167
This item was submitted to Loughborough's Research Repository by the author. Items in Figshare are protected by copyright, with all rights reserved, unless otherwise indicated. Epoxidation using dioxiranes Epoxidation using dioxiranes PLEASE CITE THE PUBLISHED VERSION PUBLISHER © L. Walton LICENCE CC BY-NC-ND 4.0 REPOSITORY RECORD Walton, Lesley. 2019. “Epoxidation Using Dioxiranes”. figshare. https://hdl.handle.net/2134/10478.

Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

This item was submitted to Loughborough's Research Repository by the author. Items in Figshare are protected by copyright, with all rights reserved, unless otherwise indicated.

Epoxidation using dioxiranesEpoxidation using dioxiranes

PLEASE CITE THE PUBLISHED VERSION

PUBLISHER

© L. Walton

LICENCE

CC BY-NC-ND 4.0

REPOSITORY RECORD

Walton, Lesley. 2019. “Epoxidation Using Dioxiranes”. figshare. https://hdl.handle.net/2134/10478.

Page 2: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

This item was submitted to Loughborough University as a PhD thesis by the author and is made available in the Institutional Repository

(https://dspace.lboro.ac.uk/) under the following Creative Commons Licence conditions.

For the full text of this licence, please go to: http://creativecommons.org/licenses/by-nc-nd/2.5/

Page 3: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Pilklngton Library

, , ' ,

Author/Filing Title I,.)lh.. -rv..l L . .......................... ) ........................................ .

~~~~~~;~~;.~~~;.~~ ............................................................ .'........ 'I

Q-tol*7°,,&7

Vol. No ............... .. Class Mark ........ ........................................

26 JUN 1998

25 JUN 1999

'~j~N~nO

- 6 ~F~ 200

8 JU\Il 2000

- 6 ocr 2000

If

1.1

\

\ r , IIIII

c 1

• ' ' '\' -, .

-~ --

,I

,.

Page 4: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and
Page 5: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Loughborough University of Technology Department of Chemistry

Epoxidation Using Dioxiranes

By

Lesley Walton.

A Doctoral Thesis.

Submitted In Partial Fulfilment Of The Requirements For The Award Of Doctor Of Philosophy Of The

Loughborough University Of Technology.

Supervisor: Professor BA Marples.

© L. Walton. August 1996

Page 6: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and
Page 7: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

This thesis is dedicated in loving memory to my grandparents

Bert and Lily

and

To my mother Joan and sister Janet with love.

Page 8: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Abstract

Chapter 1 contains a brief review of the historical background of dioxiranes

and an outline of their more recent chemistry.

The preparation of homochiral 2-fluoro-2-substituted-1-tetralones and ethyl

2-fluoro-1-indanone-2-carboxylate is described in Chapter 2 and the X-ray

crystal structure of (1' R,2' R,5' R)-( -)-menthyl 2S-fluoro-1-tetralone-2-

carboxylate presented. The synthesis of racemic 2,5,7 -trifluoro-1-indanone-2-

carboxylate is also discussed.

The dioxirane derivatives of the above ketones have been prepared in situ

and have been shown to epoxidise trans-stilbene, trans-~-methylstyrene and

6-chloro-2,2-dimethyl-2H-1-benzopyran, but not enantioselectively. The

individual results of these epoxidations are given in Chapter 3. The

incorporation of fluorine into the aromatic ring of aryl alkyl ketones is shown to

increase dioxirane reactivity.

Chapter 4 briefly describes the reactions carried out in order to ascertain the

optimum conditions for the epoxidation of alkenes by in situ derived

dioxiranes. The rate of this oxidation was shown to be highly dependent on

the concentration of the dioxirane in the organic phase.

Described in Chapter 5 are the procedures used for the isolation of the

dioxirane derivatives of dimethyldioxirane, 1,1, 1-trifluoroacetone, 2,2,2-

trifluoroacetophenone, hexafluoroacetone and cyclohexanone, in solutions of

the corresponding ketone with or without dichloromethane.

Page 9: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Contents

Acknowledgements. I - 11

11 Abbreviations. III

III Introduction ..

Chapter 1. Review of dioxirane chemistry.

1.1 A brief historical note.

1.2 Preparation of dioxiranes.

1.3 Oxygen atom transfer by dioxiranes.

1.3.1 Epoxidation reactions.

1.3.2 Asymmetric epoxidations.

1.3.3 Epoxidation of allenes.

1 - 39

1

4

7

8

11

12

1.3.4 Table 1: Recent epoxidation reactions. 13

·1.3.5 Oxidation of carbon - hydrogen bonds using dioxiranes. 20

1.3.6 Table 2: Oxidation of carbon - hydrogen bonds. 21

1.3.7 Oxidations of organic sulfur, silicon, selenium and

nitrogen containing substrates. 25

1.3.8 Table 3: Oxidations of organic sulfur, silicon, selenium

and nitrogen containing substrates. 26

1.3.9 Oxidations of metal enolates and metal complexes. 31

1.3.10 Table 4: Oxidation of metal enolates and metal

complexes. 31

1.4 Other aspects of dioxirane reactivity. 33

1.4.1 Effect of solvent on reactions. 33

1.4.2 Electrophilic character I Mode of oxygen transfer. 34

1.4.3 Decomposition of dioxiranes. 35

1.4.4 Carbonyl oxides and dioxiranes. 36

Page 10: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

IV Results and Discussion

Chapter 2. Epoxidation of alkenes using homochiral

dioxiranes.

2.1 Introduction to the Project.

2.2 Synthesis of the 1-tetralone and

40 - 57

40

1-indanone derivatives. 41

2.2.1 Synthesis of the 2,2-disubstituted tetralones. 42

Synthesis of ethyI1-tetralone-2-carboxylate. 43 Synthesis of methyI1-tetralone-2-carboxylate. 44 Synthesis of menthyI1-tetralone-2-carboxylate. 44

2.2.2 Fluorination of the alkyl 1-tetralone-2-carboxylates. 47

2.3 Separation of the diastereomers of menthyl 2-fluoro-1-

tetralone-2-carboxylate 48

2.4 Separation of the enantiomers of the tetralone derivatives

(31c) and (34) via an imine derivative. 50

Separation of the ester derivative. 50 Separation of the tertiary alcohol derivative. 51

2.5 Synthesis of the ethyl 2-fluoro-1-indanone-2-carboxylate. 51

2.6 Separation of the enantiomers of the 1-indanone derivative

via an imine derivative. 52

2.7 Synthesis of methyl 2,5,7 -trifluoro-1-indanone-2-carboxylate. 53

2.8 Synthesis of 1 S-methoxy-1 S-phenylacetone. 58

Chapter 3. Dioxirane Reactions. 59 - 62

3.1 Use of the 1-tetralone and 1-indanone dioxirane derivatives

in the oxidation of alkenes.

3.2 Results of the epoxidations.

59

59

Chapter 4. Optimizing conditions. 63 - 68

4.1 Optimizing the conditions of the dioxirane epoxidation

reaction. 63

Page 11: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

v

VI

VII

Chapter 5. Isolation of dioxiranes in solution.

5.1 Preparation of solutions of dioxiranes.

5.2 Determination of the molarity of a dioxirane solution by

titration.

Conclusion.

Experimental.

Chapter 2, Experimental.

Chapter 3, Experimental.

Chapter 4, Experimental.

Chapter 5, Experimental.

References.

69 -71

69

69

72

73 -132

73

115

117

119

133 -144

Page 12: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Acknowledgements

This research was carried out at Loughborough University of Technology

during the period 1st November 1991 to 31st October 1994 in collaboration

with Smith Kline Beecham Pharmaceuticals, New Frontiers Science Park,

Third Avenue, Harlow, Essex.

I would like to thank the following:

Smith Kline Beecham and Professor BA Marples for funding the work.

Professor BA Marples for his invaluable supervision and guidance

throughout this work at Loughborough.

Professor Curci for allowing me to work within his laboratory in Bari and for

his much needed advice. Also his research group, Caterina, Lucia, Teresa,

Anna and Michele for their help friendship and support during my month in

Bari.

Dr. P. Smith for his help and supervision at Smith Kline Beecham and his

useful discussions over the past three years.

Mr. G. Freer and Dr. G. Taylor and their colleges at the New Frontiers

Science Park for their help and friendship during my two months at

SmithKline Beecham.

Smith Kline Beecham for running some mass spectra and high field NMR

spectra.

Dr. D.S. Brown for his X-ray crystallography work.

The following technical staff at Loughborough: Mr. Alastair Daley and Mr. Paul Hartopp for their help and daily running of the

research laboratory. Mr. John Kershaw for assistance with GLC experiments, running of high field

NMR spectra and his invaluable friendship and support.

Mr. John Greenfield for running mass spectra.

Mr. Alastair Daley for running microanalysis experiments.

Page 13: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

I wish to thank the following for their friendship and help at Loughborough:

Tracey Ross, Gabrielle Loftus, Vicki Waddington, Mark Elliot, Bob and Violeta

Marmon, Sue Toh, Gi Cheung, Craig Baldwin, Jon Eddols, Dave Riddock,

Chris Frost, Jon Roffey, Dave Miller, Dave Price, Hank Cheung, Noeleen

O'Sullivan, Liz Swann, Dave Corser and all the other members of organic

research past and present 1991 -1994.

I would like to thank the following for their friendship and moral support:

Lesley Reid, Vicky Hardman and Micheal Brown, Cherry and Ken Ross,

Steve and Glenda Rooney, Penny Muncie, Jenny Glass, Julia Gunby, Andrea

Banham and Ron Mc Keating.

I extend my thanks and admiration to my mother for her constant support

throughout my endeavours.

11

Page 14: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

IR.

NMR.

MS.

ppm.

8. Hz.

J. TLC.

Rf·

GLC.

THF.

HCI.

OCM.

HMPA.

TBAHS.

Oxone®.

EOTA.Na2·

LOA.

Eu(hfch

aq.

lit.

Me.

Et.

Men.

bpt.

mpt.

ee.

eq.

Abbreviations

Infrared. Nuclear Magnetic Resonance.

Mass Spectrometry.

Parts Per Million.

Chemical Shift.

Hertz.

Coupling Constant.

Thin Layer Chromatography.

Retention Factor.

Gas Liquid Chromatography.

Tetrohydrofuran.

Hydrochloric acid.

Dichloromethane.

Hexamethylphosphoramide.

Tetrabutylammonium hydrogen sulfate.

Potassium peroxymonosulphate

(2KHS05 .KHS04.K2SO), Aldrich Chemical

Company.

Ethylenediaminetetraacetic acid disodium salt.

Lithium diisopropyl amide.

Europium D-3-heptafluorobutyrylcamphorate.

Aqueous.

Literature.

Methyl.

Ethyl.

J.Menthyl.

Boiling Point.

Melting Point.

Enantiomeric Excess.

Equivalents.

III

Page 15: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 1

Page 16: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 1.

Review of Dioxirane Chemistry.

1.1 A Brief Historical Note

Dioxiranes (1) represent the class of most strained cyclic, organic

peroxides, comprising one carbon and two oxygen atoms in a three

membered ring.

The first literature reference to a dioxirane (1) was made by Baeyer and

Villiger,1 in 1899, in a publication describing a reaction which today bears

their names. They suggested that a dioxirane was the intermediate involved

in the conversion of the cyclic ketone, menthone, to its lactone by

monoperoxysulfuric acid. This theory later lost out to a competing

mechanism, primarily as a result of an 180 study by Doering and Dorfman.2

Recent work by Rozen et a1' on the complex HOF.CH3CN has revived the

theory proposed by Baeyer and Villiger. 1 Using 180 labelled reagent, it was

found that ketone oxidation, using the HOF.CH3CN reagent, proceeds

through the original dioxirane mechanism (Scheme 1).

-- +

Scheme 1: Conversion of a ketone to a lactone using HOF.CH3CN.

The first apparent report of the preparation of a dioxirane is contained in a

1

Page 17: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

1972 patent. 4 The authors claim to have isolated perfluorodimethyldioxirane

(3) by low-temperature gas chromatography (Scheme 2). The dioxirane was

synthesised by fluorine oxidation of the precursor dialkoxide (2).

(2)

Fluorine Oxidation ..

(3)

Scheme 2: Dioxirane (3) synthesis by fluorine oxidation of the precursor dialkoxide (2).

It was Montgomery,5 in 1974, who first observed that certain ketones

enhanced the rate of decomposition of monoperoxysulphuric acid. He also

noted that a number of oxidation reactions of caroate were catalysed by the

presence of ketones. From these observations he proposed that the

monoperoxysulphate anion (5) adds to the ketone (4) to give the adduct (6).

Since a variety of ketones enhanced the rate, he suggested that the adduct

(6) further reacted to give the dioxirane (1). He did not, however, propose that

the dioxirane was acting as the oxidant. As Montgomery worked using a

slightly alkaline medium, as opposed"to the acidic conditions of the Baeyer­

Villiger reaction,1 he revived the idea that dioxiranes could be involved in

these reactions.

Edwards, Curci and co_workers6.7 ,8,9,1o,11 have since carried out kinetic

studies and 180-labelling experiments which have presented strong evidence

for the presence of dioxiranes in the ketone I caroate system. Their work also

showed that dioxirane itself was a very powerful oxidant. The reactions were

performed in an aqueous medium containing the ketone, Oxone® (caroate)

and the substrate. For efficient oxidation, the pH was strictly controlled at pH

7.5. At this pH, Edwards and co-workers6 noted that with most ketones, there

was little or no competition from the Baeyer-Villiger reaction.1 From this work

they postulated the mechanism of dioxirane formation and decomposition

(Figure 1).6

2

Page 18: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Figure 1:

Mechanism of Dioxirane Fonnation and Decompositions

5: + so ... --­+ so

BV = Baeyer-Villiger pathway

1- 50;

R 0

RXL (1)

-oosoi .. R2C=0 + 0-0

R2C=O + 0-0

+ 50;-

A recent paper by Armstrong et al,12 discusses their 180 labelling study of

dioxirane epoxidation reactions. Their results lead them to conclude "that a

dioxirane intermediate is probably not responsible for alkene epoxidation in a

ketone-accelerated axone" epoxidation system" (Scheme 3). These workers

assumed that the addition of KHSOs to the carbonyl group is non­

stereoselective and that either of the diastereotopic oxygen atoms is

geometrically capable of being transferred to the alkene. This being the case,

the resulting dioxirane intermediate would result in 50% label incorporation

into the epoxide. However, no such label was observed in the product (8).

They also noted that this in situ'system only gave the syn-isomer (8), whereas

epoxidation with isolated dimethyldioxirane (9) afforded a mixture of syn- and

anti-isomers.

3

Page 19: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Scheme 3: Epoxidation of the ketone (7) using the biphasic Oxone® system. i = Oxone®, Bu.NHSO., EDTA.Na2, aqueous NaHC03, CH2Cb, O°C to room temperature.

180 180

• 0

(7) (8)

t 180-0

16118 0

6 • 0 '6118

They postulated that the tetrahedral species (6) resulting from the addition

of HSOs- to the carbonyl group is capable of alkene epoxidation, and this

epoxidation reaction is faster than ring closure to the dioxirane.

1.2 Preparation of Dioxiranes

In 1985, Murray and Jeyaraman '3 prepared solutions of dioxiranes in their

parent ketones. They used these solutions to carry out a variety of

synthetically useful reactions.

In 1988, Curci et a/14 published the synthesis and characterisation of

methyl(trifluoromethyl)dioxirane from 1,1, 1-trifluoropropanone and potassium

peroxymonosulfate. They reported concentrations of the dioxirane in the

parent ketone ranging from 0.65 - 0.82 M. Despite the oxidising properties of

methyl(trifluoromethyl)dioxirane, its use has been somewhat restricted due to

a problem in its preparation. Some commercially available batches of 1,1,1-

trifluoropropanone failed to produce the dioxirane when treated with Oxone®

4

Page 20: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

under the conditions described by the aforementioned workers.14 Recent

studies have shown this to be due to the presence of diethyl ether. This

impurity induces the decomposition of the dioxirane in a concentration

dependent manner.15

In 1991 Adam and co-workers 16 published a simple procedure for the

preparation of dimethyldioxirane (9) in its parent ketone (Scheme 4). The

method involved the addition of solid Oxone® (caroate), in five portions, at

three minute intervals, to a vigorously stirred solution of water, acetone and

sodium hydrogen carbonate at low temperature. After three minutes from the

last addition, the cooling was removed and a slight vacuum applied. The

dimethyldioxirane (9) in acetone was then distilled into a cooled (-78°C)

receiving flask. Concentrations of up to 0.1 M were reported.

Me

Oxone ® 5 - 10°C

pH 7.4

Scheme 4: Preparation of Dimethyldioxirane (9).

)=0 Me

In 1992, Murray, Singh and Jeyaraman 17 published a method which they

had developed to prepare solutions of less volatile dioxiranes, e.g. the

dioxirane derivative of cyclohexanone. This method involved the rapid

addition of cooled Oxone® to a solution of cooled ketone, phosphate buffer

and ice. The pH was regulated by the addition of potassium hydroxide

solution. After stirring the solution for three minutes, the dioxirane I ketone

mixture was "salted out" from the aqueous layer. The two layers were then

separable, yielding a solution of the dioxirane in its ketone. Concentrations of

0.65 - O.82M were reported.

Difluorodioxirane (10), apparently the first dioxirane stable in the gas

phase at room temperature, was synthesised and characterised by Russo

and DesMarteau in 1993.18 The synthesis involved the formation of the anion

5

Page 21: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

FOCFZO- by nucleophilic attack of fluoride ion on the unstable FC(O)OF

molecule followed by its oxidation using CIF (Scheme 5). This dioxirane is

reputed to be the most reactive of all dioxiranes.

Scheme 5: The proposed, electron transfer, mechanism for the formation of difluorodioxirane. Xz = Clz, CIF, Fz.

0 F o- F O· F)lO/F

CsF X2 .. FXO-F

.. X F O-F

~ F 0 F 0

X2 +F- + XI • XI + x"2- + F F 0 F 0

(10)

In order to confirm the actual structure of difluorodioxirane (10), Burger et

aP9 carried out a detailed study of its high resolution mid-infrared FT

spectrum. Their results corroborate, unambiguously, the previously proposed

symmetrical dioxirane structure for this molecule (Figure 2).

131.5 pm",\ (1~.9 pm

F~J I

" .' /CS 157.6pm

F 109.050 0

Figure 2: The molecular structure and present best estimate of structural parameters of difluorodioxirane, as determined from its i.r. spectrum.

At the beginning of 1994, Jones et a/ published a novel procedure for the

preparation of dimethyldioxirane (9) from neutralised Care's acid, at ambient

temperatures. 20 The important aspects of the process were that Caro's acid

was partially neutralised to below a pH of 2, in order to minimise

6

Page 22: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

peroxymonosulfate losses and to avoid a raised H20 2 content. The second

neutralisation stage was carried out using NaHC03 buffer solutions. Their

procedure allowed dimethyldioxirane to be prepared in situ and ex situ,

as appropriate. They recorded ex situdimethyldioxirane concentrations of

0.12 M.

In 1994 dimesityldioxirane (12) was isolated at room temperature in both

solution and pure form. This is the first dioxirane to be isolated as a solid and

in a pure form. Sander et ar synthesised this dioxirane by irradiation of its

diazo derivative (11) in CFCI3 at 183K. The best yield was 50% on irradiation

with A. > 495 nm and up to 10 mg of dioxirane could be obtained in a single

run.

k hv Q-

c,. kp hv ~

~C-N' --- ~C' --- ~=' --- ~d °2 'I ' ° --

(11 ) (12)

Scheme 6: Synthesis of dimesityldioxirane (12).

1.3 Oxygen Atom Transfer By Dioxiranes

Dioxirane chemistry has rapidly expanded in recent years and there are

now several publications which comprehensively review the area.22 Since the

most recent review in 1995,229 literature has continued to appear which

describes the ever increasing uses of these molecules. The remaining part of

this Introduction is dedicated to the chemistry of dioxiranes. Tables 1 - 4

(Pages 13 to 32) represent some of the chemistry which has recently

emerged using these oxidants.

7

Page 23: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

1.3.1 Epoxidation Reactions of Dioxiranes

The most widely studied dioxirane in epoxidations has been

dimethyldioxirane (9). It has been shown to rapidly epoxidize alkenes in high

yields.6-11,13,23,24 Baumstark and McCloskey composed a relative reactivity

series for the epoxidation of alkenes using this reagent. 23 This work showed

the reaction to be particularly sensitive to steric factors and suggested the

involvement of a spirotransition state (Figure 3).

Figure 3: Spirotransition state for the epoxidation of alkenes by dioxiranes.

In addition to steric interactions, solvent effects are thought to play a major

role. 13,25 The presence of water has been shown to increase the rate of these

epoxidations (et Section 1.4.1).24

Scheme 7.' In situ preparation of dioxiranes.

® Oxone

Phosphate Buffer

pH7,~.

EDTA.Na2 KOH

2 _ScC

Oxone® = caroate = potassium peroxymonosulphate

+

PTC = [Me(CH2h14NHS04 = tetrabutyl ammonium hydrogen sulphate

Edwards et af have shown that their in situ method of preparing

8

Page 24: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

dioxiranes can be used to carry out a number of O-atom transfer reactions.

They demonstrated that olefins were epoxidized in a syn stereospecific

manner and in high yield. This work was extended to water-insoluble 0lefins7

by the use of a phase transfer catalyst (Scheme 7).

Denmark26 has recently looked at optimizing the conditions of the dioxirane

epoxidation reaction. He found that many of the key experimental variables in

the biphasic system are interdependent, and therefore found he was unable

to systematically examine all permutations. The results of his findings are

listed:

• The stoichiometry of the ketone has a dramatic effect on the overall

oxidation.

• The rate of Oxone® addition affects the reaction. Slower addition rates,

while maintaing a pH between 7.8 and 8.0, led to more efficient

conversions to the epoxide.

• Oxidation continues long after the addition of Oxone® is complete.

• The rate of dioxirane formation exhibited strong pH dependence, with a

maximum at pH 7.5 - 8.0. Oxidation rates were different at different pH,

and decrease in the order 8.0 > 7.8 > 7.5. The slower rate but complete

oxidation at lower pH has practical consequences for slow reacting

substrates because of the slower rate of Oxone® consumption at pH 7.5.

• The rate of oxidation increased with increasing olefin substitution: tetra>

tri > di > mono and isolated olefin reacted faster than conjugated alkenes

and allylic alcohols.

The ability of the ketone to serve as a promoter for epoxidation involves two

crucial features:

• the ability to efficiently form a dioxirane.

9

Page 25: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

• the ability to efficiently transfer the oxygen to the substrate.

The effect of the ketone structure:

• a-substitution in acyclic ketones significantly decreased epoxidising

efficiency.

• Steric congestion about the carbonyl carbon decreased the rate of

oxidation.

• Employment of electronically activated ketones increased the rate of

oxidation, provided the ketones were not too hydrophilic.

• Oxidation efficiency was strongly ring size dependent for cyclic ketones.

The trend was explained in terms of the facility of rehybridisation of the

carbonyl carbon (Sp2 to Sp3) and the eclipsing interactions with the a­

methylenes.

• Amalgamation of the carbonyl precurser of the dioxirane with the

ammonium function of the phase transfer catalyst, unless sterically

hindering, had a profound beneficial effect on the efficiency of oxidation.

• The nature of a counterion (in ketones containing a quaternary nitrogen)

greatly influenced the rate of oxidation: SF 4- > CF3S03- > N03- > CI04- (at

pH 7.S); pH was also noted to effect this influence.

• Increasing the lipophilicity of a salt was crucial for successful oxidation.

Chain length, not just carbon count, was important.

Their results confirmed the results of Edwards and Curci.6-11

The standard oxidation protocol selected by Denmark et a/ 26 consisted of :

• 10 equivalents of Oxone® (S hour addition),

• 2 equivalents of ketone,

10

Page 26: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

• 0.1 equivalent of phase transfer catalyst,

• 24 hour reaction time I 0 QC I pH 7.8

The best catalysts under biphasic conditions were found to be 4-

oxopiperidinium salts.

1.3.2 Asymmetric Epoxidation

The first report of asymmetric epoxidation using chiral dioxiranes was

published by Curci and co-workers. 10 They reported that chiral ketones can

be converted to dioxiranes by caroate. These dioxiranes reacted with

prochiral alkenes to give epoxides with enantiomeric excesses in the 9-

12.5% range. Such enantioselectivities are superior to those given by

standard reagents, such as (+)-monoperoxycamphoric acid. The optically

active parent ketone can be recovered unchanged at the end of the reaction,

hence the reaction is catalytic rather than stoichiometric. Curci has continued

to study this area of dioxirane chemistry.27 He reports asymmetric epoxidation

of the prochiral alkenes, trans-p-methylstyrene, trans-2-octene and cis-2-

methyl-2-octadecene, of 12 - 20% using (+ )-3-(trifluoroacetyl)camphor, or

R( +)- and S( -)-3-methoxy-3-phenyl-4,4,4-trifluorobutan-2-one (13) as

precursors for chiral dioxiranes generated in situ. He also reports that the

adoption of chiral ketones carrying electron-withdrawing groups as dioxirane

precursors was advantageous as far as reaction times and yields. These

results are in agreement with those previously found within this

Department.154 In the same publication he has suggested that the structure of

the chiral dioxirane oxidant and of the prochiral alkene have little influence on

the enantioselectivity. In this, dioxiranes resemble other non-metal

asymmetric epoxidising agents, relying solely on non-bonded interactions in

the transition state to achieve enantiofacial selectivity. An additional limitation

that arises with dioxiranes, is that the seemingly diastereotopic oxygens are

both available for oxygen transfer. However, given a preferential geometry of

11

Page 27: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

approach (e.g. spiro orientation) and certain steric interactions, the chiral

recognition would still have to take place via one of four diastereotopic

transition states (Figure 4).

Figure 4: The four diastereomeric transition states for the epoxidation of trans-f3-methylstyrene by R-(13)

The work in this thesis is largely concerned with the use of homochiral 2-

fluoro-2-substituted-1-tetralones and -Hndanones in asymmetric

epoxidations. Some of this work has recently been published.28

1.3.3 Epoxidation of Allenes

Crandall et a/have shown that dimethyldioxirane provides access to the

fragile diepoxides of allenes.29 They have further examined its oxidation of

allenes bearing nucleophilic groups and have reported the formation of highly

functionalised oxygen heterocycles derived from cyclisation of the

intermediate mono- and di- epoxides. 3O They have now studied the oxidative

cyclisations of allenic aldehydes (Table 1, Entry 24)51 and allenic

sulfonamides (Table 1, Entry 25),52 and report a novel and efficient route to

highly functionalised heterocyclic systems.

12

Page 28: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

1.3.4 Table 1: Recent Epoxidatlon Reactions

Entry Substrate Dioxirane31 Method32 Products No. Yield (%1

Ref.

o&=>

c")<l !'.

1 A ~ 4~,5Il-epoxide(-15%) 33

CH:; 0 4Cl.5a-epoxide (-5%) o (80%)

!'. !'.

2 O~H'OAC CH«I ~ 4a,5a-epoxide (0 - 14%)

A 1 a,2a-epoxide (0 - 21 %) 33 CH 0 1p,2fi.epoxide (0 - 4%)

o (61 - 81%)

0 ~ 1~ Major 34

3 ~

:~ A o<lo-i, J. :x Product

• lo+o

I DiD""" OMe

H

OM. H

+ 0

~ Minor

o~o+o~ Product

OMe H

-

~, 4

~" c")<l A i~tl" 35

+

CH:; 0

(50%) (45%)

Ph ro »5-Me

WM. _ M.

A R 0 0

5 CH)<j A la) IZ)-Ib)

R = H; >95% (a) 36

CH3 0 R '" OMs; >95% (b)

I I 1.2 - 1.4 equivalents

R = H, OMe + hv

R@'T~

6 .et CH,xl A $-);; CH, J 37,38

1 equivalent (a) (b) 0 ~~ 0 H, Ac: (a) > 95% ['H NMA ]) R = MeO: Ibl> 95% ['H NMA j)

7 cxS-Me

CHxY cf1M' + cf~ 37

CHl 6 A 00

,n\ (75%) M.

1.2 • 1.4 equivalents

CX:>«Me 0

(14%) .

13

Page 29: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry No.

8

9

10

11

12

13

Substrate

ct·o:C~ (E-a) 92: 8 (Z-a)

~MO

::Q5-"' M.

I M.

'" 0

"'

~H; \ COCH2X

A = Me, Et, iPr, t-Bu

X =H,Ci

Dioxirane31 Method:J2

A

CH,V!

cH(''1! A

CHi,/? cH/"'!

A

C')<i CH; 0

A

1.2 . 1.4 equivalent

A

A

Products Yield (%1

~:- [00:] (76%) (E·)

R, ...

R,~R' Me .... R,~ R, - ~~o

• (a) "' (b) (71 • :>95%); (a) fonred in each

case except where A3 -Me, OMe

~MO

I H OMo

l I m~: OEd i coc,

ICOC~ '-m""- Mo ~CN ~O+O~N

eEt Me eN

R

c:Qtc~ , COCIY(

I r.t.. >5h i •

d): , COCI-\:x

(:>95o/ •• except when A .. t·Bu,

decomposition

products only)

I

oJ<~ , COC~X

14

ReI.

39

39

39

4041

42

43

Page 30: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Substrate Dioxirane31 Method32 Products Ref. No. Yield I%l

14 aY :~ A ~ (100%)

43

\ in CH2CI 2 COCH3 COCH,

cx;vY CH,xl ~R + ~ 44 15 A CH; 0

0 0

(trans or eis, respectively) (tJans or cis) (16-41%) R (21 -51%) (+ minor products)

Q D ~ ~ R ...

16 - B 45 .. Ol~ 01

0-0 C--'O worst (27%, cis .- trans; 25 :75) X

O

best (100%. cis: trans: 4: 96)

OA AD'" &0 6·· ab.",,, 01 '01000. 00 """"0"

17 6 B + ;;;0 or opposite selectivity Observed. 45

Ol~ 01 worst yield (40%, cis: ITafl.t ; 13: 87)

0---0 0-0 worst selectivity {59%, cis: ttans: 36· 64) X

best yield and selectivity (770/., cis: rnrns; 10: 90)

~ :~ 0

18 A R"A 46 A OCONH2

A OCONH:, R - H. 0

OMe C""X

O OMe

19 ~OM. CH, ~ A ~OM. 47

in CH2CI 2

HO OAc HO OAc (100%)

OM.

~ 20 ~OM. c'Xl A

~ .• " CHQ OM.

47

CH, ~ in CH 2CI 2

HO OH 0

0 0

21 6' CH)<j A to 6::H 48 .. CH3 0 'oH

(a) (b)

A 1 : 1 mixture of (a) : (b) observed. Ills

suggested (b) possibly arises from opening of the anti-epoxlde (b) with neighbouring-group participation from the ketone carbonyl.

15

Page 31: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Substrate Dioxirane31 Method32 Products Ref. No. Yield (%)

('(~ ::><! A % % 49 22 'Vy

+ K2COS I CH, CH, CH, 'V\

(14%) 0 (13%)

~R' OH

RX~ OH

~~ RX:' CH'Xl o + +

23 CH, ! A R, R, '" 50

At constant temperature. amount of enone This paper looks at decreased with increasing substitution at epoxidation versusC-H C::C. At lower temperatures. more enona insertion. formed, except for the trimethyl derivative.

..

:?R' ::;><! y>R' 51 24 A

R, Rt 0 R3

o R, R. R. Depending on conditions used, i.e. in the presence of a nucleophile or acid, several

cyclisation products were obtained.

N >= ~:;><! ..... " ()n 52 25 . C»n A and B

TsHN TsHN n = 1,2,3 Depending on conditions used, i.e. in the

presence of a nucleophile or acid, several cyclisation products were obtained.

o CN CH)<j ltiCN

26 ~R" A (73%) 53

CH, 0 OR"

A" = (1 S)-camphanoyl R" = (1 S)-camphanoyl

27 X-COOH CHXO B k COOH

54

CH, ! Product yield decreased when two carbonyl

0.5 - 2.5 hrs groups or an amide group present.

0 0

~ 0

~ 55 28 ::;><! A ~

x--l1 Kinetics of epoxidation (>95%)

studied Reaction found to be second order

0 CH,xl o~o 29 A 56

0 CH; 0 H H

16

Page 32: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry No.

30

31

32

33

34

Ra'

Substrate

0 H

E-isomer

(Z_)

o

Dioxirane31 Method32

CH"V'?

CH{'6

CH)<j

CH3 ~ in CH:zC1 2

CH,.x1

CH3 0

in CH 2CI 2

CH3.xl

CH3 0

::t<l 3.2 equivalents

(in CH,CI,)

A

A

A

A

A

R

(73 - 86%)

~ .. O£O. o 0 0 0

(.)~ CbI

R = H: (a) (15%), (b) (10%); R = OC2H5: (b) (14%); R = Me, Cl, Br. (a) (-18%).

o l Base

Cyclization Products:

3-Hydroxy1lavones and

2-(-<t-hydroxybenlyl)-3-coumaranones.

17

Ref.

'ST.

58

58

59

60

Page 33: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Substrate Dioxirane31 Method" Products

Ref. No. Yield (%)

"''''

.1" ci CH'.Xl 35 A 60

~ CH, 0

t~ R=H,Ac, 3.0 equivalents (60%)

p-SrCoH.Co. inCH,q

RO"'U AO"~'

0 "6; -["+t; 1 - R'Xx~ oA R, oJ-N• R,

CI<I I I

36 A R R ., o I R, CH 0 ~ H2O ha

R R = Me. Sugar. R. = H. Me;

R"&R, R~= Ht Me;A.1= H, Me. R"~~ A ... ~~ A OH OH

o I R, o N R I , R R

o CH,OR, CH,.><l o CH2OR,

37 00~0,==,",R, A R,)lO~°'"\-TR' 62

CH, 0 0

~ CHX1 ~ 38 0""- A 0""-

63 CH, 0

0 R, 1.0 equivalent R2

R, = Me; A, = H,F. (85 - 96%). Dlepoxide fOlTIloo with further dloxlrane.

CH'.Xl

~ ~ CH, 0 39 0""- A 0""- 63

1.0 equivalent 0

~ or R,

A, = CFl ; R, = H. CH><l DMD: 42% (12 days); TFMD: 85% (3 min.).

CF, 0 Diepoxide fonned with further dioxirane. 1.4 equivalents

Meo::X; » 40 CH'.><l A Mea 0 64

MeO CH, 0 MeO (89%)

OM. OMe

18

Page 34: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry No.

41

42

43

44

45

46

47

Substrate

.oms

o=XH .'

T6S0'

