-62 MCF-7 SF268 NCI-H460 IA9 PTX22 A8

1 doxorubicin 0.294 + 0.141 0.056 + 0.005 0.129 + 0.048 0.008 + 0.001 0.033 + 0.007 0.201 + 0.049 0.051 + 0.017

2 Taxol 0.022 + 0.016 0.006 + 0.001 0.026 + 0.141 0.007 + 0.001 0.006 + 0.001 0.079 + 0.001 0.021 + 0.015

3 natural 94 0.057 + 0.007 0.040 + 0.007 0.030 + 0.012 0.010 + 0.001 0.038 + 0.003 0.066 + 0.007 0.018 + 0.003

4 synthetic 94 0.062 + 0.001 0.065 + 0.011 0.048 + 0.006 0.017 + 0.004 0.059 + 0.001 0.073 + 0.005 0.049 + 0.004

5 ent-94 8.077 + 0.194 6.260 + 0.171 9.475 + 0.593 6.589 + 0.054 >10 >10 8.844 + 1.301

6 109 >10 >10 >10 >10 >10 >10 >10

7 124 0.322 + 0.088 0.200 + 0.026 0.281 + 0.118 0.075 + 0.003 0.288 + 0.017 0.627 + 0.016 0.083 + 0.006

8 125 6.979 + 0.531 7.585 + 0.252 8.764 + 0.315 6.396 + 0.106 7.135 + 0.667 8.062 + 0.037 6.691 + 0.439

9 126 0.063 + 0.001 0.074 + 0.000 0.060 + 0.004 0.055 + 0.002 0.072 + 0.001 0.076 + 0.000 0.061 + 0.013

10 127 >10 >10 >10 7.291 + 0.137 7.774 + 1.094 >10 6.700 + 0.411

11 128 0.641 + 0.000 0.755 + 0.004 0.592 + 0.007 0.430 + 0.047 0.618 + 0.051 0.741 + 0.003 0.460 + 0.042

12 129 0.735 + 0.084 0.796 + 0.166 0.491 + 0.132 0.078 + 0.001 0.378 + 0.141 0.889 + 0.029 0.072 + 0.004

13 130 8.822 + 0.083 7.397 + 0.262 >10 3.796 + 0.306 7.944 + 0.430 >10 3.514 + 1.379

14 131 0.009 + 0.001 0.007 + 0.000 0.007 + 0.001 0.007 + 0.001 0.009 + 0.001 0.039 + 0.002 0.006 + 0.000

15 132 0.067 + 0.000 0.071 + 0.008 0.054 + 0.000 0.061 + 0.000 0.067 + 0.002 0.081 + 0.006 0.057 + 0.001

Page 153: Florida State University Libraries - FLVC

134

Several interesting discoveries resulted from the SAR studies performed by

Nicolaou and co-workers. As expected, both the synthetic palmerolide A and the natural

isolate demonstrated similarly potent activities across all cell lines tested, whereas the

enantiomer was more than 100-fold less active (entries 3-5). The originally proposed

diastereomer (compound 109) was practically inactive (entry 6). Removal of the

carbamate moiety from C11 oxygen (compound 124) provided a mild decrease in

activity (ca. 5-fold, entry 7). Their results also provided evidence that the C10 hydroxyl

was necessary for reactivity (entry 8, compound 125). However, the compound lacking

the C7 alcohol was comparable to palmerolide A (entry 9, compound 126).

The enamide side chain also had a substantial effect on the ability of the analogs

to inhibit the growth of the cancer cells examined. Replacing the isobutenyl group of

palmerolide A with a methyl group decreased potency by more than two orders of

magnitude, but when it was replaced by an isobutyl group, the analog retained most of

its activity (entries 10 and 15, respectively). Polar aromatic enamide analogs of

palmerolide A (compounds 128-130) retained some activity, but perhaps most

interesting was the result of analog 131; this compound was found to have a 10-fold

increase in the activity against some of the cell lines (entry 14).

Several other groups have developed synthetic methods towards the synthesis of

various fragments of palmerolide A, envisioning similar strategies to install the

stereocenters and close the macrolide as De Brabander and Nicoloau,186-188 the labs of

Maier184 and Hall185 provided some interesting alternatives.

The Maier group provided a convergent synthesis of the C3-C15 fragment (134)

beginning with ester 133 (from -valerolactone). They utilized a Noyori asymmetric

hydrogenation208 to set the C7 stereocenter, a Sharpless asymmetric dihydroxylation209

to provide the C10 and C11 oxygens, and an Ohira-Bestmann reaction210 to install the

unsaturation of the C14-C15 bond. The stereocenters of fragment 135 were generated

analogously to Nicolaou’s synthesis of the similar fragment.184 An olefination reaction

was carried out between fragments 134 and 135 establishing the C2-C3 bond

(compound 136). After having attempted an intramolecular Stille reaction which resulted

in E/Z mixture of the C14-C15 olefin, the macrolactone (137) was synthesized

stereoselectively through an intramolecular Heck reaction.211 Maier converted 137

Page 154: Florida State University Libraries - FLVC

135

through a series of steps into Nicolaou’s late stage intermediate (123), completing the

formal synthesis of palmerolide A (Figure 57).

OPMB

CO2Me

133

O

H

3

7

TBDPSO

O

OTBS

O

NH2

R = TBDPS134

+7

1115

O

I

1920

O

P(OEt)2

O

135

3

9% over 18 stepsLiCl, i-Pr2NEt,

MeCN92%

O

O

I

MeO2C

MeO2C

OTBS

RO

O

O

NH2

R = TBDPS136

Pd(OAc)2, CsCO3,

Et3N, DMF81%

O

OMeO2C

OTBS

RO

O

O

NH2

R = TBDPS136

123

6 steps(ca. 34%)

Figure 57: Key reactions in Maier’s Formal Synthesis of Palmerolide A.

The most recent synthesis, perhaps the most elegant, provided a route to

palmerolide A that incorporated asymmetric catalysis as a key feature. Hall and co-

workers185 applied an asymmetric E-crotylboration that had been developed in their lab

to install the C19 and C20 stereocenters. The reaction involved aldehyde 96, and was

catalyzed by SnCl4, using a p-F-Vivol[7] ligand (137)212 in 95% yield and 90% ee (>95:5

dr). The resulting alcohol (138) was then carried into their C16-C24 fragment (139)

(Figure 58).

Page 155: Florida State University Libraries - FLVC

136

I

O

H

Bpin

Na2CO3, 4 A MS,

toluene, -78 oC, 60 h

95% (>95:5 dr, 90% ee)

OH

I

94

137 SnCl4t-BuO2C

OH

138 139

FF

HO OH

137(R,R)-p-F-Vivol[ 7]

1920

I

Figure 58: Hall’s Asymmetric Crotylboration en Route to Palmerolide A’s C16-C24 Fragment.

The synthesis of the C1-C13 fragment (146) also employed a catalytic

enantioselective reaction developed in the Hall lab; a two step hetero [4+2]

cycloaddition / allylboration sequence involving 140 and enol ether 141 was used to set

the C7 stereocenter.213 This was followed by an esterification using acid 143 to set up

an unprecedented [3+3] B-Claisen-Ireland214 rearrangement, to provide both the C10

and C11 centers with the syn-configuration; subsequent oxidation and esterification led

to the formation of 145 (55%, 142 145), which was thus converted to the fragment

146 (Figure 59).

Page 156: Florida State University Libraries - FLVC

137

Bpin

O OEt

+

(a) Jacobsen's HDAcatalyst (ref. 214)

(1 mol %), BaO, THF, 14 h;

140 141

(b) 141, 2 h84% (96% ee)

BpinO OEt

OH(10 equiv)

HO

OOPMB

142

143

EDC-Cl, DMAP, CH2Cl2

1a. LDA (2.1 equiv)

THF, -78 oC;

1b.TMSCl, Et3N,

-100 oC, 2 h; -78 oC, 12 h;

O

pyranyl

Bpin

PMBOOTMS

i-Pr2N

1c. NaOAc, H2O2,

THF, 0 oC, 2 h

2. CH2N255% from 142

3

7

10

BpinO OEt

O

144

3

7

10

O

PMBO

O OEt

145

3

7

10

OH

OPMB

O

MeO11

MeO

O

10

11 OPMB

TIPSO

1

3

OTIPS7

146

B

Figure 59: Hall’s Unique Approach to Install the C7, C10, and C11 Stereocenters.

Hall completed the macrolide of palmerolide A through the use of a B-alkyl

Suzuki coupling reaction of vinyl iodide 139 and compound 146, followed by a

saponification of the methyl ester and a Yamaguchi macrolactonization. Hall’s lab then

installed the enamide by analogy to De Brabander’s Curtius rearrangement,182

selectively deprotected the PMB alcohol allowing for the installation of the carbamate

moiety. Palmerolide A was realized upon deprotection of the silyl ethers. Hall’s route

provided an aesthetically pleasing synthesis, incorporating three very interesting

reactions to set the required stereochemistry, and obtained palmerolide A in 0.8%

overall yield in 21 linear steps.

Palmerolide A has generated a great deal of excitement in the synthetic

community. Its promising biological activity and interesting structure have inspired

numerous labs to select palmerolide as a target, either as a testing ground for

Page 157: Florida State University Libraries - FLVC

138

methodology, or to provide efficient synthetic strategies for the benefit of the research

community at large. The efforts described above have provided instructive tools for the

synthesis of the various fragments and the core macrolide. Notably, the use of aldehyde

94 has been incorporated into each of the syntheses for the creation of the C19 and

C20 stereocenters. Moreover, the synthesis of the enamide side chain has involved

either a Curtius rearrangement strategy or a copper-catalyzed reaction with a vinyl

iodide.

The next section will provide some details of our proposed synthesis of this

promising natural product and recent developments in our fragmentation methodology.

This chemistry has allowed us to devise and carry out a concise synthesis of the C1-

C15 fragment of palmerolide A. Although there is still much work that needs to be done

before reaching our goal of the total synthesis of palmerolide A, our chemistry provides

new alternatives for the methods described above and contributes valuable information

to the synthetic community.

Synthesis of the Eastern Hemisphere of Palmerolide A

Since the expansion of the carbanion-triggered fragmentation reactions of

vinylogous acyl triflates to alkyl Grignard reagents,93 the scope of the reaction had been

increased dramatically.216,217 As discussed in Chapter 1, the bond cleavage pathway is

reminiscent of the Eschenmoser-Tanabe49-52 and related Grob-type fragmentations,39-43

but with a broader scope: the Eschenmoser-Tanabe fragmentation is limited to the

synthesis of alkynyl ketones and aldehydes from cyclic enones, whereas vinylogous

acyl triflates allow for the synthesis of a diverse range of carbonyl derivatives (Figure

60).217 The two step process—synthesis and fragmentation of vinylogous acyl triflates—

enables the conversion of symmetric 1,3-diones into acyclic, differentially functionalized

building blocks.

Page 158: Florida State University Libraries - FLVC

139

O

Me

OTf

R M

R

O Me

NHPh

O Me

Ph

O Me

CH2Ph

O Me

O MeO Me O Me

O

OEt

SS

Me

PO

MeO OMe

84% 93% 73%

21%(needed for this study)74% 88%

Bu

O Me

76%

2

Figure 60: Brief Overview of the Addition / Bond Cleavage Reactions of Vinylogous Acyl Triflate 2.

Tummatorn and Dudley recently provided access to homopropargyl alcohols

from -keto lactones (heterocyclic diones) through a related process (Figure 61).218

Conversion of heterocyclic diones (147) to the corresponding 5,6-dihydro-2-pyrone

(DHP) triflates (148) occurs in excellent yields using a similar procedure to that of their

carbocyclic analogs. Treatment of 148 with two equivalents of methyl Grignard in

toluene at -78 oC, and subsequent warming, provides good to excellent yields of the

corresponding homopropargyl alcohol 148 with retention of configuration.

Page 159: Florida State University Libraries - FLVC

140

O

O

O

R1

R2

R4

R3

MeMgBr (2.0 equiv)

- 78 oC to 60 oC

73% to quant.R2

R3

R4 OH R1O

O

OTf

R1

R2

R4

R3

Tf2O, -78 oC

Et3N, CH2Cl2

147 148

>90%

149

Figure 61: Synthesis and Nucleophile-Triggered Decomposition of DHP Triflates.

As stated above, palmerolide A is a natural product that is ideal for convergent

fragment assembly strategies. As a logical consequence, the efficient synthesis of the

key fragments becomes of a priority. We envisioned the bond cleavage methodology

developed in our lab as being well suited for synthesis of palmerolide A’s key fragments.

Our retrosynthetic analysis includes similar initial disconnections to those of De

Brabander, Nicolaou, and Maier (Figure 62). For this discussion, the focus will remain

on the eastern hemisphere (C1-C15) of palmerolide A. For the synthesis of this region,

we imagined our C1-C8 fragments originating from vinylogous acyl triflate 2 through a

Claisen-type fragmentation reaction using a phosphonate nucleophile, setting up an

olefination of aldehyde 150 to form the C8-C9 bond.

OH

O

O

HO

O

O

NH2

7

10

11

1920

O

O

OTf1920

16

16

15

O

OTf

7

3

CHOPO

OP

10

11

15

3

Palmerolide A (94)

2

150151

HN

O

PO

Figure 62: Retrosynthetic Analysis of Palmerolide A Using a Fragmentation Approach.

Page 160: Florida State University Libraries - FLVC

141

The Claisen-type condensation of vinylogous acyl triflates216 was initially

optimized by Kamijo and Dudley using the lithium enolate of acetophenone as the

nucleophile trigger. As shown in Table 7, excess enolate (2.2 equiv) was required for

complete conversion (entries 1 and 2). The direct extension of this methodology to the

synthesis of -keto phosphonates was performed, but the anion of dimethyl

methylphosphonate provided fragmentation product 153b in yields that were not

practical for complex molecule synthesis (entry 3, Kamijo and Dudley). Reoptimizing

this system for the preparation of olefination reagents revealed an advantage of the

phosphine oxide over the phosphonate (entries 3 and 4, this work, Jones and Dudley).

In contrast to enolate nucleophiles, the addition / bond cleavage reaction of a lithiated

phosphine oxide required only 1.1 equivalents of the nucleophile (entry 2 vs. entry 5).

Table 7: Claisen-Type Condensations of Vinylogous Acyl Triflate 2.

O

Me

OTfO Me

2

[EWG CH2] Li

EWG

THF

153

entry [EWG CH2] Li

(152)

Equiv of 152

153 Yield of

153

1a OLi

Ph 1.2 153a 56%

2a OLi

Ph 2.2 153a 85%

3a P

O

OMeOMe

Li

2.2 153b 21%

4b P

O

PhPh

Li

2.2 153c 75%

5b P

O

PhPh

Li

1.1 153c 81-89%

a Reproduced from reference 216.

b See Experimental information.

Page 161: Florida State University Libraries - FLVC

142

Having optimized the fragmentation reaction for the synthesis of olefination

reagents, we turned our attention towards the synthesis of the C1-C18 portion (155) of

palmerolide A (Figure 63). Lindlar hydrogenation of 153c reduced the alkyne to afford

the corresponding Z-olefin (154), which was subjected to olefin-cross metathesis35 with

ethyl acrylate. The best results for our metathesis reaction were obtained using 4 mol %

of the Grubbs’ second generation catalyst and a substoichiometric amount (15 mol %)

of titanium(IV) isopropoxide219,220 at 100 oC in a sealed tube as a solution in methylene

chloride. Ti(Oi-Pr)4 is presumed to coordinate to the -ketophosphine oxide, preventing

chelation to the ruthenium metal center which may inhibit metathesis.

O

Me

O

1. Tf2O, Pyridine95-100%

2. Ph2P(O)CH2Li

THF, -78 to 60 oC

81-89%1

H2,Pd(CaCO3-Pb)

MeOH/pyridine96%

OP

O

PhPh

153c

OP

O

PhPh

154

Ethyl Acrylate,Grubbs' II (4 mol %)

Ti(Oi-Pr)4, CH2Cl2100 oC, 89%

OP

O

PhPh

155

EtO2C

Figure 63: Synthesis of the C1-C8 Olefination Reagent for the Synthesis of Palmerolide A.

Synthesis of aldehyde partner 150 began with a Sharpless asymmetric

dihydroxylation209 of ,-unsaturated ester 156 (Figure 64).221 The syn-diol (157) was

obtained in 75% yield (99.6% ee), and was subsequently converted to acetonide 158

using acetone as the solvent (the reaction did not go to completion in CH2Cl2).

Controlled ester reduction using diisobutylaluminum hydride provided aldehyde 150,

which was used immediately in a Horner-Wittig olefination reaction (155 + 150 161).

Page 162: Florida State University Libraries - FLVC

143

OEt

O

AD-mix-, MeSO2NH2,

H2O, t-BuOH (1:1), 0 oC

75% (99.6% ee)

OEt

O

HO

OH

2,2-dimethoxypropane

CSA, acetone99%

OEt

O

O

O

DIBAL

CH2Cl2, -78 oC

(95-100%)

H

O

O

O

156 157

158 150

Figure 64: Synthesis of Aldehyde 150, the C9-C15 Fragment of Palmerolide A.

