investigations in transition metal catalysis: development ... · palladium catalyzed...

88
Investigations in Transition Metal Catalysis: Development of a Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands by Katherine Jane Jardine A thesis submitted in conformity with the requirements for the degree of Master of Science Graduate Department of Chemistry University of Toronto © Copyright by Katherine Jane Jardine, 2010

Upload: others

Post on 16-Jun-2020

26 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

Investigations in Transition Metal Catalysis: Development of a Palladium Catalyzed

Carboesterification of Olefins and

Synthesis of Chiral Sulfoxide Pincer Ligands

by

Katherine Jane Jardine

A thesis submitted in conformity with the requirements for the degree of Master of Science

Graduate Department of Chemistry University of Toronto

© Copyright by Katherine Jane Jardine, 2010

Page 2: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

ii

Investigations in Transition Metal Catalysis: Development of a

Palladium Catalyzed Carboesterification of Olefins and Synthesis

of Chiral Sulfoxide Pincer Ligands

Katherine Jane Jardine

Master of Science

Graduate Department of Chemistry

University of Toronto

2010

Abstract

The development of a palladium-catalyzed intramolecular carboesterification of unactivated

olefins is described. Olefin difunctionalization is a powerful tool for adding complexity to a

molecule, and this formal [3+2] cycloaddition generates highly functionalized fused ring

systems. Initially discovered by Dr. Yang Li in our group, it was found that when propiolic acids

with a pendant terminal olefin were treated with 1 mol % Pd(MeCN)2Cl2, 3 equivalents of

copper (II) chloride, and 3 equivalents of lithium chloride in acetonitrile at 50 °C, cyclization

occurred in up to 90% yield. The optimization of this reaction and the extension to

propiolamides and propargyl alcohols is described in this thesis. A mechanism involving a novel

palladium-carboxylate species is proposed.

Preliminary investigations into the synthesis of chiral sulfoxide pincer ligands are also described.

The nucleophilic aromatic substitution of 1,3-dibromobenzene and 2,6-dichloropyridine with

various thiols, followed by oxidation of the sulfides to sulfoxides is investigated as a route to the

desired proligands.

Page 3: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

iii

Acknowledgments

First and foremost, I would like to thank Dr. Yang Li, who developed the idea for the palladium-

catalyzed carboesterification project, and worked with me on it. I would also like to thank all the

members of the Dong group for their help, especially Peter Dornan for answering endless

questions and being an awesome cubicle buddy, and Elena Dimitrijevic and Marija Antonic for

always being there to lend a hand.

Numerous people and groups helped to make this work possible: I would like to thank the groups

of Professors Mark Taylor, Andrei Yudin, Rob Batey, and Mark Lautens for the use of chemicals

and equipment. A special thanks to Prof. Rob Batey for reading my thesis. I would also like to

thank NSERC for providing a graduate fellowship to pursue this work.

I owe a great debt of gratitude to Prof. Jamie Donaldson for all of his help, and for believing in

me despite everything.

Finally, I would like to thank my supervisor, Prof. Vy Dong, for all of her help and support.

Page 4: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

iv

Table of Contents

Acknowledgments .......................................................................................................................... iii

List of Figures .............................................................................................................................. viii

List of Tables .................................................................................................................................. x

List of Abbreviations ..................................................................................................................... xi

List of Appendices ....................................................................................................................... xiv

Chapter 1 ......................................................................................................................................... 1

1 Introduction ................................................................................................................................ 1

Chapter 2 ......................................................................................................................................... 2

2 Palladium-Catalyzed Carboesterification of Olefins ................................................................. 2

2.1 Background ......................................................................................................................... 2

2.1.1 Olefin Difunctionalization ...................................................................................... 2

2.1.2 Transition Metal-Catalyzed [3+2] Cycloadditions with Propargyl Alcohols

and Amines ............................................................................................................. 9

2.2 Plan of Study: Development of a Palladium-Catalyzed Carboesterification of Olefins ... 11

2.3 Results and Discussion: Development of a Palladium-Catalyzed Carboesterification

of Olefins .......................................................................................................................... 12

2.3.1 Synthesis of 3-(2-(allyloxy)phenyl)propiolic acid (4) .......................................... 12

2.3.2 Initial Results ........................................................................................................ 13

2.3.3 Synthesis of 3-(2-(but-3-enyloxy)phenyl)propiolic acid (10) ............................... 13

2.3.4 Solvent Screen with 3-(2-(but-3-enyloxy)phenyl)propiolic acid (10) .................. 14

2.3.5 Optimization of Catalyst Loading and Temperature with 3-(2-(but-3-

enyloxy)phenyl)propiolic acid (10) ...................................................................... 16

2.3.6 Optimization of Catalyst Loading and Temperature with 3-(2-

(allyloxy)phenyl)propiolic acid (4) ....................................................................... 17

Page 5: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

v

2.3.7 Optimization of Oxidant Loading and Chloride Source ....................................... 18

2.3.8 Base Screen ........................................................................................................... 20

2.3.9 Optimization of Substrate Concentration .............................................................. 21

2.3.10 Optimized Conditions ........................................................................................... 21

2.3.11 Scope of Carboxylic Acid Substrates ................................................................... 22

2.3.12 Extension to Non-Carboxylic Acid Substrates ..................................................... 23

2.3.13 Retrosynthetic Analysis of Propargyl Alcohol Substrates .................................... 23

2.3.14 Synthesis of Propargyl Alcohol Substrates by the Sonogashira Reaction ............ 23

2.3.15 Synthesis of Propargyl Alcohol Substrate 22 by Reduction ................................. 25

2.3.16 Cyclization of Propargyl Alchohol 22 .................................................................. 25

2.3.17 Synthesis and Reactivity of Tertiary Alcohol Substrate 24 .................................. 26

2.3.18 Synthesis of Propiolamide 25 ............................................................................... 27

2.3.19 Cyclization of Amide Substrate 25 ....................................................................... 27

2.4 Proposed Mechanism for the Palladium-Catalyzed Carboesterification of Olefins ......... 28

2.4.1 Proposed Mechanism ............................................................................................ 28

2.4.2 Mechanistic Experiments ...................................................................................... 29

2.5 Summary and Future Work: Palladium-Catalyzed Carboesterification of Olefins .......... 30

2.5.1 Summary ............................................................................................................... 30

2.5.2 Future Work .......................................................................................................... 30

Chapter 3 ....................................................................................................................................... 32

3 Development of Chiral Sulfoxide Pincer Ligands ................................................................... 32

3.1 Introduction ....................................................................................................................... 32

3.1.1 Common Properties of Pincer Ligands ................................................................. 32

3.1.2 Enantioselective Catalysis with Chiral Pincer Ligands ........................................ 33

3.1.3 Sulfoxide-Based Pincer Ligands ........................................................................... 36

3.2 Plan of Study: Development of Novel Chiral Sulfoxide-Based Pincer Ligands .............. 37

Page 6: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

vi

3.3 Results and Discussion ..................................................................................................... 37

3.3.1 Retrosynthetic Analysis of Phenyl-Based Sulfoxide Pincer Ligands ................... 37

3.3.2 Formation of a Grignard Reagent from 1,3-Dibromobenzene .............................. 38

3.3.3 Formation of a Lithiated Species from 1,3-Dibromobenzene .............................. 39

3.3.4 Nucleophilic Aromatic Substitution of 1,3-Dibromobenzene with

Cyclohexanethiol .................................................................................................. 40

3.3.5 Formation of Various 1,3-Disulfide Compounds from 1,3-Dibromobenzene by

Nucleophilic Aromatic Substitution ..................................................................... 40

3.3.6 Oxidation of Disulfide 29 to the Disulfoxide ....................................................... 41

3.3.7 Pyridine-Based Sulfoxide Pincer Ligands ............................................................ 42

3.3.8 Retrosynthetic Analysis of Pyridine-Based Sulfoxide Pincer Ligands with No

Methylene Spacer .................................................................................................. 42

3.3.9 Nucleophilic Aromatic Substitution of 2,6-Dichloropyridine with Alkyl Thiols . 43

3.3.10 Retrosynthetic Analysis of Pyridine-Based Pincer Ligands with a Methylene

Spacer .................................................................................................................... 43

3.3.11 Initial Progress Towards Compound 37 ............................................................... 44

3.4 Summary and Future Work ............................................................................................... 45

3.4.1 Summary ............................................................................................................... 45

3.4.2 Future Work .......................................................................................................... 45

Chapter 4 ....................................................................................................................................... 46

4 Experimental ............................................................................................................................ 46

4.1 General Considerations ..................................................................................................... 46

4.2 Experimental Section for the Carboesterification of Olefins ............................................ 47

4.2.1 Typical Procedures ................................................................................................ 47

4.2.2 Synthesis and Cyclization of Carboxylic Acid Substrates .................................... 49

4.2.3 Synthesis and Cyclization of Non-Carboxylic Acid Substrates ........................... 51

4.3 Experimental Section for the Preparation of Sulfoxide Pincer Ligands ........................... 54

4.3.1 Typical Procedures ................................................................................................ 54

Page 7: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

vii

4.3.2 Synthesis of Sulfoxide Pincer Ligands ................................................................. 54

References ..................................................................................................................................... 71

Page 8: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

viii

List of Figures

Figure 2.1 Sharpless asymmetric dihydroxylation of olefins. ........................................................ 2

Figure 2.2 Palladium-catalyzed diacetoxylation of olefins developed by Dong and Song. ........... 3

Figure 2.3 Mechanism of the palladium-catalyzed diacetoxylation of olefins. .............................. 3

Figure 2.4 Wolfe's palladium-catalyzed carboetherification of olefins to form substituted

tetrahydrofurans. ............................................................................................................................. 4

Figure 2.5 Mechanism of Wolfe's palladium-catalyzed synthesis of tetrahydrofurans. ................. 4

Figure 2.6 Wolfe's palladium-catalyzed carboamination of olefins to form substituted

pyrrolidines. .................................................................................................................................... 5

Figure 2.7 Sorensen's palladium-catalyzed aminoacetoxylation of olefins. ................................... 5

Figure 2.8 Mechanism of Sorensen's palladium-catalyzed aminoacetoxylation of olefins. ........... 6

Figure 2.9 Stahl's palladium-catalyzed aminoacetoxylation of olefins. ......................................... 6

Figure 2.10 Michael's palladium-catalyzed diamination of olefins. ............................................... 7

Figure 2.11 Palladium-catalyzed diamination of olefins reported by Muniz. ................................ 7

Figure 2.12 Lu's palladium-catalyzed cyclization of allylic alkynoates. ........................................ 8

Figure 2.13 Mechanism of Lu's palladium-catalyzed cyclization of allylic alkynoates. ................ 9

Figure 2.14 Palladium-catalyzed [3+2] cyclization of methylenecyclopropane with norbornene. 9

Figure 2.15 Palladium-catalyzed [3+2] cyclization of propargyl alcohols and amines with

activated olefins. ........................................................................................................................... 10

Figure 2.16 Mechanism of the palladium-catalyzed [3+2] cyclization of propargyl alcohols with

activated olefins. ........................................................................................................................... 11

Figure3.1 Development of the first pincer ligand. ........................................................................ 32

Page 9: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

ix

Figure 3.2 Sites of modification of typical pincer ligands. ........................................................... 33

Figure 3.3 Reactivity of Nishiyama's chiral PheBox rhodium pincer complexes. ....................... 34

Figure 3.4 Reactivity of Venanzi's chiral PCP platinum pincer complex..................................... 35

Figure 3.5 Longmire's chiral PCP platinum pincer complex. ....................................................... 35

Figure 3.6 Chiral at phosphorus pincer complex developed by van Koten. ................................. 36

Figure 3.7 Pincer ligands incorporating sulfoxides as chelating moieties. ................................... 36

Page 10: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

x

List of Tables

Table 2.1 Solvent optimization for the palladium-catalyzed carboesterification of (10). ............ 16

Table 2.2 Optimization of catalyst loading and temperature for the carboesterification of 10. ... 17

Table 2.3 Optimization of catalyst loading and temperature for the carboesterification of 4. ..... 18

Table 2.4 Optimization of the chloride source for the Pd-catalyzed carboesterification of 4. ..... 19

Table 2.5 Optimization of base in the Pd-catalyzed carboesterification of 4. .............................. 20

Table 2.6 Optimization of substrate concentration for the Pd-catalyzed carboesterification of 4. 21

Table 2.7 Conditions for the synthesis of 2-(3-hydroxylprop-1-ynyl)phenol by Sonogashira

coupling. ........................................................................................................................................ 24

Table 2.8 Conditions for the alkylation of 2-(3-hydroxylprop-1-ynyl)phenol with allyl bromide.

....................................................................................................................................................... 24

Table 3.1 Grignard formation from 1,3-dibromobenzene. ........................................................... 39

Table 3.2 Lithiation of 1,3-dibromobenzene. ............................................................................... 39

Table 3.3 Nucleophilic aromatic substitution of 1,3-dibromobenzene with cyclohexanethiol. ... 40

Page 11: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

xi

List of Abbreviations

A Angstrom

Ac acetate

acac acetylacetonato

aq aqueous

Ar aromatic

B base

Bn benzyl

Boc tert-butoxycarbonyl

br broad

Bu butyl

cat catalytic

COE cyclooctene

d days

d doublet

δ chemical shift (in parts per

million)

dba dibenzylideneacetone

DCE dichloroethane

DCM dichloromethane

DIBAL-H diisobutylaluminum hydride

DMA N,N-dimethylacetamide

DMF N,N-dimethylformamide

DMSO dimethyl sulfoxide

dppe diphenylphospinoethane

dppp diphenylphosphinopropane

ee enantiomeric excess

EDG electron donating group

equiv equivalents

ESI electrospray ionization

Et ethyl

EtOAc ethyl acetate

EI electron impact ionization

EWG electron withdrawing group

FT Fourier transform

g grams

GC gas chromatography

h hour

Hz Hertz

Page 12: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

xii

i iso

IR infrared spectroscopy

J coupling constant (in Hertz)

L ligand

LC liquid chromatography

M mega

M metal

m milli

m multiplet

m meta

mCPBA meta-chloroperoxybenzoic

acid

Me methyl

MeCN acetonitrile

min minutes

mol moles

MS mass spectrometry

MS molecular sieves

NMR nuclear magnetic resonance

nuc nucleophile

o ortho

p para

Ph phenyl

Phth phthalimide

ppm parts per million

Pr propyl

q quartet

quant quantitative

R alkyl group

rt room temperature

s singlet

sat saturated

t triplet

t tert

TBS tert-butyldimethylsilyl

temp temperature

TEMPO 2,2,6,6-

Tetramethylpiperidine-1-oxyl

Tf triflate

TFA trifluoroacetic acid

THF tetrahydrofuran

TLC thin layer chromatography

Page 13: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

xiii

TMS trimethylsilyl

tol toluene

Ts toluenesulfonyl

X halogen

SN2 bimolecular nucleophilic

substitution

Ns nitrobenzenesulfonyl

Page 14: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

List of Appendices

Appendix 1. NMR Spectra………………………………………………………………………58

xiv

Page 15: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

1

Chapter 1

1 Introduction

The development of new transition metal catalyzed reactions is an important goal of organic

chemistry, both for the development of new “greener” reactions that avoid producing

stoichiometric amounts of waste, and for achieving reactivity patterns that are not otherwise

possible. Transition metal catalysis has become important in industry, showcasing both its

versatility and efficiency.