~~~~o __ OR HO Jc~

R = SiMe,But; H.

Other examples of alkyne : epoxidation given, but not

shown here.

bn

BnO Bn

0=<:c!J 'Bn

0--/;-0 0=<;;J.bJ , Bn

Dioxirane31

CH,xl CH; 0

1.0 equivalent InCH,q

CH>d CH; 0

CH", A CH!\~

CH,v'i

cH/"'~

Method32

A

A

A

A

A

A

A

Products Yield (%)

o~~.o Meo..("··' ...• /; (98"10)

0 .. ·· '", ob o -.(CH,b

KJ~..oT8S

~OH mso' (41%)

Ph~H 0 ~AoO 0

; ,,--PhMo· .. · ,',0 ;: J-O : AGO

OH OCOPh

t'oO/

75 . 15 10 _ (42"10 isolated)

BoO-Z""_a r ..oq OS(Ph)"t8u

,~ a,~ BnO Bn I \ 08n t B

'. ' OB, \_

~n~ en ( ... Bn~ OHQ ~

i re. Complete inversion 01 Bn~- configuration. BoO~

0--/;-0 0=<;;J.N , Bn

2.3

0~9 +O-:'~ ,

Bn

: 1

19

Ref.

65

66 .

68

69

70

70

Page 35: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Substrate Dioxirane31 Method32 Products .No. Yield (%1

48 ~ A 0.-( CH)<j db , 0'-'< Sn CH:] 0 ,

Sn

~ON' i) -78 - r.t ~" :&: 49 :::xI A

ii) TBAF, THF, 94 6_ r.t. (~%)

CYP" CHX'1 ~ 50 A CH:] 0

0 H

-"

~ 51 :::xI A

"

1_3.5 Oxidation of Carbon - Hydrogen Bonds using Dioxiranes

The ability of dioxiranes to insert into C - H bonds is demonstrative of their

remarkable reactivityn,81,82 Dioxiranes oxidise tertiary alkanes into their

respective alcohols and secondary alkanes into ketones. Oxidations with

dimethyldioxirane proceed stereoselectively with retention of configuration, as

the reactions with cis-decalin (14) (Scheme 8) and trans-decalin (15)

(Scheme 9) demonstrate.74 This also illustrates that the oxidation of

equatorial positions is preferred to that of axial positions.

K6 17hours LT - . Hob V -(84%)

(14)

Scheme 8: Oxidation of cis-decalin (14) by dimethyldioxirane (9).

20

Ref.

70

71

72

72

Page 36: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

17 hours ~ H (20%)

(15)

Scheme 9: Oxidation of trans-decalin (15) by dimethyldioxirane (9).

The high selectivity of dimethyldioxirane in its oxidation is also illustrated by

its oxygen insertion into toluene, ethylbenzene and isopropylbenzene." The

oxidation at the benzylic position is in the order PhCH(CH,), > PhCH,CH, >

PhCH,. Similarly, in the reaction with saturated hydrocarbons, oxidation has

been shown to occur preferentially at the tertiary carbon atoms."·7'

1.3.6 Table 1: Oxidation of Carbon - Hydrogen Bonds

Entry Substrate Dioxirane" Method" Products No. Yield (%)

1 QtP 1.4 CHXi CH, 0

A ~ 1/

o 137%\

QtP ~ CF>d A i _,t,j 2 IN CH, 0 o 0 (95%)

3 C'o;9 CF,)<l A ~ Q;Y )fY 0 .... d CH, 0

(3;"') OAc os_ .... (38%) OAc cl«HatWe (21"')

4 J!to CH>d CH, ~ A ~(10%)

~ ~O-5 CH>d A _I "0< I I ~ lX_a..l1.Y.H2 CH, 0 ~ +r:r lX_a_H.Y-H, 2X_a·H. Y.O

3X .Srr:...cI.sIJ.Y -H2 X 2X_a_H, Y=O 3X.Br (!Ia'1S diasl.), Y",H 2 (>90%)

;:#~ DfW' H

~ ~ H OH

6 AcO ;: c"Xl A AcO ; H H

A,= R,=H; CH3 0 (40· 50% conversion) A, = iPr. R,= H; R, = Et. A, = H: NB. When carbonyl at C-17 position.

A. = OAc. A. = H. no reaction observed,

21

Ref.

75

75

75

75

76

n

Page 37: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Substrate Dioxirane31 Method" Products Ref.

No. Yield (%)

# ~" H

7 CH')<j A n ~ h H H

AcO ;- CH, 0 AcO i H H

A, = Me, R,l = H; (40% conversion)

mJS I) CH,)<I ~A' 8 B 78

CH:> 0

IQ H,SO •. CH,C" MoO MoO

~'~~ laYVCO''''' 0"

I I ~ :::><1 79 9 I I ~ J ~ A ]' AcO '-1: H H

A/B ring system contains Ar£) OH (35%)

the only as ring junction. NB. Retention at confiQuration

10 ~' c')<j A ~= 7.

H ~2 CH; 0 Ar£)

AoO A, A,

R, = OH. R, = H (40%); R,=R,=H. R, = H. R, = OH (11%).

~ , '(

_ J"ii';i;~ A ' (30%) 80 11

C 11J 1 'J" V; CH>d

~ CH, 0 , .,(

X. in CHllCI~

'. i (50%) 17~.21 ~·isomer

~ ~~I 'hl ':f 22 ,1.}lJ'" H

12 (11 ," c)<j A 80

~)i X. CH:! 0

Other isomers oxidised in CH2CI2 17a,21~·: (10%). separately: 17a,21!l-:

17~.21CL·: (15%). 17~.21CL·; 17a,21CL' 17a.21 CL' : rStartlno matenai).

d0-< ~H 13 CF,xl A ,& ,~,

60 0"'1 H

O( CH, 0

,?,~

AO .. ·(Jb 3.0 eqivalents

0 In CH2CI,

AO'

ReI: Table 1. Entru No. 34

22

Page 38: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Substrate Dioxirane31 Method" Products Ref. No. Yield (%)

JtR

CH')<I [~R] -H,O ~R 14 CH3 0 A - 81

":,I;' OH

CF')<I (OMO: 52 • 99% conversion [2 - 70hrs]; CH3 0 TFMO: 96 - >98% conversion [10 - 15 min.]).

15 H cYI [JiR] 'H,O R~R 8' CH:; 0 A -R A or OH

H o~ CF')<I (OMD: 50 - >96% conversion [4 - 4Bhrs]; CH3 0 TFMD: 92·95% conversion [40·100 min.]).

OH

OH &0 16 &OH

8'

CH><f A (80·90%)

CH3 0 The corresponding 1.3-d/ol also gave the 1 ,3-Acyclic 1 .2,. 1.3- and 1.4· 1.5 sq. ketoalcohol (92%) . Even with excess DMD, no

dials also looked at. Similar apprechiable formation of the diketone results observed. observed. However, oxidation of the 1,4-<:1101

gave the 1 ,4-ketoa/cohol (53%) and the diketone /37O/~)'

cI1 ::;<1 ° 17 A en: 83

1.5 sq. (85%)

~ OH OH

18 .,' 9H

.... ·OSitBuMe, CH><f A >Q ..... OSitBuMe2 84

R R' , , CH3 0 ~ :::.

0;< 6->< R = H, OH 85 - 88%

)<~~ )<~ 19 OA CH,xl A 84

0

O-\: CH, 0

R = -CH,Ph; -CH,-p-BrPh; o-\: -CH,p-CNPh;

-CH -2-Naohlhvl.

~ H~ 20 >( OCH,Ph CH')<j A 84

HO CH,Ph OH CH, 0 OH

90%

21 H'~PH CHXi A H:,~PH as

OH CH, 0

23

Page 39: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Substrate Dioxirane31 Method32 Products No. Yield (%)

Ref.

OH ffc~'dC~OH 22 xdCH,

c~xl A 86

CH, 0 (97%) "--/ (2 - 3%)

X = OMe; Me; H; F; 1.5 eq.

~>! Cl; Br, CN.

c~ 0

23 (A!t) :::xI A (A!p c~ Vf:!{:; 87

OH OH

24 t} :::xI A N°+MoH 87

2.2eq. (29%) (21%) i) 3 eq. LOA.

~~soX2: ~o~~ 25 5eq. TMSCI.

A H~ THF. -78-

88

25"c. TBOMSO' IQ CH,C~

RC COzMe 'aTMS MeCz

~'C02M8

:::xI 2.3 1 H (R=t :tratioH:TMS)

i) HBF. aq.l

C~CN

HO~~fn'N1i:2 ~C~)-;;'NH:! ii) CH,x'l

26 CF) 0 A 89

/CH,CI2 (95·99%)

Ketoneiree TFMO In keton ... free solution Is stable towards

iii) Na2COy strong acid. This allows oxyfunctlonalisation of the side-chain. provided pro1onation pro1ec1s

CH,Cl2 the amine.

i) HBF. aq./

C~CN CH3CON~

~C~C'NH:! ii) CH,x'l

( H2ln'N~ 27 CF, 0 A 89

/ CH:!Cl2 The carbocation intermediate, produced from

In this case, a brlger reaction Ketoneiree the tertiary alcohol in strong acid medium, is

time is allocated for the iii) Na2CO" trapped by acetonitrile (Ritter reaction) to give

reaction with the TFMO. the above product (96 -97%).

CH,Cl2

i) HBF. aq.l

[R~~]'~~ C~CN

ii) CH,x'l

28 R~N~ CF, 0 A 89

/ CH,Cl2 (90 - 95%)

o atom insertion occurs exclusivefy at the &-

Ketoneiree methylene group. On permethylation of the

iii) Na2COy ammonium nm-ogen centre (RNMe,), l;-oxidation becomes competaUve and no

CH,CI2 cycUsation products observed.

24

Page 40: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Substrate Dioxirane31 Method32 Products No. Yield ('Ye)

H :>1 #. 0 • R'(H) ~~'(H) 29 o • or A

~o ::xI OH

OCOCF3

30 R/'oo..R' ::xI A RAR'

R = alkyl, aryt PrIrnaIy and secondary C-H bonds were R' = H, alkyl, aryt (CF,CO),o selectively monohydroxylated when carried out

In presence of 10 fold excess of lriftuomacetlc OCM,O'C anhydride. Over oxidation to carboxytic acids or

ketones not observed.

1.3.7 Oxidation of Organic Sulfur, Silicon, Selenium and Nitrogen containing Substrates

Primary amines are oxidised by dioxiranes, rapidly and often quantitatively,

to their corresponding nitro compounds. This usually occurs via a stepwise

process, as outlined in Scheme 10.92,93 This process is supported by the fact

that both hydroxylamines and nitroso compounds are oxidised to their nitro

derivatives. It should be noted, however, that amine hydrochloride salts are

oxidised directly to their nitro compound. 94 In general secondary amines are

converted to the corresponding hydroxylamine and tertiary amines to the

amine oxides.92 In both these cases, reaction conditions are critical if the

hydroxylamine or amine oxide is to be the desired product. It should however

be noted, that in the case of tertiary amines, excess dioxirane leads to

deoxygenation of the amine oxide and regeneration of the amine'92

Scheme 10: Oxidation of primary amines.

RNH2 + 1e ..

1e -

OH R-N

/

'H

R-NO

OH , R-N

\

~20 OH

25

Ref.

90

91

Page 41: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Among the sulfur nucleophiles, sulfides are oxidised by dioxiranes to

sulfoxides and these, in turn, to sulfones (Scheme 11). The reaction can,

however, be controlled at the sulfoxide stage.95 The yields are practically

quantititative, and the oxidation of the sulfide is so rapid and complete that

sulfides such as p-tolylmethyl sulfide can be used for quantitative analysis of

the dioxirane content in solutions or to quench dioxirane during oxidations of

less reactive substrates.96

Scheme 11: Oxidation of sulfides.

[0] [0]

Dimethyldioxirane (9) is also a convenient oxidant for converting silanes to

silanols by Si - H insertion (Scheme 12)?2C When optically active

trialkylsilanes are used, the Si - H insertion proceed with complete retention

of configuration. 97 Hydroxylating silanes under these strictly neutral

conditions means that siloxanes are not produced.

Scheme 12: Oxidation of silanes.

[0]

1.3.8 Table 1: Oxidation of Organic Sulfur, Silicon, Selenium and

Nitrogen containing Substrates

Entry Substrate Dioxirane31 Method32 Products No. Yield (0/0)

0 0 0

1 c6:: ::t<! A c6:> c6:: , , H in CH2CI2

H (45%) (40%)

0 0 0

::t<! c6::. WH 2 c6:: A , , ,

InH,O H H H - (70%) (21%)

26

Ref.

98

98

Page 42: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Substrate Dioxirane" Method"

Products Ref.

No. Yield (%) 0

CH')<j 0

3 cCx: A cxXB 98 CH, 0 ,

" in EtOH (77%)

0 CH')<j 0 0

4 ~&R' A ~~;( ":((' sA CH, 0 9. •

I '" in CH2CI2

o I R, H H

(32·37%) (40·55%)

)) CH'Xl +s-O 5 A 98 CH3 0

I 2 I H in CH,CI, H m'lo)

CH'XY Ph

Phse~ \ 48 6 A ,se"---<, OCONH2 CH, 0 o CONH2

Me I

CH')<j

CS? CH, 6 Me

7 A I 99 cx;:o + ArCHO

(2.0eq) + hv

A r = Phenyl, 2·Naphthyl. Ne. 0 Chemiluminescence noted in this reaction.

8 Q ~:t<! A 0 - C).o '00

I I H O· OH

CH,OAc CHXl CH,OAc

9 ACOC~CH'N(ll'r), A AcOC~COOH 101

AcOCH, CH, 0

AcOCH, CH,OAc -4 equivalents CH20Ac

10 Ar2CN

2 CH~ A Ar2CO 102

CH 0

11 6 :::x! A Q (95%) + 6> (5%)

102

~ C~CI:!

0 CHXi

0

12 11 A HO . 11

'0' N,~P(OEt)2 >-P(OEl" (100%)

CH3 0 HO

R'~ o 0

13 CHXl A R('~~R2 ,o4 N, CH:] ! O.(H2O) (89 - 100%)

Rp R, = aliphatic or aromatic

27

Page 43: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry No.

14

15

16

17

18

19

20

21

22

Substrate

R1• R2 = aliphatic or aromatic

NOH

R1AH R1 = aliphatic or aromatic

7HX R~N

o X = Protectina Groun.

/~~1

CbZ~o,CH,

! Cbz~O'CH,

Dioxirane31

or CH,,~

CF,AJ

Method32

A

A

A

A

A

A

A

B

A

Products Yield (0/0)

~ (85-100%) R, R2

Starting material only

Ph'6 (58%)

[ ~(~ -Q ~.,O" ()--2w \\ 1 N

0" (85%)

[c:lsh]- J--~ c",)(", rH, CH,cc

cH{ '0- (66%)

NB. For easily oxidised hydrazines single electron transfer (This example) is more

rapid than the usual oxygen transfer (Table 1 Entry 1 B) shown by DMD.

t-IHX 0 NHX

R~H'-'Ry~: o 0

Predominantly in hydrated form.

28

Ref.

'04

'05

106

107

'08

108

109

110

110

Page 44: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

-

Entry Substrate Dioxirane31 Method32 Products Ref.

No. Yield (%)

C;q

23 ~(

CH; A ~/ -0 III or \

CH,.><l

CF, J C;q

~fBF3 CH; . ~tl-BF3 111 24 A or o \ CH,.><l CF, J

OC~ Ph ~O'H

25

0\ ::;<I B 112 Co,H

M , 6cH, ~ CO;zH

I OcH,

OR OR

><~' (15~~.

OR OA

26 "~ CH:xI A "ro!,~,.;..oH (33%) 1 13 ~~ CH) I '\ 1 R • SlMe2""

0, OROA ~

'r< . N"' 1 ',.I., ~b. A R (8%)

HU . [)( A

y ? (S) 27 ::;<I HJ=( (100%) o ,

N~ ~ 114 100% e,e.

Y = Br R = CO,CH,C,H, 0 Y = Cl R = CO,CH,C,H, R Y = F R = CO,CH,C,H, Y= F R=

CO,CH,OCOCICH,\,

'( X:b::X o ,

::;<I 28 fl 'A 0 '14

X = BrY = H R = CO,CH,Ph j& (100%) X = F Y = H R = CO,CH,Ph X = BrY = F R = CO,CH,Ph 100% e e X = BrY = Br R = CO,CH,Ph o I X = Br Y = Br R = CO,CH(PhJ, R X= BrY= Br R =CO,H X = BrY = Br R = CH,OH X=BrY=BrR=

CO,CH,OCOCMe,

.

29

Page 45: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Dioxirane31 Method32 Products Ref.

Entry Substrate Yield (%) No.

R'~ ~ A R2"'" . R, .. ' 0

CH)<j s-s 115 29

0' R2"'" CHa 0 S-S in CH2CI2 Up to 18: 1 diastereoselectivity observed.

Major isomer always trans.

+'9 CH)<j A +'9 116

30 CH3 0

inCH,C~ CH3-Si-CH3 CH3-Si-CH,

1 1 H OH

~'R 0 CH)<j 11 X ""- A ~S'R 117

31 CH3 0

X = H, R = CH,CO,H; Cl-2'C X = CO,H, R = Mo;

0.65 eq. (47·99 %)

X = H, R = CH,CH,OH; Smin. X = CH,OH, R = Mo;

X = H, R = ·CH=CHCO,H.

0 S'R o 0

117 CH)<j \\11 X ""- A

DS'R 32

CHa 0 X = H, R = CH,CO,H;

20'C X (35 _ 97 %)

X = CO,H, R = Mo; 1.35 eq. X = H, R = CH,CH,OH; 1-2 hrs. X = CH,OH, R = Mo;

X = H R = -CH=CHCO H. ..

[+t~-<] ! r.t.

+sJ-< CH)<j A 0

+~-s-< 1 la 33

+s-w-< CH, 0

o 53% o 30%

>-s-~-< +I-s+ o 10% 5%

X X

2 2 119 C:Xl

A and B 34 CHa 0

I' 1 N-..;-~ I' 1

0-N........"N

X - H, Me, MeO, Cl Authors suggest SN2 mechanism

30

Page 46: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

1.3.9 Oxidation Of Metal Enolates and Metal Complexes

Over the past five years the oxidation of organic substrates by dioxiranes

has been extensively investigated,22 but relatively little published on

organometallic substrates. Adam et a/have looked at the oxidation of metal

enolates. They found that silyl enol ethers afford the corresponding

epoxides120 whereas the enolates of lithium, sodium and titanium yield the «­

hydroxy carbonyl products. 121

Lluch et aJ 22 have shown that dimethyldioxirane provides an efficient

reagent for the oxidative decomplexation of Fischer carbene complexes

(Scheme 13). The reaction affords good conversion yields, is clean and

highly chemoselective. In all cases comparable or better yields of the ketone

were obtained than with conventional methods.

Scheme 13: Decomplexation of Fischer carbene complexes.

[0) -1.3.10 Table 4: Oxidation Of Metal Enolates and Metal Complexes

Entry Substrate Dioxirane31 Method32 Products No. Yield (%)

~, 0

Ph~Me 1 Ph O-T~A' A

Ph~M8 + ?- I . Me 0 ... '~ OH OH

Ar·~ C"Xl (A) (20.67 %) (S) .. "'

A, CH3 0 (5·63%ee)

Major isomer usually derived from reside attack

2 J:: A A, , A 'A,Y-yA, CH)<'f OH

ML, = Ti(OrPr),; Ti(NEt,),; CHi 0 Paper looks at diastereoselective

TiCP2C1; Na; SiM 8,. hydroxylation of chiral metal enolates.

31

Ref.

123

124

Page 47: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Substrate Dioxirane31 Method32 Products Ref. No. Yield (%)

(CO)3Cr R (CObCr . R ~R: 3 ~~ ~:><1

A 125

~R, SR, -0

Excellent yields and diastereoselectivity observed.

0 RS:Q 4 RS:Q

c"xl A \ I (-45%, -90% d.e.) 126

\1 tBu ,

ISu """ \ cHi 0 Cr(CO)3

Cr(CO), Diastereoselectivity of onho substituted complexes determined and found to be

dramatically reversed when changed from methyl to tert-butvl.

tr(CO)' H Cr(CO),

5 ~"Me c~ A ~ 127

- s NMe2 CHi E (2.4 eq.)

E

Enantiomericallv nure

6 A A ~oe 128 R2 0 Cr(CO), CH><i

R, = H: o-CH,OH: CH:! 0

m-CH,OH: p-CH,OH: 20°C OCH,CH=CH,.

OMe (CO)5Cr==< CH)<j OMe

7 R2 A O=< 128

CHi 0 R2 = H; o-CH2OH; _20°C R2

m-CH,OH: p-CH,OH: OCH,CH=CH,.

8

Me 0 J;:;<:;:R2

CH,xl A 128 (CO)3CO--CO (CO), CH, J MeO R2

R2 = H; o-CH2OH; _20°C m-CH,OH: p-CH,OH:

OCH CH=CH .

(COI3FeD"~ Cl<! -(J"'~ 9 CH 0 A (CO),Fe 129

(3 equivalents) (41% yield, 59% conversion) 56'C

Some decomposition of the iron tricarbonyl complex was noted.

OEt

::«! O=<Ph

10 (CO)5Cr~Ph

A 130 NEt2

NEt2 (3 equivalents) (52%) 2O'C Major pathway involves oxidation of the

enamino double bond.