We investigated the coupling of fragments 155 and 150 to afford the 8,9-olefin

employing several different conditions commonly used to perform olefination reactions,

including: Ba(OH)2, DBU•LiCl (Masamune-Roush conditions), and t-BuOK. In each

case, the desired product was not observed and a mysterious byproduct was observed.

Although the 1H NMR spectrum was difficult to interpret, the vinyl protons of the enoate

were no longer present. This caused us to be concerned with the potential for an

intramolecular Michael addition reaction of the -ketophosphine oxide onto the tethered

enoate (Figure 65). A common feature of all of these bases is that each is of

intermediate basicity between the initial -ketophosphine oxide anion (159) and the

enolate resulting from the undesired cyclization onto the enoate (160). The conjugate

acid may therefore play a role in promoting a cyclization.

Page 163: Florida State University Libraries - FLVC

144

OP

O

PhPh

O

EtO

HB

- B H

EtO2C

Ph2P

OOPh2P

O

EtO2C

Ph2P

EtO2C

O

155

159 160

"H+"

B: = Ba(OH)2, DBU LiCl (Masamune-Roush conditions), and t-BuOK

Figure 65: Possible Michael Addition Reaction of 155.

The combination of irreversible base (NaH) and a slight excess of aldehyde 150

provided the best results for our desired olefination, affording enone 161 in 89% yield

(Figure 66). Following olefination, a CBS-reduction222 of enone 161 provided the C7

alcohol in 89% yield, albeit with only modest diastereoselectivity (ca. 75:25 dr). The

selectivity is surprising in light of a similar CBS-reduction for which Chandrasekhar

observed 97% de.187 On the other hand, our observations are in line with the 4:1 dr

reported by De Brabander for the CBS-reduction at the C7 position of a macrocyclic

precursor to palmerolide A.182 The stereoselective reduction of the C7 ketone remains

an open challenge as we continue with our studies. TBS-protection of 162 under

standard conditions furnished our C1-C15 fragment (163) of palmerolide A that will be

utilized en route to the natural product.

155

NaH (1.0 equiv)

THF, 0 oC;

Then 150 (1.5 equiv),

0 oC to r.t., 89% O

O

O

EtO2C

161

1. (R)-CBS, THF89% (ca. 75:25 dr)

2. TBSOTf, 2,6-LutidineCH2Cl2, 94% O

O

OR

EtO2C

162: R = H163: R = TBS

Figure 66: Completion of the Eastern Hemisphere (C1-C15) Fragment Synthesis.

Page 164: Florida State University Libraries - FLVC

145

In summary, we have prepared 163, which comprises the eastern hemisphere

(C1-C15) of palmerolide A, in 7 linear steps (approximately 42% overall yield) from

unsymmetrical dione 1. The optimized addition / bond cleavage reaction (2 153c)

provides efficient entry into a comparatively short synthesis of a C1-C8 olefination

reagent for the convergent coupling with aldehyde 150. Aldehyde 150 is prepared in

three steps and >99% ee from ester 156.221 This sequence highlights yet another

example of the versatility of vinylogous acyl triflates in complex molecule synthesis, and

demonstrates a marked improvement over our previously published Claisen-type

condensations of 2.216 Chapter 5 will provide mechanistic insight into similar Claisen-

type condensation reactions, as well as additional interesting olefination reagents.

Experimental

General information:

1H NMR and 13C NMR spectra were recorded on a 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 and 77.0 ppm for 13C NMR)

for all compounds. The coupling constants (J) are reported in Hertz (Hz). IR spectra

were recorded on a Perkin-Elmer FT-IR spectrometer with diamond ATR accessory as

thin film. Mass spectra were recorded using chemical ionization (CI), electron ionization

(EI), or electrospray ionization (ESI). Melting points were taken on a MEL-TEMP melting

point apparatus and are uncorrected. All optical rotation data was recorded at 25 oC on

a Jasco P-2000 polarimeter with a 100 mm cell (concentration reported as g/100mL).

Yields refer to isolated material judged to be ≥ 95% pure by 1H NMR spectroscopy

following silica gel chromatography. All chemical were used as received unless

otherwise stated. All solvents, solutions and liquid reagents were added via syringe.

Tetrahydrofuran (THF) was purified by distillation over sodium and benzophenone.

Methylene chloride (CH2Cl2) was distilled from calcium hydride (CaH2). The n-BuLi

solutions were titrated against a known amount menthol dissolved in tetrahydrofuran

using 1,10-phenanthroline as the indicator. All reactions were carried out under an inert

Page 165: Florida State University Libraries - FLVC

146

nitrogen atmosphere unless otherwise stated. The purifications were performed by flash

chromatography using silica gel F-254 (230-499 mesh particle size).

(2-Oxo-oct-6-ynyl)-diphenylphosphine oxide (153c): To a stirred solution of

methyldiphenylphosphine oxide (201 mg, 0.92 mmol) in THF (50 mL) at –78 oC, was

added n-BuLi (0.34 mL, 0.85 mmol), as a 2.5 M solution in hexane, dropwise. The

reaction mixture was allowed to stir at –78 oC for 45 min, at which time vinylogous acyl

triflate (VAT) 261 (200 mg, 0.77 mmol) was added dropwise. The reaction mixture was

stirred at –78 oC for 15 min, 0 oC for 15 min, and finally room temperature for 45 min.

The reaction mixture was quenched with a half-saturated aqueous solution of

ammonium chloride. The product was extracted with CH2Cl2 (3 x 25 mL). The combined

extracts were washed with sat. NaHCO3, sat. brine, and were dried with MgSO4. The

dried organic solution was concentrated and purified by flash chromatography on silica

gel (60% EtOAc/Hexanes) to give 225 mg (89%) of alkyne 153c as a white solid: mp =

84 – 85 oC; 1H NMR (300 MHz, CDCl3) 7.82 – 7.69 (m, 4H), 7.60 – 7.43 (m, 6H), 3.60

(d, J = 15.0, 2H), 2.76 (t, J = 7.0, 2H), 2.06 (tq, J = 7.0, 2.5, 2H), 1.74 (t, J = 2.5, 3H),

1.65 (quintet, J = 7.0, 2H); 13C NMR (75 MHz, CDCl3) 202.27, 132.03, 131.84 (d, J =

101.9), 130.71 (d, J = 9.8), 128.56 (d, J = 12.3), 78.03, 76.07, 46.94 (d, J = 56.9), 43.97,

22.48, 17.69, 3.28; IR (thin film) 1978, 1708, 1484, 1438, 1192 cm–1; HRMS (EI): Calcd

for C20H21O2P+ [M+] 324.1279, found 324.1279.

((Z)-2-Oxo-oct-6-ene)-diphenylphosphine oxide (154): Palladium, 5 wt. % on calcium

carbonate, poisoned with lead (820 mg) was stirred in methanol/pyridine (4:1, 50 mL),

under an atmosphere of hydrogen. After 30 min, a solution of alkyne 153c (2.00 g, 6.17

mmol) in MeOH (2 mL) was added in one shot to the stirred palladium solution. The

solution was stirred for 45 min. The reaction mixture was filtered through a pad of

Celite™ and the pad was washed with CH2Cl2 ~15 mL). The filtrate was concentrated

and purified on silica gel by flash chromatography (50% EtOAc/Hexane) to afford 1.93 g

(96%) of the product 154, containing the Z-olefin as a white solid: mp = 64 – 66 oC; 1H

NMR (300 MHz, CDCl3) 7.83 – 7.68 (m, 4H), 7.61 – 7.42 (m, 6H), 5.50 – 5.35 (m, 1H),

5.32 – 5.19 (m, 1H), 3.59 (d, J = 15.0, 2H), 2.65 (t, J = 7.2, 2H), 1.94 (q, J = 7.2, 2H),

Page 166: Florida State University Libraries - FLVC

147

1.60 – 1.49 (m, 5H); 13C NMR (75 MHz, CDCl3) 202.91, 132.19, 132.09 (d, J = 102.4),

130.91 (d, J = 9.8), 129.56, 128.73 (d, J = 12.3), 124.65, 47.15 (d, J = 56.6), 44.74,

25.90, 23.14, 12.74; IR (thin film) 1708, 1438, 1193, 1120 cm–1; HRMS (ESI): Calcd for

C20H23O2PNa+ [M+Na+] 349.1333, found 349.1327.

((E)-ethyl-2-Oxo-oct-6-enoate)-diphenylphosphine oxide (155): To a solution of

olefin 154 (1.00g, 3.06 mmol) and ethyl acrylate (1.33 mL, 12.24 mmol) in CH2Cl2 (30

mL) was added freshly distilled Ti(Oi-Pr)4 (120 L, 0.46 mmol), followed by Grubbs’

second generation catalyst (76 mg, 0.09 mmol). The reaction vessel was sealed with a

Teflon screw-top with a rubber seal. The reaction mixture was placed in an oil bath

heated to 100 oC and stirred for 20 min. The solution was cooled to room temperature

and a second aliquot of Grubbs’ II catalyst was added (25 mg, 0.03 mmol). The reaction

mixture was stirred and re-heated to 100 oC; after 10 min, it was cooled to room temp.

and filtered through Celite™ and washed with CH2Cl2 (10 mL). The filtrate was

concentrated and purified on silica gel by flash chromatography (60% EtOAc/Hexane,

80% EtOAc/Hexane) to provide 1.03 g (87%) of Horner-Wittig reagent 155 as a white

solid: mp = 79 – 81 oC; 1H NMR (300 MHz, CDCl3) 7.82 – 7.68 (m, 4H), 7.62 – 7.42

(m, 6H), 6.84 (dt, J = 15.7, 6.9, 1H), 5.75 (d, J = 15.6, 1H), 4.17 (q, J = 7.1, 2H), 3.58 (d,

J = 14.9, 2H), 2.69 (t, J = 7.1, 2H), 2.09 (app. quartet, J = 6.6, 2H), 1.65 (quintet, J =

7.1, 2H), 1.28 (t, J = 7.1, 3H); 13C NMR (75 MHz, CDCl3) 202.19, 166.41, 147.94,

132.25, 131.90 (d, J = 102.4), 130.83 (d, J = 9.8), 128.74 (d, J = 12.3), 121.81, 60.09,

47.13 (d, J = 55.2), 44.30, 31.02, 21.48, 14.22; IR (thin film) 1709, 1653, 1438, 1187 cm-

1; HRMS (EI): Calcd for C22H25O4P+ [M+] 284.1490, found 384.1490.

(2R,3S)-2,3-dihydroxy-hept-6-ynoic ethyl ester (157). AD-mix-(30g, 1.6 g/mmol of

olefin) and MeSO2NH2 (1.75g, 18.4mmol) were stirred in t-BuOH/H2O (1:1, 100 mL) at 0

oC for 1 hr. To the stirred heterogeneous solution was added a solution of the known

compound 156,221 (E)--hept-2-en-6-ynoic ethyl ester (2.8 g, 18.4 mmol), in 24 mL of t-

BuOH/H2O (1:1) in one shot. The reaction mixture was stirred at 0 oC for 24 h. To the

solution was added Na2SO3 (13.4 g, 106.7 mmol) at 0 oC and stirred for an additional

hour. The reaction mixture was diluted with CH2Cl2 (100 mL) and H2O (50 mL). Product

Page 167: Florida State University Libraries - FLVC

148

extracted with CH2Cl2 (3 x 50 mL). The combined organic extracts were washed with a

saturated brine solution (150 mL), dried with Na2SO4, and concentrated under reduced

pressure. The resulting oil was purified by flash chromatography on silica gel using 30%

EtOAc/Hexanes to afford 2.57 g of diol 157 (75%, 99.6% ee as determined via chiral

HPLC on a chiracel OD column after converting the diol to the dibenzoyl ester,223 using

2% isopropanol/hexanes as the eluent at a flow rate of 1.00 mL/hr; retention times (in

minutes) of 79:30 (major) and 70.28 (minor)) as a white solid: mp = 50 - 51 oC; []D25 = -

31.7o (c = 6.7, CH2Cl2); 1H NMR (300 MHz, CDCl3) 4.31 (q, J = 7.1, 2H), 4.09 (d, J =

4.9, 1H), 4.06 (d, J = 6.3, 1H), 3.07 (d, J = 5.1, 1H), 2.39 (dt, J = 9.0, 2.7, 2H), 2.03 (d, J

= 9.3, 1H), 2.00 (t, J = 2.7, 1H), 1.94 – 1.75 (m, 2H), 1.33 (t, J = 7.1, 3H); 13C NMR (75

MHz, CDCl3) 173.22, 83.47, 73.23, 71.15, 68.95, 62.05, 32.17, 14.83, 14.04; IR (thin

film) 3444, 3291, 2111, 1732, 1214, 1118 cm-1; HRMS (ESI): Calcd for C9H14O4Na+

[M+Na+] 209.0790, found 209.0796.

(4R,5S)-5-(3-butynyl)-2,2-dimethyl-[1,3]-dioxolane-4-carboxylic acid ethyl ester

(158). To a stirred solution of diol 157 (500 mg, 2.68 mmol) in 10 mL acetone (HPLC

grade) was added 2,2-dimethoxypropane (0.4 mL, 3.22 mmol) followed by camphor-10-

sulfonic acid (CSA) (30 mg, 0.13 mmol). The solution was stirred for 24 hr at room

temperature. The reaction was diluted with CH2Cl2 (10 mL) and then quenched with

saturated NaHCO3 (20 mL). The product was extracted with CH2Cl2 (3 x 15 mL). The

combined organic layers were washed with a saturated brine solution, dried with

MgSO4, and concentrated. The resulting oil was purified by flash chromatography on

silica gel (10% EtOAc/Hexanes) to afford 603 mg (>95%) of acetonide 158: []D25 = -

23.5o (c = 4.5, CH2Cl2); 1H NMR (300 MHz, CDCl3) 4.30-4.20 (m, 3H), 4.16 (d, J = 7.6,

1H), 2.49 – 2.26 (m, 2H), 2.10 – 1.82 (m, 3H), 1.46 (s, 3H), 1.45 (s, 3H), 1.31 (t, J = 7.1,

3H); 13C NMR (75 MHz, CDCl3) 170.53, 110.98, 83.08, 78.71, 77.55, 68.86, 61.36,

32.47, 27.07, 25.61, 14.94, 14.12; IR (thin film) 3283, 2115, 1757, 1732, 1096 cm-1;

HRMS (CI): Calcd for C12H19O4+ [M+] 227.1283, found 227.1285.

(4R,5S)-5-(3-butynyl)-2,2-dimethyl-[1,3]-dioxolane-4-carboxaldehye (150). To a

solution of ethyl ester 158 (600 mg, 2.68 mmol) in CH2Cl2 (20 mL) at –78 oC was added

Page 168: Florida State University Libraries - FLVC

149

DIBAL (4.02 mL), as a 1.0M solution in toluene, dropwise. The reaction mixture was

allowed to stir at –78 oC for 1 hr. To the stirred solution was added 20 mL of a saturated

aqueous solution of sodium, potassium tartrate and 1 mL of methanol dropwise at –78

oC. The reaction mixture was warmed to room temperature and stirred for approximately

2 hrs until the biphasic solution became clear. The product was extracted with Et2O (3 x

10 mL). The combined organic layers were dried with Na2SO4 and concentrated. The

resulting oil was filtered through a plug of silica gel (10% EtOAc/Hex). The filtrate was

concentrated, leaving 480 mg of a clear oil (>95% crude yield). The crude oil was then

used immediately in the next reaction.

Horner-Wittig reaction to provide enone 161: To a solution of NaH, 60 wt. % in

mineral oil, (70 mg, 1.75 mmol) in THF (15 mL) at 0 oC was added Horner-Wittig

reagent 155 (673 mg, 1.75 mmol) at once. Upon stirring for 45min at 0 oC, aldehyde 150

(480 mg, 2.63 mmol) in 5 mL of THF was added in one shot. The reaction solution was

subsequently stirred at 0 oC for 5 min (white ppt. began to form), and was warmed and

stirred at room temperature for 1 hr. The reaction mixture was then diluted with Et2O (10

mL) and quenched with ½ sat. NH4Cl (20 mL). The product was extracted with Et2O (3 x

10 mL). The combined extracts were washed with sat. NaHCO3 (20 mL) followed by a

wash with sat. brine (20 mL). The organics were dried with MgSO4 and concentrated.

The crude yellowish oil was purified via flash chromatography on silica gel (10%

EtOAc/Hexanes to 20% EtOAc/Hexane) to afford 540 mg (89%) of enone 161 as a clear

oil: []D25 = -12.5o (c = 3.9, CH2Cl2);

1H NMR (300 MHz, CDCl3) 6.93 (dt, J = 15.6, 6.9,

1H), 6.72 (dd, J = 15.8, 5.8, 1H), 6.37 (dd, J = 15.8, 1.3, 1H), 5.83 (dt, J = 15.6, 1.5,

1H), 4.26 – 4.12 (m, 3H), 3.87 (dt, J = 5.1, 4.8, 1H), 2.60 (t, J = 7.3, 2H), 2.47 – 2.28 (m,

2H), 2.28 – 2.18 (m, 2H), 1.98 (t, J = 2.5, 1H), 1.88 – 1.74 (m, 4H), 1.43 (s, 1H), 1.42 (s,

1H), 1.29 (t, J = 7.1, 3H); 13C NMR (75 MHz, CDCl3) 198.87, 166.40, 147.87, 141.28,

130.44, 122.02, 109.64, 83.06, 79.89, 78.96, 69.07, 60.14, 39.63, 31.24, 30.81, 27.12,

26.65, 21.92, 15.06, 14.18; IR (thin film) 3275, 1714, 1677, 1651, 1371 cm-1; HRMS

(ESI): Calcd for C20H28O5Na+ [M+Na+] 371.1834, found 371.1830.