This work will investigate two areas of transition metal catalysis: the development of new

catalytic reactions, and the synthesis of chiral ligands for asymmetric transition metal catalysis.

Specifically, the first section of the thesis describes the development of an intramolecular

palladium-catalyzed formal [3+2] cycloaddition resulting in the difunctionalization of

unactivated olefins. Propiolic acids are used as the three-atom subunit to form fused 6,7,5-

tricyclic ring systems. Optimization, substrate scope, and the extension to propiolamides and

propargyl alcohols as the three-atom subunit are described. This project was initiated by Dr.

Yang Li, a postdoctoral fellow in the Dong Group, who discovered the reaction. The second part

of the thesis describes the development of novel sulfoxide-based pincer ligands that introduce

chirality at the sulfur atom. Sulfoxide groups are relatively unexplored as chelating groups in

pincer ligands, and optically pure sulfoxide-based pincer ligands have not been prepared. Initial

investigations into the synthesis of sulfoxide ligands with either phenyl or pyridyl backbones are

described.

Page 16: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

2

Chapter 2

2 Palladium-Catalyzed Carboesterification of Olefins

2.1 Background

2.1.1 Olefin Difunctionalization

Alkene difunctionalization is a powerful method of adding complexity to a molecule. The olefin

moiety is readily accessible from various reactions, and therefore is a desirable substrate for

further elaboration of complex organic molecules.1 The Sharpless asymmetric dihydroxylation

2

is perhaps the most well known example of olefin difunctionalization – the formation of a diol

from an alkene (Figure 2.1). However, the Sharpless dihydroxylation uses a toxic and expensive

osmium catalyst to effect the transformation. Recently, a number of olefin difunctionalizations

catalyzed by palladium have been achieved which broaden the scope of alkene functionalization

and have the advantage of using less toxic and less expensive reagents.

Figure 2.1 Sharpless asymmetric dihydroxylation of olefins.

2.1.1.1 Olefin Diacetoxylation

In 2008, the Dong and Song groups published a palladium-catalyzed dioxygenation of olefins

(Figure 2.2).3 This methodology utilizes an inexpensive and relatively non-toxic palladium salt

as the catalyst and hypervalent iodine as an oxidant to generate syn-diacetoxylated products.

Moreover, 1,1- and 1,2-disubstituted as well as trisubstituted olefins could be used in this

reaction. Trisubstituted olefins are challenging substrates for the Sharpless dihydroxylation.4

Promising initial investigations have shown that chiral ligands can be used to promote some

degree of enantioselectivity.5 This reaction is therefore shows promise as an alternative to the

Page 17: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

3

Sharpless dihydroxylation. Jiang and coworkers later showed that oxygen could be used as the

oxidant to perform similar chemistry.6

Figure 2.2 Palladium-catalyzed diacetoxylation of olefins developed by Dong and Song.

The reaction is thought to proceed via a novel Pd(II)/(IV) pathway (Figure 2.3). Trans-

acetoxypalladation of the olefin generates a palladium(II) alkyl species. Subsequent oxidation to

palladium(IV) occurs in the presence of hypervalent iodine. SN2-type reductive elimination to

regenerate the palladium(II) catalyst occurs via formation of an acetoxonium ion which reacts

with water to form a hydroxyacetate. Treatment with acetic anhydride generates the

diacetoxylated product.

Figure 2.3 Mechanism of the palladium-catalyzed diacetoxylation of olefins.

2.1.1.2 Carboamination and Carboetherification of Olefins

Wolfe has developed a palladium-catalyzed synthesis of tetrahydrofurans via the

carboetherification of alkenes (Figure 2.4).7 Using a palladium catalyst and tri-ortho-

Page 18: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

4

tolylphosphine as a ligand, in the presence of an aryl bromide and sodium tert-butoxide, in

toluene at 110 °C, tetrahydrofurans are formed with a moderate to excellent diastereoselectivity.

Terminal and 1,2-disubstituted olefins can be used. Electron rich or electron neutral aryl

bromides work well with this methodology; electron poor aryl bromides give moderate yields.

Figure 2.4 Wolfe's palladium-catalyzed carboetherification of olefins to form substituted tetrahydrofurans.

The reaction is proposed to proceed via initial oxidative addition of the aryl bromide to give a

Pd(II) complex (Figure 2.5).8 In the presence of sodium tert-butoxide, substitution of the

bromide by the alcohol gives a palladium alkoxide intermediate. Syn-oxypalladation of the

alkene and subsequent C–C bond forming reductive elimination generate the tetrahydrofuran

product, resulting in overall formation of a new C–C bond and a new C–O bond at the expense

of the olefin π-bond. Syn-oxypalladation, involving alkene insertion into the palladium-oxygen

bond, is uncommon, but through deuterium labeling studies Wolfe was able to provide

compelling evidence for this reaction pathway.9

Figure 2.5 Mechanism of Wolfe's palladium-catalyzed synthesis of tetrahydrofurans.

Page 19: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

5

This methodology can also be extended to the synthesis of pyrrolidines via carboamination of

olefins. Electron-rich aryl bromides give the best results. Competing N-arylation is observed for

electron-poor aryl bromides.7

Figure 2.6 Wolfe's palladium-catalyzed carboamination of olefins to form substituted pyrrolidines.

2.1.1.3 Aminoacetoxylation of Olefins

In 2005, Sorensen reported an intramolecular aminoacetoxylation of alkenes (Figure 2.7).10

Using palladium(II) acetate, hypervalent iodine, and tetrabutylammonium acetate in acetonitrile,

protected amines were converted into the corresponding nitrogen-containing heterocycles. Both

five- and six-membered rings were formed. For 1,2-disubstituted alkenes, the reaction was

shown to be stereoselective for the trans difunctionalization product.

Figure 2.7 Sorensen's palladium-catalyzed aminoacetoxylation of olefins.

The reaction is proposed to begin with trans-aminopalladation of the alkene, which is believed to

be reversible (Figure 2.8). Irreversible deprotonation of the resulting intermediate generates a

neutral palladium(II) complex. This species is further oxidized to Pd(IV) by hypervalent iodine,

and subsequent C–O bond forming reductive elimination generates the product and reforms the

active palladium(II) catalyst.

Page 20: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

6

Figure 2.8 Mechanism of Sorensen's palladium-catalyzed aminoacetoxylation of olefins.

Subsequently, Stahl showed an intermolecular variation using phthalimide as the nitrogen source

(Figure 2.9).11

The reaction is proposed to proceed through a cis-aminopalladation step and SN2-

type C–O bond forming reductive elimination from a palladium (IV) intermediate.

Figure 2.9 Stahl's palladium-catalyzed aminoacetoxylation of olefins.

2.1.1.4 Diamination of Olefins

Michael and coworkers have recently reported that N-fluorobenzenesulfonimide can be used as a

nitrogen source in the palladium-catalyzed diamination of unactivated alkenes (Figure 2.10).12

The reaction is proposed to proceed via aminopalladation of the alkene, followed by oxidative

addition of N-fluorobenzenesulfonimide to Pd(II) to generate a Pd(IV) complex. Reductive

elimination then gives the desired diamination product and regenerates the palladium(II) catalyst.

The reaction generates two differently protected amines, allowing for selective deprotection and

functionalization.

Page 21: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

7

Figure 2.10 Michael's palladium-catalyzed diamination of olefins.

Muniz reported an intramolecular diamination of olefins in 2005 (Figure 2.11).13

The reaction

uses a palladium(II) catalyst, and hypervalent iodine as the oxidant. Various fused nitrogen-

containing ring systems were formed, though the methodology worked best for the formation of

pyrrolidine rings. Piperidine formation required 25 mol % Pd(OAc)2, and formation of 7-

membered rings required 10 mol % of the catalyst. Monosubstituted and 1,1-disubstituted

olefins were effective substrates for this methodology. Deprotection of the diamine core could

be achieved in high yields by treating the products with lithium aluminum hydride, followed by

HCl.

Figure 2.11 Palladium-catalyzed diamination of olefins reported by Muniz.

2.1.1.5 Olefin Difunctionalization Initiated by Halopalladation of Propargyl Esters

Lu has developed a cyclization of allylic alkynoates that is initiated by halopalladation of a

propargyl ester (Figure 2.12).14

Treatment with 5 mol % of a palladium(II) catalyst, with

copper(II) chloride and lithium chloride in acetonitrile at ambient temperature for 72 hours gave

a mixture of two regioisomers. The major product results from initial trans-chloropalladation of

the alkyne, whereas the minor product arises from cis-chloropalladation. The selectivity for the

trans-chloropalladation product was found to increase with increased concentration of chloride

ions. A 78:22 ratio of trans- to cis-chloropalladation products was seen when 2 equivalents of

lithium chloride was used. When the chloride concentration was increased to 6 equivalents of

Page 22: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

8

lithium chloride, the ratio increased to 95:5 in favour of the trans-product.14

The solvent polarity

also affects the selectivity, with polar solvents favouring trans-halopalladation, and non-polar

solvents giving poor selectivity for either product.15

The bromo-analogues of these products can

also be formed by using a palladium(II) catalyst, copper(II) bromide, and lithium bromide.14

Figure 2.12 Lu's palladium-catalyzed cyclization of allylic alkynoates.

The reaction is proposed to proceed via initial halopalladation16

of the alkyne, with the trans-

halopalladation pathway predominating (Figure 2.13). The palladation occurs such that

palladium becomes bonded to the α-carbon of the conjugated ester. This regioselectivity is

determined by the polarity of the triple bond due to the electron-withdrawing ester moiety.15

Nucleophilic attack by the chloride anion at the most electrophilic site of the alkyne therefore

results in formation of solely the 5-membered ring products. The resulting palladium(II)

intermediate then undergoes carbopalladation to give a lactone ring with an exocyclic double

bond. The reductive elimination to form the desired product and regenerate the palladium(II)

catalyst is mediated by copper(II) chloride.14

Page 23: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

9

Figure 2.13 Mechanism of Lu's palladium-catalyzed cyclization of allylic alkynoates.

2.1.2 Transition Metal-Catalyzed [3+2] Cycloadditions with Propargyl Alcohols and Amines

Cycloaddition reactions are important methods of generating cyclic architectures from acyclic

precursors. Transition metal catalyzed [3+2] cycloadditions have been extensively investigated

as efficient methods of synthesizing highly functionalized 5-membered rings.17,18

Cycloadditions that involve olefins as the 2-atom subunit, and therefore result in an overall olefin

difunctionalization generally require highly activated Michael acceptors or strained alkenes. For

example, the formation of carbocycles can be achieved by reacting methylenecyclopropanes or

trimethylenemethanes with strained or activated alkenes (such as norbornene), using a

palladium(0) catalyst (Figure 2.14).

Figure 2.14 Palladium-catalyzed [3+2] cyclization of methylenecyclopropane with norbornene.

Page 24: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

10

Alternatively, heterocycles can be synthesized through incorporation of the heteroatom into the

3-atom subunit of the [3+2] cycloaddition.19

Balme and coworkers have demonstrated that

propargyl alcohols20

and amines21

can react with activated alkenes to form substituted

tetrahydrofurans and pyrrolidines, respectively (Figure 2.15). The reaction is catalyzed by base

and a palladium catalyst, in order to activate both the heteroatom and the alkyne, and proceeds

via a Michael addition-carbocyclization process. Treatment of an activated alkene and propargyl

alcohol or N-methyl propargyl amine with 10 mol % n-butyllithium and 5 mol %

[Pd(OAc)2(PPh3)] in tetrahydrofuran at room temperature results in formation of

tetrahydrofurans and pyrrolidines, respectively (Figure 14).

Figure 2.15 Palladium-catalyzed [3+2] cyclization of propargyl alcohols and amines with activated olefins.

The mechanism is proposed to proceed via deprotonation of the propargyl species and

subsequent Michael addition to give a stabilized enolate (Figure 2.16). Activation of the alkyne

by a palladium hydride species formed by the insertion of palladium into an alkyne C–H bond

results in carbopalladation of the alkyne. Reductive elimination reforms the palladium catalyst

and results in formation of a substituted methylenetetrahydrofuran.20

Page 25: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

11

Figure 2.16 Mechanism of the palladium-catalyzed [3+2] cyclization of propargyl alcohols with activated olefins.

2.2 Plan of Study: Development of a Palladium-Catalyzed Carboesterification of Olefins

Recently in our lab, Dr. Yang Li discovered that propiolic acids could be added across an

unactivated alkene in a formal [3+2] cycloaddition (Figure 2.17). Proparyl alcohols and amines

have been investigated in [3+2] cycloaddition chemistry,17

but propiolic acids have not

previously been used. The addition across an unactivated olefin is also of note: most palladium-

catalyzed cycloaddition reactions use the high reactivity of Michael acceptors or strained

trimethylenecyclopropanes.18,19

Figure 2.17 Formal [3+2] cycloaddition of propiolic acids with unactivated olefins.

The goal of this project is to determine the optimal conditions for the formal [3+2] cycloaddition

of propiolic acids across unactivated olefins. Furthermore, propiolic acid derivatives such as

propiolamides will be investigated to determine if this reactivity can be extended to other

Page 26: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

12

functional groups. This transformation creates fused polycyclic ring systems that could be of use

in medicinal chemistry or natural product synthesis.

2.3 Results and Discussion: Development of a Palladium-Catalyzed Carboesterification of Olefins

2.3.1 Synthesis of 3-(2-(allyloxy)phenyl)propiolic acid (4)

Initial studies began with propiolic acid 4, which was derived from salicylaldehyde in four steps

(Figure 2.18). This substrate was chosen because successful cyclization would generate a 6,7,5-

fused tricyclic framework that maps onto the core of a family of natural products with anti-HIV

activity.22

Figure 2.18 Synthesis of 3-(2-(allyloxy)phenyl)propiolic acid (4).