32

Page 48: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

1.4 Other Aspects of Dioxirane Reactivity

1.4.1 Effect of Solvent on Reactions

In 1993, Murray and Gu 131 reported that the rate of the dimethyldioxirane

epoxidation reaction is dependent on the nature of the second solvent, when

binary solvents containing 50% acetone are used for the reaction. The rate of

epoxidation was shown to be enhanced by hydrogen-bond donor solvents

and inhibited by hydrogen-bond acceptor solvents. More recently they have

shown that the dimethyldioxirane C-H insertion reaction is also favoured by

solvents with greater hydrogen-bond donor capacitY,132 although the effect is

less pronounced. Both the epoxidation 131 and C-H insertion 132 reactions were

found to follow second order kinetics, using solvents in which it could be

shown that there were no competing first or second order processes.

As previously postulated for the epoxidation reaction (16),131 these workers

believed a spirotransition state (17), for the C-H insertion reaction, 132 was

most consistent with their results.

Figure 5:

Spiro transition state for epoxidation and C-H insertion reactions.131.132.22e

(16) (H-S = Solvent) (17)

Murray has continued to study the effect of solvent on dimethyldioxira~e

(9) epoxidation and C-H insertion reactions. In a recent publication he

describes the effect of solvent on the diastereoselectivity of the oxidation. l33

He found that the diastereoselectivity of the epoxidation of cyclohex-2-en-1-01

(18) was essentially tuneable depending on solvent choice. For example,

methanol! acetone (90 !1 0) gave an epoxide distribution of 2 : 1 in favour of

33

Page 49: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

the (rans epoxide; whereas tetrachloromethane f acetone (95 f 5) gave

almost exclusively the cis epoxide. From his results he suggested that the -

OH group of (18) exerts a H-bonding effect favouring cis selectivity. Hence in

the presence of tetrachloromethane, the association between the dioxirane

and acetone would be diluted such that the intramolecular H-bonding ef:ect

was more competitive. The distribution of diastereomers in the absence of

this effect was appar.ently determined largely by steric effects. The epoxide:

enone ratio was also found to be solvent dependent.

cS (18)

1.4.2 Electrophilic character I Mode of Oxygen Transfer

While most oxygen transfer reactions of dioxiranes appear to take place by

direct substitution (SN2) on the dioxirane peroxide bond by the nucleophile,22

more complex electron transfer activity has been implicated, particularly for

substrates with low oxygen potentials. 134 Adam and Golsch 135 have,

therefore, compared the relative rates of the dimethyldioxirane oxidation of

nitrogen heteroarenes with those of methylation by methyl iodide. Their

results revealed that an SN2 mechanism rather than an electron transfer

applied for the dimethyldioxirane oxidation.

The nucleophilic (oxidation at the sulfoxide Site, i.e. SO in SSO) versus

electrophilic (oxidation at the sulfide site, i.e. Sin SSO) nature of oxidants has

previously been examined using thianthrene 5-oxide (SSO) as a mechanistic

probe. 136 The results obtained by Adam and Golsch,137 using this reagent to

assess the electrophilic character of dioxiranes, at first seemed to suggest a

dominant nucleophilic character and therefore were in contrast with the

established electrophilic character of the oxidant. 138 These results were

explained in terms of the energy requirements of the oxidations. The

activation barrier for sulfide oxidation (24.4 kcallmol) was significantly higher

34

Page 50: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

than for sulfoxide oxidation (9.6 kcallmol).138 It was also pointed out that

these were gas phase results and that the sulfide transition state has a higher

dipole moment than the sulfoxide transition state and consequently should

experience the greater stabilisation on transfer to a polar solvent. In his

review, Murray22a suggested that oxidation of sulfoxide may be initiated by

attack of dimethyldioxirane on the oxygen rather than at the sulfur (Scheme

14). However, further work by Clennan and Yang 139 showed this mechanism

(Scheme 14) to be incorrect.

Scheme 14:

o 11

(XI SY') """ S~

0-0 ..... I

(XI S+~ """ S~

Theelectrophilic nature of dimethyldioxirane, and also

(trifluoromethyl)methyldioxirane, have been examined by Ballistreri et a[14O

They, like others, looked at the relative reactivity of thioethers and sulfoxides,

and also the competitive oxidation of sulfoxides, towards oxygen-transfer by

these reagents. They concluded, as previous workers have done, that

dimethyldioxirane and (trifluoromethyl)methyldioxirane are definitely

electrophiles, the latter being less selective than the former, and the mixtures

of products obtained in the reactions, were due to the small reactivity ratio of

the sulfur centres.

1.4.3 Decomposition of Dioxiranes

Hull and Budhai142 have studied the decomposition of dimethyldioxirane

using u.v./vis. spectroscopy at 340 nm. They found that at high

concentrations (-0.1 M), decomposition had an induction period followed by

rapid decay, the latter being due to the presence of reaction products. The

phenomenon has previously been observed by Adam and Murray.143

However, Hull and Budhai also reported that at low concentrations (-0.02 M)

35

Page 51: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

decomposition was first order with an activation energy of 24.9 kcal/mol and

that its apparent thermal stability is due to the very strain energy that would,

at first, seem to make the species so unstable.

Dioxiranes have also been shown to be decomposed by ethers (cf.

Preparation of Dioxiranes, p 4).15 Photochemical and thermal initiated free­

radical activity has previously been noted in the decompositon of

dioxiranes. 141• A free-radical pathway has recently been reported for the

oxyfunctionalization of adamantane, triggered by the presence of CCbBr or

the presence of an argon blanket. 141b

1.5 Carbonyl Oxides and Dioxiranes

The recent development of dioxiranes as potent oxygen transfer agents has

stimulated much interest in the similarities, and differences, between

dioxiranes and their isomers, the carbonyl oXides. 144,145 It is now well

established that dioxirane (1) is more stable than its dioxygen ylide and the

respective barriers for their inter-conversion are sufficiently high that each

exhibits its own chemical behaviour when independently generated. Although

for many carbonyl oxides, dioxirane formation seems to be ruled out,

cyclisation of carbonyl oxides to dioxiranes have been reported. 144

Scheme 15: Proposed formation of dioxirane in gas phase ozonolysis of tetramethylethylene.

-+

In a recent publication, Murray et a/146 have suggested the involvement of a

dioxirane species in the ozonolysis of tetramethylethylene. This they have

36

Page 52: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

done in order to explain some of the observed products of its gas phase

ozonolysis (Scheme 15).

Kopecky et al, 147 in 1993, published the first example of dioxirane formation

from ozonolysis in solution. They found that addition of excess ozone to E- or

Z-1,2-dimethoxy-1,2-diphenylethene, in inert solvents, resulted in formation of

up to 0.7 mole of (methoxy)phenyldioxirane.

Further work in this field has shown that isomerisation is facilitated by the

presence of It-donors on the phenyl substituent of the carbonyl oxide

(Scheme 16).148 It has also been suggested that under some conditions, the

carbonyl oxide thus substituted could mainly isomerise into the dioxirane, the

extent of isomerisation being related to concentration and solvent polarity.

H R1 , /

C=C / , Ar-C--O C02Me

1,-o

Scheme 16: Proposed cyclisation of carbonyl oxide when Ar substituted with an electron-donating group.

Scheme 17: Cyclisation of carbonyl oxide to dioxirane during irradiation of decafluorodiphenylcarbene in an oxygen matrix.

o· 0/ F

F

F F 35 K F F

F F F F F F

F F F F

F

1 /... > 480 nm

F 0-0 F

F ,:? F

F ~ F F F

0 XXF F I A> 350 nm

F o ~ F ~

F F F

F

37

Page 53: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

A further case in which cyclisation of carbonyl oxides to dioxiranes has

been observed, was during the photolysis of carbenes in 02-doped argon

matrices. 149,150 Both dioxiranes and carbonyl oxides have been characterised

by u. v.lvis. and i.r. spectroscopy during the course of this reaction, though as

yet, neither have been isolated (Scheme 17149 and Scheme 18150).

Scheme 18: Cyclisation of carbonyl oxide to dioxirane during irradiation of propinal O-oxide in an oxygen doped argon matrix.

o· '0+

H ( N /"

H

~ q-o ( 480nm H • H .-, t. 35 K hv 590 nm H '< H 'N2 H +02

H

~ 0+-0'/

H I( H

Cremer et ar1 have investigated the equilibrium geometries, dipole

moments, infrared spectra, heats of formation and isomerisation energies of

carbonyl oxides and dioxiranes.

They conclude that:

i) the electronic structure of the carbonyl oxide is closer to that of a

zwitterion than has previously been predicted by ab initio calculations.

ii) identification of both carbonyl oxides and dioxiranes in a reaction

mixture is possible with the help of i.r. spectroscopy. The CO and 00

stretching modes of carbonyl oxides should be very intense and

appear at 1200 and 880 cm·1.

o iii) The heats of formation LlH, (298) for carbonyl oxide and dioxirane

are 30.2 and 6.0 kcall mol.

38

Page 54: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

iv) Decomposition of carbonyl oxide and dioxirane to CO, CO2, H20,

and H2 is thermodynamically favourable, while fragmentation to CH2,

O2 or ° is not.

39

Page 55: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 2

Page 56: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 2

2.1 Introduction to the Project

In the last decade or so, dioxiranes (1) have been shown to be extremely

versatile oxidants.22 However, very few homochiral dioxiranes have been

used and only low enantioselectivity has so far been reported. 10 In the course

of our research with dioxiranes,152 a new group of chiral dioxiranes generated

in situ from 1-tetralones (19) and 1-indanones (20) substituted at the C-2

position with a fluorine and either an alkoxycarbonyl or a 2-hydroxyisopropyl

group, have been looked at. To some extent these may be regarded as the

cyclic equivalents of trifluoroacetophenone and related ring fluorinated

acetophenones, which have been shown to generate a reactive dioxirane in

situ. 153,154

roe. R'

""'" . I F %.

(19)

R3 ° ~C02R2 R3~F

(20)

Rl = C02CH3, C(CH3)20H, (-)-Menthyl

R2 = CH2CH"CH,

R3 = H; F

It was thought that the planar aromatic ring and substituents, a large group

(L) and a small group (S), at the C-2 position might be important in directing

the approach of a trans-alkene to the dioxirane, which is known 152. to react in

a spiro transition state, It was initially assumed that the larger group would

take up the usually preferred equatorial conformation in the half-chair, and

therefore some selectivity of the oxygen to be delivered should be observed,

As previously described in the Introduction (p, 12, Figure 4). there are four

possible diastereomeric transition states for the epoxidation of an

unsymmetrical alkene using a homochiral ketone, The two transition states,

believed to be favoured for the epoxidation of trans-stilbene (23) using the

dioxirane derivative of a homochiral tetralone (19), are illustrated below (21)

and (22). A highly electronegative fluorine was incorporated, as the small

group (S), [j,- to the dioxirane ring system, to increase the electrophilicity of

40

Page 57: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

the oxygen to be delivered.

s

(21) (22)

It was thought that this L conformation would be

P h less energetically , favourable because of ~ the steric interaction of

the phenyl group and the large group on C-2 of the tetralone.

An alkoxycarbonyl group was introduced at position 2 of the tetralone, as

the large group (L). Since these groups are slightly electron-withdrawing, it

was anticipated that they would also help increase the electrophilicity of the

oxygen to be delivered.

Cl

(23) (24)

o (25)

Me

Me

The synthesis and separation of the diastereomers and enantiomers of

some 2-substituted-2-fluoro-1-tetralones (19) and 1-indanones (20) was

therefore undertaken. The investigation of in situdioxirane generation from

the homochiral parent ketones and their potential in enantioselective alkene

oxidation was also conducted. The alkenes used, were trans-stilbene (23),

trans-~-methylstyrene (24) and 6-chloro-2,2-dimethyl-2H-1-benzopyran (25).

2.2 Synthesis of the 1-tetralone (19) and 1-indanone (20) derivatives

The synthesis of enantiomerically pure 2-hydroxy-2-substituted-1-

tetralones 155 using oxidation of the enolates with enantiomerically pure

(camphorylsulfonyl)oxaziridines (Scheme 19), and the formation of

homochiral 2,2-disubstituted-1-tetralols 156 via Grignard reactions with optically

active (1-tetralone)-tricarbonyl chromium derivatives have been reported

(Scheme 20).

41

Page 58: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Scheme 19

Scheme 20

~A i) Base

A'U) A = Me, Et, PhCH2

X=CH2,0

ii) ~~~I ~'o 2

"w~" (5) (90 - 95%ee)

hv •

Recently, the reduction of racemic 2-alkoxycarbonyl-1-tetralones by chiral

ruthenium(ll) catalysts has been realised by kinetic resolution (Scheme

21).157

Scheme 21

o OH OH ruCOEt .7,. . 2 L'AuX

~o I ~ 5 "

W ",·.c02Et

MeO I or

" -EX)

' C02Et

MeO I "'" (S,A) (R,S)

7-MeO (R)-MeO-BiphepAuBr2 d.e. = 97% (trans) e.e. = 95% (S,A)

5-MeO (S)-BinapAuBr2 d.e_=97%(trans) e.e.=92%(A,S)

No work, however, has been reported which describes the separation of

racemic mixtures of 2,2-disubstituted-1-tetralones_ The following sections

describe the preparation and separation of the diastereomers of menthyl 2-

fluoro-1-tetralone-2-carboxylate (31 c) by chromatography, and the

preparation and separation of enantiomers of 2,2-disubstituted-1-tetralones

(19) and -1-indanones (20) via synthesis of their (RJ-(+)-a.-methylbenzylimine

(32 and 38) derivatives.

2.2.1 Synthesis of the 2,2-disubstituted-1-tetralones (19)

Several methods have been reported for the synthesis of 2-alkoxycarbonyl­

Hetralones and 1-indanones. These include the use of Mander's reagent 158

(methyl cyanoformate) and Stork 159 methodology, both of which are reported

42

Page 59: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

to alleviate the frequently encountered problem of O-acylation during

attempted C-acylation of ketone enolates. 160

Synthesis of ethyI1-tetralone-2-carboxylate (26)

Ethyl 1-tetralone-2-carboxylate (26) was the first of these compounds to be

prepared. Two methods were used as indicated (Scheme 22 and Scheme

23); the latter proved simpler and gave higher yields. I.r. and n.m.r.

spectrometry data were in agreement with those in the literature for this

compound. 161 ,162

Scheme 22

Scheme 23

i) LDA, HMPA

EtC02CN *

THF

o OH

ro~ I C02~E=t="" 0& ~ _ "" C02Et

h : I (51%)

(26)

i) NaH,(EtOhCOroO CO Et OH

I1HF -:? 2 0&"" CO2 Et -~:....--. I oo====='" I iQ t-ft ~ h '" (90%)

(26)

A general mechanism for the base promoted reactions involved in

Schemes 22 and 23 is given in Scheme 24.

Scheme 24

rod R

I +

'"

R = EtC02, MeC02 and X = CN, EtO, MeO

43

Page 60: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Synthesis of methyl 1- tetralone-2-carboxylate (27)

Since the method in Scheme 23, for the synthesis of ethyl 1-tetralone-2-

carboxylate, gave the higher yield, its methyl analogue was prepared similarly

as shown in Scheme 25. I.r. and n.m.r. spectroscopic data agree with resul;;

previously obtained within this Oepartment.161

Scheme 25

o

06 i) NaH, (MeO)2COoerO

rfl-iF ,:7 C02Me ----------~ I iQH+ ~

(27)

Synthesis of menthyI1-tetralone-2-carboxylate (28)

Oimenthyl carbonate is not commercially available, so the synthesis of

menthyl 1-tetralone-2-carboxylate was attempted using menthyl

chloroformate with sodium hydride in THF (Scheme 26). This method proved

unsuccessful.

Scheme 26

0-t i [ o i)NaH, l 0 C(Xf0 ex OH 0 X·

cD A 17 0"" CJC{0"" I X' I I THF ~

(28)

The reaction was then carried out using LOA, a stronger base, in THF, in

the presence of DMPU (Scheme 27). L-Menthyl chloroformate was again

used as the source of the ester group. Although the reaction proceeded well,

only a 40% yield of pure product was obtained because the product proved to

be difficult to purify by chromatography owing to the presence of impurities

having Rt values close to that of the product. There were two major

impurities, a carbamate (29) and the O-acylated product (30). The carbamate

44

Page 61: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

(29) was produced when LOA, still present in the reaction mixture on the

addition of the menthyl chloroform ate, reacted with the menthyl

chloroformate.

Scheme 27

~I 0 i)_~~;THF, • ~o",·6 ~ " 1: 10 , .. 6 ~ ii)DMPU,-78"C U0 ~ OJ' ~

'iii) ~ -78o

'C U (28) (40%)

, O~I A iv)H+

(29) (30)

This reaction gave a comparable yield when using HMPA rather than

OMPU, but still with minor impurities present. OMPU was preferred because

of the high toxicity of the HMPA. HMPA has been shown to encourage the

formation of O-acylated products over C-acylated products. In order to

produce the C-acylated product only, the reaction was attempted in the

absence of both OMPU and HMPA. However no observable C- or O-acylated

products were obtained.

Because of the difficulty in purifying the menthyl 1-tetralone-2-carboxylate

in the above reaction (Scheme 27), alternative methods of its preparation

were looked at.

With butyllithium in THF, and menthyl chloroform ate (Scheme 28). a 15%

yield of the desired product was obtained. No O-acylated product was

formed, hence the purification was comparatively easy. A similar experiment

was carried out using methyllithium in place of the butyllithium, but as with

butyllithium, the yield was low (10%).

45

Page 62: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Scheme 28 i :

(28) (15°/.l

A

Increasing the strength of the base did not produce sufficiently high yields.

A further alternative approach to produce the menthyl esters, via the enamine

derivatives of 1-tetralone (Scheme 29 and Scheme 30), was attempted.

Previous work, carried out by Stork and colleagues,163 had shown that ~­

ketoesters can be made by enamine acylation. These workers also found

that two equivalents of the enamine must be used, to one equivalent of the

acyl chloride in the formation of the ketoester. Methods using triethylamine to

mop up the acid produced as a by-product of the reaction, in place of the

second equivalent of enamine, proved unsuccessful.

Scheme 29

o i)A 0

o N ~)...CI oox QEt20. ~ ~ A ~I 0"" -_. ~ X 'l.vV ii) TiCI. "'- Solvent. Reflux

(77%) ii) Ho0· 11 (28) = Solvents used: Benzene ~ §

Et20 OH 0 (') ro .. ··~ Preparation of the enamines was carried out using a method devised by

White and Weingarten. 164 As shown in Scheme 29, approximately half an

equivalent of titanium tetrachloride was used to three equivalents of the

amine and one of the tetralone. The use of one equivalent of the titanium

tetrachloride and three equivalents of the amine to one equivalent of the

tetralone was also used as shown in Scheme 30. I.r and n.m.r. data were

found to be consistent with the formation of the enamines. However, as

46

Page 63: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

shown in Schemes 29 and Scheme 30, all attempts to acylate the enamines

failed.

Scheme 30

o C) C) ex) i) Et2 0. H, 001

"'"

ii) TiCI. %.

. (68%)

0~ i) :oe, ~O

-A'----'-\X!---· WO'" A ~ (28) j

Benzene, Reflux

ocYox In General and as indicated in Schemes 23, 25 and 27

Both tautomers were clearly visible in the 1H n.m.r. spectrum of the 2-

alkoxycarbonyl-1-tetralones (26), (27) and (28); the OH of the enol appearing

at approximately OH 12.5 (singlet), and the proton at the C-2 position of the

keto form at OH 3.6 (double doublet). The signals from the proton at the C-8

position of the aromatic ring were also clearly distinguishable; OH 8.0 (double

doublet) for the keto form, and OH 7.8 (double doublet) for the enol form.

These protons appeared as double doublets due to coupling with the protons

on C-7 and C-6 of the aromatic ring.

2.2.2 Fluorination of the alkyl 1-tetralone-2-carboxylates (26, 27 and 28)

In order to fluorinate, the acylated compounds, the electrophilic fluorinating

agent, N-chloromethyl-N-fluoro ethylenediamine bis(tetrafluoroborate), was

used. This compound, was supplied by Air Products as a white solid known

as Selectfluor. Previous work has shown it to be useful for the fluorination of

metal enolates. Following a method published by Banks,165 sodium hydride

was used to abstract the proton at the C-2 position of the tetralone with DMF

as the solvent. On generation of the anion, the Selectfluor reagent was added

47

Page 64: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

as a solution in DMF. The general mechanism by which this reaction occurs,

is shown in Scheme 31. All fluorination reactions resulted in yields of over

71 %. The lH n.m.r. spectra of these products no longer showed evidence of

the enol form. No signal was observed for the OH proton of the enol ester (8H

12.5) or the proton at the C-2 position of the ketoester starting material (8H

3.6). Only one signal was observed for the proton at the C-B position of the

tetralone at 8H B.1. The signal for the carbon at position 2 in the 13C n.m.r.

spectra, in each case, showed a coupling constant of approximately 200 Hz.

Coupling constants of this magnitude are often observed for C-F bonds,

suggesting that the fluorine is incorporated at this position. l66

Scheme 31

~ ~O/R ____ • ~O/R l.)lJ-~ F ~3~=Et

+ H2 ~()I - SF; ((31 b»= MMe

N 31e = ..,

+Na l' (1)_ + l.!. _ +BF4 b + BF; I + SF,

H,cl CH,ct

2.3 Separation of the diastereomers of menthyl 2-fluoro-1-tetralone-2-

carboxylate (31c)

The diastereomers of (31c) were separated by flash chromatography to

yield the homochiral tetralones. The more polar diastereomer (by tic) was

obtained as a solid. X-ray crystallograph/ 67 revealed it to be (1'R,2'R,5'R)-

(-)-menthyl 25-fluoro-Hetralone-2-carboxylate (31 c-DU) and showed that the

bulky ester residue was in an axial conformation (Figure 6).

48

Page 65: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Figure 6:

X-Ray Crystal Structure of (1'R,2'R,5'R}-(-)-Menthyl 2S­

fluoro-1-tetralone-2-carboxylate (31 c-DIJ)

(31c-DlI)

49

Page 66: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

2.4 Separation of enantiomers of (31 b) and (34) via an imine derivative

Separation of the ester derivative (31 b)

Attempts to synthesise the acetal derivatives of the enantiomers of

(31 b) 168 using the optically active alcohol, (2R,4R)-( -)-pentanediol proved

unsuccessful.'69.170.171 The imine (32), derived from (R)-(+)-a.­

methylbenzylamine, was synthesised using titanium tetrachloride as catalyst

(Scheme 32).172

Scheme 32 MeyPh

o 0 H NH

~o...Me _M_e_.·'~_'··~_· h_'_li_iC_I'., ~I ~ ~0/M_e_2_M_HC_I ... ~ Me

VU'F Benzene UJF DCM W'F 0-

(31 b) (32) (31 b)

There are four possible diastereomers of (32). These are the syn- and anti­

isomers of both the (R,R)- and the (R,S)-imine. On chromatographic

separation of the diastereomers of (32), the syn- and anti- isomers of (R,R)­

(32) eluted together, as did the syn- and anti- isomers of (R,S)-(32). This fact

was determined using 'H n.m.r. spectroscopy, and the chiral shift reagent,

Eu(hfch, on the separated enantiomers of the de protected ketone (31 b).

Figure 7: The four possible isomers of the imine (32).

Me Me Ph Ph

Y-Ph Y-Ph Me~ Me~ N 0 N 0 N '0 N 0 cx)<' ,Me I ' ""F 0 . rolo

/

Me

I S F a)<' ,Me I ' ····F 0 rolo,Me I S F

syn-(R,R) syn-(R,S) an6-(R.S) anti-(R.R)

The syn- and anti- isomers of the less polar diastereomer (32-01) were

isolated as an oil, whereas those of the more polar diastereomer (32-011)

were isolated as a solid.

Only one signal (OH 3.33)was observed in the 'H n.m.r. spectrum for the

OMe protons of the isomers in (32-01). The signals for the OMe protons of

the anti- and syn- isomers of (32-011) appeared in the 'H n.m.r. at oH 3.87 and

50

Page 67: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

3.83. It was noted that these latter syn- and anti- isomers partially

interconverted on recrystallization from methanol.

Hydrolysis of the imine (32-01) and (32-011) yielded the ketones (31 b-EI)

and (31 b-EII), respectively. Chiral shift 1 H n.m.r. spectroscopy studies

showed that (31b-EII) had been isolated in >95% e.e. (no [31 b-EI] apparently

present) and (31 b-EI) in 80% e.e .. The low e.e. in the latter case is due to

incomplete separation (32-01) from (32-011) by the chromatographic method

used, and this was confirmed by the 1H n.m.r. spectrum of the less polar

fraction.

Separation of the tertiary alcohol derivative (34)

Synthesis of the hydroxyketone, 2-fluoro-2-(2'-hydroxypropyl)-1-tetralone

(34), via the imine (32) was attempted with methyllithium and

methylmagnesium iodide using various reaction times and temperatures. The

best yield, albeit low, was obtained using methyllithlum, at O°C and quenching

after 1 - 1.5 hours (Scheme 33). Flash chromatography of the reaction

mixture containing the hydroxyimines (33) yielded the diastereomers (33-01)

and (33-011) and recovered starting material (32). The diastereomers were

deprotected with dilute acid to give the hydroxyketone enantiomers (34-EI)

and (34-EII), respectively. The imines proved fairly stable to the acid, and

required overnight stirring with 2M HCI for complete hydrolysis. Chiral shift 1H

n.m.r. spectroscopy studies showed that (34-EI) had been isolated in >95%

e.e. (2% of [34-EII]) and that (34-EII) had been isolated in 90% e.e ..

Scheme 33

MeyPh

N 0

~o/Me MeU/THF

U0 "F 1.5 hours

(32) Lt.

2MHCI

[)CM w~

(34)

2.5 Synthesis of the ethyl 2-fluoro-1-indanone-2-carboxylate (37)

EthyI2-fluoro-1-indanone-2-carboxylate was synthesised from 1-indanone

51

Page 68: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

with diethyl carbonate and sodium hydride under the same conditions as

used for the synthesis of methyI1-tetralone-2-carboxylate (31b). Fluorination

was carried out using N-fluoro-N'-chloromethyltriethylenediamine

bis(tetrafluoroborate) (Scheme 34).

Scheme 34

o orrC02Et ij NaHIDMF I •

"'" iij ~ (36) ~"\ SF;

~BF' tH,CI'

i) NaH. (EIO),CO

ITHF .. (35) (37)

2.6 Separation of enar.tiomers of (37) via imine derivative (38)

The imine (38), derived from (R)-(+)-u-methylbenzylamine, was obtained

by the method described for (32) (Scheme 35).19

Scheme 35

MeyPh

~N o/Et

:... I * F 0 2MHCI

~ DCM

(38)

The isomers of the imine (38) proved slightly more problematic to separate

than those of the tetralone imine derivatives (32).

Figure 8: The four possible isomer of the imine (38)

Me Me Ph Ph

~Ph ~Ph Me~ Me~ N 0 N 0 N 0 N 0 W/E1 ~~o/EI ~/Et ~~O/Et I ' ·· .. F 0 I S F I ' ····F 0 I S F

"'" "'" "'" syn-(R.R) syn-(R.S) anti-{R.S) anti-(R.R)

The four individual isomers were visible by tic. Two of the isomers could be

crystallised out together from the racemic mixture using ethanol. These were

the fastest running and third fastest running compounds on the tic plate. The

least polar compound was the more abundant of. the two isomers in the solid

52

Page 69: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

(38-01) (ca. 9: 1 by 1H n.m.r. spectroscopy). Removal of the imine group from

these isomers (38-01) yielded the ketone (37-EI), which was shown be one

enantiomer (> 95% e.e.) using 1H n.m.r. spectroscopy, and the chiral shift

reagent, Eu(hfch, The two isomers in the solid (38-01) must therefore have

been the syn- and anti- isomers of either (R,R)-(38) or (R,S)-(38) isomer.

The two remaining syn- and anti- isomers were isolated as an oil (38-011).

This oil contained approximately 20% of (38-01), hence the enantiomeric

excess of the deprotected ketone (37-EII) was not determined using 1H n.m.r.

spectroscopy, and the chiral shift reagent, Eu(hfch.