(R)-CBS reduction to afford C7-alcohol (162): To a stirred solution enone 161 (200

mg, 0.57 mmol) in THF (50 mL) at -40 oC was added (R)-2-methyl-CBS-oxazaborolidine

Page 169: Florida State University Libraries - FLVC

150

(1.71 mL, 1.71 mmol), 1.0 M in toluene, dropwise. The reaction mixture was stirred for

30 min at –40 oC, at which time, borane-THF complex (1.14 mL, 1.14 mmol), 1.0 M in

THF, was added dropwise. The solution was allowed to stir for an additional 45 min at –

40 oC. The reaction was quenched with Et2O/MeOH (51 mL, 50:1) at –40 oC, this was

followed by a sat. NaHCO3 solution (80 mL) once the solution reached room

temperature. The product was extracted with CH2Cl2 (3 x 35 mL), the organic layers

were combined, washed with a sat. brine solution (100 mL) and dried with Na2SO4. The

volatiles were evaporated and the crude oil purified by flash chromatography on silica

gel (20% EtOAc/Hexane) to give 178 mg (89%) of allylic alcohol 162 as a mixture of

diastereomers (3.2:1), resolved by chiral HPLC on a chiracel OD column with retention

times (in minutes) of 17:02 (major) and 20:28 (minor) using 12 % isopropanol/hexanes

as the eluent at a flow rate of 0.5mL/hr; isolated as a clear oil: []D25 = -4.9o (c = 4.2,

CH2Cl2); 1H NMR (300 MHz, CDCl3) 6.94 (dt, J = 15.6, 6.9, 1H), 5.91 – 5.77 (m, 2H),

5.67 (dd, J = 15.5, 7.5, 1H), 4.23 – 4.12 (m, 3H), 4.10 – 3.99 (m, 1H), 3.79 (dt, J = 7.1,

6.2, 1H), 2.44 – 2.28 (m, 2H), 2.28 – 2.16 (m, 2H), 1.97 (t, J = 2.6, 1H), 1.82 – 1.71 (m,

2H), 1.61-1.49 (m, 5H), 1.41 (s, 6H), 1.28 (t, J = 7.1, 3H); 13C NMR (75 MHz, CDCl3)

166.54, 148.59, 137.88, 126.96, 121.54, 108.70, 83.42, 81.25, 79.07, 71.35, 68.81,

60.08, 36.23, 31.84, 30.65, 27.10, 26.85, 23.69, 15.11, 14.15; IR (thin film) 3452, 3292,

2115, 1714, 1370 cm-1; HRMS (ESI): Calcd for C20H30O5Na+ [M+Na+] 373.1991, found

373.1984.

t-Butyldimethylsilyl Ether 163: To a stirred solution of alcohol 162, resulting from the

CBS-reduction, (100 mg, 0.28 mmol) in CH2Cl2 (40 mL) at –78 oC was added 2,6-

lutidine (190 L, 1.68 mmol), followed by the dropwise addition of TBSOTf (190 L, 0.84

mmol). The reaction mixture was stirred at –78 oC for 30min. The reaction was

quenched with 20 mL of a saturated aqueous solution of NaHCO3 at –78 oC. The

heterogeneous mixture was warmed to room temperature and stirred for 10 min. The

product was extracted with CH2Cl2 (3 x 15 mL). The combined organics were washed

with brine (40 mL), dried with Na2SO4, and concentrated. The crude oil was purified by

flash chromatography on silica gel (5% EtOAc/Hexanes) to give 122 mg (94%) of a

clear colorless oil, compound 163: []D25 = -10.2o (c = 4.1, CH2Cl2);

1H NMR (300 MHz,

Page 170: Florida State University Libraries - FLVC

151

CDCl3) 6.93 (dt, J = 15.5, 6.9 Hz, 1H), 5.86 – 5.69 (m, 2H), 5.56 (dd, J = 15.5, 7.4 Hz,

1H), 4.23 – 4.10 (m, 3H), 4.03 (t, J = 7.9 Hz, 1H), 3.78 (dt, J = 7.9, 4.0 Hz, 1H), 2.44 –

2.24 (m, 2H), 2.19 (q, J = 6.6, 2H), 1.95 (t, J = 2.4 Hz, 1H), 1.83 – 1.67 (m, 2H), 1.54 –

1.43 (m, 4H), 1.40 (s, 6H), 1.29 (t, J = 7.1, 3H), 0.89 (s, 9H), 0.05 (s, 3H), 0.03 (s, 3H);

13C NMR (75 MHz, CDCl3) 166.63, 148.85, 138.23, 126.61, 121.49, 108.73, 83.45,

81.40, 79.19, 72.33, 68.71, 60.11, 37.41, 32.08, 30.88, 27.20, 26.91, 25.83, 23.53,

18.17, 15.25, 14.25, -4.28, -4.76; IR (thin film) 3312, 2115, 1720, 1654, 1252 cm-1;

HRMS (ESI): Calcd for C26H44O5SiNa+ [M+Na+] 487.2886, found 487.2855.

Page 171: Florida State University Libraries - FLVC

152

HPLC data for dihydroxylation of 156:

Chiracel OD Column: (Standardized using products of both AD-mix- and AD-mix- dihydroxylations followed by dibenzoylation of the resulting diol).

Eluent = 2 % isopropanol/hexanes

Flow rate = 1.00 mL/hr

Detector wavelength = 240 nm

Injection time = 84.67 min

2S,3R peak elution = 155.14 min 2R,3S peak elution = 164.16 min

2S,3R peak retention = 70:28 min 2R,3S peak retention = 79:30 min

2S,3R peak area = 45144 2R,3S peak area = 23068392

2S,3R peak % area = 0.03 2R,3S peak % area = 17.93

2S,3R : 2R,3S = 1 : 511

2R,3S = 99.6 % ee

Minutes

154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169

AU

0.00

0.05

0.10

0.15

AU

0.00

0.05

0.10

0.15

Det 166

dmj-III273px2

Page 172: Florida State University Libraries - FLVC

153

HPLC data for CBS-reduction of 161:

Chiracel OD Column: (Standardized using products of both (R)-CBS and (S)-CBS catalyzed reductions).

Eluent = 12 % isopropanol/hexanes

Flow rate = 0.5 mL/hr

Detector wavelength = 225 nm

Injection time = 214.58 min

7S peak elution = 231.617 min 7R peak elution = 235.042 min

7S peak retention = 17:02 min 7R peak retention = 20:28 min

7S peak area = 10864911 7R peak area = 3371974

7S peak % area = 5.46 7R peak % area = 1.70

7S : 7R = 3.22 : 1

Minutes

228 229 230 231 232 233 234 235 236 237

AU

0.00

0.05

0.10

0.15

0.20

0.25

AU

0.00

0.05

0.10

0.15

0.20

0.25

Det 166

dmj-III250x2

Page 173: Florida State University Libraries - FLVC

154

1H NMR and 13C NMR spectra:

OP

O

PhPh

153c

Page 174: Florida State University Libraries - FLVC

155

OP

O

PhPh

153c

Page 175: Florida State University Libraries - FLVC

156

OP

O

PhPh

154

Page 176: Florida State University Libraries - FLVC

157

OP

O

PhPh

154

Page 177: Florida State University Libraries - FLVC

158

OP

O

PhPh

155

EtO2C

Page 178: Florida State University Libraries - FLVC

159

OP

O

PhPh

155

EtO2C

Page 179: Florida State University Libraries - FLVC

160

OEt

O

HO

OH

157

Page 180: Florida State University Libraries - FLVC

161

OEt

O

HO

OH

157

Page 181: Florida State University Libraries - FLVC

162

OEt

O

O

O

158

Page 182: Florida State University Libraries - FLVC

163

OEt

O

O

O

158

Page 183: Florida State University Libraries - FLVC

164

O

O

O

EtO2C

161

Page 184: Florida State University Libraries - FLVC

165

O

O

O

EtO2C

161

Page 185: Florida State University Libraries - FLVC

166

O

O

OH

EtO2C

162

Page 186: Florida State University Libraries - FLVC

167

O

O

OH

EtO2C

162

Page 187: Florida State University Libraries - FLVC

168

O

O

OTBS

EtO2C

163

Page 188: Florida State University Libraries - FLVC

169

O

O

OTBS

EtO2C

163

Page 189: Florida State University Libraries - FLVC

170

CHAPTER 5

RE-EXPLORING THE CLAISEN-TYPE CONDENSATIONS OF VINYLOGOUS ACYL TRIFLATES

New Insights into the Mechanism

First reported in 1887,224 the Claisen condensation plays an important role in

synthetic organic chemistry.225-228 The Claisen condensation involves the enolate of an

ester undergoing a reversible nucleophilic addition / elimination reaction with another

equivalent of ester in the presence of excess base. The reaction is driven to completion

due to the irreversible deprotonation of the resulting -ketoester (Figure 67).

O

OEt

O M

OEt

O M

EtO

O

OEt

O O

OEt

H

EtO M

EtOH

O O

OEt

M"H3O+" O O

OEt

irreversible deprotonation

addition elimination

Figure 67: Mechanism of the Classical Claisen Condensation of Ethyl Acetate.

As presented in previous chapters, our lab has been interested in the preparation

of alkynyl ketones using the tandem addition / C-C bond cleavage reaction of

vinylogous acyl triflates (VATs).61,93,216,217 This reaction was applied to the synthesis of

an important moth pheromone natural product (Chapter 2)93 and provided access to

substituted benzo-fused indanes (Chapter 3). The two-step conversion of cyclic diones

to tethered alkynyl ketones has been shown to be general, affording a wide variety of

differentially functionalized substrates.

Page 190: Florida State University Libraries - FLVC

171

Kamijo and Dudley were the first to examine the Claisen-type condensations of

VATs. They provided insight into the mechanism of the reaction between the lithium

enolate of acetophenone (152a) and VAT 2 (Figure 68).216 The stoichiometry played a

pivotal role in the ability of the reaction to proceed to completion. Like the classical

Claisen condensation, more than 2 equivalents of base (enolate) are needed to convert

the starting material effectively to product. According to their postulated mechanism, the

1,2-addition of the enolate to VAT 2 proceeds reversibly, leading to intermediate 164. At

elevated temperatures the fragmentation takes place, providing 1,3-diketone 153a.

However, because the enolate addition is reversible, once 153a is formed, another

equivalent of enolate (152a) deprotonates the -ketoester product. Thus, at least 2

equivalents of enolate are required for this reaction, one to undergo the addition and

another for deprotonation.

O

OTf

Ph

OLi

2

O

LiOPh

OTf

152a

164

O

O

Ph

153a

Ph

OLi

Ph

OO

O

Ph

165

Li

"H3O+"

O

O

Ph

153a

-LiOTf

Figure 68: Proposed Mechanism for the Reaction Between 2 and 152a.

With this mechanistic model in mind, the investigation into the Claisen-type

condensation reactions of VATs was carried out using the same protocol. Most of the

nucleophiles examined in this reaction gave satisfactory results (Figure 69).216 The

worst nucleophile in the series happened to be the anion of dimethyl

Page 191: Florida State University Libraries - FLVC

172

methylphosphonate, which gives rise to a potentially useful -ketophosphonate adduct

(153b).

O

OTf

2

[EWG-CH2] Li

THF

-78 oC to 60 oC

O

EWG

153

O

O

Ph

O

O

Me

O

O

OEt

O

S

O

PO

(2.2 equiv)

85% 42% 88%

O

O

Me

53% 21%

OMeMeO

Kamijo and Dudley, Org Lett. 2006, 8, 175-177

153b

Figure 69: Reported Claisen-Type Condensation Reactions of VAT 2.

In 2007, our lab became interested in the synthesis of palmerolide A.175 Our

synthetic plan called for the application of a tandem addition / C-C bond cleavage

adduct similar to 153b. In order for this to be practical, the Claisen-type condensation

reaction had to be a re-optimized for the synthesis of olefinating reagents similar to

153b. Changing the nucleophile from the lithium anion of dimethyl methylphosphonate

to the lithium anion of methyldiphenylphosphine oxide (152c) afforded Horner-Wittig

reagent 153c in the 70% yield range. Upon further optimization, we found that only 1.1

equivalents of the phosphine oxide nucleophile were necessary to convert VAT 2

effectively to the corresponding product 153c (Figure 70). This optimization culminated

in the synthesis of the C1-C15 fragment of palmerolide A (Chapter 4).

Page 192: Florida State University Libraries - FLVC

173

O

OTf

2

O

P

153c

P

O

PhPh

Li

152c

THF

-78 oC to 60 oC

2.2 equiv of 152c = 69 - 75%1.1 equiv of 152c = 81 -89%

O

PhPh

Figure 70: Observations Made During the Synthesis of the C1-C15 Fragment of Palmerolide A (Chapter 4).

The ability to decrease the loading of the phosphine oxide nucleophile provided

impetus for us to re-open the investigation into the Claisen-type condensation reactions

of vinylogous acyl triflates. We initially hypothesized that the reactivity of the nucleophile

has an important role in the reversibility of the addition step in the proposed mechanism.

We postulated that if the reactivity of the nucleophile were sufficiently high, the

reversibility of the initial addition step would be reduced. In addition, if the fragmentation

step was significantly faster than that of the retro-addition, the concentration of the

nucleophile (base) would be limited and would thus allow for the accumulation of

fragmentation product.

The pKa’s of some related pre-nucleophiles may provide insight into the relative

reactivity of their corresponding anions in the addition / fragmentation reaction (Table

8).229 The data presented in Table 8 demonstrates the similarities in acidity between

phosphonates and phosphine oxides as well as their significant difference compared to

the acidity of the acetophenone derivatives. If the elevated pKa’s of phosphine oxides

are representative of their relative reactivity, then the enhanced reactivity of the

nucleophile might be responsible for the ability to lower the number of equivalents

added.

Page 193: Florida State University Libraries - FLVC

174

Table 8: Comparison of the Acidities of Several Acetophenone, Phosphonate, and Phosphine Oxide Derivatives in DMSO.a

R

pKa

O

PhR

O

PEtOEtO

R

O

PPhPh

R

H 24.7 N/A N/A

Ph 17.7 27.6 N/A

CN 10.2 16.4 16.9

SPh 16.9 N/A 24.9 a pKa’s obtained from data presented in ref. 228.

To reiterate our previous observations, the Claisen-type ring opening of VAT 2

with the lithium enolate of acetophenone requires 2 equivalents of enolate, whereas the

similar reaction involving the lithium anion of methyldiphenylphosphine oxide is best

accomplished with 1 equivalent of the stabilized nucleophile.

Having optimized the Claisen-type condensation of VATs for the acetophenone

enolate (152a, Figure 68) and the anion of methyldiphenylphosphine oxide (152c,

Figure 70), the next logical experiment to include in our new investigation was the

addition of 1.1 equivalents of the enolate of ethyl acetate (166). Ethyl acetate was one

of the best pre-nucleophiles in our earlier study, and it is intermediate in acidity between

acetophenone and methyldiphenylphosphine (pKa of ethyl acetate in DMSO = 29.5).230

The reaction involving the enolate of ethyl acetate provided valuable data. When

1.1 equivalents of 166 were added to VAT 2, the desired fragmentation product (168)

was obtained in 56% yield (Figure 71). This result was in line with our previous

observation using 1.2 equiv. of the acetophenone enolate (152a) (56% yield).216 In this

case, however, a previously unobserved byproduct was isolated (ca. 26% yield). We

believe that the structure of this byproduct is that of alcohol 170. This byproduct proved

to be unstable even at low temperatures (-15 oC). However, when immediately

dissolved in THF, treated with excess NaH (approximately 3 equivalents) and heated to

Page 194: Florida State University Libraries - FLVC

175

60 oC for 30 min, this byproduct gave rise to the fragmentation product 168 in 78%

yield, which provides support for our proposed structure (170).

A revised mechanistic hypothesis is needed to account for the formation of -

hydroxy ester 170. We envision an effectively irreversible addition of enolate 166 to VAT

2 to provide aldolate 167. In contrast to reactions using acetophenone, the retro-aldol of

167 (167 2) does not figure prominently in our observations. Intermediate 167 begins

to undergo fragmentation upon warming, providing -ketoester 168. Subsequent

deprotonation of the -ketoester by the alkoxide, not the enolate, occurs. Thus, two

equivalents of base are still required for compete conversion of VAT 2 to 168. Although

the isolation of byproduct 170 provides evidence for the proposed reaction pathway, a

competing deprotonation of the -ketoester by the enolate resulting from a retro-addition

cannot be ruled out.

O

OTf

OEt

OLi

2

O

LiOEtO

OTf

166

167

O

O

OEt

168

O

O

OEt

169

Li"H3O+"

O

O

OEt

168

OLi

OEtO

OTf167

OH

OEtO

OTf170

THF, 78%

excess NaH

ca. 26% isolated

56%

proposed structure ofisolated byproduct

separately treated with

Figure 71: Proposed Fragmentation Reaction Pathway Between 2 and 166.