Treatment of salicylaldehyde with potassium carbonate and allyl bromide in DMF at ambient

temperature gave 1 in 95% yield after 2 days. Following purification by flash column

chromatography the aldehyde was transformed into the propiolic ester via the Corey-Fuchs

alkyne synthesis.23

Carbon tetrabromide and triphenylphosphine reacted at 0 °C under argon

atmosphere to form the phosphorus ylide, which added to the aldehyde to give vinyl dibromide 2

after 20 hours at ambient temperature. Chromatography was necessary to remove the

triphenylphosphine oxide which forms over the course of the reaction, and the vinyl dibromide

was isolated in 90% yield. Treatment with methyllithium at -78 °C for 1 hour, followed by the

addition of methyl chloroformate gave the propiolic ester 3 in 90% yield after 3 hours.

Page 27: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

13

Hydrolysis with 20% aqueous KOH in methanol gave the desired propiolic acid substrate 4 in

92% yield.

2.3.2 Initial Results

Initial investigation of the reaction was performed by Dr. Yang Li using conditions developed by

Lu for his chloropalladation of propargylic esters.14

Lu has shown that polar solvents and high

chloride concentrations favour trans-chloropalladation. Therefore, a vial was charged with 4,

copper(II) chloride, and lithium chloride. Acetic acid was added, then a stock solution of

Pd(MeCN)2Cl2 in acetic acid. The solution was heated to 50 °C and monitored by TLC and

LCMS. After 14 hours, the reaction mixture was extracted into ethyl acetate and washed with

water. It was found that with 0.5 mol% Pd(MeCN)2Cl2, three equivalents of copper(II) chloride,

and six equivalents of lithium chloride in acetic acid at 50 °C, substrate 4 cyclised give fused

tricyclic compound 5 in 50% yield by 1H NMR (using 1,3,5-trimethoxybenzene as an internal

standard). Upon increasing the amount of lithium chloride to twelve equivalents, the yield

increased to 63% (Figure 2.19).

Figure 2.19 Initial results for the palladium-catalyzed intramolecular carboesterification of (4).

While these initial results were quite promising, especially considering the low catalyst loading,

ideally the conditions could be improved by reducing the amount of copper(II) chloride and

lithium chloride required, and by moving to a solvent with greater functional group

compatibility. Therefore, the catalyst loading, oxidant, chloride source, solvent, and temperature

need to be optimized to improve the yield.

2.3.3 Synthesis of 3-(2-(but-3-enyloxy)phenyl)propiolic acid (10)

Substrate 10 was made in five steps from 3-buten-1-ol and salicylaldehyde (Figure 2.20). Its

synthesis is analogous to the synthesis of substrate 4.

Page 28: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

14

Figure 2.20 Synthesis of 3-(2-(but-3-enyloxy)phenyl)propiolic acid (10).

3-Buten-1-ol was converted to the tosylate by treatment with p-toluenesulfonyl chloride, with

pyridine functioning as both a base and the solvent. The reaction was allowed to proceed for 4

hours, warming from 0 °C to ambient temperature, and tosylate 6 was isolated in 72 % yield after

column chromatography to remove excess p-toluenesulfonyl chloride. Alkylation of

salicylaldehyde with 6 proceeded in 95 % yield after 4 days at ambient temperature in DMF,

using potassium carbonate as the base. Vinyl dibromide 8 was achieved in 90 % yield using

carbon tetrabromide and triphenylphosphine. Treatment with methyllithium in THF at -78 °C

for 1 hour, followed by the addition of methyl chloroformate afforded methyl ester 9 in 94%

yield. Hydrolysis of the ester with potassium hydroxide gave propiolic acid 10 in 80% yield

after recrystallization from diethyl ether and pentane.

2.3.4 Solvent Screen with 3-(2-(but-3-enyloxy)phenyl)propiolic acid (10)

Optimization of the solvent was initially performed using substrate 10, which cyclised to give a

mixture of two products, 11 and 12, as determined by 1H NMR spectroscopy by Dr. Yang Li

(Figure 2.21). These products arise from coordination of the olefin to palladium with different

regioselectivity, which is possible due to increased flexibility of the homoallyl group compared

with the allyl group of substrate 4. Three equivalents of copper(II) chloride and twelve

equivalents of lithium chloride were used in all cases, and the reaction was heated to 80 °C.

Page 29: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

15

Progress was monitored by TLC and LCMS, and the reactions were stopped when the starting

material was consumed or after 15 hours.

Figure 2.21 Palladium-catalyzed carboesterification of (10).

In methanol, a 13% conversion to 11 was observed (Table 2.1). In ethanol, no cyclised product

was observed, but isopropanol gave 19% conversion to 11. Therefore polar, protic solvents

show some success, and favour the formation of product 11. In acetone no product was

observed, and no starting material could be recovered. Ethyl acetate gave a 39% overall yield of

11 and 12, as did a mixture of acetic acid and acetonitrile. Acetonitrile alone gave a 45% overall

yield, with approximately a 1:1 mixture of 11 to 12. Strongly coordinating solvents such as

DMSO and DMF gave no conversion, but showed improved starting material stability.

Therefore the catalyst loading was increased to 10 mol % in DMF to see if higher catalyst

loading could improve the reactivity of the substrate in this solvent. However, complete starting

material decomposition was observed, and no desired product was formed. Chlorinated solvents

such as dichloromethane and dichloroethane gave no desired product, even at 10 mol % catalyst

loading. In tetrahydrofuran, 10 % of product 11 was observed. Acetonitrile was therefore

determined to be the best solvent for the reaction in terms of overall conversion.

Page 30: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

16

Table 2.1 Solvent optimization for the palladium-catalyzed carboesterification of (10).

2.3.5 Optimization of Catalyst Loading and Temperature with 3-(2-(but-3-enyloxy)phenyl)propiolic acid (10)

The effect of catalyst loading and temperature on the reaction was investigated with substrate 10

(Table 2.2). The reactions were performed in acetonitrile, with three equivalents of copper(II)

chloride and 12 equivalents of lithium chloride. When 0.5 mol % Pd(MeCN)2Cl2 was used, no

cyclised product was observed at 50 °C (entry 1). By NMR, 25% decomposition of the starting

material was observed under these conditions. When the temperature was increased to 80 °C,

50% of the starting material decomposed, but still no product was observed (entry 2). At 110 °C

(entry 3) no product formed and there was no starting material remaining after 14 hours.

Therefore this substrate appears to be somewhat unstable at higher temperatures. When the

catalyst loading was increased to 10 mol %, a 45% overall yield of cyclised product was

observed after 14 hours at 80 °C. Lowering the catalyst loading to 2 mol % had no effect on the

yield or product ratio. With 5 mol % catalyst loading, the temperature could also be lowered to

50 °C with no decrease in the yield.

Page 31: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

17

Table 2.2 Optimization of catalyst loading and temperature for the carboesterification of 10.

2.3.6 Optimization of Catalyst Loading and Temperature with 3-(2-(allyloxy)phenyl)propiolic acid (4)

Further optimization of the catalyst loading was performed using substrate 4 (Table 2.3). The

screen was performed using three equivalents of copper(II) chloride and 12 equivalents of

lithium chloride in MeCN. At 80 °C with 0.5 mol % Pd(MeCN)2Cl2, the cyclised product was

observed in 70% conversion by 1H NMR (entry 1) with 1,3,5-trimethoxybenzene as an internal

standard. Increasing the catalyst loading to 2 mol % increased the conversion to 75% (entry 2).

By lowering the temperature to 50 °C, 80% conversion was observed, and a 76% isolated yield

(entry 3). It was found that the catalyst loading could be lowered to 1 mol % without affecting

the conversion (entry 4). In accordance with the results found from optimization with substrate

10, increasing the catalyst loading does not improve the yield significantly. Higher temperatures

can also lead to increased decomposition and therefore lower yields.

Page 32: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

18

Table 2.3 Optimization of catalyst loading and temperature for the carboesterification of 4.

2.3.7 Optimization of Oxidant Loading and Chloride Source

Next, the effect of the oxidant loading and chloride source was investigated (Table 2.4). The

oxidant, copper(II) chloride, is needed to regenerate the palladium(II) catalyst and turn over the

catalytic cycle. The chloride source and concentration of chloride ions is also important, as high

concentrations of chloride ions in solution have been found to facilitate trans-chloropalladation

of propargyl esters,14

and therefore may help the trans-chloropalladation of propiolic acids as

well.

Page 33: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

19

Table 2.4 Optimization of the chloride source for the Pd-catalyzed carboesterification of 4.

In acetic acid, 12 equivalents of lithium chloride was optimal and gave a 63% yield of the

desired product (Table 2.4), and lowering the chloride concentration resulted in poor yields. In

acetonitrile, when 12 equivalents of lithium chloride were added, the conversion increased to

81%. However, lithium chloride is less soluble in acetonitrile than in acetic acid. It was

observed that not all of the lithium chloride was soluble in acetonitrile at this loading. When the

amount of lithium chloride was lowered to three equivalents, the salt was fully soluble and 83%

conversion was observed. With three equivalents of copper(II) chloride and no lithium chloride,

only 48% conversion was observed. Therefore a source of chloride ions in solution is important

for achieving high yield of the desired product. Tetrabutylammonium chloride was also

investigated as a chloride source, as it is more soluble in acetonitrile than lithium chloride and

therefore might be a better chloride source. With three equivalents of copper(II) chloride and

one equivalent of tetrabutylammonium chloride, 76% conversion was observed. Therefore,

tetrabutylammonium chloride can successfully be used as a chloride source in this reaction.

However, lithium chloride is less expensive and gave higher conversion.

Decreasing the copper(II) chloride loading was also attempted. With two equivalents of

copper(II) chloride and three equivalents of lithium chloride, the conversion dropped slightly to

Page 34: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

20

77% (entry 5). Using tetrabutylammonium chloride in combination with lithium chloride was

also attempted, but the conversion dropped to 71% (entry 6). Reducing the copper(II) chloride

loading further to 1.5 equivalents led to decreased yields, even with increased catalyst loading

(entries 7-9).

2.3.8 Base Screen

Over the course of the reaction, HCl is produced as a byproduct of the cyclization. The addition

of base might therefore improve the yield of the reaction by neutralizing the acid, or by initial

deprotonation of the carboxylic acid moiety prior to cyclization (Table 2.5).

Table 2.5 Optimization of base in the Pd-catalyzed carboesterification of 4.

Using 2 mol % Pd(MeCN)2Cl2, three equivalents of copper(II) chloride, and 12 equivalents of

lithium chloride in acetonitrile, addition of one equivalent of triethylamine decreased the

conversion from 80% to 32% (entry 2). One equivalent of potassium carbonate gave a 51%

conversion (entry 3). Using a different set of conditions for the copper(II) chloride and lithium

chloride loading (1.5 equivalents of CuCl2, 1 equivalent of LiCl, and 0.2 equivalents of

nBu4NCl), it was found that sodium acetate, sodium bicarbonate, potassium carbonate, and

cesium carbonate all decreased the conversion significantly. Addition of 1 equivalent of

Page 35: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

21

potassium hydrogen phosphate gave a comparable yield, but was not a significant improvement.

Therefore it was determined that addition of base did not improve the yield of the reaction.

2.3.9 Optimization of Substrate Concentration

A final experiment was carried out to determine the optimal substrate concentration at which to

perform the reaction. A concentration range of 0.025 M to 0.1 M in acetonitrile was investigated

(Table 2.6).

Table 2.6 Optimization of substrate concentration for the Pd-catalyzed carboesterification of 4.

Using 2 mol % Pd(MeCN)2Cl2, three equivalents of copper(II) chloride and twelve equivalents

of lithium chloride, a substrate concentration of 0.025 M gave 74% yield (entry 1). In

comparison, increasing the concentration to 0.05 M in substrate increased the yield to 80% (entry

2). The same trend was seen when tetrabutylammonium chloride was used as the chloride source

instead of lithium chloride, with a concentration of 0.025 M giving 64% conversion, while 0.05

M gave 76% conversion. However, further increasing the concentration to 0.1 M did not

increase the yield, resulting in 68% conversion. Therefore 0.05 M was determined to be the

optimal concentration of substrate in MeCN for the reaction.

2.3.10 Optimized Conditions

The highest yield of cyclised product was obtained using 1 mol % Pd(MeCN)2Cl2, 3 equivalents

of copper(II) chloride, and 3 equivalents of lithium chloride, in acetonitrile at 0.05 M in

substrate. An 83% conversion to 5 was obtained when 3-(2-(allyloxy)phenyl)propiolic acid (4)

Page 36: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

22

was subjected to these conditions (0.2 mmol scale). When performed on 1 mmol scale with

respect to substrate, the product was isolated in 82% yield as an off-white solid.

Figure 2.22 Optimized conditions for the palladium-catalyzed carboesterification of olefins.

2.3.11 Scope of Carboxylic Acid Substrates

Dr. Yang Li performed a scope study with the optimized conditions, to show that a variety of

propiolic acid derivatives could be used in this reaction. The results are shown in Figure 2.23.

All reactions were performed on a 0.2 mmol scale.

Figure 2.23 Scope of propiolic acid substrates performed by Dr. Yang Li.

Page 37: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

23

2.3.12 Extension to Non-Carboxylic Acid Substrates

The addition of the propiolic acid moiety across an olefin results in the formation of a lactone

ring. The extension of this methodology to substrates other than propiolic acids would allow for

the formation of different types of heterocycles. For instance, propargyl alcohols would generate

fused tetrahydrofurans, while amides could give rise to lactams. Therefore, formation of

substrates of these types is of interest in extending the scope of this methodology.

2.3.13 Retrosynthetic Analysis of Propargyl Alcohol Substrates

Substrate 22 could be made via a Sonogashira reaction24,25

from 2-bromophenol (Figure 2.24).

The allyl group could be attached by alkylation either before or after the Sonogashira coupling.

Selective allylation of the phenol group in the presence of a primary alcohol should be possible

since the phenolic proton is more acidic and therefore easier to deprotonate. However, palladium

catalysts are known to remove allyl groups, so performing the Sonogashira reaction prior to

allylation may be a better route.

Figure 2.24 Retrosynthetic analysis of propargyl alcohol 22.