2.7 Synthesis of methyl 2,5,7 -trifluoro-1-indanone-2-carboxylate (47)

In order to further increase the activity of the dioxirane derivatives of the 2-

acyl-2-fluoro-1-indanone, fluorine was incorporated into the aromatic ring.

Previous work in this Department154 has shown that fluorine positioned ortho­

or para- to the carbonyl substituent increases the oxidising ability of the

dioxirane derivative to a greater extent than if incorporated at a meta­

position. 3,5-Difluorobenzaldehyde (39) was chosen as the starting ketone

because the cyclization of the 3-(3,5-difluorophenyl)propionic acid (42), in a

later step, produces only one product, the 5,7 -difluoro-1-indanone (43) in

which the fluorines are situated ortho- and para- to the carbonyl substituent.

Ethyl 3-(3,5-difluorophenyl)-2,3-dehydropropionate (40) was synthesised

by two methods:

i) from 3,5-difluorobenzaldehyde (39) and triethyl phosphonoacetate using

the Wadsworth-Emmons Reaction (Scheme 36).173 This reaction involved the

generation of the phosphonate anion using sodium hydride. Because of the

thermal instability of this anion, and the necessity for heat to aid its formation,

the 3,5-difluorobenzaldehyde was added before heat was applied.

ii) from 3,5-difluorobenzaldehyde (39) and (carbethoxymethylene)triphenyl­

phosphorane using the Wittig Reaction (Scheme 37).174 The latter method

proved to be much quicker and simpler.

53

Page 70: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Scheme 36

Scheme 37

F

,;?

+ F ~

H

(39) 0

o H 11 1

OeNa'" 1

(EtO)2 P-C-G02Et __ e

~H (EtO)2i1 +

H,~/C02Et

A(C'H

, , , , , t

Ar

40 H "-11 1

(EtO)2 P-C-C02Et

~-6-H I Ar

(C6H5hP=CHC02C2f-is

-----.

• F

o

+ (40)

e9 ~ (EtO)2 P-rC-C02Et ,:,)""

O-C-H I Ar

F

,;?

~ ..? CO2 Et

(40)

The reaction resulted in a 91 % yield of the alkene (40). The coupling

constant of the alkene protons in the lH n.m.r. spectrum was 16.0 Hz. A

vicinal coupling constant of this magnitude for protons of an alkene show the

protons to be trans. This alkene (40) was then hydrogenated using 10%

palladium on carbon catalyst, in ethyl acetate, under an atmosphere of

hydrogen, (Scheme 38). This gave a 97% yield of ethyl 3-(3,5-

difluorophenyl)propionate (41).

Scheme 38

F F 10% Pd/C

H2 EtOAc F (40) F (41)

This ester (41) was then hydrolyzed using potassium hydroxide in ethanol

and water. The resulting intermediate then undergoes acyl-oxygen cleavage

54

Page 71: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

to give the acid (42) (Scheme 39). This mechanism is generally referred to as

BAC2 (base-catalyzed, acyl-oxygen cleavage, bimolecular).

I Scheme 39

F K'

YJI cr-OH ~ / ~

EtO 0 . F (41)

Fyjl + EtO---> C + K;

~ , o OH F (42)

A cyclization step, analogous to the one required in this synthesis, has

been reported by Metz.'7S Following his procedure, using an excess of

polyphosphoric acid and a reaction temperature of 45°e, the cyclized product

(43) was obtained in 97% yield (Scheme 40). The reaction was carried

initially out at a temperature of 100oe, but only a 41 % yield was obtained.

Scheme 40

® = phosphate group

F~ F 0

-(43)

Since only low yields, typically 5% or less, were obtained for acylation of

the indanone (43) using the method used previously, namely sodium hydride

and diethyl carbonate, an alternative approach was needed. The next

approach was to attempt the synthesis of the trimethylsilyl enol ether, which

then could be reacted with lithium diisopropylamide followed by methyl

cyanoformate to give the desired product. Olah et al 176 have successfully

synthesised the TMS enol ether of 1-tetralone (44) using

trimethylchlorosilane, triethylamine and lithium sulfide in acetonitrile, at

ambient temperature (Scheme 41).

55

Page 72: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Scheme 41 '76

U2S TMSCI Et3N

Idry MeCN

16 hrs 81 QC

OTMS

~(75%) VU (44)

Attempts at using this method to produce the TMS enol ether of the 1-

inganone derivative proved unsuccessful. The reaction was difficult to follow

by proton n.m.r. because the triethylamine masked the alkyl proton signals.

To overcome this the reaction was carried out using an ion exchange resin

(Amberlyst A-21, supplied as the free base, -N[CH3h) to mop up the

hydrochloric acid produced in the reaction. In both cases the major product

was identified as the dimer (45). The characteristic m\z values of 73 and 75,

attributable to TMS were not present in the mass spectrum of this compound.

However, the m\z value of 331 was present, which corresponds to the

molecular ion, MH+', of the sulfur-bridged dimer. The signal in the 13C n.m.r. '-'

spectrum for the carbon at position 7 of the indanone (43) appears at Qc

159.9, as a double doublet. The coupling constants, 266.6 and 14.1 Hz,

correspond to coupling to the fluorine to which it bonded, and to the fluorine

at the C-5 position, respectively. In the 13C n.m.r. spectrum of the product

(45), the signal for the carbon at position 7 appeared as a doublet. The larger

coupling had been lost, and only the smaller coupling (10.1 Hz) with the

fluorine at the C-5 position was present. Attempts to synthesise the TMS enol

ether of 1-tetralone (44) following the procedure published by Olah,176 only

gave a 17 % yield, much less than was quoted in the paper.

f

(45)

56

Page 73: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

In the reaction with diethyl carbonate, the difluorinated indanone (43)

appeared to be more open to nucleophilic attack by the difluorinated

indanone anion than the diethyl carbonate since only what appeared to be

the dimeric products were produced. In order to determine the conditions for

base-initiated deprotonation at the C-2 position, without reaction of the

produced anion with the starting indanone, a deuteration experiment was

carried out. A small quantity of 5,7 -difluoro-1-indanone (43) was added

slowly to a solution LOA at "78 ac. The reaction was quenched with d4-acetic

acid and the deuterium incorporation determined from comparison of the

electron impact mass spectrum of the crude product and that of the starting

material. It was deduced from these results that thirty eight percent of the

indanone molecules contained at least one deuterium atom. The integration

(I) in the proton n.m.r. spectrum of the crude product for the protons on C-2

(IiH 2.67 - 2.62; I = 26 mm) was markedly reduced compared to that for the

protons on C-3 (OH 3.11 - 3.07; I = 40 mm). However the proton n.m.r.

spectrum showed no evidence of any dimer formation and it was decided to

proceed using this method. Oropwise addition of the indanone (43) to LOA at

-78°C, followed by the addition of Mander's reagent, 158 gave the required

ester (46) in 64% yield (Scheme 42). Fluorination was achieved using the

usual method and produced the methyI2,5,7-trifluoro-1-indanone-2-

carboxylate (47) in high yield (Scheme 43).

Scheme 42

FM (43)

i) LDA/THF -7aoC ..

ii) MeOCOCN

F o

F

57

Page 74: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

F 0

~ C02Me

FAJ-)<F

(47)

2.8 Synthesis of 1S-methoxy-1S-phenylacetone (50)

Curci et aiD found that the in situ generated dioxirane derivative of the

methyl ketone of Moshers acid (a-methoxy-a-trifluoromethylphenylacetone)

gave low enantioselectivity in its epoxidation of alkenes. The synthesis of 1 S­

methoxy-1 S-phenylacetone (50) was therefore undertaken and achieved from

commercially available (S)-(+)-a-methoxyphenylacetic acid using a method

developed by Nahm and Weinreb (Scheme 44).191

Scheme 44 Ph (COCQ2 Ph

MeO~C02H • MeO~COCI DCM

+ Me 1 H N' CI-2 , OMe

pyridine

Ph Ph OMe

~Me MeU ~~ MeO • MeO .... Me

(50) 0 H+ (49)

Epoxidation of trans-stilbene using the in situ generated dioxirane of this

ketone (50) rapidly converted the alkene to the epoxide. However due to the

small scale of the reaction, chiral shift studies were unsuccessful on the

epoxide.

58

Page 75: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 3

Page 76: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 3

3.1 Use of the 1-tetralone (19) and 1-indanone (20) dioxirane derivatives

in the oxidation of alkenes.

The dioxiranes were generated in situ from the tetralone (19) and

indanone (20) derivatives. The concentration of ketone in the

dichloromethane appears to have a dramatic effect on the rate of the

reaction, therefore this was standardised using a 0.5 M solution of the ketone

in the dichloromethane. Potassium peroxymonosulphate (Oxone®) was

added in small portions. Reactions were monitored by 'H n.m.r or G.C.

analysis and sufficient Oxone® added until enough epoxide had been formed

for chiral shift analysis to be performed. The reaction times of the

experiments carried out, ranged from a few hours to several days.

The alkenes were first epoxidised using an isolated solution of

dimethyldioxirane in acetone.'6. 20 Chiral shift studies were then carried out on

the resulting epoxides to determine the amount of chiral shift reagent,

Eu(hfch, required for baseline separation of the peaks of the enantiomers in

the 'H n.m.r. spectrum (et. p. 116, Chapter 3, Experimental). The

enantiomeric ratio of the epoxide produced in each of the in situ reactions

was then determined using these results as a standard.

Each of the alkenes were subjected to a blank run, in which no ketone was

used in the reaction. After the addition of sixty equivalents of Oxone®, no

epoxidation was noted for trans-stilbene (23) or the chromene (24), and only

1.5% epoxidation for trans-~-methylstyrene (25). From these results we can

conclude that the dioxirane derived from the ketone is playing the major role

in the alkene epoxidations.

3.2 Results of Epoxidations

No enantioselectivity was observed for the epoxidation reactions using the

dioxirane derivatives of (31 b) and (31 c). It was thought that this may have

been due to conformational mobility of the saturated ring in the tetralone

59

Page 77: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

system. The fact that, in the solid state, the ester residue prefers to take up

the axial conformation supports this view (Figure 1).

Table 5: Results of Epoxidations using the Isolated Diastereomers of (31c)

Olastereomer 11 Olastereomer I -Enantiorrer Ratio ~

%Cooversion of ciJo2 o 0 2 Alkene to EpOlGde WO'" "~F

.

? 50:50 50:50

14% 19% RI

240

/ 50:50 50:50

84% 76%

120

~ 50:50 50:50

56% 49%

180

Equivalents of Oxone

Table 6: Results of Epoxidations using the Isolated Enantiorners of (31b)

~

Enantiomer I Enantiomer 11 Enantiomer Ratio 0 0 0 0

% Conversion of Alkene cXi/~ CJCf, /Me

to Epoxide I • F 0 I • F 0

rPh 50: 50 50: 50

Me 100% 98%+ 120

Equivalents of Ox one

60

Page 78: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

No enantioselectivity was observed in these epoxidation reactions. This

may be due to conformational mobility of the saturated ring in the tetra lone

system. We thought that incorporating a hydroxy group beta to the ketone,

may confer some conformational stability via hydrogen bonding to one of the

oxygen in the dioxirane ring. The results that we obtained for the epoxidation

of alkenes using the dioxirane derivative (48) of such a ketone (34) did not

show any improvement on our previous results.

Table 7: Results of Epoxidations using the Isolated Hydroxy­Enantiomers of (34)

Enantiomer Ratio

% Conversion to Alkene

Me

r' Ph

240

Equivalents of Ox one

Enantiomer I o OH

Me • Me

F

50 :50

96",(,

Enantiomer 11 o OH

50 :50

95%

From models ofthe 1-tetralones (19) and 1-indanones (20), it appeared

that the indanone (37) had less conformational flexibility than that of the

tetralone (31). We, therefore, hoped that with this more conformationally.

stable structure, the dioxirane derivative of the indanone (37) would show

some enantioselectivity in its epoxidation of alkenes. From our results, it can

be seen that no enantioselectivity was observed.

61

Page 79: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Table 8: Results of Epoxidations using the Isolated Enantiomers {,f (37)

EnantiornerJ Enantiornerll

Enantiomer Ratio 0 0 0 0

~~Et ~o~Et % Conversion of I • F 0

Alkene to Epoxide

Me 50::'50 ;I

Ph 98% 60

Equivalents of Ox one

Since no enantioselectivity was observed, when using the above

homochiral ketones, in the epoxidation of the alkenes, we deemed it

unnecessary to separate the enantiomers ofthe methyI2,5,7-trifluoro-1-

indanone-2-carboxylate (47), We have included this molecule in this report to

show the influence of fluorine on the rate of the reaction,

Table 9: In Summary -

Ketone % Conversion of alkene to epoxide

(Equivalents of Oxone used)

Alkene 31b-E1 31c-DII 31c-D1 34-E1 37-E1 47

- 14% 19% - - -21 - (240) (240) - - -

100% 84% 76% 98% 98% 100% 22 (120) (120) (120) (240) (60) (30)

- 56% 49% - - -23 (180) (180) - - - -

Table 9 shows the general reactivity of the dioxirane derivatives of the

ketones, Le" in decreasing order of reactivity: (47), (37), (31b), (31c), (34),

The result of the epoxidation reaction using the dioxirane derivative of methyl

2,5,7 -trifluoro-1-indanone-2-carboxylate (47) confirms our previous findings.

62

Page 80: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 4

Page 81: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 4

4.1 Optimising the conditions of the dioxirane epoxidation reaction

In order to acquire experience of the two phase dioxirane epoxidation

reaction, two reactions involving the in situ generation of the dioxirane

derivative of 1 ,1 ,1-trifluoroacetophenone were initially carried out. These

reactions are described in Chapter 4, Experimental (p117). The first

reaction involved the epoxidation of cholesterol using this dioxirane. The

epoxide was obtained in thirty percent yield. Trans-stilbene (23) was then

used in the reaction, forty percent being converted to the epoxide. These

results were in agreement with those previously obtained in this

Departmenl,l54 Since trans-stilbene (23) proved more susceptible to

epoxidation, it was initially used as the substrate in the dioxirane reactions.

The dioxirane derivatives of the 2-substituted tetralones (19) were used to

epoxidise trans-stilbene (23) using the conditions described on p118. The

results of these epoxidations are shown in Table 10.

Table 10

Ketone Precursor Number of Days Percentage Epoxide

26a 3 40

31a 2 45

31b 3 100

31 c-DII 7 0

37 6 6

As anticipated, the dioxirane derivative of ethyI2-methyl-1-tetralone-2-

carboxylate (26a) proved slower at epoxidations than the ethyl 2-fluoro-1-

tetralone-2-carboxylate (31 a). The methyl 2-fluoro-1-tetralone-2-carboxylate

(31b) app.eared to be a more efficient epoxidizing agent than the ethyl

analogue (31a). No evidence of epoxide formation was noted, using this

63

Page 82: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

method, with the menthyl 2-fluoro-1-tetralone-2-carboxylate (31c-01l). On

optimising the conditions of the dioxirane reaction, alkene epoxidation using

the dioxirane derivative of this tetralone (31-011) was achieved (refer to

Chapter 3, Table 9).

The epoxidizing ability of ethyl 2-fluoro-1-indanone-2-carboxylate (37) was

also assessed. After six days only 6% of the epoxide and 94% of the starting

alkene was observed. None of the original indanone was found by GC-MS.A

series of experiments was carried out to find the best conditions for the

dioxirane reaction. These included changing the pH, the amount of phase

transfer catalyst, Oxone®, ketone and dichloromethane (refer to p 65 - 68).

The variables having the greatest effect were:

• pH: optimum approximately 7.5

• the concentration of the ketone in the dichloromethane: decreasing the

volume of dichloromethane, but using the same quantity of ketone,

significantly increased the rate at which the epoxidation reaction

proceeded.

• the concentration of phase transfer catalyst: increasing the quantity of

phase transfer catalyst increased the rate at which the epoxidation

reaction proceeded, but not to a very significant extent.

These results are in agreement with those found by Denmark et al. 26 The

experimental procedure which was eventually used is described in the

Experimental, Chapter 3.

64

Page 83: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Ol (Jl

Entry

1

2

3

4

5

6

7

8

9

10

11

12

13

Alkene Ketone

trans- J

31b stilbene trans- 31b

stilbene trans- 31b

stilbene trans- 31b

stilbene trans- 31b

stilbene trans-

31b stilbene trans-

31b stilbene trans- 31b

stilbene trans- 31b

stilbene trans- 31b

stilbene trans- 31b

stilbene trans- 31b

stilbene trans- 31b

stilbene

Ketone TBAHS Phosphate (Quantity) (eq.) Buffer (mll

1.95 9 1.0 40 2.0 ~g. 0.27 9 1.0 40 2.0eq. 0.19 9 1.0 40 2.0eQ. 0.19 9 1.0 40 2.0eq. 0.27 9 1.0 40 2.0eq. 0.10 9

1.0 10 2.0eq. 0.20g

5.0 14 2.0eq. 0.30g

5.0 14 3.0 eg, 0.48 9 5.0 66 10.0

0.07g 10.0 20

5.0eq. 0.07g

10.0 20 5.0eQ. 0.07g

10.0 20 5.0~

0.07g 10.0 20 5.0eq.

pH Oxonee Number DCM EDTA.Na, Epoxide

(eQ.) of days (ml) (Q) (%)

7.5 50x 3 3 40 1 0

7.5 50x 3 3 40 1 100

7.5 50x4 4 40 1 0

7.5 50x 3 3 40 1 0

7.5 50x 5 5 30 1 0

7.5 50x7 7 10 0.33 0

7.5 50 x 1 1 10 0.35 0

7.5 50 x 1 1 10 0.35 -2

7.5 50x7 7 22 18mg -2

6.5 50x 2 1 6 10mg 0

6.5 50 x 16 B 6 10mg 0

6.5 50xB 4 6 10 mg 2

6.5 50x 10 5 6 10 mg 92

Page 84: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

(J) (J)

Entry

14

15

16

17

18

19

20

21

22

23

Alkene

ll-methyl-stvrene

p-methyl-stvrene Irans-

stilbene Irans-

stilbene Irans-

stilbene ~-methyl-styrene

trans-stilbene

Irans-stilbene

Irans-stilbene

Irans-stilbene

Ketone Ketone TBAHS

(Quantity) (eQ.)

31b 1.009 0.02 2.0 eq.

31b 1.009 0.02 2.0eq.

31a 0.28 9 5.0 4.0 eQ.

31a 0. 14 9 1.0 2.0eQ.

31a 0.14 9 1.0 2.0eQ.

37 1.209 0.02 2.0eQ.

0 0.679 ctCF3

1.0 1.0 eq.

0 0.67 9 ctCF3

1.0 1.0 eq.

0 0.67 9 ctCF3

1.0 3.0 eq.

0 0.67 9 ctCF

,

1.0 3.0eq.

Phosphate pH

Oxone" Number DCM EDTA.Na, Epoxide Buffer (ml) (eQ.) of days (ml) (g) (%)

15 7.5 30x3 1 25 * 0

25 7.5 47.5 x 1 1 10 * 78

15 7.5 50x 5 5 10 0.4 <2

15 7.5 60x2 2 10 0.4 0

15 7.5 60x2 3 10 0.4 0

20 7.5 30x5 3 15 * 14

40 7.5 50 xl 1 40 1 11

40 7.5 50 Xl 1 30 1 40

40 7.5 50 xl 1 30 1 <10

40 10.5 50x 1 1 30 1 0

Page 85: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Entry Alkene Ketone Ketone TBAHS Phosphate

pH Oxonee Number DCM EDTA.Na, Epoxide

(quantity) (eq.) Buffer (ml) (eq.) of days (ml) (g) (%)

trans-0

0.67g 24 ciCF3

1.0 40 7.5 50 x 1 1 30 1 7 stilbene 3.0eq.

25 trans- d CF3

0.67 9 1.0 40 7.5 50 x 1 1 30 1 32 stilbene 3.0eq.

trans-0 0.67g

26 ciCF3 5.0 40 7.5 50 x 1 1 30 1 77

stilbene 3.0eq.

trans- :>=0 0.22 9 27 stilbene 3.0eq.

1.0 40 7.5 50 x 1 1 30 1 10

28 trans- 43 0.02 9 5.0 20 7.5 50 x 2 1 6 . 0.5 mg 29

stilbene 2.0 eq.

29 trans- 43 0.02g

5.0 20 7.5 50 x 1 1 6 0.5 mg 8 stilbene 2.0 eq.

30 trans- 43 0.02 9 5.0 20 7.0 50 x 1 1 6 0.5 mg -1

stilbene 2.0eq.

31 trans- 43 0.05 9 10.0 20 6.5 50x 6 6 6 0.5 mg 7

stilbene 5.0eq.

32 trans- 43 0.05 9 10.0 20 7.0 50 x6 6 6 0.5 mg 19

stilbene 5.0eq.

33 trans- 43 0.05 9 10.0 20 6.5 50 x 2 1 6 10mg 2

stilbene 5.0eq.

34 trans- 43 0.05g 10.0 22 6.5 50 x 10 5 6 10mg 2

stilbene 5.0

Page 86: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

m Cl)

Entry

35

36

37

Alkene

trans-stilbene trans-

stilbene p-methyl-styrene

Ketone Ketone TBAHS

. (quantity) (eq.)

43 0.05g 10.0

5.0

43 0.05 9 10.0 5.0

43 1.00 9 0.04 2.0 eq.

Phosphate pH

Oxone® Number DCM EDTA.Na, Epoxide Buffer (ml) (eq.) of days (ml) (q) . (%)

22 7.5 50 x2 1 6 10 mg -1

22 7.5 50x 2 1 6 10 mg 2

20 7.5 30x 6 2 15 • 39

Page 87: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 5

Page 88: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 5

5.1 Preparation of solutions of dioxiranes

The dioxirane derivatives of acetone, 1,1,1-trifluoroacetone and

cyclohexanone were prepared in solutions of their ketones, with molarities of

0.09 M, 0.68 M and 0.09 M being obtained, respectively. Attempts to prepare

the dioxirane derivatives of hexafluoroacetone and 1,1,1-

trifluoroacetophenone in solutions of their ketones with dichloromethane as

an additional solvent were carried out, albeit with little success. The molarities

obtained for these were 0.02 M and 0 M, respectively.

The resulting dimethyldioxirane solution was used to epoxidise various

alkenes. A slight excess of the dioxirane was used in each epoxidation. The

results were as follows:

trans-stilbene oxide 100%

trans-~-methylstyrene oxide 89%

6-acetyl-2,2-dimethyl-2H-1-benzopyran 85%

6-chloro -2,2-dimethyl-2H-1-benzopyran 89%

6-cyano -2,2-dimethyl-2H-1-benzopyran 89%

The solution of the dioxirane derivative of cyclohexane was used to oxidise

Metaclopramide. This gave the N-oxide in 73% yield (32% after

recrystallisation).

5.2 Determination of the molarity of a dioxirane solution by titration

The molarity of a dioxirane solution was determined by titration of the

iodine, which a known volume of the dioxirane solution liberated from

potassium iodide, against a thiosullate solution 01 known molarity.

69

Page 89: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chemical equations for the titration

Reaction of the dioxirane with iodine:

o R 2KI + IX + H20 o R

R 12 + O=< + 2KOH

R

Reaction of the iodine with sodium thiosu/fate:

The molarity of the dioxirane solution is calculated:-

Number of moles of sodium

thiosulfate solution used = V Na2S,03 X M Na2Sp3

1000

Since 2 moles of potassium iodide react with 1 mole of the dioxirane to give 1

mole of iodine, and the iodine reacts with 2 moles of sodium thiosulfate:

1 mole dioxirane == 2 moles sodium thiosulfate

Number of moles of

dioxirane used =

1000 x 2

Mo larity of the

dioxirane solution

V Na2Sp3 X M Na2SP3 x..J.Oe(J

..J.Ge(} x 2 x V dioxirane

where M Na2S

20

3 = molarity of the sodium thiosulfate solution

V N so = volume of sodium thiosulfate used a 2 2 3

V dioxirane = volume of dioxirane solution used

Example Calculation:

Titration of a freshly prepared solution of dimethyldioxirane:

Volume of dioxirane solution taken: 0.2 ml

Molarity of the sodium thiosulfate solution: 0.01182 N

70

Page 90: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Volume of sodium thiosulfate solution needed (titre): 2.56 ml

= 2.56 x 0.01182 = 0.076 M 0.2 x 2

71

i

/' .

Page 91: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Conclusion

Page 92: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Conclusion

Although no enantioselectivity was observed for the epoxidation reactions

using the dioxirane derivatives of (19) and (20). these compounds have been

shown to efficiently epoxidise a variety of alkenes. The rate of epoxidation

was shown to be dependent on steric factors and the incorporation of fluorine

into the aromatic ring of arylalkyl ketones. The rate of epoxidation was also

shown to be highly dependent on dilution factors. i.e. on the concentration of

the dioxirane in the organic phase.

72

Page 93: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Experimental

Page 94: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Experimental

General

1H NMR spectra were recorded on a Bruker 250 MHz NMR Spectrometer. All

samples were recorded with tetramethylsilane as internal standard. Chemical

shifts (0) are reported in ppm. Multiplicities are reported as follows: br

(broad), s (singlet), d (doublet), t (triplet), q (quartet), m (multiplet).

13C NMR spectra were recorded on a Bruker 250 MHz NMR Spectrometer

19F NMR spectra were obtained on a 400 MHz NMR Spectrometer.

IR spectra were recorded on a Nicolet 205 series FT-IR Spectrometer.

Mass spectra were obtained on a Kratos MS80 Spectrometer with OS-55

data system.

Column chromatography was carried out using the flash chromatography

technique with Matrex Silica 60, 35-70 micron (Fisons Scientific Equipment)

unless otherwise noted.

Thin layer chromatography was carried out using aluminium sheet silica gel

60 F254, 0.2 mm layer thickness (Merck) or aluminium sheet aluminium oxide

60 F254 neutral (type E), 0.2 mm layer thickness (Merck).

Preparative TLC was performed on 0.2 metre glass backed plates coated

with Kieselgel 60PF254 (Merck).

Optical Rotation. The [0.]0 values were measured using an Optical Activity

AA 100 polarimeter.

Melting points were obtained on a Reichert Manual Melting Point Apparatus.

pH Stat. experiments were controlled using a Radiometer autotitrator. Type

ABU 11b.

Microanalysis was performed at Medac Ltd. (Brunei University), or within this Department.

73

Page 95: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 2, Experimental

EthYI-1-tetralone-2-carboxylate (26)161, 162 (Ethyl 1 ,2,3,4-tetrahydro-1-oxonaphthalene-2-carboxylic acid) cD . 0 °o,Et

Method 1158, 161, 162

To a stirred solution of diisopropylamine (1.27 ml, 9.03 mmole) at 0 cC,

under argon, was added butyllithium (2,5 M solution in hexane, 3.61 ml, 9.03

mmole). This gave a thick gel, to which was added dry THF (20 ml). The

resulting solution was stirred for twenty minutes then the temperature was

lowered to -78 cC. A solution of 1-tetralone (1.00 ml, 7.52 mmole) was added

and stirring continued for 1 hour at 0 cC. The temperature was lowered again

to -78 cC, and HMPA (1.31 ml, 7.52 mmole) followed by ethyl cyanoformate

(0.89 rnl, 9.03 mmole) added. The reaction mixture was allowed to warm to

room temperature and left to stir for ten minutes. After this time, the solution,

was poured into cold water (80 ml), extracted with dichloromethane (2 x 80

ml), dried (Na2S04) and concentrated in vacuo to yield a brown liquid (3.43 g).