Page 195: Florida State University Libraries - FLVC

176

This new byproduct, tentatively assigned as 170, provided a more defined

understanding of the reaction pathway, which enabled us to reconsider the reaction

between phosphine oxide nucleophile 152c and vinylogous acyl triflates (Figure 72). We

propose that the anion adds irreversibly at cold temperatures and the resulting oxy-

anion coordinates to the phosphine oxide to provide intermediate 171. This intermediate

is envisioned to resemble an oxaphosphetane intermediate, much like that formed

during a Wittig olefination reaction.231-233 Such an intermediate would reduce the oxy-

anion’s ability to deprotonate the -ketophosphine oxide product, and would reduce the

possibility of a retro-addition, thus allowing for the use of one equivalent of nucleophile

to consume the starting material. When the reaction mixture is subsequently warmed,

the postulated oxaphosphetane-like intermediate collapses and provides the

fragmentation product 153c, instead of undergoing the classical—retro-[2+2]—

olefination reaction to provide 172 (not observed).

O

OTf

PO

PhPh

Li

152c OP

OPhPh

2 171

Li

O

PO

PhPh

PO

PhPh

OLi_ OTf

153c

172

X(not observed)

Figure 72: Proposed Mechanism of the Reaction Between VAT 2 and 152c.

The results of the Claisen-type condensation reactions of VAT 2 and the various

stabilized anions (cf. 152a, 152c, and 166) have provided us with a better

understanding of their reaction mechanisms. Although it is not necessarily the reactivity

of the nucleophile that determines the ability to use fewer equivalents, the isolation of

Page 196: Florida State University Libraries - FLVC

177

the alcohol intermediate 170 was very informative as to the intermediates involved in

these reactions. The more detailed study of these reactions allowed for the expansion of

the methodology to the synthesis of -ketophosphonates. -Ketophosphonates provide

reactivity similar to -ketophosphine oxides (both are olefinating reagents), but the

phosphonates provide some distinct advantages; they are cheaper, more widely

available, and easier to work with than their phosphine oxide analogs. The next section

addresses the conversion of VATs to novel phosphonate-based olefinating reagents.

Synthesis of -Ketophosphonates

The use of phosphonates in organic chemistry has revolutionized the synthesis

of alkenes.231-244 The ability to generate E- and Z- alkenes selectively, the mild

conditions required for reaction, and the ease of their synthesis provides the distinct

advantages of phosphonates as olefination reagents over their phosphorane (Wittig

reagents) or phosphine oxide (Horner-Wittig) counterparts. Common methods for the

synthesis of phosphonates have relied on a two general strategies (Figure 73): (1) the

Arbuzov reaction,245,246 which involves the alkylation of the corresponding trialkyl

phosphite to prepare alkyl-, benzyl-, and allylphosphonates as well as phosphonate

esters; or (2) a Claisen-type condensation between esters and a dialkyl

methylphosphonates to prepare -ketophosphonates.247-251 Synthesis of -

ketophosphonates using the Arbuzov reaction is also known, but one must recognize

the potential for the competing Perkow reaction, which gives rise to enol phosphates.245

Page 197: Florida State University Libraries - FLVC

178

OR1

PR1O OR1

Trialkyl Phosphite

R2H2C X

OR1

PCH2R2

R1O

R1O

X

-R1X

O

PR1OR1O

R2

R1 = 1o Alkyl

R2 = Alkyl, Vinyl, Aryl, Ester

Dialkyl Alkylphosphonate

(1) Arbuzov Reaction

(2) Claisen-Type Condensation Reaction

O

PR1OR1O

Me

1. Base, -78 oC

2.

OR3,

O

R2

-78 oC to r.t.

O

PR1OR1O

O

R2

Base = n-BuLi, LDA, LiHMDS

Figure 73: Common Methods for the Preparation of Phosphonates.

The synthesis of -ketophosphonates was of particular interest to us. Having

obtained a poor yield (21%) of phosphonate product 153c upon treating VAT 2 with 2.2

equivalents of dimethyl lithiomethylphosphonate (152b) under our original conditions,216

we were interested to determine if the conditions optimized for the

lithiomethyldiphenylphosphine oxide (152c) nucleophile would provide increased yields

of the -ketophosphonates. The use of lithiomethyldiphenylphosphine oxide as the

nucleophile trigger provided excellent yields (up to 89%). However, the use of

phosphonates for alkene synthesis is much more common.231,238,241 What’s more, the

use of dimethyl methylphosphonate has a distinct advantage for large scale synthesis,

its cost is far lower than that of methyldiphenylphosphine oxide (ca. 66 mmol / $1 vs. 1

mmol / $1, respectively).94

Table 9 summarizes the data resulting from the Claisen-type addition / bond

cleavage reactions of various VATs and 1.1 equiv. of dimethyl lithiomethylphosphonate.

Page 198: Florida State University Libraries - FLVC

179

The reaction between VAT 2 and 1.1 equiv. of 152b proceeded in excellent yield (entry

1). This result was nearly a 5-fold increase compared to our previous report, in which

2.2 equivalents of nucleophile were used.216 VAT 173, which is similar to 2, but lacks

the -methyl substituent, provided a messy reaction. Although the product was present

in the 1H NMR spectrum, it could not be obtained in acceptable purity (entry 2). The

vinylogous acyl triflates derived from dimedone and 1,3-cycloheptanedione (175 and

177, respectively) both provided their respective phosphonate products, 176 and 177, in

acceptable yields. Interestingly, in the case of 175, an unstable byproduct was isolated

(ca. 4%), whose 1H NMR spectrum is consistent with diene 179. Such a byproduct

would support our proposed oxaphosphetane-like intermediate (cf. structure 171).

Page 199: Florida State University Libraries - FLVC

180

Table 9: Reactions of Vinylogous Acyl Triflates with 1.1 Equivalents of Dimethyl Lithiomethylphosphonate (152b).a

Entry VAT Product Yield, %b

1

O

OTf2

PO

O

OMeMeO

153b

97

2

O

OTf173

PO

O

OMeMeO

174

—c

3

O

OTf175

PO

O

OMeMeO

176

41d

4

O

OTf

177

PO

O

OMeMeO

178

78e

a Triflate (0.5 mmol) reacted with nucleophile (0.55 mmol, generated from 0.6

mmol of dimethyl methylphosphonate and 0.55 mmol n-BuLi) at -78 oC to 60

oC

over 80 min. b Isolated yields.

c Product detected by

1H NMR, however not

obtained in acceptable purity, all attempts to purify failed. d Obtained a byproduct

proposed to be diene 179. e decomposition of 177 was observed after purification

and had to be used immediately.

PMeO

O

Li

MeO

152b

OTf

179

Vinylogous acyl triflates 173, 175, and 177, which lack the -methyl substituent,

are relatively unstable when compared to their analog, VAT 2. Vinylogous acyl triflate 2

can be stored under an inert atmosphere for several months at -10 oC without any

Page 200: Florida State University Libraries - FLVC

181

observable decomposition by 1H NMR spectroscopy, whereas VATs 173 and 174 begin

to discolor after 1 to 2 days.

VAT 177 is even less stable; it began to decompose upon removal of solvent and

had to be used immediately. In addition, 1,3-cycloheptanedione, the precursor to VAT

177, is extremely cost prohibitive (1 gram / $264.50, approximately 30 mol / $1).94 For

these reasons, the two-step conversion from 1,3-cycloheptanedione to -

ketophosphonate 178 is less than ideal.

We desired an alternative strategy for accessing olefination reagents

homologated tethered alkynes (cf. 178) using the KAPA acetylene zipper reaction.252

The KAPA acetylene zipper reaction rearranges internal alkynes to terminal alkynes.

Rearrangement of phosphonate 153b was not effective (Figure 74, eq. 1), likely due to

competing amidation of the phosphonate with 1,2-propanediamine. This technical

problem was easily overcome by switching to the corresponding phosphine oxide

(153c). Fragmentation of VAT 2 with lithiomethyldiphenylphosphine oxide provides

153c, and carrying out a subsequent KAPA zipper (alkyne isomerization) reaction

provides Horner-Wittig reagent 180 (ca. 44% over two steps), an analog of phosphonate

178 (Figure 74, eq. 2).

O

OTf

2

OPPh

PhLi

152c

THF, -78 to 60 oC

81-89%

O

PO

PhPh

153c(Chapter 4)

1,3-Propanediamine,

0 oC, 12 h

49% Unoptimized

10 equiv KH

PO

O

PhPh

180

PO

O

OMeMeO

178

O

PO

OMeMeO

153c

1,3-Propanediamine,

0 oC, 12 h

10 equiv KH

not observed(1)

(2)

Figure 74: Synthesis of a 180, an Analog of Phosphonate 178.

Page 201: Florida State University Libraries - FLVC

182

Having demonstrated the ability to fragment various vinylogous acyl triflates to

provide dimethyl -ketophosphonates, we turned our attention to determining if the

fragmentation reaction could be expanded to the use of other phosphonate

nucleophiles. Table 10 provides the results of this series of experiments.

Table 10: Fragmentation of VAT 2 Using 1.1 Equivalents of Various Phosphonate

Derived Nucleophiles.a

entry Phosphonate Product Yield, %b

1 PEtO

O

EtO

Me

181

PO

O

OEtEtO

182

Me

94

2 PEtO

O

EtO

Bn

183

PO

O

OEtEtO

184

Ph

70c

3 PEtO

O

EtO

Ph

185

PO

O

OEtEtO

186

Ph

0d,e

4 PF3CH2CO

O

F3CH2COMe

187

PO

O

OCH2CF3F3CH2CO

188

0d,f

a Triflate 2 (0.50 mmol) reacted with nucleophile (0.55 mmol, generated from 0.6

mmol of phosphonate and 0.55 mmol of n-BuLi) in THF at -78 to 60 oC over 80 min.

b Isolated yields.

c Obtained byproduct, proposed to be 189 (ca. 8% yield).

d

decomposition of starting VAT 2 observed. e 20% recovered VAT 2.

f 35%

recovered VAT 2.

O

OTf

2 OTf

Ph

189

Page 202: Florida State University Libraries - FLVC

183

The reaction between VAT 2 and the anion of diethyl ethylphosphonate (181)

proceeded cleanly in 94% yield (entry 1). This result is remarkable. In our previous

studies of the Claisen-type condensation reactions, substitutions at the -position of the

nucleophile led to decomposition of the starting VAT. In the case of the nucleophile

derived from diethyl 2-phenylethylphosphonate (183) (entry 2), the phosphonate product

184 was obtained in 70% yield. This reaction provided an unstable byproduct consistent

with an E/Z- mixture of dienes 189, in a roughly 1:1 ratio (ca. 8% yield). Again, alkene

byproducts are consistent with a postulated oxaphosphetane-like intermediate (cf. 171,

Figure 71).

The more stabilized nucleophiles derived from phosphonates 185 and 187 failed

to produce any discernable products, and small amounts of starting VAT 2 were

recovered. In addition, to the electronic stabilization provided by the phenyl substituent,

the increased steric profile may also inhibit the desired reaction with VAT 2. The anion

of phosphonate 187, which would give rise to a Still-Gennari-type242 olefination reagent,

is prone to homo-condensation,250 thus hampering its viability in Claisen-type

condensation reactions.

In summary, this work has provided valuable insight into the mechanism of the

Claisen-type fragmentation of vinylogous acyl triflates. The observance of the suspected

alcohol byproduct 170 allowed for a better understanding of the mechanism involving

phosphine oxide derived nucleophiles. Ultimately, the results obtained during our

synthesis of the C1-C15 fragment of palmerolide A allowed for the expansion of the

method to the synthesis of -ketophosphonates and a better understanding of these

reactions. The ability of the phosphorus atom to coordinate to the resulting alkoxyanion

after addition, perhaps forming an oxaphosphetane-like intermediate, is a key feature.

This coordination allows for the reduction in the equivalencies of nucleophile required

and provides the desired reactivity. If correct, the proposed structure of the olefinated

byproducts 179 and 189 would support the transient formation of a true

oxaphosphetane intermediate.

We have demonstrated throughout the course of our extensive research into the

tandem nucleophilic addition / C-C bond cleavage reactions of vinylogous acyl triflates

that this class of compounds can give rise to interesting and synthetically useful

Page 203: Florida State University Libraries - FLVC

184

compounds. Tethered alkynyl ketones, alkynyl -ketoesters, alkynyl -ketophosphine

oxides, and now, through re-optimized conditions, alkynyl -ketophosphonates are

available from these easily prepared VAT substrates. The synthetic utility of such

compounds has been demonstrated in the preparation of (Z)-6-heneicosen-11-one

(Chapter 2), penta- and hexasubstituted indanes (Chapter 3), and the C1-C15 fragment

of palmerolide A (Chapter 4). The new addition described in this chapter has led to the

synthesis of some potentially useful -ketophosphonates. Their utility in synthesis has

yet to be explored. Future endeavors into the chemistry and application of vinylogous

acyl triflates and these -ketophosphonates are currently underway in the Dudley

laboratory.

Experimental

General information:

1H NMR and 13C NMR spectra were recorded on a Varian 300 MHz spectrometer or a

Bruker 600 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 and 77.0 ppm for 13C NMR for all compounds. The coupling constants

(J) are reported in Hertz (Hz). IR spectra were recorded on a Perkin-Elmer FT-IR

spectrometer with diamond ATR accessory as thin film. Mass spectra were recorded

using electron ionization (EI) or fast-atom bombardment (FAB) on a JEOL JMS600H

spectrometer. Melting points were taken on a MEL-TEMP melting point apparatus and

are uncorrected. Yields refer to isolated material judged to be ≥ 95% pure by 1H NMR

spectroscopy following silica gel chromatography. All chemical were used as received

unless otherwise stated. All solvents, solutions and liquid reagents were added via

syringe. Tetrahydrofuran (THF) was purified by distillation over sodium and

benzophenone. Methylene chloride (CH2Cl2) was distilled from calcium hydride (CaH2).

The n-BuLi solutions were titrated against a known amount menthol dissolved in

tetrahydrofuran using 1,10-phenanthroline as the indicator. All reactions were carried

out under an inert nitrogen atmosphere unless otherwise stated. The purifications were

Page 204: Florida State University Libraries - FLVC

185

performed by flash chromatography using silica gel F-254 (230-499 mesh particle size).

Vinylogous acyl triflates were prepared from the corresponding 1,3-dione according to

our published procedure.61

Standard Procedure for the Claisen-type Condensation of the Vinylogous Acyl

Triflates with Phosphonate Nucleophiles: To a THF solution (2 mL) of phosphonate

153b (0.6 mmol) was added n-BuLi (0.22 mL, 0.55 mmol; 2.5 M solution in hexanes) at

-78 oC. After being stirred for 20 minutes at -78 oC, was added the vinylogous acyl

triflate 2 (0.50 mmol) was added dropwise to the resulting solution. The mixture stirred

at -78 oC for 10 min, at 0 oC for 10 min, at r.t. for 30 min, and 60 oC for 30 min; during

the course of the reaction the solution changed from clear to yellow, and then a yellow

to a reddish solution. The solution was diluted with 3 mL of Et2O. A half saturated

aqueous NH4Cl solution was used to quench the reaction and the mixture was extracted

3 times with 5 mL portions of EtOAc. The combined organic layers were washed with 5

mL of NaHCO3(aq), 5 mL of saturated brine, dried with MgSO4, filtered and concentrated.

The residual oil was purified on silica gel column chromatography (EtOAc/Hexanes =

10% - 40%) to afford 112 mg of -ketophosphonate 153b (97% yield).

Procedure for Converting -Ketophosphine Oxide 153c into -Ketophosphine

Oxide 180 Through KAPA Zipper Reaction: To potassium hydride (307mg, 2.3 mmol;

30 % by wt.), freshly washed 3 times with hexane, was added 1,3-diaminopropane (2

mL). The heterogeneous mixture was stirred at room temperature for one hour; during

which, the solution changed from clear to opaque orange/brown in appearance. The

solution was then cooled to 0 oC and a solution of 153c (71 mg, 0.22 mmol; in 1 mL of

1,3-diaminopropane) was added dropwise. The reaction mixture stirred at 0 oC for

approximately 12 hrs, at which time, it was quenched with 2 mL of water, followed by 2

mL of a sat. aqueous solution of ammonium chloride. The mixture was warmed to rt.

and the product was extracted with EtOAc (3 x 5 mL). The combined organics were

dried with MgSO4 and concentrated. The crude residue was purified by flash column

chromatography on silica gel (EtOAc/Hexanes = 40 % to 50 %). 35 mg of 4 was

obtained as a white solid (49% yield).