2.3.14 Synthesis of Propargyl Alcohol Substrates by the Sonogashira Reaction

Initially, 2-bromophenol was treated with Pd(PPh3)2Cl2, copper(I) iodide and triethylamine in

THF at ambient temperature, with tetrahydropyran-protected propargyl alcohol as a coupling

partner (Table 2.7). The Sonogashira cross coupling was unsuccessful, so the more reactive 2-

iodophenol was used instead. However, this resulted only in some decomposition of the starting

materials. Next, the cross coupling was attempted with unprotected propargyl alcohol.26

Using

2-bromophenol, no product was observed under Sonogashira conditions at either ambient

Page 38: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

24

temperature or 80 °C. However, with 2-iodophenol, the desired product was isolated in 70%

yield after 16 hours at ambient temperature.

Table 2.7 Conditions for the synthesis of 2-(3-hydroxylprop-1-ynyl)phenol by Sonogashira coupling.

Alkylation of the phenol group with allyl bromide would give the desired substrate 22.

However, a number of alkylation methods were attempted without success (Table 2.8).

Table 2.8 Conditions for the alkylation of 2-(3-hydroxylprop-1-ynyl)phenol with allyl bromide.

Treatment with allyl bromide and potassium carbonate in DMF at ambient temperature resulted

in no reaction. Increasing the temperature to 50 °C had no effect, and using acetone as a solvent

was also unsuccessful. In a refluxing mixture of ethanol and water, the reaction likewise failed.

The addition of sodium iodide to create a better electrophile in situ was not effective. Lastly, a

stronger base was used for the deprotonation. With sodium hydride in DMF, however, no

desired product was observed.

Page 39: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

25

2.3.15 Synthesis of Propargyl Alcohol Substrate 22 by Reduction

Another method to synthesize substrate 22 could be via reduction of methyl ester 3, which is an

intermediate in the synthesis of propiolic acid 4 (Figure 2.25). Originally, it was thought that

selective reduction of the ester moiety in the presence of the alkyne could be challenging.

However, treatment of 3 with 2 equivalents of DIBAL-H resulted in a 60% yield of the desired

propargyl alcohol after 9.5 hours at -78 °C.27

The reaction did not go to completion despite the

addition of a second aliquot of DIBAL-H after 4 hours, so the product was isolated from the

starting material by chromatography. The yield based on recovered starting material was 97%.

Figure 2.25 DIBAL-H reduction of methyl 3-(2-(allyloxy)phenyl)propiolate.

2.3.16 Cyclization of Propargyl Alchohol 22

Initial cyclization experiments were performed using the conditions optimized for the carboxylic

acid substrates. Consequently, substrate 22 was treated with 1 mol % Pd(MeCN)2Cl2, three

equivalents of copper(II) chloride and three equivalents of lithium chloride in acetonitrile, and

the reaction mixture was heated to 50 °C for 13 hours. The reaction mixture was concentrated,

taken up into chloroform, and the insoluble material was filtered off with a plug of cotton.

Purification by column chromatography on silica in 10 % ethyl acetate in hexanes gave a yellow

oil which was isolated in 45 %.

Figure 2.26 Palladium-catalyzed cyclization of 3-(2-(allyloxy)phenyl)prop-2-yn-1-ol (22).

Page 40: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

26

Further optimization of the reaction conditions was unsuccessful. Increasing the temperature to

80 °C resulted in decomposition, as did increasing the catalyst loading to 5 mol %. When one

equivalent of either potassium carbonate or potassium hydrogen phosphate was added, the

reaction did not go to completion. Addition of 10 mol % 12-crown-4 as a chloride phase transfer

agent did not improve the yield.

The exact structure of the product remains unclear. By analogy to the carboxylic acid substrates,

product 23a would be expected. However, the regioselectivity of chloropalladation of the alkyne

is determined by the polarity of the alkyne,15

with the chloride anion attacking the most

electrophilic site. In the case of the propiolic acid substrates, the presence of the electron

withdrawing carboxyl group directs the chloropalladation such that nucleophilic attack of the

chloride ion occurs at the β-carbon of the propiolic acid. With 22, there is no strongly electron-

withdrawing substituent, and the regioselectivity of the chloropalladation is more difficult to

determine. Chloropalladation with the opposite regioselectivity would generate product 23b.

Therefore, further investigation is needed to determine which of the two products, 23a or 23b, is

formed in the course of the reaction.

2.3.17 Synthesis and Reactivity of Tertiary Alcohol Substrate 24

Figure 2.27 Synthesis and cyclization of 4-(2-(allyloxy)phenyl-2-methylbut-3-yn-2-ol (24).

Tertiary alcohol substrate 24 was also synthesized. Treatment of methyl ester 3 with three

equivalents of methyllithium in THF at -78 °C gave substrate 24 in 58% yield after 4.5 hours.

This substrate was synthesized in the hope that the cyclised product would be less prone to

decomposition than primary alcohol 22. However, using the conditions optimized for the

cyclization of the propiolic acids, no reaction was observed and only starting material was

recovered.

Page 41: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

27

2.3.18 Synthesis of Propiolamide 25

Figure 2.28 Synthesis of 3-(2-(allyloxy)phenyl)-N-phenylpropiolamide (25).

Substrate 25 was made from vinyl dibromide 2. 2 was treated with methyllithium in THF at -78

°C for one hour, followed by the addition of phenyl isocyanate to afford 25 in 34% yield after

one hour (Figure 2.28).

2.3.19 Cyclization of Amide Substrate 25

Cyclization of substrate 25 was first attempted using the standard conditions of 1 mol %

palladium catalyst, 3 equivalents of copper(II) chloride and 3 equivalents of lithium chloride in

acetonitrile (Figure 2.29). The product was isolated in 48 % yield after 20 hours. Furthermore,

when 0.5 equivalents of potassium hydrogen phosphate was added to the reaction mixture, a 67

% conversion was observed by 1H NMR, using 1,3,5-trimethoxybenzene as an internal standard.

Figure 2.29 Palladium-catalyzed cyclization of 3-(2-(allyloxy)phenyl)-N-phenylpropiolamide (26a).

Some lactone was also observed and isolated by preparatory TLC following the cyclization of

25. Therefore, a secondary reaction pathway could be cyclization through the amide oxygen,

resulting in 26b, which could hydrolyze either under the reaction conditions or on silica to give

lactone 5 (Figure 2.30). Further experiments are necessary to determine how predominant this

pathway is compared to cyclization via nitrogen, and whether the major product isolated from

the reaction is 26a, or in fact 26b.

Page 42: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

28

Figure 2.30 Cyclization of 3-(2-(allyloxy)phenyl)-N-phenylpropiolamide (25) through oxygen.

2.4 Proposed Mechanism for the Palladium-Catalyzed Carboesterification of Olefins

2.4.1 Proposed Mechanism

Figure 2.31 Proposed mechanism of the palladium-catalyzed carboesterification of olefins.

We propose that the reaction is initiated by trans-chloropalladation of the alkyne with the

palladium(II) catalyst (Figure 2.31). Both trans- and cis-chloropalladation of alkynes is well

precedented;16

however, propiolic acids are unexplored substrates for halopalladation. A high

concentration of choride anions in solution has been shown to favour trans-chloropalladation

over cis-chloropalladation for propargylic esters.14

The regioselectivity of the trans-

chloropalladation is controlled by the polarity of the alkyne induced by the electron-withdrawing

carboxylic acid group.15

Therefore, we propose that trans-chloropalladation would give a

Page 43: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

29

palladium-carboxylate species (I), capable of coordinating to the unactivated alkene. Following

this, two pathways can be imagined. In Path A, oxypalladation28

of the alkene would give

palladacycle IIA. Subsequent C–C bond forming reductive elimination would generate the

product. Alternatively, carbopalladation of the alkene would generate palladacycle IIB.

Reductive elimination could generate the product; however, direct C–O bond forming reductive

elimination from a palladium(II) species is not precedented. SN2-type reductive elimination via

formation of chloride intermediate III could also give the product. However, III was never

observed as a side product of the reaction for any substrate.

2.4.2 Mechanistic Experiments

In order to further investigate the mechanism, as well as the scope of the reaction, substrate 27

was synthesized by Dr. Yang Li, and subjected to the reaction conditions (Figure 2.32). Both

Path A and Path B would be expected to give trans-28 as the only product, so the experiment

cannot distinguish between the two pathways. However, formation of trans-28 would provide

support for one of these mechanisms, whereas formation of the cis isomer could indicate that

both of the proposed mechanisms are incorrect.

Figure 2.32 Palladium-catalyzed carboesterification of a 1,2-disubstituted olefin.

In fact, when subjected to the reaction conditions by Dr. Yang Li, a 3:1 mixture of trans-28 and

cis-28 was isolated in 69% overall yield. Two possibilities are evident from this result. First,

multiple pathways might be operating, and a minor pathway gives the cis-isomer. Alternatively,

substrate 27 could isomerize to the Z-isomer under the reaction conditions,29

resulting in a

product mixture, or the product itself could isomerize.

Therefore, in order to test for isomerization of the product, trans-28 was resubjected to the

reaction conditions. No isomerization was observed. To ensure that the presence of HCl formed

over the course of the reaction did not facilitate isomerization, trans-28 was also subjected to the

Page 44: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

30

reaction conditions in the presence of substrate 4. Over the course of the cyclization of propiolic

acid 4, no isomerization of trans-28 was observed. Therefore the products do not isomerize

under the reaction conditions, and any isomerisation must occur before or during the cyclization.

An isomerization experiment was performed in an attempt to determine whether the olefin of the

starting material isomerizes. Substrate 27 was subjected to the reaction conditions, and the

reaction was halted at partial completion. The starting material was reisolated to determine if

any had isomerized to the Z-isomer. No olefin isomerization products of the starting material

could be observed with certainty by 1H NMR or LC-MS. However, it is possible that the starting

material isomerizes only to a very small degree, but the Z-isomer reacts faster, resulting in the

formation of a significant amount of cis-28.30

Therefore, while there is no direct evidence for formation of cis-28 by olefin isomerization, it

cannot be ruled out, and one of the proposed mechanisms may be the sole reaction pathway.

Alternatively, there may be another minor reaction pathway operating which gives cis-28

directly from 27 without isomerization.

2.5 Summary and Future Work: Palladium-Catalyzed Carboesterification of Olefins

2.5.1 Summary

A novel formal [3+2] cycloaddition resulting in the carboesterification of unactivated olefins was

investigated. Optimized conditions were found for the cyclization of propiolic acids in which the

substrates were treated with 1 mol % Pd(MeCN)2Cl2, 3 equivalents of copper(II) chloride and 3

equivalents of lithium chloride in acetonitrile at 50 °C. Isolated yields up to 90% were achieved.

The reaction is proposed to proceed via initial chloropalladation of the alkyne, followed by

oxypalladation of the alkene and subsequent C–C bond forming reductive elimination to generate

a fused ring system incorporating a 5-membered lactone. Preliminary results for propargyl

alcohols and amides indicate that they are also promising substrate classes for this reaction.

2.5.2 Future Work

Future work should extend the scope of this transformation beyond propiolic acids. The

preliminary results with propargyl alcohols and amides are promising, but further investigation is

Page 45: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

31

necessary to confirm the structures of the products formed from these substrates. The conditions

also need to be further optimized for these substrates, and a scope study performed. Propargyl

amines could also be investigated in this transformation.

More mechanistic experiments are needed to support the proposed mechanism, such as

deuterium labeling studies and kinetics experiments. Calculations could also be used to help

determine if the proposed mechanistic pathway is the most likely. Synthesis and cyclization of

the Z-olefin analogue of substrate 27 would give results that would be interesting to compare to

those obtained for the cyclization of 27.

Finally, this reaction creates one new stereogenic carbon centre in the case of terminal olefins,

and two for 1,2-disubstituted alkenes. Therefore, the potential for asymmetric induction could be

investigated by testing the effect of performing the reaction in the presence of a chiral ligand.

Page 46: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

32

Chapter 3

3 Development of Chiral Sulfoxide Pincer Ligands

3.1 Introduction

A pincer ligand is a tridentate ligand with meridional binding. The first pincer ligand was

synthesized by Moulton and Shaw in 1976 (Figure 3.1).31

Tridentate metal complexes were

synthesized by reacting 1,3-bis(bromomethyl)benzene with di-tert-butylphosphine, followed by

heating the proligand to reflux in ethanol with a metal salt. The ligand was called a PCP pincer

ligand, since it was coordinated to the metal centre through phosphorous, carbon, and another

phosphorus atom. New pincer ligands have continued to be developed because of the unusual

reactivity and stability that this ligand class has shown.

Figure3.1 Development of the first pincer ligand.

3.1.1 Common Properties of Pincer Ligands

Since the development of the first PCP pincer ligand by Shaw, the synthesis of new pincer

ligands has been a growing field.32

The tridentate scaffold imparts rigidity, which stabilizes the

metal-ligand bond, limits the number of active sites, and prevents ligand exchange. 33

The

resultant high stability of the metal-ligand complex allows it to function at high temperatures and

therefore catalyze challenging reactions under harsh conditions. The metal centre in a pincer

complex is often quite electron rich, imparting a nucleophilic character to the metal.33

As well,

in pincer ligands with an aryl backbone, the aryl ring is nearly coplanar with the d8 metal

coordination plane, which allows for an unusual overlap between the filled metal dxz orbital and

the antibonding π* orbital of the arene.33

Importantly, there is high correlation between ligand

Page 47: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

33

modifications and the properties of the metal centre, therefore allowing systematic improvements

in reactivity and stability.33

Figure 3.2 Sites of modification of typical pincer ligands.

Figure 3.2 shows a standard pincer ligand with an aryl backbone.34

The metal centre, M, is

bound to the ligand in three locations, and often the metal will bind at least one counterion.

Common metals include, but are not limited to, palladium, platinum, rhodium, iridium, and

ruthenium. The metal binding affinity of the ligand can be tuned by changing Y or the donor

sites E. Y is most commonly carbon or nitrogen. This has a significant effect on the electronics

of the ligand , as if Y=C, the ligand is monoanionic, but if Y=N, the ligand is neutral. The size

of the cavity is also affected by substituents on the arms A, and the donor sites E. The donor

sites E can affect the reactivity and stability through both electronic and steric factors:

substitutents on E will block approach to the metal, and the hardness or softness of E will affect

the binding affinity for different metals. The metal’s electronic properties will be affected by the

electron donating or electron withdrawing nature of E. Commonly, E is phosphorus (PR2),

nitrogen (NR2), sulphur (SR), or oxygen (OR). The arms of the pincer ligand, A, are commonly

modified by adding substituents to change steric constraints or introduce chirality. Electronic

properties can also be modified by changing the nature of A. Most often, CR2, NR, and O are

used at this position. Finally, R can be changed to effect remote electronic modulations, or can

be used as an anchoring site for solid supports.