Tic analysis (silica with 9 : 1 light petroleum [b,p. 40 - 60 cC] : ethyl acetate)

showed three spots; 0.75 (desired product [26]), 0.51 (1-tetralone) and

baseline material. The product was isolated by column chromatography (silica

with 9 : 1 light petroleum [b.p. 40 - 60 cC] : ethyl acetate) yielding a pale

yellow liquid (0.835 g, 51%); vmax (neat) 3060,2976 (C-H), 1738 (ester C=O),

1684 (a.-arylketone C=O), 1642 (enol ester C=O), 1618 (enol ester C=O),

1596, 1568 and 1452 cm,1 (arene C-C); OH (250 MHz, CDCla) 12.49 (1 H, s, J

enolic H), 8.04 (1 H, d,d, J7.8, 1.2 Hz, ketoester 8-H), 7.80 (1 H, d,d, J7.5,

1.6 Hz, enol 8-H), 7,52 - 7.17 (6 H, m, ketoester and enol 5-H, 6-H, 7-H), 4.27

(4 H, q, J7.2 Hz, ketoester and enol CH2CHa), 3.59 (1 H, d,d, J10.3, 4.9 Hz,

ketoester 2-H), 3.04 - 2.48 (8 H, m, ketoester and enol3-H and 4-H), 1.34 (3

H, t, J7.1 Hz, ketoester CH2CHa), 1.30 (3 H, t, J7.2 Hz, enol CH2CHa); Oc

74

Page 96: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

(62.5 MHz, CDCb) 193.2 (ketone 1-C), 172.7 and 170.2 (ketone and enol2-

C02Et), 165.0 (enoI1-C), 143.7 and 139.4 (ketone and enoI8a-C), 133.8 and

130.5 (ketone and enoI8-C), 131.8 and 130.0 (ketone and enoI4a-C), 128.8

and 127.6 (ketone and enoI7-C), 127.4 and 126.8 (ketone and enoI6-C),

126.5 and 124.3 (ketone and enol 5-C), 97.0 (enoI2-C), 54.4 (ketone 2-C),

61.2 and 60.5 (ketone and enol 2-C02CH2CH3), 27.7 and 27.6 (ketone and

enol 4-C), 26.4 and 20.5 (ketone and enol 3-C), 14.3 and 14.2 (ketone and

enol 2-C02CH2CH3). Ratio of ketoester : enol was 1 : 1. After chromatography

with petroleum ether only, no ketoester was observed, but gradually the ratio

moved towards 1 : 1 on standing at room temperature.

Method 2161

1-Tetralone (4.00 g, 3.88 ml, 27.36 mmole) was added to a stirred

suspension of sodium hydride (60% dispersion in oil, 1.32 g, 32.84 mmole) in

dry diethyl carbonate (19.50 g, 20 ml, 165 mmol), under an atmosphere of

" argon. On application of heat for five minutes, a solid formed. The reaction

mixture was allowed to cool. Tic analysis of the solid showed no evidence of

1-tetralone. The solid was dissolved in 2 M HCI (100 ml) and extracted with

ethyl acetate (3 x 100 ml). The organic extracts were dried (Na2S04) then

concentrated in vacuo to yield a brown liquid (12.006 g). The product was

purified by column chromatography (silica with 9 : 1 light petroleum [b.p. 40 -

60 ·C] : ethyl acetate) followed by Kugelrohr distillation to yield a pale yellow

liquid (5.483 g, 92%); b.p. 190·C (1 mm Hg). I.r. and n.m.r. data as above.

75

Page 97: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Methyl 1-tetralone-2-carboxylate (27) 161.177

(Methyl 1.2,3,4-tetrahydro-1-oxonaphthalene-2-carboxylic acid)

o

1-Tetralone (2.50 g, 2.78 ml, 17.10 mmole) was added to a stirred

suspension of sodium hydride (60% dispersion in oil, 0.821 g, 20.52 mmole)

in dry dimethyl carbonate (19.50 g, 20 ml, 165 mmole), under an atmosphere

of nitrogen. The reaction mixture was heated for fifteen minutes during which

time a lilac and white solid formed. Further dry dimethyl carbonate (5 ml) was

added via syringe. Tic analysis of the solid showed no evidence of 1-

tetralone. The solid was allowed to cool to room temperature, dissolved in

hydrochloric acid (aqueous 2M, 50 ml) and extracted with ethyl acetate (4 x

50 ml). The combined extracts were dried (MgS04) and concentrated in

vacuo to yield a brown oil (4.328 g). This crude product was purified by flash

chromatography (silica with 9 : 1 light petroleum [b.p. 40 - 60 QC] : ethyl

acetate) to yield a colourless crystalline solid (27) (3.411 g, 98%); m.p. 61.5

QC - 69.5 QC; b.p. 200 QC (2 mmHg); vmax (neat) 2951 (C-H), 1745 (ester

C=O), 1685 (a-aryl ketone C=O) 1650, 1648, 1620, 1600, 1570, 1453, 1441,

1363 cm'1; OH (250 MHz, CDCI3) 12.41 (1 H, s, enol OH), 8.01 (1 H, d,d, J7.8,

1.4 Hz, ketoester 8-H), 7.77 (1 H, d,d, J7.0, 1.6 Hz), 7.45 - 7.13 (6 H, m,

keto ester and enol 7-H, 6-H and 5-H, 3.78 (3 H, s, enol CH3), 3.74 (3 H, s,

ketoester CH3), 3.01 - 2.91 (2 H, m, ketoester 4-H), 2.79 - 2.73 (2 H, s, enol

4-H), 2.63 - 2.08 (4 H, m, ketoester and enol 3-H); Bc (62.5 MHz, CDCI3)

192.9 (ketone 1-C), 173.0 and 170.6 (ketone and enol 2-C02Me), 165.0 (enol

1-C), 143.7 and 139.4 (ketone and enoI8a-C), 133.9 and 130.5 (ketone and

enoI7-C), 131.6 and 129.9 (ketone and enoI4a-C), 128.8 and 127.7 (ketone

and enol 6-C), 127.4 and 126.9 (ketone and enol 5-C), 126.5 and 124.3

(ketone and enol 8-C), 96.8 (enol 2-C), 54.4 (ketone 2-C), 52.3 and 51.6

(ketone and enoI2-C02 CH3), 27.7 and 27.6 (ketone and enoI4-C), 26.3 and

20.5 (ketone and enol ~-C).

76

Page 98: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

L-Menthyl 1-tetralone-2-carboxylate (28)

(L-Menthyl 1 ,2,3,4-tetrahydro-1-oxonaphthalene-2-carboxylic acid)

o

Method 1158

o ..... Men o

To stirred diisopropylamine (0.977 g, 9.65 mmole) under argon, was added

butyllithium (2.3M in hexane, 4.1 ml, 9.65 mmole) at -20 QC. This immediately

formed a clear gel, to which was added tetrahydrofuran (dry, 3 ml) and the

resulting solution left to stir at -20 QC for thirty minutes. The temperature was

lowered then to -78 QC and 1-tetralone (1.200 g, 1.112 ml, 8.00 mmole)

added. The flask was warmed to 0 QC and stirred for one hour, after which

time the temperature was lowered to -78 QC and HMPA (1.44 g, 1.40 ml, 8.00

mmole) followed by L-menthyl chloroformate (2.11 g, 2.07 ml, 9.65 mmole)

added: The reaction mixture was left to stir overnight at room temperature,

then poured into a separating funnel containing distilled water (100 ml). The

aqueous fraction was then extracted with dichloromethane. The organic

extracts were dried (MgS04) and concentrated in vacuo to yield a brown oil

. "(4.289 g). This crude material was purified by flash chromatography (flash

silica with 1:1 dichloromethane : light petroleum [b.p. 40 - 60 QC] followed by

7:1 light petroleum [b.p. 40 - 60 QC] : diethyl ether) to yield the desired

material as a clear oil (1.100 g, 42%). Three products were produced in this

reaction. These appeared (tic analysis using silica with 1:1 dichloromethane :

light petroleum [b.p. 40 - 60 QC]) at R, values of 0.75 (L-Menthyl 3,4-

dihydronaphthalene-1-carbonic acid [30]),.0.57 streaking (L-MenthyI1-

tetralone-2-carboxylate [28]), 0.57 (N,N-diisopropyl L-menthyl carbamate

[29]). Tic analysis using silica with 7 : 1 light petroleum [b.p. 40 - 60 QC] :

diethyl ether gave R, values of 0.65 streaking (L-Menthyl 1-tetralone-2-

carboxylate [28]) and 0.40 (N,N-diisopropyl L-menthyl carbamate [29]).

77

?

Page 99: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

L-Menthyl Hetralone-2-carboxylate (28): vmax (neat) 2955 (aromatic C-H),

2932 (C-H), 2870 (saturated C-H), 1734 (ester C=O of ketoester), 1690 (a.­

aryl ketone C=O), 1643 (enol ester C=O), 1618 (enol ester C=O), 1570, 1455,

1389,1268,1213,1085 cm-'; OH (250 MHz, CDCI3) 12.60 (1 H, s, OH), 8.05

(1 H, d,d, J7.7, 1.2 Hz, ketoester 8-H), 7.80 (1 H, d,d J6.9, 1.7 Hz, enol ester

8-H), 7.47 (1 H, t,t, J7.5, 1.4, ketoester 7-H), 7.35 -7.15 (5 H, m, enol ester 7-

H, 6-H, 5-H, ketoester 6-H, 5-H), 4.89 -4.79 (2 H, m, enol ester and ketoester

CHO-CO), 3.59 -3.55 (1 H, m, ketoester 2-H), 3.00 - 2.99 (2 H, m, ketoester

4-H), 2.81 (2 H, t, J7.8 Hz, enol ester 4-H), 2.59 - 2.52 (4 H, m, enol ester

and ketoester3-H), 2.25 - 2.12 (2 H, m, menthyl C-H), 1.97 -1.83 (2 H, m,

menthyl C-H), 1.73 - 1.24 (10 H, m, menthyl C-H), 1.20 - 0.69 (7 H, m,

menthyl C-H), 0.94 - 0.90 (12 H, m, enol ester and ketoester CH3 of iso­

propyl group), 0.80 (6 H, d, J7.0 Hz, enol ester and ketoester Me); Oc (62.5

MHz, CDCI3) 193.5 (ketone 1-C), 172.4 and 169.6 (ketone and enoI2-C02),

164.9 (enoI1-C), 143.6 and 139.3 (ketone and enoI8a-C), 133.7 and 130.1

(ketonEl and enoI7-C), 130.2 and 131.8 (ketone andenoI4a-C), 128.8 and

127.7 (ketone and enol 6-C), 127.4 and 126.9 (ketone and enol 5-C), 126.5

and 124.2 (ketone and enol 8-C), 97.3 (enol 2-C), 75.4 and 74.5 (ketone and

enol 1 '-C), 47.1 and 46.8 (ketone and enol 2'-C), 41.8 and 40.7 (ketone and

enol 6'-C), 34.3 and 34:2 (ketone and enol 4'-C), 31.4 (CC1--l(CH3M 27.8 and

27.5 (ketone and enol 4-C), 26.6 (5'-C), 26.4 and 20.5 (ketone and enol 3-C),

23.7 (3'-C), 22.0 and 20.7 (2C, CCH( C1--l3)2) , 16.7 (5-CCl-I3).

N,N-diisopropyl L-menthyl carbamate (29): V max (neat) 2956 (aromatic C­

H), 2933 (C-H), 2871 (saturated C-H), 1686 (C=O), 1456, 1433, 1368, 1328,

1287,1269 cm-'; OH (250 MHz, CDCI3) 4.66 (1 H, t,t J10.8, 4.3 Hz, CH-O­

CO), 3.91 (2 H, broad s, NCH), 2.08 - 1.92 (2 H, m, menthyl C-H), 1.70 - 1.65

78

Page 100: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

(2 H, m, menthyl C-H), 1.48 - 1.27 (2 H, m, menthyl C-H), 1.20 (12 H, d, J6.8

Hz, NCH(CH3h), 1.14 - 0.79 (3 H, m, menthyl C-H), 0.90 (6 H, d, J6.9 Hz,

CCH(CH3)2), 0.80 (3 H, d,d, J6.9, 1.7 Hz, Me); oe (62.5 MHz, CDCI3) 155.5

(OCON), 74.2 (1-C), 47.4 (2-C), 45.7 (2C, NQ-l), 41.6 (6-C), 34.5 (4-C), 34.5

and 34.3 (CQ-l(CH3H 26.2 (5'-C), 23.4 (3-C), 22.1 and 20.7 (2C,

CCH(Q-l3h), 21.0 (4C, (Q-l3h CHN), 16.3 (5-CQ-l3); m/z (electron impact)

283.2515 (M+, 2.5%, C17H33N02 requires 283.21129), 284 (4, M++1), 146

(57), 84 (100), 69 (25),55 (26),43 (43).

L-MenthyI3,4-dihydronaphthalene-1-carbonic acid (30): V max (neat) 2956

(aromatic C-H), 2933 (C-H), 2871 (saturated C-H), 1758 (C=O), 1733, 1456

(arene C-C), 1286, 1253, 1226, 1182, 1012 cm-1; OH (250 MHz, CDCI3) 7.20-

7.10 (4 H, m, ArH), 5.78 (1 H, t, J4.7 Hz, 2-H), 4.56 (1 H, t,t, J10.9, 4.4 Hz,

CH-O-CO), 2.84 (2 H, t, J7.9 Hz, 4-H), 2.46 - 2.38 (2 H, m, 3-H), 2.15 - 1.97

(2 H, m, menthyl C-H), 1.72 - 1.67 (1 H, m, menthyl C-H), 1.48 - 1.42 (1 H, m,

menthyl C-H), 1.19 - 0.80 (5 H, m, menthyl C-H), 0.94 - 0.91 (6 H, m, CH3 of

iso-propyl group), 0.85 - 0.82 (3 H, m, Me); Oe (62.5 MHz, CDCI3) 153.2

(OC02), 146.1 (1-C), 136.4 (8a-C), 130.4 (4a-C) , 128.2 (8-C), 128.0 (7-C),

127.5 (6-C), 120.5 (5-C) , 115.0 (2-C), 79.2 (1 '-C), 47.0 (2'-C), 40.6 (6'-C),

34.1 (4'-C), 31.4 (CQ-l(CH3h), 27.4 (4-C), 26.5 (5'-C), 23.4 (3-C), 22.0 (3'-C),

22.0 and 20.7 (2C, CCH(Q-l3h) 16.3 (5-CQ-l3); m/z (electron impact)

328.2051 (M+, 3.0%, C21H2803 requires 328.203845), 146 (100, M+­

C02Menthyl), 139 (47), 118 (16), 95 (41), 83 (80), 69 (28), 55 (39), 41 (34),

27 (13).

79

Page 101: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Method 2

To a stirred solution of butyllithium (2.4 M in THF, 1.66 mmole, 3.98

mmole) in THF (0.5 ml), at -78 cC, under nitrogen, was added Hetralone

(0.540 g, 0.500 ml, 3.63 mmole). The solution was left to stir at 0 cC for one

hour. This resulted in an orange coloured solution. The reaction mixture was

returned to -78 cC and L-menthyl chloroformate added to it. This solution was

allowed to warm slowly to room temperature and left to stir overnight. The

solution was poured into a separating funnel containing distilled water (100

ml), extracted with dichloromethane (3 x 100 ml) and dried (MgS04).

Concentration of the dried organic extracts in vacuo yielded a yellow oil. The

desired product was obtained from this oil by column chromatography (silica

with 1:1 light petroleum [b.p. 40 - 60 CC]) as a clear oil (0.177 g, 15%). N.m.r

and Lr. data as above.

Method i64.178

Step I: To a stirring mixture of 1-tetralone (0.500 g, 0.455 ml, 3.42 mmole)

and pyrrolidine (0.730 g, 0.856 ml, 10.26 mmole) in dry diethyl ether (5 ml).

under nitrogen and at 0 cC, was added slowly titanium tetrachloride (0.357 g,

0.206 ml, 1.88 mmole). After one hour, the reaction was allowed to warm to

room temperature and left to stir overnight. Further diethyl ether (5 ml) was

,· .. added, then the solution filtered through celite to remove the titanium (IV)

oxide produced in the reaction. The organic phase was then concentrated in

vacuo to yield a yellow oil (0.528 g, 77%); V max (neat) 3060 (aromatic C-H),

2965, 2931,2877,2826, 1687, 1620 (enamine C=C), 1485, 1372, 1354,

1321, 1287,1186, 1135 cm·1; OH (250 MHz, CDCI3) 7.41 (1 H, d, J7.5 Hz, 8-

H), 7.24 - 7.08 (3 H, m, 7-H, 6-H, 5-H), 5.16 (1 H, t, J4.7 Hz, 2-H), 2.97 - 2.92

(4 H, m, N(CH2h), 2.67 - 2.61 (2 H, m, 3-H), 2.23 - 2.15 (2 H, m, 4-H), 1.96 -

1.82 (m, 4H, NCH2 (CH2)).

Step 11: To a stirring solution of the enamine (0.200 g, 1.00 mmole), in dry

benzene (15 ml), at room temperature and under nitrogen, was added L­

menthyl chloroformate (0.110 g, 0.108 ml, 0.50 mmole). The mixture was

heated to reflux for four days. Tic analysis was carried out on the solution, but

80

Page 102: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

no evidence of L-menthyI1-tetralone-2-carboxylate was observed. After the

four days, the reaction mixture was concentrated in vacuo to yield a brown oil

(0.437 g). The crude n.m.r. of this material showed no evidence of the ~­

ketoester.

Method 4164• 178

Step I: Titanium tetrachloride (0.636 g, 0.368 ml, 3.35 mmole) was added

slowly to a stirring solution of 1-tetralone (0.500 g, 0.455 ml, 3.35 mmole) and

morpholine (0.876 g, 0.877 ml, 10.05 mmole) in dry diethyl ether (5 ml), under

nitrogen and at 0 cC. The reaction was left to stir for one hour at this

temperature. It was then allowed to warm to room temperature and left to stir

for a further six hours. After this time, the titanium (IV) oxide was removed by

filtration through celite and the resulting solution concentrated in vacuo to

yield an orange residue (0.691 g, 96%). The n.m.r. spectrum of this

compound showed the presence of unreacted 1-tetralone. The 1-tetralone

was removed by distillation of the residue (100 cC, 2.5 mm Hg). The

remaining residue was then purified by Kugelrohr distillation (200 cC, 2.5 mm

Hg) to give an orange-tinged gel (0.482 g, 67%); vmax (neat) 2956 (aromatic

C-H), 2933, 2887, 2850, 2827, 1624 (enamine C=C), 1442 (arene C-C),

1370, 1259, 1237, 1119, 1018 cm-'; OH (250 MHz, CDCI3) 7.42 (1 H, d, J7.0

Hz, 8-H);.T22 -7.10 (3 H, m, 7-H, 6-H, 5-H), 5.27 (1 H, t, J4.7 Hz, 2-H), 3.84

(4 H, t, J4.6 Hz, OCH2), 2.82 (4 H, t, J4.6 Hz, NCH2), 2.70 - 2.64 (2 H, m, 4-

H), 2.26 - 2.13 (2 H, m, 3-H).

Step 11: as for Step 11 in Method 3.

81

Page 103: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Ethyl 2-methyl-1-tetralone-2-carboxylate (26a)

o o o o

Sodium hydride (60% dispersion in oil, washed with dry petroleum ether,

0.263 g, 6.57 mmole) was added to a dry 100 ml round-bottomed flask,

followed by dry THF (25 ml). To the stirred suspension, under argon, was

added HMPA (0.981 g, 0.952 ml, 5.47 mmole) and a solution of ethyl 1-

tetralone-2-carboxylate (26) (0.800 g, 5.47 mmole) in dry THF (5 ml). The

reaction mixture was stirred for one hour at room temperature after which

time methyl iodide (1.553 g, 0.681 ml, 10.94 mmole) was added. The course

of the reaction was followed by tic chromatography (silica with 9: light

petroleum [b.p. 40 - 60 QC] : ethyl acetate). No starting material (Rt 0.74) was

visible after five minutes, but a new compound had appeared at Rt 0.49. The

reaction was carefully quenched with saturated ammonium chloride solution

(15 ml). This mixture was then poured into a separating funnel containing

distilled water (15 ml) and diethyl ether (30 ml). The layers were separated

and the aqueous layer further extracted with diethyl ether (3 x 30 ml). The

organic extracts were combined, washed with saturated sodium chloride

solution (4 x 15 ml), dried (MgS04) and concentrated in vacuo to give a deep

yellow liquid (1.195 g). The crude product was purified by flash

chromatography using silica with dichloromethane (Rt of product using this

system was 0.44). The product was obtained with a slight impurity as a yellow

orange liquid (0.150g, 18%) and in pure form as a yellow orange liquid (26a)

(0.553g, 65%); Vmax (neat) 3068,2980 (C-H), 1726 (ester C=O), 1684 (ketone

C=O), 1600 and 1452 cm-1 (arene C-C); OH (250 MHz, CDCI3) 8.06 (1 H, d,d,

J7.9, 1.4 Hz, 8-H), 7.50 - 7.20 (3 H, m, 5-H, 6-H, 7-H), 4.14 (2 H, q, J 14.1,

7.0 Hz, CH2CH3), 3.06 - 2.89 (2 H, m, 4-H), 2.66 - 2.56 (1 H, m, , 3-H), 2.11 -

2.00 (1 H, m, 3-H), 1.51 (3 H, s, 2-Me), 1.19 (3 H, t, J7.1 Hz, CH2CH3)·

82

Page 104: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Ethyl 2-fluoro-1-tetralone-2-carboxylate (31a)

o o o o

.. F

To sodium hydride (60% dispersion in oil, 0.092 g, 2.29 mmole), under dry

nitrogen, at 0 QC, was added ethyl 1-tetralone-2-carboxylate (26) (0.500 g,

2.29 mmole) in dry DMF (6 ml) in a dropwise manner. The reaction mixture

was left to stir for thirty minutes at 20 QC, after which time it was cooled to

-50 QC. To the cooled reaction mixture was added N-fluoro, N-chloromethyl

triethylenediamine bis(tetrafluoroborate) (80% active ingredient, 0.922 g, 2.08

mmole) in dry DMF (6 ml). The reaction was left to stir for fifteen minutes at -

50 QC, then allowed to warm slowly to room temperature.'65 The reaction was

followed by tic chromatography (silica with 9 : 1 light petroleum [b.p. 40 - 60

QC] : ethyl acetate). The reaction appeared to be complete after five minutes,

the starting material (Rr 0.53) having given rise to a new compound (Rr 0.23).

The reaction mixture was diluted with diethyl ether (100 ml). This organic

phase then was washed with NaHC03 solution (10% aqueous, 100 ml) and

NaCI solution (saturated aqueous, 100 ml), dried (MgS04) overnight, then

concentrated in vacuo to yield an orange oil (0.799 g). The desired product

was obtained by flash chromatography (silica with 9 : 1 light petroleum [b.p.

40 - 60 QC] : ethyl acetate). Ethyl 2-fluoro-1-tetralone-2-carboxylate (31 a)

was obtained as a clear oil (0.442 g, 90%); vmax (neat) 2984,2944 (C-H),

1763, 1739 (ester C=O), 1697 (ketone C=O), 1603 and 1310 cm-1; OH (250

MHz, CDCI3) 8.07 (1 H, d,d, J7.9, 1.3 Hz, 8-H), 7.55 (1 H, t,t, J7.5, 1.4 Hz, 7-

H), 7.36 (1 H, t, J7.4 Hz, 6-H), 7.28 (1 H, d, J7.7 Hz, 5-H), 4.29 (2 H, q, J7.1

Hz, C02CH2), 3.16 - 3.09 (2 H, m, 4-H), 2.78 - 2.14 (2 H, m, 3-H) and 1.27 (3

H, t, J7.1 Hz, CH3); OF (376.5 MHz, CDCI3, external reference CF3C02H) -

163.7, three minor impurities, all less than 1 %, also noted at -163.6, -164.1

and -189.8; m/z (electron impact) 237.0930000 (mH+, 3%, C'3H'3F03

requires 237.0926889) 216 (18), 170 (76), 163 (17),143 (12),133 (20),118

83

Page 105: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

(100),109 (7). 90 (61),63 (11); Analysis Calculated for C'3H'3F03 (236.245):

C (66.09), H (5.55); Found: C (66.06), H (5.57).

o F

Occasionally, 2-fluoro-1-tetralone was found in the crude product. This was

obtained in purified form as a colourless crystalline solid; V max (neat) 2944 (C­

H), 1703 (C=O), 1603, 1274, 1227,929 cm-'; OH (250 MHz, CDCI3) 8.07 (1 H,

d,d, J7.9, 1.4 Hz, 8-H), 7.53 (1 H, t,t, J7.5, 1.5 Hz, 7-H), 7.36 (1 H, t, J7.8

Hz, 6-H), 7.27 (1 H, d, J7.6 Hz, 5-H), 5.15 (1 H, 2d,d, J47.9, 12.7,5.2 Hz, 2-

H), 3.17 (2 H, m, 4-H), 2.64 - 2.30 (2 H, 2m, 3-H); 8c (62.5 MHz, CDCI3) 193.4

(d, J14.8 Hz, 1-C), 143.1 (8a-C), 134.2 (8-C), 131.2 (4a-C) , 128.7 (7-C),

127.8 (6-C), 127.1 (5-C), 91.2 (d, J187.9 Hz, 2-C), 30.3 (4-C), 30.0 (4-C'),

27.1 (3-C), 26.9 (3-C'); Analysis Calculated for C,oHgFO (164.180): C(73.16),

H(5.53); Found: C(73.09), H(5.50).

84

Page 106: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Methyl 2-fluoro-1-tetralone-2-carboxylate (31 b)

.... Me o

To a stirred suspension of sodium hydride (60% dispersion in oil, 0.098 g,

2.45 mmole) in DMF (2.5 ml), at 0 QC, under nitrogen, was added methyl 1-

tetralone-2-carboxylate (27) (0.500 g, 2.45 mmole) in DMF (2.5 ml). This was

left to stir at 0 °c for twenty minutes. A solution of N-fluoro, N-chloromethyl

triethylenediamine bis(tetrafluoroborate (0.943 g, 2.13 mmole) in DMF (5 ml)

was added to the above solution at -50°C. '65 The reaction mixture was then

allowed to slowly warm to room temperature and left to stir overnight. After

this time, the reaction mixture was poured into a separating funnel containing

diethyl ether (100 ml). This solution was washed with NaHC03 solution (10%

aqueous, 100 ml) and NaCI solution (saturated aqueous, 100 ml), dried

(MgS04 ) and concentrated in vacuo to yield a yellow oil (0.775 g). Tic

analysis (9 : 1 light petroleum [b.p. 40 - 60°C] : ethyl acetate) of the crude

reaction mixture showed the reaction to be complete after approximately five

minutes (starting material: Rt 0.38, product: R, 0.13). The crude product was

then purified by flash chromatography (silica with 9 : 1 light petroleum [b.p. 40

- 60°C] : ethyl acetate, gradually moving to 4 : 1) followed by recrystalization

from ethanol to give clear columnar-shaped crystals (31 b) (0.334 g, 71 %);

m.p. 75°C; vmax (neat) 2955 (aromatic C-H), 1766 (ester C=O), 1744, 1697

(ketone C=O), 1603, 1457 (arene C-C), 1438, 1312, 1281, 1228, 1199,1088

cm-'; OH (250 MHz, CDCI3) 8.06 (1 H, d,d, J7.9, 1.3 Hz, 8-H), 7.55 (1 H, t,t, J

7.5, 1.5 Hz, 7-H), 7.40 - 7.28 (2 H, m, 5-H, 6-H), 3.62 (3 H, s, Me), 3.18 - 3.09

(2 H, m, 3-H), 2.77 - 2.52 (2 H, m, 4-H); Oc (62.5 MHz, CDCI3) 188.5 (d, J

18.8 Hz, 1-C), 167.7 (2-OJ2Me), 143.2 (8a-C), 134.7 (8-C), 130.5 (4a-C),

128.8 (7-C), 128.5 (6-C), 127.3 (5-C), 93.4 (d, J201 Hz, 2-C), 53.1 (CH3),

31.9 (d, J22.5 Hz, 3-C), 24.8 (d, J7.0 Hz, 4-C); m/z (electron impact)

222.0680 (M+, 7.1 %, C'2H"F03 requires 222.06923), 202 (10), 170 (47), 163

(12), 133 (15), 118 (100), 90 (51), 63 (6); Analysis Calculated for C'2H"F03

(222.218): C(64.86), H(4.99); Found: C(64.74), H(4.93).