Page 205: Florida State University Libraries - FLVC

186

Analytical Data:

Ethyl 3-oxo-7-nonynoate (168): pale yellow oil; 1H NMR (300 MHz, CDCl3) 4.19 (q, J

= 7.0 Hz, 2H), 3.45 (s, 2H), 2.39 (t, J = 7.2 Hz, 2H), 2.17 (tq, J = 6.8, 2.5 Hz, 2H), 1.76

(t, J = 2.5 Hz, 3H), 1.76 (app. quintet, J = 7.0 Hz, 2H) 1.28 (t, J = 7.0 Hz, 3H); 13C NMR

(75Hz, CDCl3) 202.35, 167.0, 77.9, 76.4, 61.2, 49.3, 41.6, 22.5, 17.8, 14.0, 3.3; IR

(thin film) 1745, 1742, 1651, 1415, 1242, 1027 cm-1; HRMS (FAB) Calcd for

C11H16O3Na [M+] 219.0097. Found 219.0097. Spectroscopic data in consistent with

previous report.216

Proposed Structure (170): yellow oil that quickly decomposed upon isolation; 1H NMR

(300 MHz, CDCl3) 4.20 (q, J = 7.1 Hz, 2H), 3.84 (s, 1H), 2.77 (d, J = 15.4 Hz, 1H),

2.49 (d, J = 15.4 Hz, 1H), 2.41-2.30 (m, 2H), 1.99-1.67 (m, 7H), 1.29 (t, J = 7.1, 3H).

Diagnostic peaks are circled.

1-(dimethylphosphonato)-2-oxo-6-octyne (153b): pale yellow oil; 1H NMR (300 MHz,

CDCl3) 3.78 (d, J = 11 Hz, 6H), 3.10 (d, J = 22 Hz, 2H), 2.73 (t, J = 7.2 Hz, 2H), 2.16

(tq, J = 6.9, 2.5 Hz, 2H), 1.76 (t, J = 2.5 Hz, 3H), 1.75 (app. quintet, J = 7.0 Hz, 2H); 13C

NMR (75 MHz, CDCl3) 201.4, 78.0, 76.3, 52.9 (d, J = 6.5), 42.8, 41.3 (d, J = 128 Hz),

22.6, 17.8, 3.3; IR (thin film) 1712, 1449, 1254, 1025, 810 cm-1; HRMS (EI+) Calcd for

C10H17O4P+ [M+] 232.0864. Found 232.0860. Spectroscopic data in consistent with

previous report.216

1-(dimethylphosphonato)-4,4-dimethyl-2-oxo-6-heptyne (176): pale yellow oil; 1H

NMR (300 MHz, CDCl3) 3.78 (d, J = 11.3, 6H), 3.08 (d, J = 22.7 Hz, 1H), 2.67 (s, 1H),

2.28 (d, J = 2.5 Hz, 1H), 2.01 (t, J = 2.5 Hz, 1H), 1.09 (s, 3H); 13C NMR (75 MHz,

CDCl3) 200.90 (d, J = 5.8 Hz), 81.91, 70.45, 52.96 (d, J = 5.8 Hz) 52.72, 42.76 (d, J =

128.1 Hz), 33.34, 31.01, 26.79; IR (thin film) 1714, 1465, 1366, 1249, 1024, 811 cm-1;

HRMS (EI+) Calcd for C11H20O4P+ [[M+H]+] 247.1099. Found 247.1096.

Page 206: Florida State University Libraries - FLVC

187

1-(dimethylphosphonato)-2-oxo-7-nonyne (178): clear oil; 1H NMR (300 MHz, CDCl3)

3.76 (d, J = 11.2 Hz, 6H), 3.07 (d, J = 22.8 Hz, 2H), 2.62 (t, J = 7.0 Hz, 2H), 2.17 (dt, J

= 7.0, 2.6 Hz, 2H), 1.92 (t, J = 2.6 Hz, 1H), 1.68 (app quintet, J = 7.6 Hz, 2H), 1.50 (app

quintet, J = 7.6 Hz, 2H); 13C NMR (75 MHz, CDCl3) 201.39 (d, J = 5.1 Hz, 1C), 83.88,

68.55, 52.98 (d, J = 4.5 Hz, 1C), 43.38, 41.25 (d, J = 128.3 Hz, 1C), 27.51, 22.35,

18.13; IR (thin film) 1712, 1456, 1249, 1021, 806 cm-1; HRMS (EI+) Calcd for

C10H18O4P+ [[M+H]+] 233.0943. Found 233.0943.

Proposed structure (179): yellow oil that quickly decomposed; 1H NMR (300 MHz,

CDCl3) 6.19 (s, 1H), 5.03 (apparent doublet, J = 7.1 Hz, 2H), 2.24 (s, 2H), 2.08 (s,

2H), 0.99 (s, 6H). Diagnostic peaks are circled.

(2-oxo-7-octynyl)-diphenylphosphine oxide (180): white solid; mp = 68-71 oC; 1H

NMR (300 MHz, CDCl3) 7.95 – 7.66 (m, 4H), 7.66 – 7.34 (m, 5H), 3.58 (d, J = 15.0 Hz,

2H), 2.68 (t, J = 7.1 Hz, 2H), 2.12 (dt, J = 7.0, 2.5 Hz, 2H), 1.91 (t, J = 2.5 Hz, 1H), 1.66

– 1.50 (app. quintet, J = 7.2 Hz, 2H), 1.41 (app. quintet, J = 7.2 Hz, 2H); 13C NMR (150

MHz, CDCl3) 202.58 (d, J = 5.2 Hz), 132.27 (d, J = 2.9 Hz), 131.99 (d, J = 102.2 Hz),

130.92 (d, J = 5.2 Hz), 128.82 (d, J = 7.9 Hz), 84.06, 68.44, 47.14 (d, J = 56.1 Hz),

44.64, 29.70, 27.55, 22.35, 18.18; IR (thin film) 2232, 1709, 1438, 1187, 907, 725, 693

cm-1; HRMS (EI+) Calcd for C20H21O2P+ [M+] 324.1279. Found 324.1282.

2-(diethylphosphonato)-3-oxo-7-nonyne (182): clear oil; 1H NMR (300 MHz, CDCl3)

4.20 – 4.03 (m, 4H), 3.22 (dq, J = 24.9, 7.1 Hz, 2H), 2.91 (dt, J = 18.0, 7.2 Hz, 1H), 2.65

(dt, J = 18.0, 7.1 Hz, 1H), 2.16 (m, 2H), 1.82 – 1.68 (m, 5H), 1.35 (m, 9H); 13C NMR (75

MHz, CDCl3) 205.43 (d, J = 3.9 Hz), 78.06, 75.97, 62.45 (d, J = 7.3 Hz), 62.35 (d, J =

7.7 Hz), 46.45 (d, J = 127.1 Hz), 41.72, 22.65, 17.79, 16.15 (d, J = 5.6 Hz), 10.75 (d, J =

6.4 Hz), 3.24; IR (thin film) 1713, 1448, 1245, 1048, 1018, 956, 791 cm-1; HRMS (EI+)

Calcd for C13H23O2P+ [M+] 274.1334. Found 274.1338.

2-(diethylphosphonato)-3-oxo-1-phenyl-7-nonyne (184): clear colorless oil; 1H NMR

(300 MHz, CDCl3) 7.20 (m, 5H), 4.24 – 4.05 (m, 4H), 3.52 (ddd, J = 23.2, 11.3, 3.2 Hz,

Page 207: Florida State University Libraries - FLVC

188

1H), 3.30 (ddd, J = 13.6, 11.6, 7.4 Hz, 1H), 3.09 (ddd, J = 13.6, 10.6, 3.0 Hz, 1H), 2.75

(dt, J = 17.9, 7.1 Hz, 1H), 2.27 (dt, J = 17.9, 7.1 Hz, 1H), 1.98 (m, 2H), 1.72 (t, J = 2.5

Hz, 3H), 1.64 – 1.49 (m, 2H), 1.35 (m, 6H); 13C NMR (75 MHz, CDCl3) 204.72, 138.81

(d, J = 16.4 Hz), 128.46, 126.49, 78.06, 75.91, 62.76 (d, J = 6.6 Hz), 62.55 (d, J = 6.6

Hz), 54.37 (d, J = 123.9 Hz), 43.64, 32.30 (d, J = 3.9 Hz), 22.45, 17.68, 16.27 (d, J = 5.8

Hz), 3.29; IR (thin film) 1713, 1455, 1247, 1047, 1019, 960, 699 cm-1; HRMS (EI+) Calcd

for C19H27O4P+ [M+] 350.1647. Found 350.1657.

Proposed Structure (189): yellow oil that quickly decomposed; 1H NMR (300 MHz,

CDCl3) 7.39 – 7.12 (m, 5H), 5.66 (dt, J = 70.5, 7.4 Hz, 1H), 3.55 (dd, J = 33.8, 7.5 Hz,

2H), 2.51 (s, 2H), 2.46 – 2.39 (m, 1H), 2.30 – 2.21 (m, 1H), 2.17 – 1.78 (m, 5H).

Diagnostic peaks are circled.

Page 208: Florida State University Libraries - FLVC

189

1H NMR and 13C NMR Spectra:

OH

OEtO

OTf

170proposed

Page 209: Florida State University Libraries - FLVC

190

PO

O

OMeMeO

176

Page 210: Florida State University Libraries - FLVC

191

PO

O

OMeMeO

176

Page 211: Florida State University Libraries - FLVC

192

PO

O

OMeMeO

178

Page 212: Florida State University Libraries - FLVC

193

PO

O

OMeMeO

178

Page 213: Florida State University Libraries - FLVC

194

OTf179

proposed

Page 214: Florida State University Libraries - FLVC

195

PO

O

PhPh

180

Page 215: Florida State University Libraries - FLVC

196

PO

O

PhPh

180

Page 216: Florida State University Libraries - FLVC

197

PO

O

OEtEtO

182

Me

Page 217: Florida State University Libraries - FLVC

198

PO

O

OEtEtO

182

Me

Page 218: Florida State University Libraries - FLVC

199

PO

O

OEtEtO

184

Ph

Page 219: Florida State University Libraries - FLVC

200

PO

O

OEtEtO

184

Ph

Page 220: Florida State University Libraries - FLVC

201

OTf

Ph

189proposed

Page 221: Florida State University Libraries - FLVC

202

LIST OF REFERENCES

1. Diels, O.; Alder, K. Cause of the ―azo-ester‖ reaction. Ann. 1926, 450, 237-254.

2. Diels, O.; Alder, K. Synthesis in the hydroaromatic series. I. Addition of ―diene‖ hydrocarbons. Ann. 1928, 460, 98-122.

3. Diels, O.; Alder, K. Synthesis in the hydroaromatic series. V. -4-Tetrahydro-o-phthalic acid (reply to communication of E. H. Farmer and F. L. Warren: Properties of conjugated double bonds. (7)). Ber. 1929, 62B, 2087-2090.

4. Diels, O.; Alder, K.; Pries, P. Synthesis in the hydroaromatic series. IV. Addition of maleic anhydride to arylated dienes, trienes, and fulvenes. Ber. 1929, 62B, 2081-2087

5. Komnenos, T. The reaction of unsaturated aldehydes with malonic acid and ethyl malonate. Liebigs Ann. Chem. 1883, 218, 145-169.

6. Michael, A. J. Prakt. Chem./Chem.-Ztg. 1887, 35, 349.

7. Michael, A. Am. Chem. J. 1887, 9, 115.

8. Michael, A. Addition of sodium acetoacetate and sodium diethyl malonate to unsaturated acids. J. Prakt. Chem./Chem.-Ztg. 1994, 49, 20.

9. Evans, D. A.; Bartroli, J.; Shih, T. L. Enantioselective aldol condensations. 2. Erythro-selective chiral aldol condensations via boron enolates. J. Am. Chem. Soc. 1981, 103, 2127-2109.

10. Cassar, L. Synthesis of aryl- and vinyl-substituted acetylene derivatives by use of nickel and palladium complexes. J. Organomet. Chem. 1975, 93, 253-257.

11. Dieck, H. A., Heck, F. R. Palladium catalyzed synthesis of aryl, heterocyclic, and vinylic acetylene derivatives. J. Organomet. Chem. 1975, 93, 259-263.

12. Sonogashira, K.; Tohda, Y.; Hagihara, N. Convenient synthesis of acetylenes. Catalytic substitutions of acetylenic hydrogen with bromo alkenes, iodo arenes, and bromo pyridines. Tetrahedron Lett. 1975, 4467-4470.

13. Mukaiyama, T. The directed aldol reaction. Org. React. 1982, 28, 203-331.

14. Heathcock, C. H. The aldol reaction: acid and general base catalysis. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I. Eds.; Pergamon: Oxford, 1991; Vol. 2, pp 133-238.

Page 222: Florida State University Libraries - FLVC

203

15. Rickborn, B. The retro-Diels-Alder reaction. Part I. C-C Dienophiles. Org. React.1998, 52, 1-393.

16. Sweger, R. W.; Czarnik, A. W. Retrograde Diels-Alder reactions. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I. Eds.; Pergamon: Oxford, 1991; Vol. 5, pp 551-592.

17. Bergmann, E. D.; Ginsburg, D.; Pappo, R. The Michael reaction. Org. React. 1959, 10, 179-555.

18. Jung, M. E. Stabilized nucleophiles with electron deficient alkenes and alkynes. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon: Oxford, 1991; Vol. 4, pp 1-67.

19. Boger, D. L.; Ichikawa, S.; Jiang, H. Total synthesis of the rubrolone aglycon. J. Am. Chem. Soc. 2000, 122, 12169-12173.

20. Takasu, K.; Mizutani, S.; Noguchi, M.; Makita, K.; Ihara, M. Total synthesis of (+)-culmorin and (+)-longiborneol: an efficient construction of the tricycle[6.3.0.03,9]undecan-10-one by intramolecular double Michael addition. J. Org. Chem. 2000, 65, 4112-4119.

21. Furstner, A.; Ruiz-Caro, J.; Prinz, H.; Waldmann, H. Structure assignment, total synthesis, and evaluation of the phosphate modulating activity of glucolipsin A. J. Org. Chem. 2004, 69, 459-467.

22. Paterson, I.; Davies, R. D. M.; Marquez, R. Total synthesis of the callipeltoside aglycon. Angew. Chem., Int. Ed. Engl. 2001, 40, 603-607.

23. Jacobi, P. A; Craig, T. A.; Walker, D. G.; Arrick, B. A.; Frechette, R. F. Bis heteroannulation. 4. Facile synthesis of methylene acids, methylbutenolides, a- methyl-g-lactones, and related materials. Total synthesis of (+)-lingularone and (+)-petasalbine. J. Am. Chem. Soc. 1984, 106, 5585-5594.

24. Itagaki, N.; Sugahara, T.; Iwabuchi, Y. Expedient synthesis of potent cannabinoid receptor agonist (-)-CP55,940. Org. Lett. 2005, 7, 4181-4183.

25. Hill, R. K. Cope, oxy-Cope, and anionic oxy-Cope rearrangements. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds; Pergamon: Oxford, 1991; Vol. 5, pp 785-826.

26. Lutz, R. P. Catalysis of the Cope and Claisen rearrangements. Chem. Rev. 1984, 84, 205-247.

27. Lee, D. G.; Chen, T. Cleavage reactions. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon,: Oxford, 1991; Vol. 7, pp 541-591.

Page 223: Florida State University Libraries - FLVC

204

28. Shing, T. K. M. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon: Oxford, 1991; Vol. 7, pp 703-717.

29. Hundlicky, T.; Reed, J. E. Rearrangement of vinylcyclopropanes and related systems. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon: Oxford, 1991; Vol. 5, pp 899-870.

30. Jun, C. –H. Transition metal-catalyzed carbon–carbon bond activation. Chem. Soc. Rev. 2004, 33, 610-618 and references therein.

31. Murakami, M.; Ito, Y. In Activation of Unreactive Bonds and Organic Synthesis; Murai, S., Ed.; Springer: Berlin, 1999; pp 97-129.

32. Jennings, P. W.; Johnson, L. L. Metallacyclobutane complexes of the group eight transition metals: synthesis, characterizations, and chemistry. Chem. Rev. 1994, 94, 2241-2290.

33. Tsuji, J. In Handbook of Organopalladium Chemistry for Organic Synthesis; Negishi, E. –i., Ed.; Wiley: New York, 2002; Vol. 2. Pp 2643-2653.

34. Handbook of Metathesis; Grubbs, R. H., Ed.; Wiley-VCH: Weinheim, 2003.

35. Grubbs, R. H. Olefin metathesis. Tetrahedron 2004, 60, 7117-7140

36. Buchmeiser, M. R. Homogeneous metathesis polymerization by well-defined group VI and group VIII transition-metal alkylidenes: fundamentals and applications in the preparation of advanced materials. Chem. Rev. 2000, 100, 1565-1604.

37. Matsuda, T.; Shigeno, M.; Murakami, M. Asymmetric synthesis of 3,4-dihydrocoumarins by rhodium-catalyzed reaction of 3-(2-hydroxyphenyl)cyclobutanones. J. Am. Chem. Soc. 2007, 129, 12086-12087.

38. Chandler, C. L.; Phillips, A. J. A total synthesis of (±)-trans-kumausyne. Org. Lett. 2005, 7, 3493-3495.

39. Grob, C. A., Baumann, W. 1,4-Elimination reaction with simultaneous fragmentation. Helv. Chim. Acta 1955, 38, 594-610.