3.1.2 Enantioselective Catalysis with Chiral Pincer Ligands

Enantioselective catalysis with chiral pincer ligands is less well developed than the use of

bidentate ligands for asymmetric catalysis. Often, higher enantioselectivities can be achieved

with more traditional bidentate ligands than with the chiral pincer ligands that have currently

Page 48: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

34

been developed. However, promising developments have recently been made, and the continued

investigation of new chiral pincer ligands promises further improvements.

Nishiyama and coworkers recently reported a series of PheBox-type pincer ligands that have

shown both high reactivity and enantioselectivity in a number of different reactions.35

The

ligand is an NCN pincer ligand, and the rhodium complex has chirality imparted by the

bis(oxazoline) groups on the ligand backbone (Figure 3.3). The asymmetric allylation of

aldehydes with methallylstannane gave the methallylated products in 90-99% enantioselectivity

and 82-97% yield. The reaction is not air sensitive, and the catalyst can be recovered at the end

of the reaction. The same catalyst system was also applied to the hetero-Diels-Alder reaction.

With only 2 mol % of the catalyst, yields up to 90% and enantioselectivities up to 82% were

observed. These types of catalysts have also been applied to asymmetric conjugate reductions

(with yields up to 99% and enantioselectivities up to 98%), and the Michael addition of α-

cyanocarboxylates and acrolein (with up to 99% yield and up to 86% enantioselectivity).

Figure 3.3 Reactivity of Nishiyama's chiral PheBox rhodium pincer complexes.

The first chiral PCP platinum complex was reported by Venanzi and coworkers in 1994 (Figure

3.4).36

Chirality was introduced into the ligand backbone by adding substituents to the

methylene arms, giving the ligand C2 symmetry. The PCP-platinum complex was used to

catalyze the aldol condensation of aldehydes with methyl isocyanoacetate. Using 1.5 mol % of

catalyst and 12 mol % diisopropylethylamine in dichloromethane at room temperature, yields of

the cyclic products ranged from 84 to 97%. The trans product was favoured in most cases. The

catalyst system achieved up to 65% enantioselectivity for the trans product and up to 32% for the

Page 49: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

35

cis product. While the yields were excellent, enantioinduction was fairly low for this catalyst

system, and a long, challenging ligand synthesis also detracts from its appeal.

Figure 3.4 Reactivity of Venanzi's chiral PCP platinum pincer complex.

Longmire reported the synthesis of a similar PCP platinum complex37

with simple methyl groups

on the pendant arms of the ligand (Figure 3.5). This chiral catalyst was able to perform the same

aldol condensation as Venanzi’s catalyst with comparable yields and enantioselectivities, and

with a significantly simpler synthesis.

Figure 3.5 Longmire's chiral PCP platinum pincer complex.

The first pincer ligand that was chiral at phosphorus was developed by van Koten in 2001

(Figure 3.6).38

The complex was also investigated as a catalyst for the aldol condensation of

aldehydes with methyl isocyanoacetate. While the diastereoselectivity of 98:2 favouring the

trans product was much higher than that achieved by Venanzi’s system, which had a 70:30 ratio

of trans to cis, the enantioselectivity was less than 11% for all cases.

Page 50: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

36

Figure 3.6 Chiral at phosphorus pincer complex developed by van Koten.

3.1.3 Sulfoxide-Based Pincer Ligands

Sulfoxides are versatile moieties to incorporate into ligands. They can bind to metals through

oxygen or sulfur,39

they are configurationally stable and chiral at sulfur, and there are many

literature reports on the synthesis of optically pure sulfoxides.40

Despite the advantages of

incorporating sulfoxide groups into ligands, only a few examples of sulfoxide-based pincer

ligands have been reported. In 1986, Riley and Oliver reported two alkyl sulfoxide pincer

ligands complexed to ruthenium (Figure 3.7, I and II).41,42

In 2002, Evans and coworkers

designed an aryl sulfoxide pincer ligand in which the sulfoxides were coordinated to palladium

through the sulfur (III).43

The most recent example of a pincer ligand incorporating a sulfoxide

moiety was in 2008, when Milstein reported a pyridine-based mixed SNN pincer ligand which

was complexed to rhodium and iridium (IV).44

The catalytic activity of these metal complexes

has not been investigated, and synthesis has been limited to the racemic complexes.

Figure 3.7 Pincer ligands incorporating sulfoxides as chelating moieties.

Page 51: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

37

3.2 Plan of Study: Development of Novel Chiral Sulfoxide-Based Pincer Ligands

Given the limited investigation of sulfoxide-based pincer ligands, we proposed to design new

pincer ligands that use sulfoxides as chelating moieties (Figure 3.8). We envisioned variations to

our ligands in a number of areas. We wanted to first investigate whether we could achieve O-

coordination in an aryl- or pyridine-based pincer complex by changing the length of the ligand

arm. The nature of the metal could also affect the coordination mode. The R group on sulfur

would affect the reactivity of the metal complex both electronically and sterically. Finally, there

have been no attempts to synthesize a chiral sulfoxide-based pincer ligand, despite many reports

on the preparation of optically pure sulfoxides.40

Therefore, the preparation of these complexes

asymmetrically would be of interest, as well as testing their reactivity as chiral catalysts.

Figure 3.8 Target chiral sulfoxide pincer ligands.

3.3 Results and Discussion

3.3.1 Retrosynthetic Analysis of Phenyl-Based Sulfoxide Pincer Ligands

Figure 3.9 Retrosynthetic analysis of phenyl-based sulfoxide pincer ligands.

Page 52: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

38

Retrosynthetically, two potential routes to the desired sulfoxide appeared possible (Figure 3.9).

In route A, treatment of 1,3-dibromobenzene with either n-butyllithium45,46

or magnesium47

to

give the dimetallated species, followed by treatment with a commercially available chiral

sulfinate ester could afford the desired chiral proligand in a single step. Alternatively, a two step

procedure would allow for more extensive variation of the R group on the sulfoxide: nucleophilic

aromatic substitution of 1,3-dibromobenzene with a thiol and an appropriate base could generate

the disulfide compound, and subsequent asymmetric oxidation would give the desired chiral

sulfoxide.

3.3.2 Formation of a Grignard Reagent from 1,3-Dibromobenzene

Initial attempts were directed at the formation of a Grignard reagent from 1,3-dibromobenzene,47

from which treatment with a chiral sulfinate would give the desired proligand (route A). In order

to preserve the chiral reagent, conditions for the formation of the Grignard species were tested

using other electrophiles (Table 3.1). First, 1,3-dibromobenzene was dissolved in anhydrous

THF under an argon atmosphere. Magnesium metal was added, and iodine to initiate formation

of the Grignard reagent. Heating to 70 °C resulted in some disappearance of magnesium. The

Grignard reagent was added to benzaldehyde, but no desired product was observed by 1H NMR

or LC-MS (entry 1). The reaction was repeated at 95 °C, but again no product formation was

observed (entry 2). Using DMF as the electrophile also did not result in formation of the desired

product, as confirmed by 1H NMR, despite disappearance of the magnesium (entry 3). Attempts

to form the Grignard reagent in diethyl ether as the solvent were unsuccessful, as no

disappearance of magnesium was observed, indicating formation of the Grignard reagent was

unsuccessful. As expected, treatment of this mixture with DMF resulted in no desired product

(entry 4).

Page 53: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

39

Table 3.1 Grignard formation from 1,3-dibromobenzene.

3.3.3 Formation of a Lithiated Species from 1,3-Dibromobenzene

Next, formation of a dilithiated species45,46

was attempted (Table 3.2). 1,3-dibromobenzene was

treated with four equivalents of n-butyllithium in THF at -78 °C under an argon atmosphere for

one hour, after which DMF was added. After 30 minutes at -78 °C, less than 10% of the desired

dialdehyde was observed by 1H NMR, though some monosubstituted product was observed, as

well as unidentified byproducts (entry 1). With longer reaction times, no desired product was

observed (entry 2). Attempts to form the monosubstituted product by treatment with 1.1

equivalents of n-butyllithium resulted in no desired product (entry 3).

Table 3.2 Lithiation of 1,3-dibromobenzene.

Page 54: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

40

3.3.4 Nucleophilic Aromatic Substitution of 1,3-Dibromobenzene with Cyclohexanethiol

As formation of the proligand in a single step via route A proved challenging, route B was

investigated. When treated with NaH, cyclohexanethiol underwent nucleophilic aromatic

substitution with 1,3-dibromobenzene48,49,50

to give mainly the monosubstituted product 30 by

GCMS when heated to 50 °C for 14 h (Table 3.3, entry 1). However, increasing the temperature

to 110 °C gave a 3:1 ratio of the desired disubstituted product 29 to monosubstituted 30 after 18

h. Using KOH as the base in DMA at 160 °C for 3 days51

gave a 2.7:1 ratio of 29:30, with a 55

% isolated yield of the desired product 29. However, when the scale was increased from 0.42

mmol to a 2 mmol scale, only 18 % of the desired product was isolated after 5 days at 160 °C

(entry 4). Interestingly, increasing the temperature to 175 °C resulted in a switch in the

selectivity, favouring 30. Despite challenges with this route, the desired product could be

isolated from the monosubstituted product in acceptable yield on small scale, so this method was

pursued for the formation of a variety of disulfide compounds.

Table 3.3 Nucleophilic aromatic substitution of 1,3-dibromobenzene with cyclohexanethiol.

3.3.5 Formation of Various 1,3-Disulfide Compounds from 1,3-Dibromobenzene by Nucleophilic Aromatic Substitution

The nucleophilic aromatic substitution of 1,3-dibromobenzene with various thiols was then

performed. When 1,3-dibromobenzene was treated with n-butylthiol and potassium hydroxide in

dimethylacetamide at 160 °C for 4 days, an approximately 1:1 mixture of the desired product 31

and the monosubstituted product was observed (Figure 3.10). The same result was seen when

Page 55: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

41

thiophenol was used as the nucleophile, resulting in a 1:1 mixture of 32 to monosubstituted

product. For both of these substrates, it is possible that increased reaction time, optimized

reaction temperature, or sequential addition of thiol could increase the yield of the disubstituted

product. In the case of t-butylthiol, only the monosubstituted product 33 was observed by 1H

NMR, in 92 % conversion.

Figure 3.10 Nucleophilic aromatic subsitution of 1,3-dibromobenzene with various thiols.

3.3.6 Oxidation of Disulfide 29 to the Disulfoxide

Figure 3.11 Oxidation of 1,3-bis(cyclohexylthio)benzene with mCPBA.

1,3-bis(cyclohexylthio)benzene (29) was treated with meta-chloroperoxybenzoic acid in

methylene chloride at 0 °C for 15 minutes. After aqueous workup, 34a was isolated in 27%

yield. Analysis of the 13

C NMR spectrum showed a statistical 1:1 mixture of the racemic and

meso diastereomers.

Page 56: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

42

Figure 3.12 Asymmetric oxidation of ,3-bis(cyclohexylthio)benzene.

Asymmetric oxidation of 1,3-bis(cyclohexylthio)benzene was performed using hydrogen

peroxide and a vanadium-Schiff base catalyst system that has been shown to work well for the

oxidation of sulfides to sulfoxides.52,53

The desired product 34b was isolated in 37% yield after

1.5 h at 0 °C in dichloromethane. Analysis of the 13

C NMR spectrum showed a 3:1 mixture of

diastereomers. This indicates that the first oxidation affects the second oxidation in this system,

and a non-statistical mixture of meso and racemic diastereomers was achieved.

3.3.7 Pyridine-Based Sulfoxide Pincer Ligands

The pyridine moiety is a common backbone for many pincer ligands. While phenyl-based pincer

ligands are anionic, pyridine-based pincer ligands are neutral. This has implications for both the

reactivity and stability of the metal-ligand complex. Therefore, investigation of both phenyl-

based and pyridine-based pincer ligands is of interest.

3.3.8 Retrosynthetic Analysis of Pyridine-Based Sulfoxide Pincer Ligands with No Methylene Spacer

Figure 3.13 Retrosynthetic analysis of pyridine-based sulfoxide pincer ligands.

Treatment of 2,6-dichloropyridine with a thiol could result in the formation of the corresponding

disulfide via nucleophilic aromatic substitution. Asymmetric oxidation would generate the

desired proligand, which could be complexed to a metal to give the disulfoxide catalyst.

Oxidation to the sulfoxide could also result in oxidation of the pyridine to the N-oxide; however,

N-oxides can be reduced in the presence of sulfoxides by acetic acid and elemental iron.54

Page 57: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

43

3.3.9 Nucleophilic Aromatic Substitution of 2,6-Dichloropyridine with Alkyl Thiols

Figure 3.14 shows the nucleophilic aromatic substitution of 2,6-dichloropyridine with various

thiols. The reactions were performed using the same conditions as for the phenyl analogues: 2,6-

dichloropyridine was dissolved in DMA, thiol and KOH were added, and the reaction mixture

was heated to 160 °C. However, the reactions proceeded much faster, and were only left for one

day. When cyclohexanethiol was used, the desired product 35 was isolated in quantitative yield.

The reaction was also carried out with n-butylthiol (36) and t-butylthiol (37), resulting in 100%

and 80% yields, respectively. Nucleophilic aromatic substitution at the 2- and 6-position of the

pyridine ring is facile compared with the corresponding reaction performed on a phenyl ring.

This is evident in the high yields and shorter reaction times. Indeed, formation of 1,3-bis(t-

butylthio)benzene was not possible via SNAr, resulting in only the monosubstituted product

when attempted. However, the corresponding pyridine analogue was isolated in 80% yield.

Figure 3.14 Nucleophilic aromatic substitution of 2,6-dichloropyridine with various thiols.

3.3.10 Retrosynthetic Analysis of Pyridine-Based Pincer Ligands with a Methylene Spacer

Figure 3.15 Retrosynthetic analysis of pyridine-based sulfoxide pincer ligands with a methylene spacer.

Page 58: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

44

Figure 3.15 shows the retrosynthetic analysis of a pyridine-based proligand that incorporates a

methylene group into the arms of the pincer. The compound could be synthesized via lithiation

of 2,6-lutidine and subsequent treatment with a commercially available chiral sulfinate ester.54

This route has the benefit of introducing the chiral moiety from a commercially available chiral

source, therefore avoiding the potential for poor enantioselectivity in the chiral oxidation step.

3.3.11 Initial Progress Towards Compound 37

Figure 3.16 Initial progress towards pyridine-based sulfoxide pincer ligands with a methylene spacer.