85

Page 107: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Menthyl 2-fluoro-1-tetralone-2-carboxylate (31 c)

To a stirred suspension of sodium hydride (60% dispersion in oil, 0.091 g,

2.28 mmole) in DMF (1 ml), at 0 QC and under nitrogen, was added menthyl

1-tetralone-2-carboxylate (28) (0.750 g, 2.28 mmole in DMF (4 ml). This was

left to stir at 0 QC for twenty minutes. A solution 01 N-fluoro, N-chloromethyl

triethylenediamine bis(tetralluoroborate) (80% active ingredient, 0.879 g, 1.99

mmole) in DMF (5 ml) was then added dropwise to the above solution at

-50 QC. 165 The solution was allowed to warm slowly to room temperature and

left to stir for ten minutes, after which time the reaction appeared to be

complete by tic chromatography (silica with 20 : 1 light petroleum [b.p. 40 - 60

QC] : ethyl acetate) (Starting material: RIO.80, product: RI 0.61). The reaction

mixture was poured into a separating lunnel containing diethyl ether (100 ml).

This solution was washed with NaHC03 solution (10% aqueous, 100 ml) and

NaCI solution (saturated aqueous, 100 ml), dried (MgS04) and concentrated

in vacuo to yield a yellow oil (0.802 g). Column chromatography (silica with 1 :

.... 1 dichloromethane : light petroleum [b.p. 40 - 60 QC]) yielded a colourless oil

(31c) (0.647 g, 94%).

Separation 01 the diastereomers was carried out by flash chromatography

using silica with 2 : 1 dichloromethane : light petroleum [b.p. 40 - 60 QC]. The

RI values 01 the diastereomers, using this solvent system, were 0.49

(diastereomer I) and 0.41 (diastereomer 11).

86

Page 108: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Diastereomer I (31c-D1): m.p. 101°C; Optical rotation: [alD -61.5 (22°C,

dichloromethane, 0.260 g 1100 ml); vmax (neat) 2956 C-H), 2931,2871

(saturated C-H), 1759 (ester C=O), 1739, 1698 (ketone C=O), 1603, 1456,

1310, 1277, 1227, 1190 cm-1; OH (250 MHz, COCI3) 8_07 (1 H, d,d, J7.8, 1.1

Hz, 8-H), 7.54 (1 H, t,t, J7.5, 1.4 Hz, 7-H), 7.39 (1 H, t, J7.3 Hz, 6-H), 7.27 (1

H, d, J7.6 Hz, 5-H), 4.78 (1 H, t,t, J10.9, 4.4 Hz, C02CHj, 3.24 - 3.05 (2 H,

m, 4-H), 2.76 - 2.51 (2 H, m, 3-H), 2.10 - 2.06 (1 H, m, 2'-H), 1.69 - 0.85 (8 H,

m, 3'-H, 4'-H, 5'-H, 6'-H, CHPr), 0.90 (3 H, d, J6.5 Hz, 5'-Me), 0.77 (1 H, d, J

7.0 Hz, Me of Pr), 0.69 (1 H, d, J7.0 Hz, Me' of Pr); 3c (62.5 MHz, COCI3) .

188.7 (d, J15.1 Hz, 1-C), 167.0 (d, J25.2 Hz, C02), 142.8 (8a-C), 134.4 (7-

C), 130.8 (4a-C) , 128.7 (6-C), 128.3 (5-C), 127.2 (8-C), 93.4 (d, J193.5 Hz,

2-C), 77.0 (1'-C), 46.7 (2'-C), 40.2 (6'-C), 34.0 (3'-C), 31.7 (d, J21.9 Hz, 3-C),

31.4 (5'-C), 25.9 (a-lPr), 25.0 (d, J7.7 Hz, 4-C), 23.0 (4'-C), 21.9 (Me), 20.7

(Me'), 15.8 (Me"); OF (376.5 MHz, COCI3 , 1H decoupled, external reference

CF3C02H) -163.87 ; mlz (electron impact) no M+ found. Peaks present

suggest compound has undergone a MCLafferty Type Rearrangement (~­

cleavage with y-hydrogen transfer) to give mass ion peak of 208 (89, M+-

138),170 (42),164 (21, M+-138 - C02), 138 (15),118 (27), 95 (31),83

(100), 69 (37), 55 (43), 41 (25); Analysis Calculated for C21H27F03 (346.445):

C(72.81), H(7.86); Found: C(72.56), H(7.98). X-ray crystallographic data:167

Crystal (0.50 x 0.21 x 0.20 mm) grown from methanol, C21H27F03, monoclinic,

space group P21, a =11.878(9), b = 12.694(7), c = 6.43(1) A, ~ = 90.2(2)°, Z =

2, Ox = 1.186 gcm-3, T = 293 K. MoKa radiation (/0.. = 0.7169 A), 11 = 0.48 cm-1.

Stoe Stadi-2 diffractometer, 1857 reflections measured of which 1360 with

F/cr(F»5. Structure solved by direct methods and refined to a final R = 0.070,

Rw = 0.070. All non-H atoms treated anisotropically. Phenyl and one

87

Page 109: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

methylene H atoms placed in calculated positions. Other H atoms found from

difference map and not refined. The final maximum shift/error = 0.39, IIp

excursions -0.2 to 0.2.

Diastereomer 11 (31c-DII): V max (neat) 2956 (aromatic C-H), 2930, 2871

(saturated C-H), 1760 (ester C=O), 1734, 1698 (ketone C=O), 1601, 1456,

1311,1275,1227,1189 cm"; OH (250 MHz, CDCI3) 8.07 (1 H, d, J7.8 Hz,

8-H), 7.55 (1 H, t, J7.5 Hz, 7-H), 7.36 (1 H, t, J7.5 Hz, 6-H), 7.29 (1 H, d, J

6.8 Hz, 5-H), 4.80 (1 H, t,t, J10.9, 4.4 Hz, C02CH), 3.23 - 3.09 (2 H, m, 4-H),

2.73 - 2.53 (2 H, m, 3-H), 2.06 - 2.02 (1 H, m, 2'-H), 1.75 - 0.85 (8 H, m, 3'-H,

4'-H, 5'-H, 6'-H, CHFr), 0.89 (3 H, d, J6.5 Hz, 5'-Me), 0.81 (3 H, d, J7.0 Hz,

Me of Fr), 0.74 (3 H, d, J7.0 Hz, Me' of Pr); Oc (62.5 MHz, CDCI3)188.6 (d, J

15.1 Hz, 1-C), 167.0 (d, J25.2 Hz, CO2), 143.0 (8a-C), 134.4 (7-C), 130.9

(4a-C), 128.7 (6-C), 128.3 (5-C), 127.2 (8-C), 93.1 (d, J 193.8 Hz, 2-C), 77.0

(1'-C), 46.8 (2'-C), 40.5 (6'-C), 34.0 (3'-C), 32.0 (d, J22.4 Hz, 3-C), 31.4 (5'­

C), 25.9 (Q-iFr), 25.1 (d, J7.7 Hz, 4-C), 23.2 (4'-C), 21.9 (Me), 20.8 (Me'),

. J 6.0 (Me").

88

Page 110: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Methyl 2-fluoro-1-[( RH + )-a-methylbenzyli mi no ]-tetralone-2-carboxylate

(32)

o o

F ..

To a solution of methyl 2-fluoro-1-tetralone-2-carboxylate (31 b) (2.780 g,

12.510 mmole) and (R}(+)-a-methylbenzylamine (4.548 g, 4.838 ml, 37.531

mmole) in dry benzene (100 ml), at 0-5 QC, was added TiCI4 (1.0 M in

dichloromethane, 6.88 ml, 6.881 mmole) dropwise, via syringe, under

nitrogen. 17o The reaction was left to stir at 0-5 QC for one hour, then allowed

to warm to room temperature and left to stir for a further one hour. After this

time, the reaction mixture was filtered and the filtrate concentrated in vacuoto

yield a reddish white solid (still containing Ti02). This material was purified by

flash chromatography using alumina (basic, activated, Brockmann Type I,

standard grade, 150 mesh) with 19 : 1 light petroleum [b.p. 40 - 60°C] : ethyl

acetate. The solvent was removed in vacuo from the desired fractions to yield

a sticky, pale yellow solid (3.651 g, 90%). Two spots were visible by tic (silica

with 7 : 3 light petroleum [b.p. 40 - 60 QC] : diethyl ether) appearing at Rf 0.53

and 0.45 (cream ish yellow colour when visualised byanisaldehyde).

Recrystallization from methanol yielded a white solid [Diastereomer 11 (32-

011)]. Further purification of the mother liquor using dry flash chromatography

(tic alumina, neutral, with light petroleum [b.p. 40 - 60 QC]) and

recrystallizations using methanol yielded the isolated diastereomers.

Diastereomer I (32-01): colourless oil (0.196 g, 11 %); V max (film) 2926

(aromatic C-H), 1758 (ester C=O), 1634, 1456, 1267, 1089 cm·1; OH (250

MHz, CDCI3 ) 8.24 (1 H, d,d, J6.5, 1.6 Hz, ArH), 7.42 -7.14 (8 H, m, ArH),

5.13 - 5.03 (1 H, m, NCH), 3.33 (3 H, s, OMe protons), 3.00 - 2.77 (2 H, m, 4-

H), 2.48 - 2.15 (2 H, m, 3-H), 1.55 (3 H, d, J6.2 Hz, CMe protons); Oc (62.5

89

Page 111: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

MHz, CDCI3) 170.4 (d, J26.1 Hz, C02Me), 154.2 (d, J 15.6 Hz, 1-C), 145.4

(4a-C), 139.1 (NCH[Me)C), 134.0 (d, J3.9 Hz, 8a-C) , 130.2, 128.2 (2 C),

127.5, 127.4, 127.1, 126.6, 126.2 (2 C), 91.1 (d, J 192.2 Hz, 2-C), 60.5 (d, J

3.4 Hz, C02CH3), 52.4 (NCH), 34.6 (d, J23.3 Hz, 3-C), 26.0 (CHCH3). 24.8

(d, J 4.5 Hz, 4-C). Evidence of syn and antiisomers present in the 13C n.m.r.

spectrum. The aromatic tertiary carbons of the minor isomer are visible at Oc 130.4,128.8,128.6,126.9,126.7, 126.0, 125.7. Also visible are Oc 60.0

(C02 a-i3) , 52.3 (NCH), 31.4 (d, 3-C), 26.5 (CHCH3), 25.1 (d, 4-C); m/z

(electron impact) 325.1486 (M+, 6.4%, C2oH20FN02 requires 325.14780),77

(12), 105 (100),118 (9),146 (4),170 (4), 246 (4).

Diastereomer 11 (32-011): colourless crystalline solid (0.281 g, 16%); m.p.

101°C; vmax (film) 2963 (aromatic C-H), 1761 (ester C=O), 1627, 1450, 1265,

1071 cm-1; 8H (250 MHz, CDCI3) 8.26 (1 H, d,d, J6.7, 2.5 Hz, 6-H), 7.47 (2 H,

d, J7.2 Hz, ArH), 7.33 -7.12 (6 H, m, ArH), 5.13 - 5.06 (1 H, m, NOt), 3.87*

& 3.83 (1 H, singlets, OMe protons of syrr and anti- isomers), 2.91 - 2.85 (2

H, m, 4-H), 2.44 - 2.29 (2 H, m, 3-H), 1.60 & 1.42' (3 H, doublet & double

doublet, J6.4 & 6.3, 1.2 Hz, respectively, CMe protons); Oc (62.5 MHz,

CDCI3) 170.7* (d, J25.5 Hz, C02Me), 170.1 (d, J27.1 Hz, C02Me), 159.2 (d,

J20.7 Hz, 1-C), 153.7* (d, J15.7 Hz, 1-C), 145.9 (4a-C) , 145.1' (4a-C) ,

142.4 (NCH[Me)C), 140.8' (NCH[Me)C), 138.1' (d, J3.9 Hz, 8a-C), peaks of

tertiary aromatic carbon of syn and anti isomers: 130.5, 130.1, 128.8, 128.7,

128.4,128.5,127.9,127.8,127.0,126.8,126.7,126.2, 126.0, 125.9,96.1 (d,

J188.8 Hz, 2-C), 90.7* (d, J192.7 Hz, 2-C), 60.6' (d, J4.4 Hz, C02CH3),

59.7 (C02CH3), 52.9' (NCH), 52.3 (NCH), 34.4' (d, J23.4 Hz, 3-C), 32.0 (d, J

24.0 Hz, 3-C), 25.9 (CHCH3), 25.4 (d, J3.0 Hz, 4-C), 24.9 (d, J5.8 Hz, 4-C),

23.9' (CHCH3); m/z (electron impact) 325.1491 (M+, 5.5%, C2oH2QFN02

requires 325.14780),77 (10). 105 (100),117 (6),146 (3),246 (3)

, Peaks of most abundant isomer.

90

Page 112: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Regeneration Of Methyl 2-fluoro-1-tetralone-2-carboxylate (31 b-EI and

31 b-EII)

MeyPh

N 0 0 0

o .... Me o .... Me

F • F

A heterogeneous mixture of hydrochloric acid (aq. 2 M, 15 ml) and methyl

2-fluoro-1-[(R)-( + )-a-methylbenzylimino]-tetralone-2-carboxylate (32) (0.218

g, 0.670 mmole) in dichloromethane (5 ml) was vigorously stirred for 1 hour at

room temperature. After this time, no starting material was visible by tic. The

organic layer was isolated, made to 20 ml, washed with brine (saturated, 20

ml), dried (Na2S04) and concentrated in vacuo to yield a colourless solid

(0.147 g, 99%). This material was run through a silica column to give a

colourless solid (31 b) (0.143 g, 96%).

Removal of the imine from (32-011) yielded the enantiomerically pure

ketone (31 b-EII), and deprotection of (32-01) gave the other enantiomer

(31 b-EI). Both compounds had spectral and physical data as previously

described for (31 b).

Chiral shift studies were carried out on the isolated enantiomers, using the

chiral shift reagent Eu(hfch The proton n.m.r. spectrum for (31b-EII) showed

no splitting of peaks with the chiral shift reagent. On addition of a racemic

mixture of the ketone to the n.m.r. sample, splitting was observed. From this it

can be deduced that the parent imine consists of the syn and anti isomers of

the same diastereomer, either (R,R) or (R,S).

NB! The Eu(hfcb shifts the 6-H proton of (31 b-EII) further downfield than that

of (31 b-EI). Baseline separation was observed for the most downfield proton

with 2.7 equivalents of the chiral shift reagent.

91

Page 113: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

2-FI uoro-2-(2-hydroxyisopropyl)-1-[( RH. + )-a.-methylbenzylimino]­

tetralone (33)

"Me o

To a solution of the iminoester (32) (1.836 g, 5.643 mmole) in

tetrahydrofuran (dry, 25 ml) was added methyllithium (1.5 M in diethyl ether,

5.64 ml, 8.464 mmole) at 0 QC under nitrogen. The reaction was allowed to

stir at 0 QC for 1.5 hours, after which time it was quenched with an aqueous

solution of ammonium chloride (saturated, 50 ml). The resulting mixture was

extracted with diethyl ether. The ethereal extracts were combined, dried

(Na2S04) and concentrated in vacuo to yield a dark orange oil. This oil was

subjected to flash chromatography (aluminium oxide, Brockmann Grade 11,

standard grade, -150 mesh; 7 : 3 light petroleum [b.p. 40 - 60 QC] : ethyl

acetate). This gave 0.399 g (22%) of the desired hydroxyimine

diastereomers. This initial column afforded some separation of the

diastereomers: the more polar diastereomer (33-01) (0.093 g, 5%), the less

polar diastereomer (33-011) (0.152 g, 8%). Some staring material was

recovered as a colourless oil (0.894 g, 49%).

Spectral data for the more polar diastereomer (33-01): vmax (neat) 3354

(broad, OH), 2975, 2926,1641,1453,1073,1058,964 cm·\ Optical rotation

[ajo +37.0 (20 QC, chloroform, 0.568 g / 100 ml); OH (250 MHz, CDCI3) 7.51 -

7.15 (9 H, m, ArH), 6.20 (1 H. broad s, OH), 504 (1 H, q, J6.4 Hz, NCh),

303 - 2.78 (2 H, m, 4-H), 2.47 - 2.15 (2 H, m, 3-H), 1.51 (3 H, s, C(OH)CH3),

1.49 (3 H, d, J3.4 Hz, NCC}-6'), 1.38 (3 H, S, C(OH)CH3"); Oc (62.5 MHz,

CDCI3 ) 166.6 (d, 24.4 Hz, C=N), 145.1 (NCH[MejC), 141.2 (8a-Cl. 130.3,

129.6 (4a-C), 128.7, 128.4, 127.8, 126.9, 126.1, 125.5,96.9 (d, J188.8 Hz,

92

Page 114: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

2-C), 74.2 (d, J24.4 Hz, COH), 60.2 (Q-iMe), 29.6 (d, J23.9 Hz, 3-C), 25.9

(C[OH]CH3), 25.5 (4-C), 25.5 (NCQ-i3'), 24.8 (C[OH]CH3"); m/z (electron

impact) 325.1836 (M+, 0.7%, C2QH2QFN02 requires 325.1842).

Spectral data for the less polar diastereomer (33-DII): vmax (neat) 3379

(broad, OH), 2976, 2940, 1455, 1367, 1072, 964 cm-'; Optical rotation [a]D

+200.7 (20°C, chloroform, 0.304 9 1100 ml); OH (250 MHz, CDCI3) 7.38 - 7.22

(9 H, m, ArH), 6.25 (1 H, broad s, OH), 5.03 (1 H, q, J6.4 Hz, NCh), 3.04-

2.66 (2 H, m, 4-H), 2.46 - 1.83 (2 H, m, 3-H), 1.58 (3 H, d, J6.2 Hz,

NCCH3 ) , 1.43 (3 H, s, C(OH)Cff:l'), 1.13 (3 H, s, C(OH)Cff:l"); Oc (62.5 MHz,

CDCI3) 166.0 (d, 24.4 Hz, C=N), 145.1 (NCH[Me]C), 142.0 (8a-C), 130.4,

129.3 (4a-C) , 128.8, 128.2, 128.0, 127.0, 126.0, 125.7,96.8 (d, J181.9 Hz,

2-C), 74.4 (d, J23.4 Hz, COH), 61.0 (Q-iMe), 30.5 (d, J23.8 Hz, 3-C), 26.2

(C[OH] Q-i3') , 25.7 (4-C), 25.7 (NCQ-i3)' 24.3 (C[OH]Q-i3"); m/z (electron

impact) 325.1828 (M+, 1.1 %, C2QH20FN02 requires 325.1842).

93

Page 115: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Regeneration of 2-fluoro-2-{2-hydroxyisopropyl)-1-tetralone (34-EI and

34-EII)

..

A heterogeneous solution of the hydroxyimine (33-01) (0.301 g, 0.925

mmole) in dichloromethane (10 ml) and dilute hydrochloric acid (2 M, 20 ml)

was vigorously stirred for 20 hours. After this time no starting material was

apparent by tic analysis (7 : 3 light petroleum [b.p. 40 - 60 QC] : diethyl ether,

starting material Rt 0.30 and product Rt 0.19). The aqueous layer was

extracted with dichloromethane (2 x 50 ml). The organic extracts were dried

(Na2S04) and concentrated in vacuo to yield an orange oil. this was purified

by flash chromatography ( basic alumina, Brockmann Grade 11 with 7 : 3 light

petroleum [b.p. 40 - 60 QC] : diethyl ether) The desired fractions were

combined to give a colourless solid (34-EI); m.p. 74 QC; Optical rotation: [alo -

7.6 (25 QC, chloroform, 0.724 g 1100 ml); V max (film) 3781 (broad, OH, H­

bonded), 2980 (Aryl C-H), 1684 (C=O), 1601, 1456 (Arene C-G) and 1229

cm-'; 8H (250 MHz, CDCI3) 8.01 (1 H, d,d, J8.0, 1.0 Hz, 8-H), 7.52 (1 H, t,d, J

7.3, 1.3 Hz, 7-H), 7.34 (1 H, t, J7.5 Hz, 6-H), 7.25 (1 H, d, J7.6 Hz, 5-H),

3.54 (1 H, broad s, OH), 3.27 - 2.96 (2 H, m, 4-H), 2.58 - 2.21 (2 H, m, 3-H),

1.38 (3 H, s, Me), 1.37 (3 H, s, Me); 8c (62.5 MHz, CDCI3) 201.4 (1-G), 143.8

(8a-C), 134.2 (7-C), 132.0 (4a-C), 128.5 (6-C), 128.2 (5-C) , 126.9 (8-C), 95.3

(d, J182.6 Hz, 2-C), 74.0 (d, J22.6 Hz, COH), 30.0 (d, J22.8 Hz, 3-C), 25.8

(d, J3.6 Hz, Me), 24.9 (d, J7.8 Hz, 4-C), 24.1 (d, J4.1 Hz, Me); m/z (electron

impact) 222.1061 (M+, 0.4%, C"H,oF20 2 requires 222.1056), 164 (100), 144

(13),131 (9),118 (20), 90 (22),77 (6),63 (8), 59 (33),51 (7),43 (73),31 (8).

94

Page 116: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Ethyl 1-indanone-2-carboxylate (36)161, m

.. ..... Et

o

1-lndanone (35) (0.200 g, 1.51 mmole) in dry diethyl carbonate (distilled

from sodium hydride, 5.5 ml, 45.4 mmole) was added, via syringe, to a stirred

suspension of sodium hydride (0.073 g, 1.82 mmole) in dry diethyl carbonate

(5.5 ml, 45.5 mmole). The solution was heated to reflux for five minutes.

During this time a dark green solid was formed (note that on a larger scale

using 2.00 g of 1-indanone (35). an exotherm was noted and no additional

heat was required for the formation of this solid). Further diethyl carbonate (5

ml) was added. Tic analysis of the solid (silica with 9 : 2 light petroleum [b.p.

40 - 60 QC] : ethyl acetate) showed no evidence of any 1-indanone (35) (Rt

0.47), only the product (36) (Rt 0.56). The reaction mixture was heated to

reflux for a further fifteen minutes, after which time it was allowed to cool to

room temperature. The cooled solid was dissolved in HCI (aqueous 2 M, 50

ml). The aqueous phase was extracted with ethyl acetate (4 x 50 ml). The

combined organic extracts were dried (MgS04) and concentrated in vacuo to

yield a brown oil (0.611 g). The sample was purified by column

chromatography (silica with 9 : 1 light petroleum [b.p. 40 - 60 QC] : ethyl

-·acetate) to yield a colourless oil (36) (0.265 g, 86%); V max (neat) 3009

(aromatic C-H), 1740 (ester C=O of ketoester), 1714 (a-aryl ketone C=O),

1650 (enol ester C=O), 1573, 1465 (arene C-C), 1370, 1258, 1208, 1187,

1153, 1093 cm"; OH (250 MHz, CDCI3) 12.5 (1 H, broad s, OH), 7.74 - 7.32 (8

H, m, keto ester and enol ArH), 4.74 - 4.12 (4 H, m, ketoester and enol

CH2CH3), 3.67 (1 H, d,d, J8.2, 4.1 Hz, 2-H), 3.68 - 3.29 (4 H, m, ketoester

and enol 3-H), 1.38 - 1.26 (6 H, m, ketoester and enol CH2CH3); Oc (62.5

MHz, CDCI3) 199.5 (ketoester 1-C), 169.1 (ketoester 2-C02Et), 153.7 (enoI1-

C), 143.0 (ketoester 7a-C), 135.3 (ketoester 7 -C), 135.2 (ketoester 3a-C),

129.3 (enoI7-C), 127.7 (ketoester 6-C), 126.8 (enoI6-C), 126.6 (ketone 5-C),

124.7 (enoI5-C), 124.4 (ketone 4-C), 120.5 (enoI4-C), 102.5 (enoI2-C), 61.5

(ketoester CH2CH3), 60.0 (enol CH2CH3), 53.3 (ketoester 2-C), 32.5 (enol 3-

C), 30.3 (ketone 3-C), 14.4 (enol CH2 CH3), 14.2 (ketoester CH2 a-i3)·

95

Page 117: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Ethyl 2-fluoro-1-indanone-2-carboxylate (37)165

o ....-Et o

o ....-Et o

To a stirred suspension of sodium hydride (60% dispersion in oil, 0.098 g,

2.45 mmole) in DMF (0.5 ml), at 0 QC and under nitrogen, was added ethyl 1-

indanone-2-carboxylate (36) (0.500 g, 2.45 mmole in DMF (2.5 ml). This was

left to stir at 0 QC for thirty minutes. A solution of N-fluoro, N-chloromethyl

triethylenediamine bis(tetrafluoroborate) (80% active ingredient, 0.986 g, 2.23

mmole) in DMF (5 ml) was then added dropwise to the above solution at -50

QC. The solution was allowed to warm slowly to room temperature and left to

stir for five minutes, after which time the reaction appeared to be complete by

tic chromatography (silica with 9 : 2 light petroleum [b.p. 40 - 60 QC] : ethyl

acetate) (Starting material: R, 0.48 and 0.17, with streaking between; Product:

R, 0.30). The reaction was left to stir for a further twenty five minutes. After

this time, the reaction mixture was poured into a separating funnel containing

diethyl ether (100 ml). This solution was washed with NaHC03 solution (10%

aqueous, 100 ml) and NaCI solution (saturated aqueous, 100 ml), dried

(MgS04

) and concentrated in vacuoto yield a yellow oil (37) (0.635 g). The

crude product was then purified by flash chromatography (silica with 9 : 1 light

petroleum [b.p. 40 - 60 QC] : ethyl acetate) to yield a very pale yellow oil

(73%); V max (neat) 2986 (aromatic C-H), 1766 (ester C=O), 1728 (ketone

C=O), 1608, 1467, 1298, 1216, 1193,1074 cm-1; OH (250 MHz, CDCI3) 7.82

(d, 1 H, J7.6 Hz, 7-H), 7.72 (t, 1 H, J7.6 Hz, 6-H), 7.54 - 7.47 (m, 2H, 5-H, 4-

H), 4.27 (q, 2H, J7.1 Hz, CH2CH3), 3.81 (d,d, 1 H, J 17.7, 11.7 Hz, 3-H), 3.43

(d,d, 1H, J23.4, 17.7 Hz, 3-H'), 1.24 (t, 3H, J7.1 Hz, CH3); lie (62.5 MHz,

CDCI3

) 195.3 (d, J17.5 Hz, 1-C), 167.3 (d, J27.2 Hz, G02Et) , 150.9 (d, J3.2

Hz, 7a-C), 136.7, 133.3 (3a-C), 128.6, 126.6, 125.6,94.5 (d, J200.3 Hz, 2-

C), 62.6 (CH2CH3), 38.3 (d, J23.5 Hz, 3-C), 14.0 (CH2 CH3)·

96

Page 118: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Ethy 1 2-11 uoro-1-[( RH + )-a-methylbenzylimino ]-indanone-2-carboxylate

(38)

o o

F

...-Et o

To a stirring solution of ethyl 2-fluoro-1-indanone-2-carboxylate (37) (3.973

g, 17.88 mmole) and a-methylbenzylamine (6.500 g, 6.92 ml, 53.64 mmole)

in benzene (dry, 100 ml) was added titanium tetrachloride (1.0 M in

dichloromethane, 9.83 ml, 9.83 mmole), at 0 QC and under nitrogen. After one

hour atO QC, the reaction was allowed to warm to room temperature and left

to stir for a further one hour. After this time, the benzene was removed in

vacuo, dichloromethane (100 ml) was added, followed by diethyl ether (50

ml). The resulting suspension was filtered through glasswool. Solvent was

removed in vacuo from the resulting filtrate, which was then purified by flash

chromatography (basic alumina, Brockmann Grade I with 9 : 1 light petroleum

[b.p. 40 - 60 QC]). The imine diastereomers eluted together and were

concentrated in vacuo to yield a colourless residue.