40. Grob, C. A. The principle of ethylogy in organic chemistry. Experientia 1957, 13, 126-129.

41. Grob, C. A. Heterolytic fragmentation. A class of organic reactions. Angew. Chem., Int. Ed. Engl. 1967, 6, 1-15.

42. Grob, C. A. Mechanisms and stereochemistry of heterolytic fragmentation. Angew. Chem., Int. Ed. Engl. 1969, 8, 535-546.

Page 224: Florida State University Libraries - FLVC

205

43. Weyerstahl, P., Marschall, H. Fragmentation reactions. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon: Oxford, 1991; Vol. 6, pp 1041-1070.

44. Wharton, P. S. Stereospecific synthesis of 6-methyl-trans-5-cyclodecenone. J. Org. Chem. 1961, 26, 4781-4782.

45. Wharton, P. S.; Hiegel, G. A.; Coombs, R. V. trans-5-cyclododecenone. J. Org. Chem. 1963, 28, 3217-3219.

46. Wharton, P. S.; Hiegel, G. A. Fragmentation of 1,10-decalindiol monotosylate. J. Org. Chem. 1965, 30, 3254-3257.

47. Wharton, P. S.; Baird, M. D. Conformation and reactivity in the cis,trans-2,6-cyclodecadienyl system. J. Org. Chem. 1971, 36, 2932-2938.

48. Njardarson, J. T.; Wood, J. L. Evolution of the synthetic approach to CP-243,114. Org. Lett. 2001, 3, 2431-2434.

49. Eschenmoser, A.; Felix, D.; Ohloff, G. New fragmentation reaction of ,-unsaturated carbonyls. Synthesis of exaltone and rac-muscone from cyclododecanone. Helv. Chim. Acta 1967, 50, 708-713.

50. Felix, D.; Schreiber, J.; Ohloff, G. Eschenmoser, A. Synthetic Methods. 3. ,- Epoxyketone alkynone fragmentation. I. Synthesis of exaltone and (+/-)-muscone from cyclododecanone. Helv. Chim. Acta 1971, 54, 2896-2912.

51. Tanabe, M.; Crowe, D. F.; Dehn, R. L. Novel fragmentation reaction of ,-epoxyketones. Synthesis of acetylenic ketones. Tetrahedron Lett. 1967, 3943-3946.

52. Tanabe, M.; Crowe, D. F.; Dehn, R. L.; Detre, G. Synthesis of secosteroid acetylenic ketones. Tetrahedron Lett. 1967, 3739-3743.

53. Murakami, Y.; Shindo, M.; Shishido, K. Enantiocontrolled synthesis of a chiral building block via diastereoselective ring-closing metathesis. Synlett 2005, 664-666.

54. Fehr, C.; Galindo, J.; Etter, O. An efficient enantioselective synthesis of (+)-(R,Z)-5-Muscenone and (-)-(R)-Muscone – an example of a kinetic resolution and enantioconvergent transformation. Eur. J. Org. Chem. 2004, 1953-1957.

55. Mander, L. N.; McLachlan, M. M. The total synthesis of the galbulimima alkaloid GB 13. J. Am. Chem. Soc. 2003, 125, 2400-2401.

56. Dai, W.; Katzenellenbogen, J. A. New approaches to the synthesis of alkyl-substituted enol lactone systems, inhibitors of serine protease elastase. J. Org. Chem. 1993, 58, 1900-1908.

Page 225: Florida State University Libraries - FLVC

206

57. Trost, B. M.; Chang, V. K. An approach to botrydianes: on the steric demands of a metal catalyzed enyne metathesis. Synthesis 1993, 824-832.

58. Gordon, D. M.; Danishefsky, S. J.; Schulte, G. K. Studies in the benzannulation of cycloalkynone: an approach to the synthesis of antibiotics containing the benz[a]anthracene core structure. J. Org. Chem. 1992, 57, 7052-7055.

59. Woods, G. F.; Tucker, I. W. The reaction of b-cyclohexanedione (dihydroresorcinol) and its ethyl enol ether with phenylmagnesium bromide. J. Am. Chem. Soc. 1948, 70, 2174-2177.

60. Zimmerman, H. E.; Nesterov, E. E. An experimental and theoretical study of the type C enone rearrangement: mechanistic and exploratory organic photochemistry. J. Am. Chem. Soc. 2003, 125, 5422-5430.

61. Kamijo, S.; Dudley, G. B. A tandem carbanion addition/carbon-carbon bond cleavage yields alkynyl ketones. J. Am Chem. Soc. 2005, 127, 5028-5029.

62. Saito, Y.; Yamamoto, H.; Kageyama, H.; Nakamura, O.; Miyoshi, T.; Matsuoka, M. Investigation of the solution condition of lithium electrolyte solutions with LiCF3SO3 salt. J. Mater. Sci. 2000, 35, 809-812.

63. Hori, M.; Mori, M. Synthesis of heterocycles utilizing N2—TiCl4—Li—TMSCl. J. Org. Chem. 1995, 60, 1480-1481.

64. Wickman, B. E.; Mason, R. R.; Trostle, G. C. Douglas-fir tussock moth. Forest and Insect and Disease Leaflet 86; United States Department of Agricultural Forest Services, 1981.

65. Federal Register 39; 1974, no. 44, pp 8377-8381.

66. Smith, R. G.; Daterman, G. E.; Davies, G. D., Jr. Douglas-fir tussock moth pheromone identification and synthesis. Science 1975, 188, 63-64.

67. Beckwith, R. C. Influence of host foliage on the Douglas-fir tussock moth. Environ. Entomol. 1976, 5, 73-77.

68. Wickman, B. E. In The Douglas-Fir Tussock Moth: A Synthesis; Brooks, M. H.; Stark, R. W.; Campbell, R. W. Eds.; U.S. Forest Serv. Tech. Bull., 1978; Vol. 1585, pp 66-75.

69. Wickman, B. E.; Mason, R. R.; Paul, H. G. Flight, attraction, and mating behavior of the Douglas-fir tussock moth in Oregon. Environ. Entomol. 1975, 4, 405-408.

70. Torgerson, T. R. Identification of parasites of the Douglas-fir tussock moth based on adults, cocoons, and puparia. United States Department of Agriculture Forest Service

Page 226: Florida State University Libraries - FLVC

207

Research Paper; Pacific Northwest Forest and Range Experiment Station, 1977, PNW-215; pp 28.

71. Evenden, J.; Jost, E. J. A report of the tussock moth control. United States Department of Agriculture: North Idaho, 1947; pp 51.

72. Daterman, G. E. In The Douglas-Fir Tussock Moth: A Synthesis; Brooks, M. H.; Stark, R. W.; Campbell, R. W. Eds.; U.S. Forest Serv. Tech. Bull., 1978; Vol. 1585, pp 902-102.

73. Kelleher, K. Palmer Mountain Douglas-Fir Tussock Moth Control; United States Department of the Interior: Bureau of Land Management, Wenatchee Field Office, 2009, Project No. 09-0026; website: http://www.blm.gov/or/districts/spokane/plans/files/Palmer_Mtn_Tussock_Moth_EA_Final.pdf

74. Welter, S. C.; Pickel, C.; Millar, J. Cave, F.; Van Steenwyk, R. A; Dunley, J. Pheromone mating disruption offers selective management options for key pests. Calif. Agric. 2005, 59, 16-22.

75. Hulme, M.; Gray, T. G. Mating disruption of the Douglas-fir tussock moth (Lepidoptera: Lymantriidae) using a sprayable formulation of (Z)-6-heneicosen-11-one. Environ. Entomol. 1994, 23, 1097-1100.

76. Hulme, M.; Gray, T. G. Pheromone dosage on mating disruption of the Douglas-fir tussock moth. J. Entomol. Soc. B. C. 1997, 93, 99-104.

77. Environmental Protection Agency, Registration of Lepidopteran Pheromones Federal Register Notices, (Z)-6-Heneicosen-11-one (129060), 22 Jun 2005, website: http://www.epa.gov/pesticides/biopesticides/ingredients/fr_notices/frnotices_lep_pheromones.htm#129060

78. Smith, R. G.; Davies, G. D., Jr.; Daterman, G. E. Synthesis of (Z)-6-heneicosen-11-one. Douglas-fir tussock moth sex attractant. J. Org. Chem. 1975, 40, 1593-1595.

79. Kocienski, P. J.; Cernigliaro, G. J. Pheromone synthesis. II. A synthesis of (Z)-6-heneicosen-11-one. The sex pheromone of the Doulas-fir tussock moth. J. Org. Chem. 1976, 41, 2927-2928.

80. Mori, K.; Uchida, M.; Matsui, M. Synthesis of aliphatic insect pheromones from alicyclic starting materials: (Z)-6-heneicosen-11-one and (Z)-8-dodecenyl acetate. Tetrahedron 1977, 33, 385-387.

81. Fetizon, M.; Lazare, C. A rapid and convenient synthesis of the sex pheromone of the Douglas-fir Lepidoptera Orgyia pseudotsugata. J. Chem. Soc., Perkin Trans. 1 1978, 842-844.

Page 227: Florida State University Libraries - FLVC

208

82. Akermark, B.; Ljungqvist, A. Eutectic potassium-sodium-aluminum chloride as a mild catalyst for ene reactions: simple synthesis of the sex pheromone from the Douglas-fir tussock moth. J. Org. Chem. 1978, 43, 4387-4388.

83. Trost. B. M.; Ornstein, P. L. A chemoselective desulfurization method via homogeneous nickel catalysis. Tetrahedron Lett. 1981, 22, 3464-3466.

84. Murata, Y.; Inomata, K.; Kinoshita, H.; Kotake, H. A convenient method for the synthesis of acyclic ketones. Synthesis of sex pheromone of Doulas-fir tussock moth. Bull. Chem. Soc. Jpn. 1983, 56, 2539-2540.

85. Reddy, G. B.; Mitra, R. B. Acetone dimethyl hydrazone: useful synthon for the synthesis of pheromones with keto and spiro ketal functions. Synth. Commun. 1983, 16, 1723-1729.

86. Pramod, K.; Ramanathan, H.; Subba Rao, G. S. R. Strategic synthesis based on cylcohexadienes: preparation of 2-, 2,3-, and 2,4-substituted cyclohexenones; synthesis of (Z)-heneicosen-11-one. J. Chem. Soc., Perkin Trans. 1 1983, 7-10.

87. Sodeoka, M.; Shibasaki, M. New functions of (arene)tricarbonylchromium(0) complexes as hydrogenation catalysis: stereospecific semihydrogenation of alkynes and highly chemoselective hydrogenation of ,-unsaturated carbonyl compounds. J. Org. Chem. 1985, 50, 1147-1149.

88. Wenkert, E.; Ferreira, V. F.; Michelotti, E. L.; Tingoli, M. Synthesis of acyclic, cis olefinic pheromones by way of nickel-catalyzed Grignard reactions. J. Org. Chem. 1985, 50, 719-721.

89. Ballini, R. New and convenient synthesis of (Z)-6-heneicosen-11-one, the Douglas-fir tussock moth (Orgyia pseudotsugata) sex pheromone, and (Z)-non-6-en-2-one, the immediate precursor for the synthesis of brevicomin, the sex attractant of the western pine beetle Dentroctonus brevicomis. J. Chem. Soc. Perkin Trans. 1 1991, 1419-1421.

90. Stowell, J. C.; Polito, M. A. A .delta.-keto aldehyde synthesis: application to the preparation of the sex pheromone of the Douglas-fir tussock moth. J. Org. Chem. 1992, 57, 4560-4562.

91. White, J.; Whiteley, C. G. Use of organoboranes in the synthesis of pheromones: a convenient synthesis of (Z)-6-heneicosen-11-one and (Z)-5-undecen-11-one, pheromones of the Douglas-fir tussock moth and the bontebok, respectively. Synthesis 1993, 1141-1144.

92. Hayes, J. F.; Shipman, M.; Twin, H. Multicomponent reactions involving 2-methyleneaziridines: rapid synthesis of 1,3-disubstituted propanones. J. Org. Chem. 2002, 67, 935-942.

Page 228: Florida State University Libraries - FLVC

209

93. Jones, D. M.; Kamijo, S.; Dudley, G. B. Grignard-triggered fragmentation of vinylogous acyl triflates: synthesis of (Z)-6-heneicosen-11-one, the Douglas fir tussock moth sex pheromone. Synlett 2005, 936-938.

94. Prices Made Available from Sigma-Aldrich: http://www.sigmaaldrich.com/united-states.html.

95. Rosenmund, K. W.; Bach, H. Chem. Ber. 1961, 94, 2394.

96. Piers, E.; Grierson, J. R. Alkylation of 1,5-dimethoxy-1,4-cyclohexadiene. A convenient synthesis of 2-alkyl- and 2-alkenyl-1,3-cyclohexanediones. J. Org. Chem. 1977, 42, 3755-3757.

97. Fraga, B. M. Natural sesquiterpenoids. Nat. Prod. Rep. 2008, 25, 1180-1209.

98. Ayer, W. A.; McCaskill, R. H. The cybrodins, a new class of sesquiterpenes. Can. J. Chem. 1981, 59, 2150-2158.

99. Arnone, A.; Cardillo, R.; Di Modugno, V.; Nasini, G. Seconday mould metabolites. Part 29. Isolation and structure elucidation of candicansol, 3-epi-illudol and 1-O-acetyl-3-epi-illudol, novel sesquiterpenoids from Clitocybe candicans, and absolute configuration of 3-epi-illudol. J. Chem. Soc., Perkin Trans. 1 1989, 1995-2000.

100. Clericuzio, M.; Pan, F.; Han, F.; Pang, Z.; Sterner, O. Stearoyldelicone, an unstable protoilludane sesquiterpenoids from intact fruit bodies of Russula delica. Tetrahedron Lett. 1997, 38, 8237-8240.

101. Saito, K.; Nagao, T.; Matoba, M.; Koyama, K.; Natori, S.; Murakami, T.; Saiki, Y. Chemical assay of ptaquiloside, the carcinogen of Pteridium aquilinum, and the distribution of related compounds in the pteridaceae. Phytochemistry 1989, 28, 1605-1611.

102. Hikino, H.; Takahashi, T.; Arihara, S.; Takemoto, T. Structure of Pteroside B, glycoside of Pteridium aquilinum var latiusculum. Chem. Pharm. Bull. 1970, 18, 1488-1489.

103. Yoshihira, K.; Fukuoka, M.; Kuroyanagi, M.; Natori, S. Chemical and toxicological studies on bracken fern, Pteridium aquilinum var. latiusculum. I. Introduction, extraction, and fractionation of constituents, and toxicological studies including carcinogenicity tests. Chem. Pharm. Bull. 1978, 26, 2346-2364.

104. Fukuoka, M.; Kuroyanagi, M.; Yoshihira, K.; Natori, S. Chemical and toxicological studies on bracken fern, Pteridium aquilinum var. latiusculum. II. Structures of pterosins, sesquiterpenes having 1-indanone skeleton. Chem. Pharm. Bull. 1978, 26, 2365-2385.

105. Yamada, K.; Ojika, M.; Kigoshi, H. Ptaquiloside, the major toxin of braken, and related terpene glycosides: chemistry, biology, and ecology. Nat. Prod. Rep. 2007, 24, 798-813.

Page 229: Florida State University Libraries - FLVC

210

106. Suzuki, S.; Muriayama, T.; Shiono, Y. Illudalane sesquiterpenoids, echinolactones A and B, from a mycelia culture of Echinodontium japonicum. Phytochemistry 2005, 66, 2329-2333.

107. Muriayama, T.; Tanaka, N. Prog. Chem. Org. Nat. Prod. 1988, 54, 1-355 and references therein.

108. Palmero, J. A.; Brasco, M. F. R.; Spagnuolo, C. Seldes, A. M. Illudalane sesquiterpenoids from the soft coral Alcyonium paessleri: the first natural nitrate esters. J. Org. Chem. 2000, 65, 4482-4486.

109. Carbone, M.; Nunez-Pons, L.; Castellucio, F.; Avila, C.; Gavagnin, M. Illudalane sesquiterpenoids of the alcyopterosin series from the Antarctic marine soft coral Alcyonium grandis. J. Nat. Prod. 2009, 72, 1357-1360.

110. Sengupta, P.; Sen, M.; Niyogi, S. K.; Pakrashi, S. C.; Ali, E. Isolation and structure of wallichoside, a novel pteroside from Pteris wallichiana. Phytochemistry 1976, 6, 995-998.

111. Castillo, U. F.; Sakagama, Y.; Alonso-Amelot, M. Ojiva, M. Pteridanoside, the first protoilludane sesquiterpene glucoside as a toxic component of neotropical bracken fern Pteridium aquilinum var. caudatum. Tetrahedron 1999, 55, 12295-12300.

112. Sheridan, H.; Lemon, S.; Frankish, N.; McArdle, P.; Higgins, T.; James, J.; Bhandari, P. Synthesis and antispasmodic activity of nature and identical substituted indanes and analogs. Eur. J. Med. Chem. 1990, 25, 603-608.

113. Finkielsztein, L. M; Bruno, A. M.; Renou, S. G.; Iglesias, G. Y. M. Design, synthesis, and biological evaluation of alcyopterosin and illudalane derivatives as anticancer agents. Bioorg. Med. Chem. 2006, 14, 1863-1870.