Proligand 40 (Figure 3.16) has been previously investigated as an organocatalyst for the

asymmetric allylation of N-benzoylhydrazones.54

However, it has not previously been

coordinated to a metal and used as a ligand for transition metal catalysis. The reported synthesis

begins with the oxidation of the pyridine nitrogen to the N-oxide with meta-chloroperoxybenzoic

acid. This reaction was performed, and the desired N-oxide 38 was generated in quantative

yield. The 1H NMR spectrum was consistent with the literature data. From here, treatment of

the N-oxide with n-butyllithium forms the dilithiated species, which gives the chiral disulfoxide

39 upon addition of the sulfinate ester. The N-oxide can be reduced to the corresponding

pyridine in the presence of the sulfoxide groups using iron and acetic acid to generate the desired

proligand 40.

Page 59: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

45

3.4 Summary and Future Work

3.4.1 Summary

Initial steps towards the synthesis of novel chiral sulfoxide pincer ligands were performed. A

series of phenyl- and pyridyl-based bis(sulfides) were successfully synthesized by SNAr, using

1,3-dibromobenzene and 2,6-dichloropyridine, respectively. Both racemic and asymmetric

oxidation of 1,3-bis(cyclohexylthio)benzene (29) was performed. The synthesis of a pyridyl-

based sulfoxide ligand that incorporates methylene groups into the arms of the pincer ligand was

also investigated and the first step was performed.

3.4.2 Future Work

Oxidation of the synthesized phenyl- and pyridyl-based bis(sulfides), both racemically and

asymmetrically remains to be completed, as well as the synthesis of the pyridyl-based sulfoxide

with methylene groups in the arms of the pincer. From there, these compounds need to be

complexed to different metals. Following characterization of the metal complexes, it would be

interesting to test their catalytic activity and enantioselectivity in various reactions.

Page 60: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

46

Chapter 4

4 Experimental

4.1 General Considerations

Commercial reagents were purchased from Sigma Aldrich, Strem or Alfa Aesar and used without

further purification unless otherwise noted. All reactions were carried out under an atmosphere

of air unless otherwise indicated. Reactions were carried out at ambient temperature unless

otherwise noted. Reactions were monitored using thin-layer chromatography (TLC) on EMD

Silica Gel 60 F254 plates. Visualization of the developed plates was performed under UV light

(254 nm) or KMnO4 stain. Organic solutions were concentrated under reduced pressure on a

Büchi rotary evaporator. Column chromatography was performed with Silicycle Silia-P Flash

Silica Gel. All salts were purchased from Aldrich and used without purification. Solvents were

purchased from Caledon. Dichloromethane and tetrahydrofuran were sparged with argon and

passed through two alumina columns to remove water.

1H and

13C NMR spectra were recorded using a Varian Mercury 300, Varian Mercury 400, or a

VRX-S (Unity) 400 spectrometer. NMR spectra were obtained in CDCl3 and internally

referenced to the residual solvent signal. Data for 1H NMR are reported as follows: chemical

shift (δ ppm), multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, br =

broad), integration, coupling constant (Hz). Data for 13

C NMR are reported in terms of chemical

shift (δ ppm). High resolution mass spectra (HRMS) were obtained on a micromass 70S-250

spectrometer (EI) or an ABI/Sciex Qstar Mass Spectrometer (ESI). Low resolution mass spectra

(LRMS) were obtained on a Waters 2795 LC with a Waters Micromass ZQ. Infrared (IR) spectra

were obtained on a Perkin-Elmer Spectrum 1000 FT-IR Systems spectrometer and are reported

in terms of frequency of absorption (cm-1

). Melting point ranges were determined on a Fisher-

Johns Melting Point Apparatus.

Page 61: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

47

4.2 Experimental Section for the Carboesterification of Olefins

4.2.1 Typical Procedures

Method A – A typical procedure for the allylation of phenols: To a stirred solution of the phenol

derivative in N,N-dimethylformamide (0.7 M) was added allyl bromide (1.5 equiv) and K2CO3

(1.5 equiv). The resulting mixture was stirred for 24 hours or until complete consumption of the

starting material was achieved as determined by TLC. The crude reaction mixture was then

diluted with ethyl acetate and successively washed with aqueous 20% KOH (2x10 mL), water

(3x10 mL) and brine (1x10 mL). The organic layer was dried over Na2SO4, filtered, and

concentrated to dryness in vacuo. The product was isolated by flash column chromatography in

ethyl acetate and hexanes.

Method B – A typical procedure for the dibromovinylation of an aldehyde: To a flame dried

round-bottom flask was added triphenylphosphine (3 equiv) and anhydrous dichloromethane (0.3

M). The resulting solution was maintained under argon and cooled to 0 °C. Carbon tetrabromide

(1.5 equiv) was added and the orange solution was stirred at 0 °C for 15 min. A solution of the

aldehyde in dichloromethane (2 M) was added. The resulting mixture was warmed to ambient

temperature and stirred until TLC showed complete consumption of the starting material.

Hexanes (30 mL per 10 mmol aldehyde) was added and the dark brown slurry was stirred for 15

min, then filtered. The filtrate was concentrated in vacuo. The product was isolated by flash

column chromatography in ethyl acetate and hexanes.55

Page 62: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

48

Method C – A general procedure for the formation of methyl propiolates: A solution of the

dibromide in anhydrous tetrahydrofuran (0.1 M) in a flame dried round-bottom flask under argon

was cooled to –78 °C. Methyllithium (3 equiv, 1.6M in hexanes) was added and the resulting

solution was stirred at –78 °C for 1 hour. Methyl chloroformate (6 equiv) was added quickly and

the solution was maintained at –78 °C for 1 hour or until TLC showed complete consumption of

the starting material. Water (10 mL per 10 mmol scale) was added and the mixture was stirred at

room temperature for 10 minutes, then concentrated in vacuo to remove the tetrahydrofuran. The

residue was taken up in diethyl ether and the organic layer was washed with water (3x10mL) and

brine (1x10mL), dried over Na2SO4, filtered, and concentrated in vacuo. The product was

isolated by flash column chromatography in ethyl acetate and hexanes.55

Method D – General procedure for the hydrolysis of methyl propiolates: To a solution of the

methyl propiolate derivative in methanol (0.2 M) was added aqueous 20% KOH (20 equiv), and

the resultant solution was stirred until TLC showed complete consumption of the starting

material. The reaction mixture was cooled to 0 °C, and concentrated H2SO4 was added slowly

until the pH value was less than 1. H2O (30 mL) was added and this mixture was extracted with

EtOAc (3x50 mL). The combined organic layers were then washed with water (15 mL) and brine

(15 mL), dried over Na2SO4, filtered, and concentrated in vacuo. The product was isolated via

recrystallization from diethyl ether and hexanes.

Page 63: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

49

Method E – General procedure for the cyclization of propiolic acids with Pd(MeCN)2Cl2: To a

solution of the propiolic acid derivative (0.2 mmol), LiCl (26 mg, 0.6 mmol, 3 equiv) and CuCl2

(80 mg, 0.6 mmol, 3 equiv) in acetonitrile (3.6 mL) was added a stock solution of Pd(MeCN)2Cl2

(0.52 mg in 0.4 mL MeCN, 0.002 mmol). The mixture was heated at 50 oC for 14 to 20 hours.

The resulting solution was concentrated directly in vacuo. The product was isolated by flash

column chromatography in diethyl ether and hexanes.

4.2.2 Synthesis and Cyclization of Carboxylic Acid Substrates

2-allyloxybenzaldehyde (1)

The title compound was prepared according to Method A from salicylaldehyde

(2.0 g, 16.4 mmol) to yield the product (2.7 g, 100%) as a pale yellow oil after

flash column chromatography in 10% ethyl acetate in hexanes. This compound is

also commercially available. 1H NMR (300 MHz) δ 4.66 (d, J = 5.1 Hz, 2H), 5.34 (dd, J1 = 10.6

Hz, J2 = 1.3 Hz, 1H), 5.46 (dd, J1 = 17.3 Hz, J2 = 1.4 Hz, 1H), 6.02-6.15 (m, 1H), 6.98 (d, J = 8.5

Hz, 1 H), 7.03 (t, J = 7.6 Hz, 1H), 7.50-7.56 (m, 1H), 7.84 (dd, J1 = 7.7 Hz, J2 = 1.7 Hz, 1H),

10.54 (s, 1H) ppm; 13

C NMR (100 MHz) δ 69.1, 112.8, 118.0, 120.8, 125.0, 128.3, 132.3, 135.8,

160.9, 189.6 ppm; IR (neat) 3078, 2862, 2762, 1683, 1598, 1482, 1456, 1285, 1239, 994, 757

cm-1

; LRMS (ESI+) m/z 163.0 (M+1), 185.1 (M+23).

1-(allyloxy)-2-(2,2-dibromovinyl)benzene (2)

The title compound1 was prepared according to Method B from 2-

allyloxybenzaldehyde (2.7 g, 16.4 mmol) to yield the product (4.7 g, 90%) as a

yellow oil after flash column chromatography in 5% ethyl acetate in hexanes.

Spectral data was consistent with literature data. 1H NMR (400 MHz) δ 4.57 (ddd, J1 = 5.1 Hz, J2

= 1.6 Hz, J3 = 1.6 Hz, 2H), 5.30 (tdd, J1 = 1.4 Hz, J2 = 1.4 Hz, J3 = 10.5 Hz, 1H), 5.41 (tdd, J1 =

Page 64: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

50

1.6 Hz, J2 = 1.6 Hz, J3 = 17.3 Hz, 1H), 6.06 (tdd, J1 = 5.1 Hz, J2 = 10.3 Hz, J3 = 17.2 Hz, 1H),

6.86 (d, J = 8.3 Hz, 1H), 6.97 (ddd, J1 = 0.5 Hz, J2 = 1.0 Hz, J3 = 7.7 Hz, 1H), 7.30 (ddd, J1 = 7.9

Hz, J2 = 7.9 Hz, J3 = 1.7 Hz, 1H), 7.64 (s, 1H), 7.70 (dd, J1 = 1.5 Hz, J2 = 7.7 Hz, 1H) ppm; 13

C

NMR (100 MHz) δ 69.1, 89.7, 111.9, 117.5, 120.3, 124.7, 129.2, 129.8, 132.9, 132.9, 155.5

ppm; IR (neat) 3023, 2867, 1597, 1483, 1449, 1243, 1226, 1108, 749 cm-1

; LRMS (ESI+) m/z

317.0 (M+1), 339.0 (M+23).

methyl 3-(2-(allyloxy)phenyl)propiolate (3)

The title compound1 was prepared according to Method C from 1-(allyloxy)-

2-(2,2-dibromovinyl)benzene (5.1 g, 15.9 mmol) to afford the product (3.1 g,

90%) as a pale yellow oil after flash column chromatography in 5-10% ethyl

acetate in hexanes. Spectral data was consistent with literature data. 1

H NMR (400 MHz) δ 3.81

(s, 3H), 4.61 (ddd, J1 = 4.8 Hz, J2 = 1.7 Hz, J3 = 1.7 Hz, 2H), 5.29 (tdd, J1 = 1.5 Hz, J2 = 1.5 Hz,

J3 = 10.6 Hz, 1H), 5.49 (tdd, J1 = 1.6 Hz, J2 = 1.6 Hz, J3 = 17.3 Hz, 1H), 6.03 (tdd, J1 = 4.8 Hz,

J2 = 10.5 Hz, J3 = 17.3 Hz, 1H), 6.87 (d, J = 8.4 Hz, 1H), 6.92 (ddd, J1 = 7.6 Hz, J2 = 7.6 Hz, J3

= 0.8 Hz, 1H), 7.33-7.38 (m, 1H), 7.50 (dd, J1 = 7.6 Hz, J2 = 1.7 Hz, 1H) ppm; 13

C NMR (100

MHz) δ 52.8, 69.3, 83.7, 84.5, 109.4, 112.5, 117.6, 120.8, 132.3, 132.6, 135.0, 154.8, 160.8 ppm;

IR (neat) 2920, 2851, 2219, 1707, 1490, 1447, 1434, 1300, 1280, 1175, 994, 750 cm-1

; LRMS

(ESI+) m/z 217.1 (M+1).

3-(2-(allyloxy)phenyl)propiolic acid (4)

The title compound was prepared according to Method D from methyl 3-(2-

(allyloxy)phenyl)propiolate (3.0 g, 13.9 mmol) to yield the product (2.6 g,

92%) as an off-white solid after recrystallization from diethyl ether and

pentane. 1H NMR (400 MHz) δ 4.65 (ddd, J1 = 4.8 Hz, J2 = 1.7 Hz, J3 = 1.7 Hz, 2H), 5.33 (tdd,

J1 = 1.5 Hz, J2 = 1.5 Hz, J3 = 10.6 Hz, 1H), 5.52 (tdd, J1 = 1.7 Hz, J2 = 1.7 Hz, J3 = 17.3 Hz, 1H),

6.06 (tdd, J1 = 4.8 Hz, J2 = 10.6 Hz, J3 = 17.3 Hz, 1H), 6.90 (d, J1 = 8.4 Hz, 1H), 6.96 (ddd, J1 =

7.6 Hz, J2 = 7.6 Hz, J3 = 0.9 Hz, 1H), 7.41 (ddd, J1 = 1.7 Hz, J2 = 7.5 Hz, J3 = 8.4 Hz, 1H), 7.55

Page 65: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

51

(dd, J1 = 7.6 Hz, J2 = 1.7 Hz, 1H), 9.04-10.82 (bs, 1H) ppm; 13

C NMR (100 MHz) δ 69.2, 84.0,

86.2, 108.9, 112.4, 117.6, 120.7, 132.3, 132.7, 135.1, 158.3, 160.9 ppm; IR (neat): 2912.6,

2596.9, 2205.1, 1665.9, 1310.4, 1285.2, 1253.6, 1231.1, 1200.5, 989.4, 917.7, 755.8, 746.1 cm-1

;

HRMS (ESI) calc’d for C12H9O3 [M-1]- 201.0557, Found 201.0563; mp 83-85 °C.