Diastereomer I (38-01) colourless crystalline solid (1.334 g, 23%); m.p. 105

QC; Optical rotation: [alo +13.6 (25 QC, chloroform, 0.568 g 1100 ml); Vrnax

(neat) 2981, 1756, 1659, 1190, 1062 cm-1; 8H (250 MHz, CDCI3) 7.93* (1 H, d,

J7.4 Hz, ArH), 7.77 (1 H, d, J7.7 Hz, ArH), 7.44 -7.19 (8 H from synand 8 H

from antiisomer, m, ArH), 5.42 (1 H, d,q, J2.3, 6.5 Hz, NCH), 5.10* (1 H, d,q,

J2.2, 6.2 Hz, NCH), 4.47 - 4.22 (2 H, m, CH2CH3), 4.07 - 3.89* and 3.81 -

3.68* (2 H, 2 m, CH2CH3), 3.68 - 3.26 (2 H from syn and 2 H from antiisomer,

m, 3-H), 1.60* (3 H, d, J6.3 Hz, Me protons), 1.59 (3 H, d, J6.4 Hz, Me

protons), 1.30 (3 H, t, J7.1 Hz, CH2CH3), 0.84* (3 H, t, J7.2 Hz, CH2CH3);

The isomers of this compound are in a ratio of 9 : 1. This proton spectrum

was decoupled at the methyl group (CH 2CH3 ) appearing at 8H 0.82, causing

97

Page 119: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

the multiplets of the methylene protons (CI-kCH3)at &; 4.07 - 3.89 and 3.81 -

3.68 to become doublets; lie (62.5 MHz, CDCI3) 169.6* (d, J28.7 Hz,

C02Me). 162.4* (d, J15.3 Hz, N=C), 146.4 (N-CH(CH3)C), 145.3* (N­

CH(CH3)C), 143.7* (d, J3.7 Hz, 7a-C) , 137.6* (d, J3.0 Hz, 3a-C), 132.2:

128.7,128.5,* 128.2,* 128.0,127.7,126.7,* 126.6,126.2: 125.1: 123.6:

94.3* (d, J197.4 Hz, 2-C), 62.0* (CH2CH3), 61.7 (CH2CH3). 60.8* (NCH), 59.4

(NCH), 42.2' (d, J25.4 Hz, 3-C), 39.6 (d, J24.9 Hz, 3-C), 25.9' (Me carbon),

25.1 (Me carbon), 14.3 (CH2CH3), 13.3* (CH2CH3). NB. The quaternary and

aromatic carbons of the minor isomer are not all visible in the carbon

spectrum. Seven peaks are present for the aromatic CH carbons of the major

isomer; only four peaks can be clearly identified for those of the minor isomer;

Analysis Calculated for C20H2oFN02 (325.380): C(73.81), H(6.20), N (4.31);

Found: C(74.14), H(6.41), N (4.30).

'Peaks of most abundant isomer.

Diastereomer (38-011): this contained approximately 20% of (38-01).

Identifiable peaks in the 'H n.m.r. spectrum are OH (250 MHz, CDCI3) 7.90' (1

H, d, J7.9 Hz, ArH), 7.52 - 7.21 (8 H from syn and 8 H from antiisomer, m,

ArH), 5.50 - 5.40 (1 H, NCH), 5.12* (1 H, d,q, J 2.2, 6.4 Hz, NCH), 4.41 -

4.26* (2 H, m, CH2CH3), 4.07 - 3.68 (2 H, 2 m, CH2CH 3), 3.68 - 3.26 (2 H from

syn and 2 H from antiisomer, m, 3-H), 1.68 (3 H, d, J6.6 Hz, NCHCH3), 1.49'

(3 H, d, J6.5 Hz, NCHCH3), 1.32* (3 H, t, J7.1 Hz, CH2 CH3), 1.19 (3 H, t, J

7.2 Hz, CH2 CH3).

98

Page 120: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Regeneration Of Ethyl 2-fluoro-1-indanone-2-carboxylate (37)

...-Et o

o o

F

Ethyl 2-fluoro-1-[(R)-(+)-a-methylbenzylimino]-indanone-2-carboxylate (38-

01) (0.338 g, 1.038 mmole) was dissolved in dichloromethane (7.5 ml). To

this was added 2 M HCI (7.5 ml) and the solution vigorously stirred at room

temperature for one and a half days. The aqueous layer was then extracted

with dichloromethane (3 x 100 ml). The organic extracts were dried (Na2S04),

and concentrated in vacuo to yield an orange oil (0.229 g, 99%). this oil was

purified by flash chromatography (flash silica with 9 : 1 light petroleum [b.p.

40 - 60 cC] : diethyl ether) to yield a colourless solid (37-EI)(0.185 g, 80%);

Optical rotation [a.]D -86.4 (25 cC, chloroform, 0.428 g 1100 ml). Yield not

optimised since still column fractions containing desired material in them, but

in impure form. Spectral data as for racemic (37).

Removal of the imine from (38-01) yielded the enantiomerically pure ketone

(37-EI), and deprotection of (38-011) gave the other enantiomer (37-EII). Both

compounds had spectral and physical data as previously described for the

racemic material (37).

Chiral shift studies were carried out on the isolated enantiomers, using the

chiral shift reagent Eu(hfch The proton n.m.r. spectrum for (37-EI) showed

no splitting of peaks with the chiral shift reagent. On addition of a racemic

mixture of the ketone to the n.m.r. sample, splitting was observed. From this it

can be deduced that the parent imine consists of the syn and anti isomers of

the same diastereomer, either (R,Rj or (R,S).

The Eu(hfcb shifts the 6-H proton of (37-EI) further up field than that of

(37-EII). Baseline separation was observed for the most downfield proton with

2.7 equivalents of the chiral shift reagent.

99

Page 121: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Synthesis of methyl 2,5,7-trifluoro-1-indanone-2-carboxylate (47)

Ethyl 3-(3,5-difluoro)-2,3-dehydropropionate (40)

o F ., F

H

F F

A dry three-necked flask (100 ml), equipped with a magnetic stirrer,

thermometer, condenser and dropping funnel was purged with dry nitrogen,

then charged with sodium hydride (60% dispersion in oil, 0.464 g, 11.61

mmole) and dry benzene (5 ml, filtered from sodium hydride immediately

before use). To this stirred suspension, in an ice-water bath, was added

triethyl phosphonoacetate (2.082 g, 9.29 mmole) in dry benzene (7.5 ml),

over a twenty minute period, in a dropwise manner.'73 After this addition was

complete, the mixture was stirred for twenty minutes at 0 QC. 3,5-Difluoro­

benzai'dehyde (39) (1.100 g, 7.74 mmole) was added dropwise to this

solution, resulting in the formation of a gummy precipitate of sodium ethyl

phosphate. A further portion of benzene (5 ml) was added to aid stirring. The

solution was allowed to warm slowly to room temperature and left to stir

·overnight. Water (50 ml) was then added carefully to the reaction mixture

followed by ethyl acetate (35 ml). The organic layer was isolated, then the

aqueous layer further extracted with ethyl acetate (2 x 50 ml). The combined

organic layers were washed with water (100 ml), dried (MgS04) and

concentrated in vacuo to yield a yellow residue (1.693 g). The product was

purified by column chromatography (silica with 19 : 1 light petroleum [b.p. 40 -

60 QC] : ethyl acetate). Tic analysis of the fractions was carriec;i out using silica

with 9 : 1 light petroleum [b.p. 40 - 60 QC] : ethyl acetate, the product having

an RI value of 0.62 with this system. The resulting solid was recrystallized

from light petroleum [b.p. 40 - 60 QC] to give a colourless solid (40) (1.496 g,

91%); mpt. 38.5 - 39.0 QC; V max (neat) 2985 (alkene C-H), 1716 (C=O), 1645,

1621,1593 (a-carbonyl C=C), 1453 (arene C-C), 1440, 1311, 1280, 1181,

100

Page 122: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

1121 cm"'; 8H (250 MHz, CDCla) 7.57 (1 H, d, J 16.0 Hz, Cf£0 2) , 7.03 (2 H,

d,d, J8.1, 1.9 Hz, Cf£FCHCFCH), 6.87 - 6.78 (1 H, m, CFCf£F), 4.28 (2 H,

q, J7.2 Hz, CH2CHa), 1.34 (3 H, t, J7.2 Hz, CH2CHa). Coupling constant of

alkene protons shows double bond to be trans.; m/z (electron impact)

212.0502 (M+, 9.4%, C"H,oF20 2 requires 212.06489), 167 (100, M+ - OEt),

139 (22, M+ - [OEt + CO]), 127 (38),101 (11),63 (4).

101

Page 123: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Ethyl 3-(3,5-difluoro)-2,3-dehydropropionate (40)

o F F

H •

F F

To 3,5-difluorobenzaldehyde (39) (4.1134 g, 28.95 mmole) in dry

dichloromethane (12 ml), under nitrogen and at room temperature, was

carefully added (carboxymethylene)triphenylphosphorane (12.10 g, 34.74

mmole) in dry dichloromethane (12 ml).174 This produced an immediate

exotherm. The reaction was left to cool to room temperature, after which time

it appeared to have gone to completion by tic. Tic analysis was carried out

using silica with 9 : 1 light petroleum [b.p. 40 - 60°C) : ethyl acetate, the

product having an R, value of 0.62 with this system. The dichloromethane

was removed in vacuo to give a colourless solid. To this was added hexane

(3 x 100 ml) and the mixture stirred for several minutes before the hexane

was collected by filtration. The hexane fractions were concentrated in vacuo

to yield an off-white solid (6.796 g). The product was purified by column

chromatography (silica with 19 : 1 light petroleum [b.p. 40 - 60°C] : ethyl

acetate) to give a colourless solid (40) (6.121 g, 99.7%); mpt. 38.5 - 39.0 °C

(recrystallized from light petroleum [b.p. 40 - 60 0c); Vmax (neat) 2985

(alkene C-H), 1716 (C=O), 1645, 1621, 1593 (a-carbonyl C=C), 1453 (arene

C-C), 1440, 1311, 1280, 1181, 1121 cm'1; OH (250 MHz, CDCI3) 7.57 (1 H, d,

J 16.0 Hz, CHC02) , 7.03 (2 H, d,d, J8.1, 1.9 Hz, ChCFCHCFCh), 6.87 - 6.78

(1 H, m, CFCHCF), 4.28 (2 H, q, J7.2 Hz, CH2CH3), 1.34 (3 H, t, J7.2 Hz,

CH2CH3). Coupling constant of alkene protons shows double bond to be

trans.; (Found: m/z 212.0502 (M+·). C11H10F202 requires 212.0649).

102

Page 124: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Ethyl 3-(3,5-difluorophenyl)propionate (41)

F F

F F

To a round bottomed flask (250 ml) was added ethyl 3-(3,5-difluorophenyl)-

2,3-dehydropropionate (40) (1.464 g, 6.90 mmole), 10% Pd/C (0.150 g), ethyl

acetate (100 ml) and a magnetic stirrer. The flask was evacuated, then filled

with hydrogen. This procedure was repeated six times to ensure an

atmosphere of hydrogen. The solution was then left to stir overnight at room

temperature. After this time, the reaction appeared to have gone to

completion by tic analysis (silica with 9 : 1 light petroleum [b.p. 40 - 60 QC] :

ethyl acetate) : The Rt value of the starting material was 0.61 and that of the

product 0.54. The reaction mixture was filtered through celite to remove the

palladium catalyst, then concentrated in vacuo to yield a clear liquid (1.461 g,

99%). This material was purified by column chromatography (silica with 19 : 1

light petroleum [b.p. 40 - 60 QC] : ethyl acetate) to give a clear oil (41) (1.428

g, 97%); vmax (neat) 3093 (aromatic C-H), 2984 (alkene C-H), 2938 (saturated

C-H), 1736 (C=O), 1629, 1597, 1462 (arene C-C), 1376, 1186, 1161 cm-'; OH

(250 MHz, CDCI3) 6.76 (2 H, d,d, J6.4, 2.1 Hz, CFChCCH2), 6.72 - 6.59 (1

H, m, CFChCF), 4.14 (2 H, q, J7.0 Hz, CH2CH3), 2.93 (2 H, t, J7.3 Hz,

CH2C02Et) , 2.61 (2 H, t, J7.3 Hz, ArCH2 ), 1.24 (3 H, t, J7.0 Hz, CH2CH3); Oc

(62.5 MHz, CDCI3) 172.3 (C02), 162.9 (2 C, d,d, J246.5, 12.7 Hz, C-F),

144.6 (t, J9.2 Hz, QCH2h C02Et), 111.2 - 110.9 (2 C, m, CF CHCCH2), 101.6

(t. J25.3, CFCHCF), 60.5 (CH2CH3), 35.0 (ArCH2), 30.5 (CH2C02), 14.0

(CH2CH3); m/z (electron impact) 214.0812 (M+, 33.6%, C"H'2F202 requires

214.08054),169 (19),151 (2), 140 (100),127 (32),101 (13),95 (2),84 (1.5),

43 (2), 29 (2.5).

103

Page 125: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

3-(3,5-Difluorophenyl)propionic acid (42) 175

F F

F F

EthyI3-(3,5-difluorophenyl)propionate (41) (0.700 g, 3.27 mmole) was

dissolved in a mixture of ethanol (100 ml) and water (50 ml). To this was

added potassium hydroxide (1.834 g, 32.68 mmole). The mixture was then

heated to reflux for two hours with stirring. Af1er this time, tic analysis of the

reaction mixture (silica with 9 : 1 light petroleum [b.p. 40 - 60°C] : ethyl

acetate) showed baseline material only. The reaction was allowed to cool to

room temperature, then the ethanol removed in vacuo. The aqueous fraction

(made to 100 ml) was washed with dichloromethane (2 x 100 ml), acidified

with hydrochloric acid (aqueous 2 M) until a cloudy white suspension

remained on shaking, then extracted with dichloromethane (3 x 100 ml). The

combined extracts were dried (MgS04 ) and concentrated in vacuo to yield a

pale yellow solid (42) (0.557 g, 92%). This was recrystallized from light

petroleum [b.p. 40 - 60°C] to give a colourless crystalline solid (0.507 g,

84%); mpt. 59 - 60°C (Iiterature175 m.p. 58 -59°C from hexane); Ymax (neat)

3099,3057,2958 (C02H), 1703 (C=O), 1629, 1598, 1462, 1438, 1310, 1116

Cm'l; OH (250 MHz, CDCI3 ) 11.15 (1 H, broad s, C02H), 6.79 - 6.61 (3 H, m,

ArH), 2.94 (2 H, t, J7.6 Hz, ArCH2), 2.68 (2 H, t, J7.5 Hz, CH2C02); lie (62.5

MHz, COCl3) 178.9 (C02H), 166.0 (2 C, d,d, J246.6, 12.8 Hz, C-F), 143.8

(d, CFCHCF), 111.3 - 1109 (2 C, m, CFa-lCCH2), 101.9 (t, J24.9 Hz,

CFa-lCF), 34.8 (ArC1-i2), 30.1 (d, J1.7 Hz, a-l2C02); m/z (electron impact)

186.0461 (M+, 52.0%, CgHaF202 requires 186.04924), 168 (4), 140 (100),

127 (75), 122 (12), 114 (12), 101 (27), 95 (6), 81 (7), 70 (8), 63 (7), 51 (18),

44 (7), 40 (8),29 (2.5); Analysis Calculated for CgHsF202 (186.160): C(58.07),

H(4.33); Found: C(57.97), H(4.28).

104

Page 126: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

5,7 -Difluoro-1-indanone (43) 175

F 0

F

.. F

F

Polyphosphoric acid (6.0 g) was added to a round bottomed flask (25 ml),

equipped with a magnetic stirrer and condenser. To this was added 3-(3,5-

difluorophenyl)propionic acid (42) (0.314 g, 1.69 mmole). The mixture was

heated at 45 QC for forty eight hours, with an attempt at stirring. After this

time, the reaction mixture was allowed to cool to room temperature. Water

(distilled, 100 ml) was mixed carefully with the reaction mixture to give a

cloudy white suspension. This aqueous fraction was extracted with

dichloromethane (3 x 100 ml). The organic extracts were reduced in volume,

washed with sodium hydrogen carbonate solution (saturated, 3 x 100 ml),

dried (MgS04), concentrated in vacuo and recrystallised from methanol to

yield a colourless crystalline solid (43) (0.276 g, 97%); m.p. 81 - 81.5 QC

(literature175 m.p. 81 - 82 QC from hexane - chloroform); V max (neat) 3085

(aromatic C-H), 2925 (saturated C-H), 1711 (ester C=O of ketoester), 1679

(a-aryl ketoneC=O), 1619 (enol esterC=O), 1593, 1439, 1326, 1208, 1124

cm-1; OH (400 MHz, CDCI3) 6.96 (1 H, d,d, J7.9, 1.8 Hz, 4-H), 6.74 (1 H, d,t, J

2.0,9.2 Hz, 6-H), 3.15 (2 H, d,t, J6.1, 0.5 Hz, 2-H), 2.75 - 2.72 (2 H, m, 3-H).

No evidence of enol form in n.m.r.; 8c (100 MHz, CDCI3 ) 201.5 (d, JCF7 2.0

Hz, 1-C), 167.8 (d,d, JCF5 257.8 Hz, JcF7 11.1 Hz, 5-C), 159.9 (d,d, JCF7 266.6

Hz, JCF5 14.1 Hz, 7-C), 159.4 (d,d, J 11.1, 4.0 Hz, 3a-C) , 121.1 (d,7a-C),

109.8 (d,d, JCF5 22.14 Hz, JCF7 5.0 Hz, 4-C), 103.8 (d,d, J27.2, 23.1 Hz, 6-C),

37.2 (2-C), 26.4 (d, J2.0 Hz, 3-C) Proton and carbon n.m.r. spectra in

agreement with the Iiterature;175 m/z (electron impact) 168.0351 (M+', 100%,

CgHsF20 requires 168.03868),151 (7),140 (53),132 (4),120 (7),114 (10),

101 (24), 99 (6), 74 (4),63 (7),51 (8),40 (5),31 (3); Analysis Calculated for

CgHsF20 (168.144): C(64.28), H(3.57); Found: C(64.11), H(3.52).

105

Page 127: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

2-Deutero-5, 7 -difluoro-1-indanone (43a)

F

H ----l .....

H F

D

To a stirred solution of LOA (0.223 ml, 0.446 mmole, 2.0 M solution in

heptane / tetrahydrofuran / ethyl benzene) in THF (0.5 ml) was added 5,7-

difluoro-1-indanone (0.050 g, 0.297 mmole) in THF (10 ml) over forty five

minutes, at -78°C and under nitrogen. The solution gradually took on a

yellow colour. After thirty minutes stirring at -78 QC, the reaction mixture was

quenched with deuterated acetic acid (0.034 ml, 0.038 g, 0.595 mmole). A

further portion of the deuterated acetic acid was added then the solution

allowed to warm to room temperature. During this period the yellow colour of

the solution disappeared. After reaching room temperature, the reaction

mixture was concentrated in vacuo to yield a yellow residue (0.370 g). The

crude reaction mixture was taken up in CDCI3 and undissolved solid removed

by filtration. Removal of the solvent, after n.m.r. analysis, yielded a yellow

solid (0.048 g). This solid contained both (43a) and (43). Using mass

spectroscopyand 'H n.m.r spectroscopy, the approximate percentage of

molecules having deuterium incorporated at the C-2 position was determined.

Determination of deuterium incorporation in (43a) using electron impact mass spectroscopy.

Proteo(43)

M+ M+' M:t:2 . : M+3 M+4 .<,'-"- ., , .

168 169 170 171 172 86.9 65.0 6.2 1.0 0.5

1 0.748 0.071 0.012 0.006 '. "

106

Page 128: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Deutero (43a)

d" . -d, . --- . d . '.:: ' .. Jo. .d" .d .... , .'

168 169 170 171 172 Mass' - - .;'-.

87.5 85.8 51.9 19.2 4.0 _ Peak Height,": . ;".' Ax

87.5 64.2 3.7 0.2 (0.02) Peak'Height dox Peak HeightM""\ " ' .- -, ',By

21.6 48.2 19.0 4.0 21.6 16.2 1.5 0.3

32.0 17.5 3.7 32.0 23.9 2.3

-6.4 1.4

do + d, + d2

87.5 + 21.6 + 32 = 141.1

(62%) (15%) (23%) (100%)

Therefore there was approximately 38% (Le. 15% + 23%) of the molecules

containing at least one deuterium atom.

The total deuteration was:

N.m.L results

N.m.r

(43a) 400 MHz

For (43a):

(d, x 1) +(d2 x2)

15 + 46

OH of 2-H OH of 3-H

2.77 - 2.73 3.18 - 3.15

= 61 %

Integral Height for protons on C-2

26mm

• 1 proton was equivalent to 20 mm (Le. 40 mm 12).

Integral Height for protons on C-3

40mm

• The integral height for the second proton at the C-2 position, not replaced

by a deuterium was 26 mm - 20 mm = 6 mm.

• The loss of integral height was therefore 20 mm - 6 mm = 14 mm.

• It follows that the percentage of deuterium incorporated at position C-2 of

(43a) was:

14 x 100 = 70 % 20

107

Page 129: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Attempted synthesis of the TMS enol ether of 5,7-difluoro-1-indanone 176

F

To a well stirred suspension of Li2S (0.041 g, 0.892 mmole) in dry

acetonitrile (1 ml) in a 25 ml round-bottomed flask fitted with a water

condenser and under nitrogen was added trimethylsilyl chloride (0.189 ml,

0.162 g, 1.487 mmole). To this was added 5,7-difluoro-1-indanone (43)

(0.100 g, 0.595 mmole) followed by Amberlyst resin A-21 (1 ml of beads

previously washed with dry acetonitrile [15 ml) under nitrogen). [ N.B. The

reaction was also carried out using triethylamine (0.083 ml, 0.060 g, 0.595

mmole) in place of the Amberlyst resin. In this case the reaction proved

difficult to follow because the proton signals of the triethylamine, in the n.m.r.

spectrum of the reaction mixture, obscured the alkyl proton signals of the

indanone). After stirring for five hours at room temperature, no reaction

appeared to have taken place. A further portion of acetonitrile (5 ml) was

added and the reaction heated to reflux overnight. The reaction was then

quenched with d.-acetic acid, the solvent removed in vacuo. The resulting

solid was taken up in d-chloroform and filtered.

The two main compounds in the crude reaction mixture (RI 0.17 and 0.42)

were then isolated by preparative thin layer chromatography (silica with 2 : 1

light petroleum [b.p. 40 - 60 CC) : ethyl acetate). The compound with the RI

value of 0.42 was isolated and shown to be the 5,7 -difluoro-1-indanone

(0.013 g). The compound having the RI value of 0.17 was isolated as

colourless crystals (0.018 g). This was identified as the sulfur-bridged dimer

(45); Vmax (KBr disc) 3068, 2923,1709 (C=O), 1579, 1445, 1290, 1259 cm'l;

108

Page 130: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

OH (400 MHz, CDCb) 7.03 (1 H, d,d, J 1.1, 8.0, Hz, 4-H), 6.76 (1 H, d,d, J 2.0,

9.2 Hz, 6-H), 3.15 (2 H, t, J 6.2, 0.5 Hz, 2-H), 2.75 - 2.72 (2 H, m, 3-H). Oc

(100 MHz, CDCI3) 202.9 (1-C), 167.8 (d, J CFS 257.6 Hz, 5-C), 159.7 (d, JCF5

11.1 Hz, 3a-C), 121.1 (d, JCF510.1 Hz, 7-C), 128.2 (d, JCFS 4.0 Hz, 7a-C),

117.4 (d, JCF5 25.2 Hz, 4-C), 112.2 (d, J CFS 22.1 Hz, 6-C), 37.2 (2-C), 25.8 (d,

JCFS 2.0 Hz, 3-C); m\z (electron impact) 330 (M+', 100%), 313 (38), 297 (8),

257 (10), 244 (9),182 (34),165 (10),149 (22), 83 (10), N.B. The

characteristic m/z values of 73 and 75, attributable to TMS, not present; m/z

(chemical impact) 348 (MNH/', 41%),331 (MH+', 100%), 182 (10),168 (24),

151 (30), 149 (22), 120 (7), 109 (6) and 83 (5).

109

Page 131: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Methyl 5,7-difluoro-1-indanone-2-carboxylate (46)

F o F

F F

To a stirred solution of LOA (0.892 ml, 1.78 mmole, 2.0 M solution in

heptane / tetrahydrofuran / ethylbenzene) in THF (dry, 2 ml) was added

slowly dropwise (over forty five minutes), 5,7 -difluoro-1-indanone (43) (0.200

g, 1.19 mmole) in THF (dry, 40 ml), at -78°C and under nitrogen. Stirring was

continued for a further thirty minutes, after which time methyl cyanoformate'58

(0.111 g, 0.104 ml, 1.308 mmole). After stirring for half an hour at -78°C, the

reaction mixture was allowed to warm to room temperature. After five

minutes, at which time the reaction appeared to be complete by tic, the

reaction was poured into cold water (50 ml). The product was then extracted

into diethyl ether (4 x 50 ml). The ethereal extracts were then dried (Na2S04)

and concentrated in vacuo to yield a brown solid (0.968 g). The crude

material was purified by column chromatography (silica with 1 : 1 hexane : .

ethyl acetate) to give a colourless crystalline solid (46) (0.173 g, 64%); m.p.

94 - 95°C; Vmax (KBr disc) 3096 (aromatic C-H), 2934 (saturated C-H), 1735

..... (ester C=O of ketoester), 1700 (a-aryl ketone C=O), 1631 (ester C=O of

ketoester), 1593, 1444, 1222, 1123, 888 cm-'; vmax (neat), 3097, 3050, 2964,

2931,1731,1701,1632 (medium-weak). 1595, 1124 cm-'; OH (400 MHz,

COCI3) 6.99 (1 H, d,d, J8.0, 0.8 Hz, keto 4-H [enol 4-H double doublet

obscured underneath with the most downfield peak just visible at OH 7.01]),

6.78 (1 H, d,t, J2.0, 9.0 Hz, keto 6-H [triplet of enol 6-H appears at OH 6.86]),

3.86 (3 H, s, enol OMe). 3.80 (3 H, s, keto OMe), 3.78 (1 H, d,d, J8.4, 4.0

Hz, keto 2-H), 3.60 - 3.33 (2 H, ABX type system, keto 3-H [peaks of enol 3-H

obscured]). Ratio of keto: enol tautomers is approximately 5: 1; Bc (100

MHz, COCI3) 193.8 (d, J2.0 Hz, keto 1-C), 168.8 (keto C02Me), 168.0 (d,d J

260.6,12.1 Hz, 5-C), 160.2 (d,d, J268.6, 15.1 Hz, keto 7-C), 157.5 (d,d, J

12.1, 3.0 Hz, keto 7a-C), 147.3 (3a-C) , 120.1 (enol 2-C), 109.6 (d,d, J23.1,

110

Page 132: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

4.0 Hz, keto 4-C), 108.7 (d, J27.2 Hz, enoI4-C), 104.1 (d,d, J27.2, 23.1, Hz,

keto 6-C), 103.3 (d,d, J 27.2, 23.1 Hz, enol 6-C), 53.8 (keto OMe), 53.0 (keto

2-C), 33.2 (enoI3-C), 30.4 (d, J2.0 Hz, keto 3-C), NB. Quaternary carbons of

enol form not visible in spectrum; m/z (electron impact) 226 (M+, 56%), 195

(24),166 (100),147 (17),138 (56),119 (83),112 (32),99 (42), 93 (20),87

(27),74 (24),69 (23),63 (38), 59 (57), 55 (30), 44 (31); Analysis Calculated

for Cll

HsF20

3 (226.182): C(58.41), H(3.57); Found: C(58.03), H(3.49).

111

Page 133: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Methyl 2,5,7-trifluoro-1-indanone-2-carboxylate (47) .