114. Tanaka, R.; Nakano, Y.; Suzuki, D.; Urabe, H.; Sato, F. Selective preparation of benzyltitanium compounds by the metalative Reppe reaction. Its application to the first synthesis of alcyopterosin A. J. Am. Chem. Soc. 2002, 124, 9682-9683.

115. Witulski, B.; Zimmermann, A.; Gowans, N. D. First total synthesis of marine illudalane sesquiterpenoids alcyopterosin E. Chem. Comm. 2002, 2984-2985.

116. Nakao, Y.; Hirata, Y.; Ishihara, S.; Oda, S.; Yukawa, T.; Shirakawa, E.; Hiyama, T. Stannylative cycloaddition of enynes catalyzed by palladium-iminophosphine. J. Am. Chem. Soc. 2004, 126, 15650-15651.

117. Jones, A. L.; Snyder, J. K. Intramolecular rhodium-catalyzed [2+2+2] cyclizations of diynes with enones. J. Org. Chem. 2009, 74, 2907-2910.

Page 230: Florida State University Libraries - FLVC

211

118. Reppe, W.; Schweckendiek, W. J. Cyclisierende polymerization von acetylene. III. Benzol, benzolderivate und hydroaromatische verbindungen. Justus Liebigs Ann. Chem. 1948, 560, 104-116.

119. Saito, S.; Kawasaki, T.; Tsuboya, N.; Yamamoto, Y. Highly regioselective

cyclotrimerization of 1-perfluoroalkynes catalyzed by nickel. J. Org. Chem. 2001, 66, 796-802.

120. Ozerov,O. V.; Patrick, B. O.; Ladipo, F. T. Highly regioselective [2+2+2] cycloaddition of terminal alkynes catalyzed by 6-arene complexes of titanium supported dimethylsilyl-bridged p-tert-butyl calyx[4]arene ligand. J. Am. Chem. Soc. 2000, 122, 6423-6431.

121. Ozerov, O. V.; Ladipo, F. T.; Patrick, B. O. Highly regioselective alkyne cyclotrimerization catalyzed by titanium complexes supported by proximally bridged p-tert-butyl calyx[4]arene ligand. J. Am. Chem. Soc. 1999, 121, 7941-7942.

122. Chio, K. S.; Park, M. K.; Han, B. H. J. Chem. Res. (S) 1998, 518-519.

123. Hill, J. E.; Balaich, G.; Fanwick, P. E.; Rothwell, I. P. The chemistry of titanacyclopentadiene rings supported by 2,6-diphenoxide ligation: stoichiometric and catalytic reactivity. Organometallics 1993, 12, 2911-2924.

124. Kataoka, Y.; Takai, K.; Oshima, K.; Utimoto, K. Selective reduction of alkynes to (Z)-alkenes via niobium- or tantalum-alkyne complexes. J. Org. Chem. 1992, 57, 1615-1618.

125. Jhingan, A. K.; Maier, W. F. Homogeneous catalysis with a heterogeneous palladium catalyst. An effective method for the cyclotrimerization of alkynes. J. Org. Chem. 1987, 52, 1161-1165.

126. Mori, N.; Ikeda, S.; Odashima, K. Chemo- and regioselective cyclotrimerization of monoynes catalyzed by a nickel(0) and zinc(II) phenoxide system. Chem. Commun. 2001, 181-182.

127. Sato, Y.; Ohashi, K.; Mori, M. Synthesis of biaryls using nickel-catalyzed [2+2+2] co-cyclization. Tetrahedron Lett. 1999, 40, 5231-5234.

128. Takahashi, T.; Tsai, F.-Y.; Li, Y.; Nakajima, K.; Kotora, M. Carbon-carbon bond formation reaction of zirconacyclopentadienes with alkynes in the presence of Ni(II)-complexes. J. Am. Chem. Soc. 1999, 121, 11093-11100.

129. Takahashi, T.; Xi, Z.; Yamazaki, A.; Liu, Y.; Nakajima, K.; Kotora, M. Cycloaddition reaction of zirconacyclopentadienes to alkynes: highly selective formation of benzene derivatives from three different alkynes. J. Am. Chem. Soc. 1998, 120, 1672-1680.

Page 231: Florida State University Libraries - FLVC

212

130. Kotha, S.; Mohanraja, K.; Durani, S. Constrained phenallyl peptides via a [2+2+2]-cycloaddtition strategy. Chem. Commun. 2000, 1909-1910.

131. Witulski, B.; Stengel, T.; Fernandez-Hernandez, J. M. Chemo- and regioselective crossed alkyne cyclotrimerization of 1,6-diynes with terminal monoalkynes mediated by Grubbs’ catalyst or Wilkinson’s catalyst. Chem. Commun. 2000, 1965-1966.

132. Yamamoto, Y.; Ogawa, R.; Itoh, K. Highly chemo- and regioselective [2+2+2] cycloaddition of unsymmetrical 1,6-diynes with terminal alkynes catalyzed by Cp*Ru(cod)Cl under mild conditions. Chem. Commun. 2000, 549-550.

133. Witulski, B.; Stengel, T. Rhodium(I)-catalyzed [2+2+2] cycloadditions with N-functionalized 1-alkynyl amides: a conceptually new strategy for the regiospecific synthesis of indolines. Angew. Chem., Int. Ed. 1999, 38, 2426-2430.

134. Yamamoto, Y.; Nagata, A.; Itoh, K. Pd(0)-catalyzed [2+2+2] cycloaddition of dimethyl nona-2,7-diyne-1,9-dioate derivatives with dimethyl acetylendicarboxylate. Tetrahedron Lett. 1999, 40, 5035-5038.

135. Negishi,E.; Harring, L. S.; Owczarczyk, Z.; Mohamud, M. M.; Ay, M. Cyclic cascade carbopalladation reactions as a route to benzene and fulvene derivatives. Tetrahedron Lett. 1992, 33, 3253-3256.

136. Suzuki, D.; Urabe, H.; Sato, F. Metalative Reppe reaction. Organized assembly of acetylene molecules on titanium template leading to a new style of acetylene cyclotrimerization. J. Am. Chem. Soc. 2001, 123, 7925-7926.

137. Asis, S.; Bruno, A. M.; Martinez, A. R., Sevilla, M. V.; Gaozzo, C. H.; Romano, A. M.; Coussio, J. D.; Ciccia, G. Diarylsemicarbazones: synthesis, antineoplastic activity and topoisomerase I inhibition assay. Il Farmaco 1999, 54, 517-523.

138. Obtained from the Dudley Laboratory Research Webpage: http://www.chem.fsu.edu/dudley/overview.html

139. Asao, N.; Takahashi, K.; Lee, S.; Kashahara, T.; Yamamoto, Y. AuCl3-catalyzed benzannulation: synthesis of naphthyl ketone derivatives from o-alkynylbenzaldehydes with alkynes. J. Am. Chem. Soc. 2002, 124, 12650-12651.

140. Asao, N.; Nogami, T.; Lee, S.; Yamamoto, Y. Lewis acid-catalyzed benzannulation via unprecedented [4+2] cycloaddition of o-alkynyl(oxo)benzenes and enynals with alkynes. J. Am. Chem. Soc. 2003, 125, 10921-10925.

141. Asao, N.; Aikawa, H.; Yamamoto, Y. AuBr3-catalyzed [4+2] benzannulation between an enynal unit and enol. J. Am. Chem. Soc. 2004, 126, 7458-7459.

Page 232: Florida State University Libraries - FLVC

213

142. Asao, N.; Sato, K.; Menggenbateer; Yamamoto, Y. AuBr3- and Cu(OTf)2-catalyzed intramolecular [4+2] cycloaddition of tethered alkynyl and alkenyl enynones and enynals: a new synthetic method for functionalized polycyclic hydrocarbons. J. Org. Chem. 2005, 70, 3682-3685.

143. Barluenga, J.; Vasquez-Villa, H.; Ballesteros, A.; Gonzalez, J. M. Regioselective synthesis of substituted naphthalenes: a novel de novo approach based on metal-free protocol for stepwise cycloaddition of o-alkynylbenzaldehyde derivatives with either alkynes or alkenes. Org. Lett. 2003, 5, 4121-4123.

144. Yue, D.; Ca, N. D.; Larock, R. C. Efficient syntheses of heterocycles and carbocycles by electrophilic cyclization of acetylenic aldehydes and ketones. Org. Lett. 2004, 6, 1581-1584.

145. Sato, K.; Asao, N.; Yamamoto, Y. Efficient method for synthesis of angucyclinone antibiotics via gold-catalyzed intramolecular [4+2] benzannulation: enantioselective total synthesis of (+)-ochryomycinone and (+)-rubiginone B2. J. Org. Chem. 2005, 70, 8977-8981.

146. Lemhadri, M.; Doucet, H.; Santelli, M. Sonogashira reaction of aryl halides with propiolaldehyde diethyl acetal catalyzed by a tetraphosphine/palladium complex. Tetrahedron 2005, 61, 9839-9847.

147. Kehoe, J. M.; Kiley, J. H.; English, J. J.; Johnson, C. A.; Peterson, R. C.; Haley, M. M. Carbon networks based on dehydrobenzoannulenes. 3. Synthesis of graphyne substructures. Org. Lett. 2000, 2, 969-972.

148. Kimball, D. B.; Haley, M. M. Triazenes: a versatile tool in organic synthesis. Angew. Chem. Int. Ed. 2002, 41, 3338-3351.

149. Moore, J. S.; Weinstein, E. J.; Wu, Z. A convenient masking group for aryl iodides. Tetrahedron Lett. 1991, 22, 2465-2466.

150. Satyamurthy, N.; Barrio, J. R. Cation exchange resin (hydrogen form) assisted decomposition of 1-aryl-3,3-dialkyltriazenes. A mild and efficient method for the synthesis of aryl iodides. J. Org. Chem. 1983, 48, 4396-4397.

151. Lertpibulpanya, D.; Marsden, S. P. Concise access to indolizidine and pyrroloasepine skeletal via intramolecular Schmidt reactions of azido 1,3-diketones. Org. Biomol. Chem. 2006, 4, 3498-3504.

152. Wang, C.; Tobrman, T.; Xu, Z.; Negishi, E. Highly regio- and stereoselective synthesis of (Z)-trisubstituted alkenes via propyne bromoboration and tandem Pd-catalyzed cross-coupling. Org. Lett. 2009, 11, 4092-4095.

Page 233: Florida State University Libraries - FLVC

214

153. American Cancer Society website review of melanoma: http://www.cancer.org/docroot/CRI/content/CRI_2_2_1X_What_is_melanoma_skin_cancer_50.asp?sitearea.

154. Cancer.netTM, oncologist-approved cancer information from the American Society of Clinical Oncology: http://www.cancer.net/patient/Cancer+Types/Melanoma?sectionTitle=Overview.

155. Tucker, M. A. Melanoma epidemiology. Hematol. Oncol. Clin. N. Am. 2009, 23, 383-395.

156. National Cancer Institute, from the U.S. National Institutes of Health: http://www.cancer.gov/cancertopics/wyntk/melanoma/page1.

157. Atzpodien, J.; Morawek, L.; Fluck, M. Bleomycin, vinorelbine and trofosamide in relapsed stage IV cutaneous malignant melanoma patients. Cancer Chemother. Pharmacol. 2009, 64, 901-905.

158. Yang, A. S.; Chapman, P. B. The history and future of chemotherapy for melanoma. Hematol. Oncol. Clin. N. Am. 2009, 23, 583-597.

159. Bedikan, A. Y.; Papadopoulos, N. E.; Kim, K. B.; Hwe, W. –J.; Homsl, J.; Glass, M. R.; Cain, S.; Rudewicz, P.; Vernillet, L.; Hwe, P. A phase IB trial of intravenous INO-1001 plus oral temozolomide in subjects with unresurrectable stage—III or IV melanoma. Cancer Investigation 2009, 27, 756-763.

160. Bakus, B. J.; Targett, N. M.; Schulte, B. Chemical ecology of marine organisms: an overview. J. Chem. Ecol. 1986, 12, 951-987.

161. In Ecological roles of marine natural products. Paul, V. J., Ed.; Comstock: Ithaca, New York, 1992.

162. Pawlik, J. R. Marine invertebrate chemical defense. Chem. Rev. 1993, 93, 1911-1922.

163. Hay, M. E. Marine chemical ecology: what’s known and what’s next. J. Exp. Mar. Biol. Ecol. 1996, 200, 103-134.

164. McClintock, J. B.; Baker, B. J. A review of the chemical ecology of Antarctic marine invertebrates. Amer. Zool. 1997, 37, 329-342.

165. Bakus, G. J.; Green, G.; Toxicity in sponges and holothurians: a geographic pattern. Science 1974, 185, 951-953.

166. Green, G. Ecology of toxicity of marine sponges. Mar. Biol. 1977, 40, 207-215.

Page 234: Florida State University Libraries - FLVC

215

167. Bakus, G. J. Chemical defense mechanisms in the Great Barrier Reef, Australia. Science 1981, 211, 497-498.

168. Dayton, P. K; Robilliard, G. A.; DeVries, A. L. Anchor ice formation in McMurdo Sound, Antarctica and its biological effects. Science 1969, 163, 273-274.

169. Dayton, P. K.; Robilliard, G. A.; Paine, R. T.; Dayton, L. B. Biological accommodation in the benthic community at McMurdo Sound, Antarctica. Ecol. Monogr. 1974, 44, 105-128.

170. Dayton, P. K.; Mordida, B. J.; Bacon, F. Polar marine communities. Amer. Zool. 1994, 34, 90-99.

171. Eastman, J. T. In Antarctic fish biology: evolution in a unique environment. Academic Press, Inc.: New York, 1993.

172. McClintock, J. B. Trophic biology of shallow-water Antarctic echinoderms. Mar. Ecol. Prog. Ser. 1994, 111, 191-202.

173. Amsler, C. D.; Iken, K. B.; McClintock, J. B.; Baker, B. J. In Marine Chemical Ecology, McClintock, J. B.; Baker, B. J. Eds.; CRC Press: Boca Raton, FL, 2001; 267-300.

174. Information available through the Antarctic Treaty Secretariat: http://www.ats.aq/.

175. Diyabalanage, T.; Amsler, C. D.; McClintock, J. B.; Baker, B. J. Palmerolide A, a cytotoxic macrolide from the Antarctic tunicate Synoicum adareanum. J. Am. Chem. Soc. 2006, 128, 5630-5631.

176. Martinez-Ziguilan, R.; Lynch, R. M.; Martinez, G. M.; Gillies, R. J. Am. J. Physiol. Cell Physiol. 1993, 265, C1015-C1029.

177. Boyd, M. R.; Farina, C.; Belfiore, P.; Gagliardi, S.; Kim, J. W.; Hayakawa, Y.; Beutler, J. A.; McKee, T. C.; Bowman, B. J.; Bowman, E. J. J. Pharmacol. Exper. Ther. 2001, 297, 114-120.

178. Xie, X. S.; Pardon, D.; Liao, X.; Wang, J. Roth, M. G.; De Brabander, J. K. Salicylihalamide A inhibits the V0 sector of the V-ATPase through a mechanism distinct from Bafilomycin A1. J. Biol. Chem. 2004, 279, 19755-19763.

179. Bowman, E. J.; Siebers, A.; Karlheinz, A. Bafilomycins: a class of inhibitors of membrane ATPases from microorganisms, animal cells, and plant cells. Proc. Natl. Acad. Sci. U.S.A. 1988, 85, 7972-7976.

180. Kim, J. W.; Shin-ya, K.; Furihata, K.; Hayakawa, Y.; Seto, H. Oximidines I and II: novel antitumor macrolides from Pseudomonas sp. J. Org. Chem. 1999, 64, 153-155.

Page 235: Florida State University Libraries - FLVC

216

181. Obtained through personal communications with Bill Baker.

182. Jiang, X.; Liu, B.; Lebreton, S.; De Brabander, J. K. Total synthesis and structural revision of the marine metabolite palmerolide A. J. Am. Chem. Soc. 2007, 129, 6386-6387.

183. Nicolaou, K. C.; Guduru, R.; Sun, Y. –P.; Banerji, B.; Chen, D. Y. –K. Total synthesis of the originally proposed and revised structure of palmerolide A. Angew. Chem. Int. Ed. 2007, 46, 5896-5900.

184. Jagel, J.; Maier, M. E. Formal synthesis of palmerolide A. Synthesis 2009, 2881-2892.

185. Penner, M.; Rauniyar, V.; Kaspar, L. T.; Hall, D. G. Catalytic asymmetric synthesis of palmerolide A via organoboron methodology. J. Am. Chem. Soc. 2009, 131, 14216-14217.

186. Kaliappan, K.; Gowrisankar, P. Synthetic studies on the marine natural product, palmerolide A: synthesis of C1-C9 and C15-C21 fragments. Synlett 2007, 1537-1540.

187. Chandrasekhar, S.; Vijeender, K. Chandrasekhar, G.; Reddy, Ch. R. Towards the synthesis of palmerolide A: asymmetric synthesis of C1-C14 fragment. Tetrahedron: Asymmetry 2007, 18, 2473-2478.

188. Cantagrel, G.; Meyer, C.; Cossy, J. Synthetic studies towards the marine natural product palmerolide A: synthesis of the C3-C15 and C16-C23 fragments. Synlett 2007, 2983-2986.