(Z)-10-chloro-3a,4-dihydrobenzo[b]furo[3,4-e]oxepin-1(3H)-one (5)

The title compound was prepared according to Method E from 3-(2-

(allyloxy)phenyl)propiolic acid (202.2 mg, 1 mmol) and the reaction mixture

was heated at 50 °C for 14 hours to yield the product (193 mg, 82%) as an off-

white solid after flash column chromatography in diethyl ether and hexanes. 1H NMR (400

MHz) δ 3.32-3.39 (m, 1H), 3.98 (dd, J1 = 9.5 Hz, J2 = 4.4 Hz, 1H), 4.40 (dd, J1 = 9.4 Hz, J2 = 8.7

Hz, 1H), 4.48 (dd, J1 = 11.0 Hz, J2 = 10.4 Hz, 1H), 4.68 (dd, J1= 10.3 Hz, J2 = 6.2 Hz, 1H), 7.12

(dd, J1 = 8.1 Hz, J2 = 1.2 Hz, 1H), 7.22-7.26 (m, 1H), 7.39-7.43 (m, 1H), 7.83 (dd, J1 = 8.0 Hz,

J2 = 1.7 Hz, 1H) ppm; 13

C NMR (100 MHz) δ 40.5, 65.1, 81.0, 122.7, 124.1, 124.5, 129.8, 130.6,

132.2, 138.8, 156.1, 166.7 ppm; IR (neat): 2913, 1749, 1617, 1478, 1229, 1087, 1027, 763 cm-1

;

HRMS (ESI) calc’d for C12H10ClO3 [M+H]+ 237.0312, Found 237.0321; mp 62-66 °C.

4.2.3 Synthesis and Cyclization of Non-Carboxylic Acid Substrates

3-(2-(allyloxy)phenyl)prop-2-yn-1-ol (22)

A solution of methyl 3-(2-(allyloxy)phenyl)propiolate (1.08 g, 5.0 mmol) in

anhydrous tetrahydrofuran (50 mL) in a flame dried round bottom flask under

argon was cooled to -78 °C. Diisobutylaluminum hydride (1.0 M in THF, 10

mL, 10 mmol) was added slowly and the reaction was stirred for 6 h. A second aliquot of

DIBAL-H (1.0 M in THF, 3 mL, 3 mmol) was added and the reaction was stirred a further 3 h.

The mixture was quenched with aqueous saturated ammonium chloride (100 mL). The solution

was extracted with ethyl acetate (3 x 100 mL). The organic layer was washed with 1 M HCl ( 1

Page 66: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

52

x 100 mL) to break up the aluminum salts, aqueous saturated sodium carbonate (1 x 100 mL) to

neutralize the solution, then water (2 x 100 mL) and brine (1 x 100 mL). The organic layer was

dried over sodium sulfate, filtered, and concentrated in vacuo to afford the title compound (566

mg, 60 %) as a pale yellow oil after column chromatography in 20 % ethyl acetate in hexanes.

1H NMR (400 MHz) δ 7.41 (dd, J = 1.7 Hz, J = 7.6 Hz, 1H), 7.29-7.24 (m, 1H), 6.90 (dt, J = 0.9

Hz, J = 7.5 Hz, 1H), 6.87 (d, J = 8.4 Hz, 1H), 6.12-6.02 (m, 1H), 5.47 (ddd, J = 1.7 Hz, J = 3.3

Hz, J = 17.3 Hz, 1H), 5.30 (qd, J = 1.5 Hz, J = 10.6 Hz, 1H), 4.62 (td, J = 1.6 Hz, J = 5.0 Hz,

1H), 4.54 (s, 1H) ppm.

23

The product was prepared according to Method E from 3-(2-(allyloxy)phenyl)prop-2-yn-1-ol

(37.6 mg, 0.2 mmol) and the reaction mixture was heated at 50 °C for 23 h to yield the product

(20 mg, 45 %) as a yellow oil after flash column chromatography in 10 % ethyl acetate in

hexanes.

1H NMR (400 MHz) δ 7.76 (dd, J = 1.6 Hz, J = 8.0 Hz, 1H), 7.24 (m, 1H), 7.11 (ddd, J = 1.4 Hz,

J = 7.4 Hz, J = 8.0 Hz, 1H), 7.04 (dd, J = 1.3 Hz, J = 8.0 Hz, 1H), 4.72 (dd, J = 2.2 Hz, J = 15.0

Hz, 1H), 4.61 (dd, J = 1.2 Hz, J = 15.0 Hz, 1H), 4.55 (dd, J = 5.1 Hz, J = 10.7 Hz, 1H), 4.14 (m,

1H), 3.63 (dd, J = 7.5 Hz, J = 8.7 Hz, 1H), 3.15 (m, 1H); 13

C NMR (100 MHz) δ 155.9, 142.4,

129.4, 129.2, 126.8, 122.9, 121.6, 120.1, 74.4, 72.7, 70.8, 45.3 ppm.

4-(2-(allyloxy)phenyl)-2-methylbut-3-yn-2-ol (24)

A solution of methyl 3-(2-(allyloxy)phenyl)propiolate (50 mg, 0.23 mmol) was

dissolved in 1 mL anhydrous tetrahydrofuran in a flame dried flask under

argon was cooled to -78 °C. Methyllithium (1.6 M in THF, 0.43 mL, 0.69

mmol) was added slowly and the reaction mixture was stirred for 4.5 h. Water (2 mL) was added

and the reaction was stirred for 10 minutes, then extracted with diethyl ether (3 x 5 mL), dried

over sodium sulfate, filtered, and concentrated to afford the product (28.9 mg, 58 %) as a pale

yellow oil after purification by preparatory TLC using 20 % ethyl acetate in hexanes.

Page 67: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

53

1H NMR (300 MHz) 7.38 (dd, J = 1.7 Hz, J = 7.6 Hz, 1H), 7.25 (ddd, J = 1.7 Hz, J = 7.5 Hz, J =

8.3 Hz, 1H), 6.89 (m, 2H), 6.07 (tdd, J = 4.8 Hz, J = 10.5 Hz, J = 17.2 Hz, 1H), 5.52 (qd, J = 1.7

Hz, J = 17.2 Hz, 1H), 5.29 (qd, J = 1.6 Hz, J = 10.6 Hz, 1H), 4.59 (td, J = 1.7 Hz, J = 4.7 Hz,

2H), 2.07 (s, 1H), 1.64 (s, 6H) ppm.

3-(2-(allyloxy)phenyl)-N-phenylpropiolamide (25)

A solution of 1-(allyloxy)-2-(2,2-dibromovinyl)benzene (500 mg, 1.6 mmol)

in anhydrous tetrahydrofuran (25 mL) in a flame dried round bottom flask

under argon was cooled to -78 °C. Methyllithium (1.6 M in Et2O, 3 mL, 4.7

mmol) was added dropwise. The yellow-orange solution was stirred for 1 h at -78 °C, then

phenyl isocyanate (0.24 mL, 3.1 mmol) was added. The reaction was stirred for a further 30

minutes, then quenched with water (10 mL) and concentrated to remove most of the

tetrahydrofuran. The organic material was extracted with diethyl ether (3 x 25 mL). The organic

layer was washed with water (2 x 25 mL) and brine (1 x 25 mL), dried over sodium sulfate,

filtered, and concentrated in vacuo to yield the product (148 mg, 34 %) as a pale yellow solid

after flash column chromatography in 20 % ethyl acetate in hexanes followed by recrystallization

in dichloromethane/ethyl acetate/pentane.

1H NMR (400 MHz) δ 7.60 (s, 1H), 7.54 (m, 3H), 7.37 (m, 3H), 7.14 (t, J = 7.4 Hz, 1H), 6.95 (t,

J = 7.5 Hz, 1H), 6.91 (d, J = 8.4 Hz, 1H), 6.09 (m, 1H), 5.51 (dd, J = 1.1 Hz, J = 17.2 Hz, 1H),

5.34 (dd, J = 0.8 Hz, J = 10.6 Hz, 1H), 4.65 (d, J = 4.9 Hz, 2H) ppm; 13

C NMR (75 MHz) δ

160.2, 137.5, 134.5, 132.7, 131.8, 129.1, 124.7, 120.8, 119.8, 117.7, 112.4, 109.7, 87.5, 82.6,

69.3 ppm.

26

The product was synthesized according to Method E from 3-(2-

(allyloxy)phenyl)-N-phenylpropiolamide (13.9 mg, 0.05 mmol) and the

reaction mixture was heated at 50 °C for 20 h to yield the product (7.6 mg, 48

%) as a pale yellow solid after purification by preparatory TLC in 20% ethyl acetate in hexanes.

Page 68: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

54

1H NMR (300 MHz) δ 7.93 (dd, J = 1.6 Hz, J = 8.1 Hz, 1H), 7.61-7.58 (m, 2H), 7.41-7.31 (m,

3H), 7.21-7.15 (m, 2H), 7.10 (dd, J = 1.2 Hz, J = 8.0 Hz, 1H), 4.72 (dd, J = 5.4 Hz, J = 11.8 Hz,

1H), 4.21 (dd, J = 3.4 Hz, J = 11.7 Hz, 1H), 3.88-3.82 (m, 1H), 3.71 (dd, J = 8.4 Hz, J = 11.1 Hz,

1H), 3.45-3.37 (m, 1H) ppm; LRMS (ESI+) m/z 312 (M+1), 350 (M+39).

4.3 Experimental Section for the Preparation of Sulfoxide Pincer Ligands

4.3.1 Typical Procedures

Method F – General procedure for the SNAr reaction: 1,3-dibromobenzene (100 mg, 0.42

mmol) or 2,6-dichloropyridine (100 mg, 0.68 mmol) was dissolved in DMA (1 mL or 1.6 mL).

Potassium hydroxide (2.2 equivalents) was added, then thiol (2 equivalents). The vessel was

sealed and the reaction mixture was heated to 160 °C for 3-6 days. The solution was allowed to

cool, then diluted with diethyl ether, washed with water (2 × 3 mL), brine (1 × 3 mL), dried over

MgSO4, and concentrated. The resulting oil was purified by preparatory TLC in hexanes.

4.3.2 Synthesis of Sulfoxide Pincer Ligands

1,3-bis(cyclohexylthio)benzene (29)

The title compound was prepared according to Method F from 1,3-

dibromobenzene (100 mg, 0.42 mmol) and cyclohexanethiol (98.8 mg, 0.85

mmol) to yield the product (72 mg, 55%) as a pale yellow oil after 3 days. 1H NMR (400 MHz) δ

1.23-1.41 (m, 10 H), 1.61-1.63 (m, 2H), 1.75-1.77 (m, 4H), 1.97-1.99 (m, 4H), 3.08-3.14 (m,

2H), 7.16-7.23 (m, 3H), 7.40 (t, J = 1.5 Hz, 1H) ppm; 13

C NMR (100 MHz) δ 25.7, 26.0, 46.5,

128.915, 129.7, 134.1, 135.9 ppm; IR (neat) 2926.60, 2851.47, 1568.40, 1460.72, 1447.59,

1262.09, 997.07, 780.38 cm-1

; HRMS (EI+) calc’d for C18H26S2 306.1490, found 306.1476.

1,3-bis(butylthio)benzene (31)

Page 69: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

55

The title compound was prepared according to Method F from 1,3-

dibromobenzene (100 mg, 0.42 mmol) and 1-butanethiol (76.5 mg, 0.85

mmol) to yield the product as yellow oil after 3 days. 1H NMR (400 MHz)

0.85 (t, J = 7.3 Hz, 6H), 1.37 (dddd, J1 = 7.3 Hz, J2 = 14.4 Hz, 4H), 1.56 (ddd, J1 = 7.3 Hz, J2 =

15.0 Hz, 4 H), 2.84 (t, J = 7.4 Hz, 4H), 7.00-7.03 (m, 2H), 7.08-7.12 (m, 1H), 7.17 (t, J = 1.7 Hz,

1H) ppm.

1,3-bis(phenylthio)benzene (32)

The title compound was prepared according to method F from 1,3-

dibromobenzene (100 mg, 0.42 mmol) and thiophenol (93.4 mg, 0.85 mmol) to

afford the title compound after 3 days. Spectral data was consistent with literature data.51

(3-bromophenyl)(tert-butyl)sulfane (33)

The title compound was prepared according to Method F from 1,3-

dibromobenzene (100 mg, 0.42 mmol) and 2-methyl-2-propanethiol (76.5 mg,

0.85 mmol) to yield the product (92 % conversion) as a pale yellow oil. 1H NMR (400 MHz) δ

1.29 (s, 9H), 7.20 (dd, J1 = 7.8 Hz, J2 = 7.8 Hz, 1H), 7.49 (ddd, J1 = 1.1 Hz, J2 = 2.0 Hz, J3 = 8.0

Hz, 1H), 7.46 (ddd, J1 = 1.1 Hz, J2 = 1.5 Hz, J3 = 7.7 Hz, 1H), 7.70 (dd, J1 = 1.8 Hz, J2 = 1.8 Hz,

1H) ppm.

2,6-bis(cyclohexylthio)pyridine (35)

The title compound was prepared according to Method F from 2,6-

dichloropyridine (100 mg, 0.68 mmol) and cyclohexanethiol (157 mg, 1.35

mmol) to yield the product (209 mg, 100%) as a pale yellow oil after flash column

chromatography (20% ethyl acetate in hexanes) to remove trace DMA. 1H NMR (400 MHz) δ

7.22 (t, J = 7.8 Hz, 1H), 6.80 (d, J = 7.8 Hz, 2H), 3.90-3.84 (m, 2H), 2.11-2.09 (m, 4H), 1.81-

1.24 (m, 16H).

2,6-bis(butylthio)pyridine (36)

The title compound was prepared according to Method F from 2,6-

dichloropyridine (100 mg, 0.68 mmol) and 1-butanethiol (122 mg,

Page 70: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

56

1.35 mmol) to yield the product (173 mg, 100%) as a pale yellow oil after flash column

chromatography (20% ethyl acetate in hexanes). 1H NMR (400 MHz) δ 7.28 (d, J = 8.1 Hz, 1H),

6.87 (d, J = 7.8 Hz, 2H), 3.20 (t, J = 7.3 Hz, 4H), 1.77-1.69 (m, 4H), 1.54-1.45 (m, 4H), 0.98 (t, J

= 7.3Hz, 6H) ppm.

2,6-bis(tert-butylthio)pyridine (37)

The title compound was prepared according to Method F from 2,6-

dichloropyridine (100 mg, 0.68 mmol) and 2-methyl-2-propanethiol (122

mg, 1.35 mmol) to yield the product (139 mg, 80%) after flash column chromatography. 1H

NMR (400 MHz) δ 7.35 (t, J = 7.7 Hz, 1H), 7.13 (d, J = 7.7 Hz, 2H), 1.52 (s, 18H) ppm.