189. Shirokawa, S. –i.; Kamiyama, M.; Nakamura, T.; Okada, M.; Nakazaki, A.; Hosokawa, S.; Kobayashi, S. Remote asymmetric induction with vinylketene N,O-acetal. J. Am. Chem. Soc. 2004, 126, 13604-13605.

190. Marshall, J.; Eidam, P. Diastereoselective additions of vinylzinc reagents to -chiral aldehydes. Org. Lett. 2004, 6, 445-448.

191. Urbansky, M.; Davis, C. E.; Surjan, J. D.; Coates, R. M. Synthesis of enantiopure 2-C-methyl-D-erythritol-4-phosphate and 2,4-cyclodiphosphate from D-arabitol. Org. Lett. 2004, 6, 135-138.

192. Takai, K.; Shinomiya, N.; Kai, A.; Saigo, K. Transformation of aldehydes into (E)-1-alkenylboronic esters with a germinal dichromium reagent derived from a dichloromethylboronic ester and CrCl2. Synlett 1995, 963-964.

193. Govoni, M.; Lim, H. D.; El-Atmioui, D.; Menge, W. M. P. B.; Timmerman, H.; Bakker, R. A.; Leurs, R.; De Esch, I. J. P. A chemical switch for the modulation of the functional activity of higher homologues of histamine on the human histamine H3 receptor: effects

Page 236: Florida State University Libraries - FLVC

217

of various substitutions at the primary amino function. J. Med. Chem. 2006, 49, 2549-2557.

194. Inanaga, J.; Hirata, K.; Saeki, H.; Kastuki, T.; Yamaguchi, M. A rapid esterification by means of mixed anhydride and its application of large-ring lactonization. Bull. Chem. Soc. Jpn. 1979, 52, 1989-1993.

195. De Mico, A.; Margarita, R.; Parlanti, L.; Vescovi, A.; Piancatelli, G. A versatile and highly selective hypervalent iodine (III)/2,2,6,6-tetramethyl-1-piperidinyloxy-mediated oxidation of alcohols to carbonyl compounds. J. Org. Chem. 1997, 62, 6974-6977.

196. Stork, G.; Nakamura, E. Large-ring lactones by internal ketophosphonate cyclizations. J. Org. Chem. 1979, 44, 4010-4011.

197. Evans, D. A.; Takacs, J. M.; McGee, L. R.; Ennis, M. D.; Mathre, D. J.; Bartrolli, J. M. Chiral enolate design. Pure Appl. Chem. 1981, 53, 1109-1127.

198. Makabe, h.;, Higuchi, M.; Konno, H.; Murai, M.; Miyoshi, H. Synthesis of (4R,15R,16R,21S)- and (4R,15S,16S,21S)-rollicosin. Tetrahedron Lett. 2005, 46, 4671-4675.

199. Brown, H. C.; Jadhav, P. K.; Bhat, K. S. Chiral synthesis via organoboranes. 13. A highly diastereoselective and enantioselective addition of [(Z)--alkoxyallyl]-diisopinocampheylboranes to aldehydes. J. Am. Chem. Soc. 1988, 110, 1535-1538.

200. Kocovsky, P. Carbamates: a method of synthesis and some synthetic applications. Tetrahedron Lett. 1986, 27, 5521-5524.

201. Boden, C. D. J.; Pattenden, G.; Ye, T. Palladium-catalyzed hydrostannylations of 1-bromoalkynes. A practical synthesis of (E)-1-stannylalk-1-enes. J. Chem. Soc., Perkin Trans. 1 1996, 2417-2419.

202. Schaus, S. E.; Brandes, B. D.; Larrow, J. F.; Tokunaga, M.; Hansen, K. B.; Gould, A. E.; Furrow, M. E.; Jacobsen, E. N. Highly selective hydrolytic kinetic resolution of terminal epoxides catalyzed by chiral (salen) Co(III) complexes. Practical synthesis of enantioenriched terminal epoxides and 1,2-diols. J. Am. Chem. Soc. 2002, 124, 1307-1315.

203. Yin, L.; Liebscher, J. Carbon—carbon coupling reactions catalyzed by heterogeneous palladium catalysts. Chem. Rev. 2007, 107, 133-173.

204. Takai, K.; Utimoto, N. K. Simple and selective method for aldehydes (E)-haloalkenes (RCH=CHX) conversion by means of a haloform-chromous chloride system. J. Am. Chem. Soc. 1986, 108, 7408-7410.

Page 237: Florida State University Libraries - FLVC

218

205. Jiang, L.; Job, G. E.; Klapars, A.; Buchwald, S. L. Copper-catalyzed coupling of amides and carbamates with vinyl halides. Org. Lett. 2003, 5, 3667-3669.

206. Nicolaou, K. C.; Sun, Y. –P.; Guduru, R.; Banerji, B.; Chen, D. Y. –K. Total synthesis of the original proposed structures of palmerolide A and isomers thereof. J. Am. Chem. Soc. 2008, 130, 3633-3644.

207. Nicolaou, K. C.; Leung, G. Y. C.; Dethe, D. H.; Guduru, R.; Sun, Y. –P.; Lim, C. S.; Chen, D. Y. –K. Chemical synthesis and biological evaluation of palmerolide A analogues.

208. Matsumura, K.; Hashiguchi, S.; Ikariya, T.; Noyori, R. Asymmetric transfer hydrogenation of ,-acetylenic ketones. J. Am. Chem. Soc. 1997, 119, 8738-8739.

209. Kolb, H. C.; VanNieuwnhze, M. S.; Sharpless, K. B. Catalytic asymmetric dihydroxylation. Chem. Rev. 1994, 94, 2483-2547.

210. Roth, G. J.; Liepold, B.; Muller, S. G.; Bestmann, H. J. Further improvement of the synthesis of alkynes from aldehydes. Synthesis 2004, 59-62.

211. Negishi, E. –i. Bull. Chem. Soc. Transition metal-catalyzed organometallic reactions that have revolutionized organic synthesis. Bull. Chem. Soc. Jpn. 2007, 80, 233-257.

212. Rauniyar, V.; Hall. D. G. Rationally improved chiral Brønsted acid for catalytic enantioselective allylboration of aldehydes with an expanded reagent scope. J. Org. Chem. 2009, 74, 4236-4241.

213. Gao, X.; Hall, D. G. 3-Boronoacrolein as an exceptional heterodyne in the highly enantio- and diastereoselective Cr(III)-catalyzed three component [4+2]allylboration. J. Am. Chem. Soc. 2003, 125, 9308-9309.

214. Burke, S. D.; Fobare, W. F.; Pacofsky, G. J. Chelation control of enolate geometry. Acyclic diastereoselection via the enolate Claisen rearrangement. J. Org. Chem. 1983, 48, 5221-5228.

215. Dossetter, A. G.; Jamison, T. F.; Jacobsen, E. N. Highly enantio- and diastereoselective hetero-Diels-Alder reactions catalyzed by new chiral tridentate chromium(III) catalysts. Angew. Chem. Int. Ed. 1999, 38, 2398-2400.

216. Kamijo, S.; Dudley, G. B.; Claisen-type condensation of vinylogous acyl triflates. Org. Lett. 2006, 8, 175-177.

217. Kamijo, S. Dudley, G. B. Tandem nucleophilic addition/fragmentation reactions and synthetic versatility of vinylogous acyl triflates. J. Am. Chem. Soc. 2006, 128, 6499-6507.

Page 238: Florida State University Libraries - FLVC

219

218. Tummatorn, J.; Dudley, G. B. Ring opening/fragmentation of dihydropyrones for the synthesis of homopropargyl alcohols. J. Am. Chem. Soc. 2008, 130, 5050-5051.

219. Furstner, A.; Langemann, K. Total synthesis of (+)-ricinelaidic acid lactone and (-) –gloesporone based on transition-metal-catalyzed C-C bond formations. J. Am. Chem. Soc. 1997, 119, 9130-9136.

220. Michaunt, A.; Boddaert, T.; Coquerel, Y.; Rodriguez, J. Reluctant cross-metathesis reactions: the highly beneficial effect of microwave irradiation. Synthesis 2007, 18, 2867-2871.

221. Prepared in two steps (one pot) from 4-pentynol (85% yield), see: Vatale, J. –M. One-pot selective oxidation/olefination of primary alcohols using TEMPO-BAIB system and stabilized phosphorus ylides. Tetrahedron Lett. 2006, 47, 715-718.

222. Corey, E. J.; Helal, C. J. Reduction of carbonyl compounds with chiral oxazaborolidine catalysts: a new paradigm for enantioselective catalysis and powerful new synthetic method. Angew. Chem. Int. Ed. 1998, 37, 1986-2012.

223. For a similar procedure, see: Kumar, P.; Naidu, V. Enantio- and diastereocontrolled total synthesis of (+)-boronolide. J. Org. Chem. 2006, 71, 3935-3941.

224. Claisen, L., Lowman, O. A new method for the synthesis of benzoyl acetic esters. Ber.

1887, 20, 651-654.

225. Davis, B. R.; Garratt, P. J. Acylation of esters, ketones, and nitriles. In Comprehensive Organic Synthesis; Trost, B. M.; Flemming, I. Eds.; Pergamon: Oxford, 1991; Vol. 2, pp 795-863.

226. Hauser, C. R.; Hudson, D. E. Jr. Acetoacetic ester condensation and certain related reactions. Org. React. 1942, 1, 266-302.

227. Hauser, C. R.; Swamer, F. W.; Adams, J. T. 1,3,5-Triacetylbenzene. Org. React. 1954, 8, 59.

228. For Diekmann condensation see: Schaefer, J. P.; Bloomfield, J. J. Diekmann condensation (Including the Thorpe-Ziegler condensation). Org. React. 1967, 15, 1-203.

229. Bordwell, F. G. Equilibrium acidities in dimethyl sulfoxide solution. Acc. Chem. Res. 1988, 21, 456-463.

230. Zhang, X. –M.; Bordwell, F. G.; Van Der Puy, M.; Fried, H. E. Equilibrium acidities and hemolytic bond dissociation energies of the acidic C—H bonds in N-substituted trimethylammonium and pyridinium cations. J. Org. Chem. 1993, 58, 3060-3066.

Page 239: Florida State University Libraries - FLVC

220

231. Kelly, S. E. Alkene synthesis. In Comprehensive Organic Synthesis; Trost, B. M.; Flemming, I. Eds.; Pergamon: Oxford, 1991; Vol. 1, pp 729-818.

232. Maryanoff, B. E.; Reitz, A. B. The Wittig olefination and modifications involving phosphoryl-stabilized carbanions. Stereochemistry, mechanism, and selected synthetic aspects. Chem. Rev. 1989, 89, 863-927.

233. Vedejs, E.; Peterson, M. J. Stereochemistry and mechanism in the Wittig reaction. Top. Stereochem. 1994, 21, 1-157.

234. Boutagy, J.; Thomas, R. Olefin synthesis with organic phosphonate carbanions. Chem. Rev. 1974, 74, 87-99.

235. Heron, B. M. Heterocycles from intramolecular Wittig, Horner and Wadsworth-Emmons reactions. Heterocycles 1995, 41, 2357-2386.

236. Ernst, H.; Muenster, P. Carotenoid Synthesis. Wittig and Horner-Emmons reactions. Carotenoids 1996, 2, 307-310.

237. Lawrence, N. J. The Wittig reaction and related methods. Preparation of Alkenes 1996, 19-58.

238. Nicolaou, K. C.; Harter, M. W.; Gunzer, J. L.; Nadin, A. The Wittig and related reactions in natural products synthesis. Liebigs Annalen/Recueil 1997, 1283-1301.

239. Motoyoshiya, J. Recent developments in the Z-selective Horner-Wadsworth-Emmons reactions. Trends in Organic Chemistry 1998, 7, 63-73.

240. Minami, T.; Okauchi, T.; Kouni, R. -Phosphonovinyl carbanions in organic synthesis. Synthesis 2001, 349-357.

241. Prunet, J. Recent methods for the synthesis of (E)-alkene units in macrocyclic natural products. Angew. Chem. Int. Ed. 2003, 2, 2826-2830.

242. Still, W. C.; Gennari, C. Direct synthesis of Z-unsaturated esters. Tetrahedron Lett. 1983, 24, 4405-4408.

243. Blanchette, M. A.; Choy, W.; Davis, J. T.; Essenfeld, A. P.; Masamune, S.; Roush, W. R.; Sakai, T. Horner-Wadsworth-Emmons reaction: use of lithium chloride and an amine for base-sensitive compounds. Tetrahedron Lett. 1984, 25, 2183-2186.

244. Iorga, B.; Eymery, F.; Mouries, V.; Savignac, P. Phosphorylated aldehydes: preparation and synthetic uses. Tetrahedron 1998, 54, 14637-14677.

245. Arbuzov, B. A. Michaelis-Arbuzov and Perkow reactions. Pure Appl. Chem. 1964, 9, 307-353.

Page 240: Florida State University Libraries - FLVC

221

246. Bhattacharya, A. K.; Thyagarajan, G. Michaelis-Arbuzov rearrangement. Chem. Rev. 1981, 81, 415-430.

247. Recent Example: Somu, R. V.; Boshoff, J.; Quao, C.; Bennet, E. M.; Barry, C. E. Aldrich, C.C. Rationally designed nucleoside antibiotics that inhibit siderophore biosynthesis of Mycobacterium tuberculosis. J. Med. Chem. 2006, 49, 31-34.

248. Palacios, F.; Ochoa de Retana, A. M.; Alonso, J. M. Regioselective synthesis of fluoroalkylated -aminophosphorus derivatives and aziridines from phosphorylated oximes and nucleophilic reagents. J. Org. Chem. 2006, 71, 6141-6148.

249. Wetermann, J.; Schneider, M.; Platzek, J.; Petrov, O. Practical synthesis of a heterocyclic immunosuppressive vitamin D analog. Org. Process Res. Dev. 2007, 11, 200-205.

250. Maloney, K. M.; Chung, J. Y. A general procedure for the preparation of -ketophosphonates. J. Org. Chem. 2009, 74, 7574-7576.

251. Milburn, R. R.; McRae, K.; Chan, J.; Tedrow, J.; Larsen, R.; Faul, M. A practical synthesis of -ketophosphonates. Tetrahedron Lett. 2009, 50, 870-872.

252. Yamashita, A.; Brown, C. A. Saline hydrides and superbases in organic reactions. IX. Acetylene zipper. Exceptionally facile contrathermodynamic multipositional isomerization of alkynes with potassium 3-aminopropylamide. J. Am. Chem. Soc. 1975, 97, 891-892.

Page 241: Florida State University Libraries - FLVC

222

BIOGRAPHICAL SKETCH

Birth Place

Melrose, Massachusetts

February 3rd, 1981

Educational Background

Florida State University, Tallahassee, FL

August 2004 to December 2009 Ph.D. in Organic Chemistry (anticipated completion in December 2009) Research Advisor: Professor Gregory B. Dudley

Barry University, Miami Shores, FL

August 1999 to December 2003 B.S. degree in Chemistry, B.S. degree in Biology – cum laude Research Advisor: Professor Paul I. Higgs

The Canterbury School, Ft. Myers, FL

August 1995 to June 1999

Future Position

University of Pennsylvania, Philadelphia, PA

Beginning January 2010 Postdoctoral Research Associate Under the supervision of Professor Amos B. Smith, III

Awards and Honors

Gamma Sigma Epsilon, National Chemistry Honors Society (2002). Polymer Chemist Societies Award for Outstanding Performance in Organic

Chemistry (2002). Outstanding Graduating Senior for Performance in Physical Sciences,

Mathematics, and Computer Sciences, School of Arts and Sciences, Barry University (2003.

Golden Key, Graduate Student Honor Society (2007-2009).

Page 242: Florida State University Libraries - FLVC

223

Publications

(2) Jones, D. M.; Dudley, G. B. Synthesis of the C1-C15 region of palmerolide A using a refined Claisen-type addition / bond cleavage methodology. Synlett, in press.

(1) Jones, D. M; Kamijo, S.; Dudley, G. B. Grignard triggered fragmentation of vinylogous acyl triflates: synthesis of (Z)-6-heneicosen-11-one, the Douglas-fir tussock moth. Synlett 2006, 936-938.

Presentations

(2) ―Organic Synthesis and Methodology: Towards the Illudalane Sesquiterpenoids.‖ Jones, D. M.; Dudley, G. B. Presented at the Florida Annual Meeting and Exposition (FAME), Orlando, FL, Summer 2007.

(1) ―Grignard triggered fragmentation of vinylogous acyl triflates: synthesis of (Z)-6-

heneicosen-11-one, the Douglas-fir tussock moth.‖ Jones, D. M.; Kamijo, S.; Dudley, G. B. Presented at the 231st ACS Annual Meeting, Atlanta, GA, March 28th, 2006.

Posters

(2) ―An Addition / Fragmentation Approach to Palmerolide A.‖ Jones, D. M.; Jeong-Im, J.; Dudley, G. B. Presented at the Gordon Research Conference on Natural Products, Tilton, NH, July 26th-31st, 2009.

(1) ―Progress Towards Palmerolide A.‖ Jeong, J.; Jones, D. M.; Dudley, G. B. Presented at

The 236th ACS National Meeting, Philadelphia, PA, August 17th-21st, 2008.