2,6-bis(cyclohexylsulfinyl)pyridine (34)

Racemic oxidation:

To a solution of 1,3-bis(cyclohexylthio)benzene (67 mg, 0.22 mmol) in dichloromethane (2 mL)

was added mCPBA (70%, 75 mg, 0.44 mmol). The reaction mixture was stirred for 15 minutes,

then quenched with water (2 mL). The organic layer was extracted with dichloromethane (3 x 2

mL), dried over magnesium sulfate, filtered, and concentrated to give the title compound (20 mg,

27%) as an off-white solid after purification by preparatory TLC (20% ethyl acetate in hexanes)

as a mixture of isomers. 1H NMR (400 MHz) δ 7.78-7.76 (m, 1H), 7.74-7.64 (m, 3H), 2.63-2.57

(m, 2H), 1.91-1.83 (m, 6H), 1.71-1.64 (m, 4H), 1.45-1.36 (m, 4H), 1.28-1.46 (m, 6H) ppm; 13

C

NMR (100 MHz) δ 143.6, 129.6, 127.2, 121.5, 121.5, 63.3, 63.2, 26.4, 26.4, 25.3, 25.3, 25.3,

25.2, 23.7, 23.5 ppm; LCMS (ESI+) m/z 361 (M + 23).

Asymmetric oxidation:

To a solution of L* (see Figure 3.12, 21.3 mg, 0.045 mmol) in dichloromethane (3.5 mL) was

added VO(acac)2 (4.0 mg, 0.015 mmol). The solution changed from yellow to blue-green, and

was stirred at ambient temperature for 45 minutes. 1,3-bis(cyclohexylthio)benzene (115 mg,

0.375 mmol) was added and the reaction mixture was stirred for 1 hour at ambient temperature,

then cooled to 0 °C. Hydrogen peroxide (30%, 0.115 mL, 1.125 mmol) was added and the

reaction was stirred for 1.5 hours. Aqueous saturated sodium thiosulfate (2 mL) was added to

Page 71: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

57

quench the reaction, and the organic layer was extracted with dichloromethane (3 x 5 mL),

washed with brine (5 mL), dried over magnesium sulfate, filtered, and concentrated to yield the

product (37 mg, 11%) after purification by preparatory TLC (20% ethyl acetate in hexanes). 1H

NMR (400 MHz) δ 7.78-7.76 (m, 1H), 7.74-7.64 (m, 3H), 2.63-2.57 (m, 2H), 1.91-1.83 (m, 6H),

1.71-1.64 (m, 4H), 1.45-1.36 (m, 4H), 1.28-1.46 (m, 6H) ppm; 13

C NMR (100 MHz) δ 143.6,

129.6, 129.4, 127.2, 121.4, 121.4, 63.2, 63.2, 26.4, 26.3, 25.3, 25.3, 25.2, 25.2, 23.6, 23.5 ppm.

Page 72: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

58

Appendix 1

NMR Spectra

3-(2-(allyloxy)phenyl)prop-2-yn-1-ol (22)

ppm (t1)0.01.02.03.04.05.06.07.08.0

0

500

1000

1500

2000

1.0

0

1.0

21

.03

1.9

3

0.9

11

.20

1.9

61

.84

Page 73: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

59

Compound 23

ppm (t1)0.01.02.03.04.05.06.07.08.0

-100

0

100

200

300

400

500

600

700

800

1.0

0

1.2

21

.07

1.0

3

1.1

21

.13

1.1

4

2.3

5

1.1

8

1.1

2

ppm (t1)50100150

0

100

200

300

400

Page 74: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

60

4-(2-(allyloxy)phenyl)-2-methylbut-3-yn-2-ol (24)

ppm (t1)0.01.02.03.04.05.06.07.08.0

0

5000

1.0

0

1.0

0

1.0

0

1.9

9

1.3

20

.93

2.0

0

0.9

3

6.3

6

Page 75: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

61

3-(2-(allyloxy)phenyl)-N-phenylpropiolamide (25)

ppm (t1)0.01.02.03.04.05.06.07.08.0

0

500

1000

1.0

21

.03

0.9

5

2.9

8

1.0

0

2.0

5

3.8

2

0.9

40

.95

ppm (t1)50100150

-1000

0

1000

2000

3000

4000

5000

6000

16

0.2

13

13

7.4

58

13

4.5

42

13

2.6

57

13

1.7

68

12

9.0

77

12

4.7

49

12

0.8

23

11

9.8

15

11

7.7

39

11

2.4

01

10

9.7

17

87

.46

6

82

.62

8

69

.31

5

Page 76: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

62

Compound 26

ppm (t1)0.01.02.03.04.05.06.07.08.0

0

500

1000

1500

2000

1.0

0

1.0

6

1.0

2

1.0

0

1.2

7

0.9

2

3.1

1

2.0

00

.98

2.3

2

Page 77: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

63

1,3-bis(cyclohexylthio)benzene (29)

ppm (t1)0.01.02.03.04.05.06.07.08.0

0

100

200

300

400

500

2.0

0

10

.21

4.0

8

4.0

0

2.3

2

2.8

3

0.8

8

ppm (t1)50100150

0

5000

10000

15000

20000

25000

13

5.8

60

13

4.1

40

12

9.6

66

12

8.9

05

46

.47

0

33

.28

3

26

.00

9

25

.72

7

Page 78: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

64

1,3-bis(butylthio)benzene (31)

ppm (t1)0.01.02.03.04.05.06.07.08.0

0

100000000

200000000

300000000

400000000

4.0

0

4.2

6

4.1

6

6.1

8

0.9

41

.03

1.8

3

Page 79: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

65

(3-bromophenyl)(tert-butyl)sulfane (33)

ppm (t1)0.01.02.03.04.05.06.07.08.0

0

10000000

20000000

30000000

40000000

1.0

0

1.9

4

0.9

1

9.6

0

Page 80: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

66

2,6-bis(cyclohexylthio)pyridine (35)

ppm (t1)0.01.02.03.04.05.06.07.08.0

-100000000

0

100000000

200000000

300000000

400000000

500000000

600000000

700000000

1.0

0

1.9

2

2.0

5

4.2

5

5.0

52

.61

9.0

13

.89

Page 81: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

67

2,6-bis(butylthio)pyridine (36)

ppm (t1)0.01.02.03.04.05.06.07.08.0

0

50000000

100000000

150000000

200000000

250000000

300000000

350000000

2.0

0

0.9

3

4.5

4

4.8

1

4.7

1

6.6

9

Page 82: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

68

2,6-bis(tert-butylthio)pyridine (37)

ppm (t1)0.01.02.03.04.05.06.07.08.0

-10000000

0

10000000

20000000

30000000

40000000

1.0

0

1.8

7

18

.99

Page 83: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

69

2,6-bis(cyclohexylsulfinyl)pyridine (34) : racemic oxidation

ppm (t1)0.01.02.03.04.05.06.07.08.0

0

5000000002

.00

5.9

9

4.1

0

4.3

0

7.4

3

0.9

42

.83

ppm (f1)50100150

0

5000

10000

14

3.5

74

12

9.5

97

12

7.2

34

12

1.4

85

12

1.4

50

63

.25

9

63

.20

3

26

.42

4

26

.35

1

25

.57

9

25

.33

3

25

.31

2

25

.26

6

25

.22

3

23

.70

0

23

.51

8

Page 84: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

70

2,6-bis(cyclohexylsulfinyl)pyridine (34) : asymmetric oxidation

ppm (t1)0.01.02.03.04.05.06.07.08.0

0

5000000002

.00

5.9

9

4.1

0

4.3

0

7.4

3

0.9

42

.83

ppm (t1)50100150

0

500000000

14

3.5

53

12

9.5

50

12

9.4

29

12

7.1

74

12

1.4

23

12

1.3

89

63

.22

6

63

.17

1

60

.31

0

26

.39

2

26

.32

0

25

.54

1

25

.30

0

25

.28

0

25

.22

8

25

.18

6

23

.63

9

23

.46

4

14

.12

9

Page 85: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

71

References

1 Jensen, K.H.; Sigman, M.S. Org. Biomol. Chem. 2008, 6, 4083.

2 Jacobsen, E. N.; Marko, I.; Mungall, W. S.; Schroeder, G.; Sharpless, K. B. J. Am. Chem. Soc.

1988, 110, 1968.

3 Li, Y.; Song, D.; Dong, V. M. J. Am. Chem. Soc. 2008, 130, 2962.

4 Kolb, H. C.; Van Nieuwenhze, M. S.; Sharpless, K. B. Chem. Rev. 1994, 94, 2483.

5 Li, Y.; Antonic, M.; Song, D.; Dong, V.M. Unpublished results.

6 Wang, A.; Jiang, H.; Chen, H. J. Am. Chem. Soc. 2009, 131, 3846.

7 Wolfe, J. P. Eur. J. Org. Chem. 2007, 571.

8 Wolfe, J. P. Synlett 2008, 19, 2913.

9 Hay, M. B.; Wolfe, J. P. J. Am. Chem. Soc. 2005, 127, 16468.

10 Alexanian, E. J.; Lee, C.; Sorensen, E. J. J. Am. Chem. Soc. 2005, 127, 7690.

11 Liu, G.; Stahl, S. S. J. Am. Chem. Soc. 2006, 128, 7179.

12 Sibbald, P. A.; Michael, F. E. Org. Lett. 2009, 11, 1147.

13 Streuff, J.; Hövelmann, C. H.; Nieger, M.; Muñiz, K. J. Am. Chem. Soc. 2005, 127, 14586.

14 Ma, S.; Lu, X. J. Org. Chem. 1993, 58, 1245.

15 Lu, X.; Zhu, G.; Wang, Z. Synlett, 1998, 115.

Page 86: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

72

16 Lu, X. Handbook of Organopalladium Chemistry for Organic Synthesis, V2 (Ed.: E. J.

Negishi), Wiley-VCH, Weinheim, 2002, pp. 2267.

17 Yamazaki, S. Chem. Eur. J. 2008, 14, 6026.

18 Lautens, M.; Klute, W.; Tam, W. Chem. Rev. 1996, 96, 49.

19 Welker, M. E. Chem. Rev. 1992, 92, 97.

20 Marat, X.; Monteiro, N.; Balme, G. Synlett 1997, 845.

21 Clique, B.; Monteiro, N.; Balme, G. Tetrahedron Lett. 1999, 40, 1301.

22 Xiao, W.-L.; Yang, L.-M.; Gong, N.-B.; Wu, L.; Wang, R.-R.; Pu, J.-X.; Li, X.-L.; Huang, S.-

X.; Zheng, Y.-T.; Li, R.-T.; Lu, Y.; Zheng, Q.-T.; Sun, H.-D. Org. Lett. 2006, 8, 991.

23 Corey, E. J.; Fuchs, P. L. Tetrahedron Lett. 1972, 36, 3769.

24 Sonogashira, K.; Tohda, Y.; Hagihara, N. Tetrahedron Lett. 1975, 4467-4470.

25 Nicolaou, K. C.; Bulger, P. G.; Sarlah, D. Angew. Chem., Int. Ed. 2005, 44, 4442.

26 Mamane, V.; Hannen, P.; Fürstner, A. Chem. Eur. J. 2004, 10, 4556.

27 Wadsworth, D. H.; Geer, S. M.; Detty, M. R. J. Org. Chem. 1987, 52, 3662.

28 Wolfe, J. P. Synlett, 2008, 19, 2913.

29 Zhu, G.; Ma, S.; Lu, X.; Huang, Q. J. Chem Soc., Chem Commun. 1995, 271.

30 Curtin, D. Y. Rec. Chem. Prog. 1954, 15, 111.

31 Moulton, C. J.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1976, 1020.

32 Singleton, J. T. Tetrahedron 2003, 59, 1837.

33 Albrecht, M.; van Koten, G. Angew. Chem. Int. Ed. 2001, 40, 3750.

Page 87: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

73

34 Benito-Garagorri, D.; Kirchner, K. Acc. Chem. Res. 2008, 41, 201.

35 Nishiyama, H. Chem. Soc. Rev. 2007, 36, 1133.

36 Gorla, F.; Togni, A.; Venanzi, L. M.; Albinati, A.; Lianza, F. Organometallics 1994, 13, 1607.

37 Zhang, X.; Longmire, J.M.; Shang, M. Organometallics 1998, 17, 4374-4379.

38 Williams, B. S.; Dani, P.; Lutz, M.; Spek, A. L.; van Koten, G. Helv. Chim. Acta 2001, 84,

3519.

39 Calligaris, M.; Carugo, O. Coord. Chem. Rev. 1996, 153, 83.

40 Fernández, I.; Khiar, N. Chem. Rev. 2003, 103, 3651.

41 Riley, D. P.; Oliver, J. D. Inorg. Chem. 1986, 25, 1821.

42 Riley, D. P.; Oliver, J. D. Inorg. Chem. 1986, 25, 1825.

43 Evans, D. R.; Huang, M.; Seganish, W. M.; Fettinger, J. C.; Williams, T. L. Organometallics

2002, 21, 893.

44 Schaub, T.; Radius, U.; Diskin-Posner, Y.; Leitus, G.; Shimon, L. J. W.; Milstein, D.

Organometallics 2008, 27, 1892.

45 Eapen, K. C.; Saba, C. S.; Tamborski, C. J. Fluorine Chem. 1987, 35, 571.

46 Mariz, R.; Luan, X.; Gatti, M.; Linden, A.; Dorta, R. J. Am. Chem. Soc. 2008, 130, 2172.

47 Bürgi, J. J.; Mariz, R.; Gatti, M.; Drinkel, E.; Luan, X.; Blumentritt, S.; Linden, A.; Dorta, R.

Angew. Chem. Int. Ed. 2009, 48, 2768.

48 Hennrich, G.; Lynch, V. M.; Anslyn, E. V. Chem. Eur. J. 2002, 8, 2274.

49 Wang, Y.; Parkin, S. R.; Gierschner, J.; Watson, M. D. Org. Lett. 2008, 10, 3307.

Page 88: Investigations in Transition Metal Catalysis: Development ... · Palladium Catalyzed Carboesterification of Olefins and Synthesis of Chiral Sulfoxide Pincer Ligands Katherine Jane

74

50 Gingras, M.; Placide, V.; Raimundo, J.-M.; Bergamini, G.; Ceroni, P.; Balzani, V. Chem. Eur.

J. 2008, 14, 10357.

51 Holmes, K. E.; Gilby, L. M.; Stonehouse, J. M.; Kelly, P. F.; Elsegood, M. R. J.

CrystEngComm 2004, 6, 34.

52 Pelotier, B.; Anson, M. S.; Campbell, I. B.; Macdonald, S. J. F.; Priem, G.; Jackson, R. F. W.

Synlett 2002, 1055.

53 Bolm, C.; Bienewald, F. Angew. Chem. Int. Ed. Engl. 1995, 34, 2640.

54 García-Flores, F.; Flores-Michel, L. S.; Juaristi, E. Tetrahedron Lett. 2006, 47, 8235.

55 Padwa, A.; Wong, G.S.K. J. Org. Chem. 1986, 51, 3125.