0

.... Me .... Me 0

• 0 F

F F

To a stirred suspension of sodium hydride (80% dispersion in oil, 0.051 g,

1.703 mmole) in DMF (dry, 2 ml), at 0 °C and under nitrogen, was added

methyI5,7-difluoro-1-indanone-2-carboxylate (46) (0.321 g, 1.419 mmole) in

DMF (dry, 10 ml).165 This was left to stir for twenty minutes at 0 °C then the

temperature was lowered to -50°C and the N-chloromethyl, N-fluoro

ethylenediamine bis(tetrafluoroborate) (0.503 g, 1.419 mmole) added. The

reaction was left to stir at -50°C for ten minutes then allowed to warm to

room temperature. After stirring for one hour at room temperature, the

reaction mixture was poured into a separating funnel containing diethyl ether

(100 ml), washed with a saturated solution of sodium hydrogen carbonate (50

ml) and a saturated solution of sodium chloride (50 ml), dried (Na2S04) and

concentrated in vacuo to yield a pale yellow solid (0.489 g). This was purified

by flash chromatography (silica with 3 : 1 light petroleum [b.p. 40 - 60°C] :

ethyl acetate) to give a colourless solid (47) (0.289 g, 83%); V max (KBr disc)

2968,1764,1733,1620,1602,1415,1327,1229,1213,1126, 854 cm'\ OH

(400 MHz, CDCI3) 7.01 ( 1 H, d,d, J7.6 0.8 Hz, 4-H), 6.85 (1 H, d,d, J8.9, 1.9

Hz, 6-H), 3.84 (3 H, s, OMe), 3.79 (1 H, d,d, J 18.0, 11.4 Hz, 3-H), 3.43 (1 H,

d,d, J22.7, 18.0 Hz, 3-H); Oc (100 MHz, CDCI3) 169.6 (d,d, J262.6, 12.1 Hz,

5-C), 167.2 (d, J28.2 Hz, 7a-C) , 161.0 (d,d, J269.6, 13.1 Hz, 7-C), 154.4 (d,

J 12.1 Hz, 3a-C), 110.1 (d,d, J 23.1, 4.0 Hz, 4-C), 105.2 (d,d, J26.2, 22.1 Hz,

6-C), 94.9 (d, J203.2 Hz, 2-C), 53.7 (OMe), 38.4 (d, J24.1 Hz, 3-C); m/z

(electron impact) 244.0348 (M+, 16.0%, CnH7F303 requires 244.03473),224

(39), 201 (14), 194 (20),185 (100),173 (29),165 (77),156 (23),137 (41),

125 (14),112 (24),107 (10),87 (31), 81 (9),59 (33).

112

Page 134: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

N-Methoxy N-methyl 25-methoxy-25-phenylacetamide (49)

~Ph ~e

S N MeO 'OMe

o

To the (S)-(+)-a-methoxyphenylacetic acid (0.258 g, 1.553 mmole),

dissolved in dichloromethane (0.5 ml), was added oxalyl chloride191

(0.493 g,

0.339 ml, 3.881 mmole). The solution was stirred for one hour at room

temperature, after which time effervesence had ceased. The excess oxalyl

chloride and solvent were removed in vacuo to yield a clear oil. To the oil was

added dichloromethane (5 ml), followed by the N,O-dimethylhydroxylamine

hydrochloride 192 (0.167 g, 1.708 mmole). The solution was cooled to OOC and

pyridine (0.270 g, 0.276 ml, 3.416 mmole) added. The mixture was stirred at

room temperature for one hour, then evaporated in vacuo. The residue was

partitioned between brine and a 1:1 mixture of diethyl ether and

dichloromethane. The organic layer was then washed washed with 10%

copper sulfate solution (50 ml), 10% sodium carbonate solution, dried

(Na2S04) and concentrated in vacuo to yield the crude product as a pale

yellow oil (0.218 g, 67 %). The oil was purified by flash chromatography (silica

iNith 1:1 ethyl acetate: dichloromethane) to give the amide (49) as a

colourless oil (0.161 g, 50 %); Vmax (neat) 3508,2938,1674 (C=O), 1456,

1386, 1110, 1089, 971 cm·1; OH (250 MHz, CDCI3) 7.39 - 7.26 (5 H, m, ArH),

5.04 (1 H, s, CH), 3.35 (CH3), 3.32 (CH3), 3.09 (3 H, s, CH3); Oc (62.5 MHz,

CDCI3

) 170.7 (NC=O), 136.2 (ArC), 128.4 (ArCH). 128.3 (ArCH), 128.2 (ArCH),

127.7 (ArCH), 126.9 (ArCH), 80.4 (2-C), 60.6 (NOCH3), 56.7 (OCH3), 32.0

(NCH3 )·

113

Page 135: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

1 5-Methoxy-1 5-phenylacetone (50)

S N JyPh ~e

MeO 'OMe

o

J; ~Me MeO' I(

o

To a solution of N-methoxy N-methyl 2S-methoxy-2S-phenylacetamide

(49) (0.059 g, 0.282 mmole) in dry tetrohyrofuran (5 ml) was added

methyllithium'92 (1.4 M solution in diethyl ether, 0.282 ml, 0.395 mmole) at

OOC. The reaction mixture was stirred for one minute, after which time no

amide was visible by tic (silica with 5% methanol in dichloromethane. Rt

[amide) 0.41, Rt [ketone) 0.86). To the reaction mixture, at O°C, was added

5% hydrochloric acid (20 ml). The reaction mixture was then extracted with a

1:1 mixture of diethyl ether and dichloromethane (3 x 20 ml). The organic

extracts were dried (Na2S04) and concentrated in vacuo to yield the crude

product (50) as a pale brown oil (0.059 g); V max (neat) 2932, 2828, 1723

(C=O), 1681, 1455, 1354, 1104 cm"; OH (250 MHz, CDCI3) 7.35 - 7.29 (5 H,

m, ArH), 4.60 (1 H, s, CH), 3.33 (3 H, s, OCH3), 2.06 (3 H, s, O=CCH3); Oc

(62.5 MHz, CDCI3) 207.0 (C=O), 135.8 (MeOCHC[aromatic)),128.8

(ArCH). 128.4 (ArCH), 128.5 (ArCH), 126.9 (2 C, ArCH), 69.3 (1-C), 57.1

(OCH3 ), 25.0 (O=CCH3).

114

Page 136: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 3, Experimental

General procedure for in situ dioxirane reactions

All apparatus was meticulously washed with EDT A. Na2 solution, 179

followed by distilled acetone, '80 then dried before use. The assembled

glassware was then covered in aluminium foil to help eliminate light.

To the reaction flask, at 0 cC, was added phosphate buffer (25 ml),'81

dichloromethane (X ml, determined by the quantity of ketone used),'82

tetrabutylammonium hydrogen sulphate (- 0.002 g), the alkene (1 equivalent)

and the ketone (2 equivalents, 0.5 M concentration in the dichloromethane).

The pH of the vigorously stirred solution was adjusted to between pH 7.0 and

7.5 by the addition of KOH solution. l83 Solid Oxone ® (30 equivalents) was

added over approximately half an hour, the pH being maintained between pH

7.0 and 7.5 throughout, by the addition of KOH solution. 183 The reaction was

left to stir for a further four hours, then a sample taken for G.C. or 1H n.m.r.

analysis. If the reaction had not gone to completion, a further portion of solid ®

Oxone (30 equivalents) was added, portionwise. The reaction left to stir for

several hours, a G.C. or 'H n.m.r. analysis carried out, and if the epoxidation ®

reaction was still incomplete, further solid Oxone added. This was repeated

until the either the reaction was complete, or, if the reaction was slow, until

sufficient epoxide had been formed for chiral shift analysis to be performed. If

the reaction needed to be left overnight, the stirring was stopped and then

resumed the following morning.

The reaction mixture was then filtered through glasswool into a separating

funnel. Dichloromethane (3 x 35 ml) was then used to rinse out the reaction

flask and extract the aqueous phase. The organic extracts were combined,

dried (Na2S04) and concentrated in vacuo. The resulting residue was

analysed by G.C. and lor 'H n.m.r. spectroscopy. The alkene and ketone

were then isolated by flash chromatography.

115

Page 137: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

For the reactions using the dioxirane derivatives of a homochiral ketone,

the optical rotation of the isolated alkene was determined and a chiral shift

experiment, using Eu(hfch, carried out.

Equivalents of chiral shift reagent needed for determination of enantiomer

ratio were:

i) Trans·stilbene (23): 0.05 equivalents.

In the presence of Eu(hfch, the ChPh proton in the enantiomers was

resolved in the lH n.m.r. spectrum from OH (250 MHz, CDCla) 3.85 (s),

to OH (250 MHz, CDCla) -3.96 and -3.95.

ii) Trans-~-methyl styrene (24): 0.15 equivalents.

The ChPh proton appeared as a doublet at OH (250 MHz, CD Cia) 3.57

(d, J2.0 Hz) in the lH n.m.r. spectrum of (24). In the presence of 0.15

equivalents of Eu(hfch, it appeared at OH (250 MHz, CDCla) -4.12 and

-4.20;

iii) 6-Chloro-2,2-dimethyl-2H-1-benzopyran (25): 0.10 equivalents.

In the presence of Eu(hfch, the proton at C-3 in the enantiomers was

resolved from OH (250 MHz, CDCla) 3.49 (d, J4.4 Hz), to OH (250 MHz,

CDCla) -3.85 and -3.77.

All chiral shift experiments revealed a 1 : 1 ratio of the alkene enantiomers,

consequently no optical rotation was seen in any of the cases.

116

Page 138: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 4, Experimental

Reaction 1: Preparation and reaction of the dioxirane derivative of

1,1, 1-trifluoroacetophenone with cholesterol

Using the above general procedure, Oxone® (39.8 g, 64.7 mmole) and

disodium ethylenediaminetetraacetate dihydrate (1 g) in distilled water (300

ml) were added, over a two and a half hour period, to cholesterol (0.500 g,

1.29 mmole), dichloromethane (30 ml), 1,1,1-trifluoroacetophenone (0.675 g,

0.545 rnl, 3.88 mmole). pH 7.5 phosphate buffer (40 ml) and

tetrabutylammonium hydrogen sulphate (0.439 g, 1.29 mmole). Work up,

using sodium sulphite (1 g), diethyl ether (100 ml) and sodium chloride

solution (aqueous, saturated, 3 x 100 ml) as in the general procedure, yielded

a crude white residue (1.018 g). Tic analysis (silica with 20 : 30 : 1 diethyl

ether: light petroleum [b.p. 40 - 60 QC] : ethyl acetate) showed cholesterol (Rt

0.35) to be still present, but also showed that formation of its epoxide (Rt

0.12) had taken place. N.m.r. spectroscopy showed there to be 70% of the

starting material [OH (250 MHz, CDCI3) 0.68 (3 H, s, 18-Me)], and 30% of the

epoxide [OH (250 MHz, CDCI3 ) 0.64 (3 H, s, 18-Me) and 0.61 (3 H, s, 18-Me)].

>.' • - "~~eaction 2: Preparation and reaction of the dioxirane derivative of 1,1,1-

trifluoroacetophenone with trans-stilbene

Using the above general procedure, Oxone® (39.2 g, 63.8 mmole) and

disodium ethylenediaminetetraacetate dihydrate (1 g) in distilled water (300

ml) were added, over a four and a half hour period, to trans-stilbene (0.230 g,

1.28 rnmole). dichloromethane (30 ml), 1,1,1-trifluoroacetophenone (0.667 g,

0.538 ml, 3.83 mmole), pH 7.5 phosphate buffer (40 ml) and

tetrabutylammonium hydrogen sulphate (0.433 g, 1.28 mmole). Work up,

using sodium sulphite (1 g), diethyl ether (100 ml) and sodium chloride

solution (aqueous, saturated, 3 x 100 ml) as in the general procedure, yielded

an off white residue (0.279 g). Tic analysis (silica with 1 : 1 toluene: light

petroleum [b.p. 40 - 60 QC]) of the solution showed trans-stilbene (Rt 0.83) to

117

Page 139: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

be still present, but also showed that formation of the trans-stilbene oxide (Rt

0.61) had taken place. N.m.r. spectroscopy showed there to be 60% of the

starting material [OH (250 MHz, CDCI3 ) 7.11 (2 H, s, ChPh)) , and 40% of the

epoxide [OH (250 MHz, CDCI3 ) 3.86 (2 H, s, ChPh)).

General procedure for the in situ dioxirane reaction.7

To a three-necked round-bottomed flask was added the alkene in

dichloromethane, followed by the ketone, pH 7.5 buffer and tetrabutyl­

ammonium hydrogen sulphate. The mixture was allowed to stir for twenty

minutes at 2 -5 QC. The buffer (pH 7.5 at 5 QC) was made from KH2P04 (1.179

g) and Na2HP04 (4.302 g) dissolved in distilled water (1000 ml). A pH of 7.5

was maintained using a pH stat (5 M KOH). The Oxone® and EDTA.Na2·2H20

were dissolved in distilled water and added dropwise to the vigorously stirring

mixture over several hours. Care was taken not to allow the pH to fluctuate

from pH 7.5. The reaction was then left to stir overnight.

Work up: Sodium sulphite was added to the reaction mixture and stirring

continued for thirty minutes. Further dichloromethane was added, then the

dichloromethane separated from the aqueous layer and concentrated in

vacuo. The resulting residue was redissolved in diethyl ether, washed with

aqueous sodium chloride solution, dried (MgS04) and concentrated in vacuo

to yield a residue containing any unreacted alkene, the ketone and any

oxidized alkene.

118

Page 140: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Chapter 5, Experimental

Preparation of Solutions of Dioxiranes

Titration of isolated dioxirane solutions

All the apparatus to be used was washed with water then acetone, '80 and

dried. An acetic acid I acetone solution'·' (3 ml) was added to a small conical

flask, followed by a solution of potassium iodide'·· (5 ml). The flask was

stoppered and shaken. A portion of the isolated dioxirane solution (0.2 ml)

was pipetted into the flask which was again stoppered and shaken. The

dioxirane oxidises the potassium iodide to produce iodine (I,), which turns the

solution yellow. The resulting solution was titrated against a solution of

sodium thiosulfate.'·7 The end-point of the titration was denoted by the loss of

the yellow colour of the iodine (refer to p.69 for the chemical equations for the

titration) .

119

Page 141: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Preparation of Dimethyldioxirane (9)16

H3CX1 H3C 0

in acetone

The apparatus was covered in foil, flushed with nitrogen and the receiving

flask cooled to -78 ac. Sodium bicarbonate (240 g) was added to the reaction

flask followed by the EDT A. Na2 solution (450 ml).I79 The mixture was stirred

using an overhead stirrer for five minutes after which time the stirring was

stopped and acetone l80 (320 ml) added. The stirring was recommenced and

continued for five minutes. The caroate (500 g) was then added in portions.

The stirring was stopped during each addition then started again when the

addition of a portion was complete. On completion of caroate addition, a

vacuum line was attached to the apparatus. A light vacuum (- 560 mmHg)

was applied using a water pump. Note that the dimethyldioxirane solution

begins to distil over before the vacuum is applied. The distillation of the

dimethyldioxirane solution is complete after approximately one hour. The

molarity of the solution was obtained by titration using the above method.

Molarity of solution: 0.09 M

120

Page 142: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Preparation of Methyl(trifluoromethyl)dioxirane (1. R'= CF3• R2= Me)14

in 1,1, 1-trifluoroacetone

After the apparatus had been covered in foil, a flow of nitrogen was put

through the apparatus and maintained for two hours. After two hours, both

the receiving flasks and trap were cooled to -78 QC. The reaction flask was

then cooled to 0 QC, the flow of nitrogen stopped and the nitrogen inlet

sealed. Sodium bicarbonate was then added to the reaction flask followed by

the EDT A. Na2 solution. 179 The solution was briefly stiired using an overhead

stirrer. The 1,1, 1-triflouoroacetone was added to the reaction flask and the

solution again stirred briefly before the addition of the caroate in one portion.

The solution was vigorously stirred and a light vacuum applied using a water

pump (- 600 - 700 mm Hg). The distillation was complete in approximately

ten minutes.

Molarity of solution: 0.68 M

121

Page 143: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Preparation of the dioxirane derivative of cyclohexanone 17

0-0

CS '0 oyclohe"""o",

A mixture of cyclohexanone (40 ml), phosphate buffer1s1 (50 ml) and ice

(14 g) was stirred at 0 cC (ice - salt bath). Oxone® (90 g, cooled) was added

as a slurry in distilled water (200 ml, cooled). Potassium hydroxide (5 M

solution in distilled water) was added simultaneously to maintain a pH of 7 -

8.5. Af1er addition of the Oxone® was complete, the reaction mixture was

stirred for two to three minutes then poured into a beaker containing a cooled

mixture of Na2S04 (32 g), NaH2P04 (16 g) and Na2HP04 (8 g). The combined

mixture was stirred vigorously in an ice - salt bath. The liquid phase was

rapidly transferred to a cooled separating funnel (500 ml) and the aqueous

phase .removed. The deep yellow organic phase (25 ml) was dried (Na2S04,

cooled) then analysed for peroxide content.

Molarity of solution: 0.09 M

122

Page 144: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Attempted preparation of the dioxirane derivative of 1,1,1-

trifluoroacetophenone (1, Rl= CF3 , R2= Ph)

o

~CF' A mixture of 1,1, 1-trifluoroacetophenone (2 ml), phosphate buffer181 (2 ml)

and ice (1 g) was stirred at 0 cC (ice-salt bath). Cooled Oxone® (ice-salt­

methanol bath) was added as a slurry in water (10 ml) over ten minutes. A

potassium hydroxide solution (5 M), cooled to 0 cC, was added

simultaneously to maintain the pH at 7 - 8.5. A yellow colour was formed

immediately on combining the reagents. The mixture was stirred vigorously

for two to three minutes and then poured into a beaker containing a cooled

mixture of anhydrous Na2S04 (2.3 g) : NaH2P04.H20 (1.1 g) : Na2HP04.7H20

(0.6 g). The combined mixture was stirred vigorously in an ice-salt bath. The

liquid phase was transferred rapidly to a cooled separatory funnel. NB. At this

point the aqueous phase was separated from rest of the reaction mixture.

The organic phase was dispersed throughout the remaining solid. Using a

cooled pipette, the organic phase was pi petted into a small cooled beaker

containing drying agent (NaS04)' The remaining organic material in the solid

was extracted with dichloromethane (10 ml) and combined with the other

organic phase. The yellow liquid was then filtered into a cooled flask and

tested for peroxide content.

Molarity of solution: None

123

Page 145: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Preparation of the dioxirane derivative of hexafluoroacetone (1. R'= CF3•

R2= CF3) using sodium bicarbonate as buffer

.. F3CX1 F3C 0

in dichloromethane

To a rapidly stirring mixture of sodium hydrogen carbonate (26.7 g, 318.1

mmole), hexafluoroacetone trihydrate (1.00 g, 4.54 mmole), EDTA.Na2

solution 179 (60 ml) and dichloromethane (5 ml) was added solid Oxone®

(41.49 g, 136.3 mmole) in portions, over a period of 4 hours. After addition of

all the Oxone®, the solution was left to stir for 2 hours. At this point no yellow

colour was seen in the dichloromethane phase ( the dichloromethane was

dispersed through the solid and on the surface of the aqueous layer as small

globules). It was noted that some liquid, which appeared to have a slight

yellow colour, was present on the surface of the first cold finger after addition

of approximately half the Oxone®. It was the liquid (- 3 ml) obtained from this

cold finger which was tested for peroxide content after all the Oxone® had

been added. Molarity of solution: 0.013 M

The reaction was continued. To the reaction flask was added further

"dichloromethane (5 ml) and a small quantity of phase transfer catalyst

(tetrabutylammonium hydrogen sulfate [TBAHS], 15 mg). Solid Oxone®

(13.83 g, 45.4 mmole) was added over ten minutes and the solution left to stir

for twenty minutes at 0 aC. The ice-bath was then removed and the mixture

allowed to warm to room temperature. After thirty minutes a slight vacuum

was applied for fifteen minutes. Approximately 4 ml of a slightly yellow

coloured solution was collected in the cold finger. Molarity of solution: 0.018

M.

124

Page 146: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Preparation of the dioxirane derivative of hexafluoroacetone (1, R'= CF3 ,

R2= CF3) using a phosphate buffer

.. F3CX1 F3C 0

in dichloromethane

Oxone® (20.74 g, 68.16 mmole) was added to a vigorously stirring mixture

of hexafluoroacetone.trihydrate (0.500 g, 2.272 mmole), TBAHS (0.015 g),

phosphate bufferl89 (145 ml) and dichloromethane (25 ml) in 2 portions. The

first portion was added over 0.5 hour then the solution left to stir for 1.5 hours

at 0 QC. After this time the remaining Oxone® was added over 0.5 hour and

the solution left to stir for 1 hour then a light vacuum applied for 0.5 hour.

Approximately 5 ml of solution was collected in the cold finger. This was then

tested for its peroxide content.

Molarity of solution: 0.023 M

125

Page 147: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Oxidations using solutions of isolated dioxiranes

In each case all the apparatus was meticulously washed with EDTANa2

solution 179 followed by acetone 180 and dried before use. The reaction flasks

were cover in aluminium foil to exclude light.

Oxidation of trans-stilbene (23)

Ph~ Ph

..

To a stirred solution of trans-stilbene (0.252 g, 1.40 mmole) in acetone180

(20 ml), at 0 °C, was added a solution of dimethyldioxirane in acetone (0.07

M, 20 ml, 1.40 mmole). The reaction was left to stir at 0 °C for 7 hours, after

which time it was stored in the freezer over night. After this time the reaction

mixture was allowed to warm to room temperature. The reaction was followed

by tic (silica with 1 : 1 toluene: light petroleum [b.p. 40 - 60°C]); alkene RI

0.86; epoxide RI 0.68. After 1 hour at room temperature, the reaction

appeared to have gone to completion. The reaction mixture was concentrated

in vacuo to yield a colourless solid (0.274 g, 100%); V max (film) 3036, 2991,

1464, 1452, 1072 cmo1; liH (250 MHz, CDCI3) 7.40 - 7.27 (10 H, m, ArH), 3.85

(2 H, s, PhCH).

Conditions used for G.C. Analysis:

G.C. Apparatus: Pye Series, 104 Chromatograph, Pye Unicam.

Column: 10% Apiezon on Chomosorb W, AW/DMCS, 100/200 mesh.

Gases: Standard machine settings.

Carrier Gas: Nitrogen.

Detector: FID.

Chart Recorder: Omniscribe Recorder, Houston Instruments.

Program: Initial oven temperature: 80°C for 0.5 minute;

Ramp rate: 3 °C I minute;

Final oven temperature: 280°C for 10 minutes.

Retention times of standard solutions:

trans-stilbene : 14 minutes;

trans-stilbene oxide: 18 minutes.

126

Page 148: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Oxidation of tran9-~-methylstyrene (24)'90

Me~ Ph

To trans-~-methylstyrene (0.200 g, 1.692 mmole) was added a solution of

dimethyldioxirane in acetone (23.3 ml, 0.087 M, 2.031 mmole). The reaction

was followed using gas chromatography. The solution was left to stir at 0 °C

for fifteen minutes after which time the reaction appeared to be complete.

Solvent was then removed in vacuo to yield a colourless oil which was taken

up in dichloromethane, dried (Na2S04) and again concentrated to yield a

colourless oil which was purified by column chromatography to yield a

colourless oil (0.202 g, 89%); Vmax (film) 3025, 2915,1496,1442 cm·'; OH (250

MHz, CDCI3) 7.36 - 7.23 (5 H, m, ArH), 3.57 (1 H, d, J2.0 Hz, ChPh), 3.03 ( 1

H, q,d, J5.1, 2.1 Hz, MeCh), 1.44 (3 H, d, J5.1 Hz, Me); Bc (62.5 MHz,

CDCI3) 137.7, 126.4, 128.0, 125.5, 59.5 (Q-iPh), 59.0 (Q-iMe), 17.8 (Q-i3).

GC-MS analysis of the purified material showed evidence of a small amount

of decomposition of the epoxide during chromatography [m/z 105, C(O)Ph;

derived from the decomposition product, H3CCH2C(O)Ph]

Conditions used for G. C. Analysis:

G.C. Apparatus: Pye Series, 104 Chromatograph, Pye Unicam.

Column:

Gases:

10% Apiezon on Chomosorb W, AW/DMCS, 100 1200 mesh.

Standard machine settings.

Carrier Gas: Nitrogen.

Detector: FID.

Chart Recorder: Omniscribe Recorder, Houston Instruments.

Program: Initial oven temperature: 80°C for 0.5 minute;

Ramp rate: 3 °C I minute;

Final oven temperature: 280°C for 10 minutes.

Retention times of standard solutions:

trans-~-methylstyrene : 5.5 minutes;

trans-~-methylstyrene oxide: 8.5 minutes.

127

Page 149: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Oxidation of 6-acetyl-2,2-dimethyl-2H-1-benzopyran

To the 6-acetyl-2,2-dimethyl-2H-1-benzopyran (0.150 g, 0.742 mmole) at 0

QC, was added dimethyldioxirane in acetone (0.037 M, 24.1 ml, 0.890

mmole). This was left to stir, at 0 QC, for 1.5 hours, after which time the

reaction appeared to be complete by tic analysis (silica with 50 : 49 : 1 diethyl

ether: light petroleum [b.p. 40 - 60°C] : dichloromethane); alkene Rt 0.91;

epoxide Rt 0.41. Solvent was removed in vacuo to yield a colourless oil which

was purified by flash chromatography (silica with 50 : 49 : 1 diethyl ether:

light petroleum [b.p. 40 - 60°C] : dichloromethane) to yield a colourless oil

(0.138 g, 85%) which solidified on storing in the fridge for one week; V max

(film) 2980, 2924, 1679 (C=O), 1610, 1272, 1192 cm-1; OH (250 MHz, CDCI3)

8.01 (1 H, d, J2.2 Hz, 5-H), 7.83 (1 H, d,d, J8.5, 2.2 Hz, 7-H), 6.85 (1 H, d, J

8.5 Hz, 8-H), 3.97 (1 H, d, J4.4 Hz, 4-H), 3.53 (1 H, d, J4.4 Hz, 3-H), 2.57 (3

H, s, CH3CO), 1.60 (3 H, S, 2-CCH3')' 1.29 (3 H, s, 2-CCf6").

Conditions used for G. C Analysis (not quantiative):

-·G.C. Apparatus: Pye Series, 104 Chromatograph, Pye Unicam.

Column:

Gases:

10% Apiezon on Chomosorb W, AW/DMCS, 100/200 mesh.

Air: 11; Hydrogen: 17.

Carrier Gas: Nitrogen.

Detector: FID.

Chart Recorder: Omniscribe Recorder, Houston Instruments.

Program: Oven temperature: 180°C;

Retention times of standard solutions:

6-cyano-2,2-dimethyl-2H-1-benzopyran: 2 minutes;

6-cyano-2,2-dimethyl-2H-1-benzopyran oxide: Decomposition noted.

2 peaks: 3 and 3.5 minutes.

128

Page 150: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Oxidation of S-chloro-2,2-dimethyl-2H-1-benzopyran (25)

Cl, Cl,

To 6-chloro-2,2-dimethyl-2H-1-benzopyran (0.075 g, 0.385 mmole) at 0 QC,

was added a solution of dimethyldioxirane in acetone (0.037 M, 12.5 ml,

0.462 mmole). This was left to stir at 0 QC for 1.5 hours after which time the

reaction appeared to have gone to completion by tic analysis (silica with 60 :

39 : 1 light petroleum [b.p. 40 - 60 QC] : diethyl ether: dichloromethane);

alkene Rt 0.85; epoxide Rt 0.62. Concentration of the reaction mixture in

vacuo yielded a colourless oil which was purified by flash chromatography

(silica with 85 : 15 light petroleum [ b. p. 40 - 60 QC] : diethyl ether) to yield a

colourless oil (0.072 g, 89%); Vmax (film) 2976, 2932, 1268, 1204 cm"; OH (250

MHz, CDCI3 ) 7.32 (1 H, d, J2.6 Hz, 5-H), 7.19 (1 H, q, Ja.6, 2.6 Hz, 7-H),

6.75 (1H, d, Ja.7 Hz, 8-H), 3.85 (1 H, d, J4.4 Hz, 4-H), 3.49 (1 H, d, J4.4

Hz, 3-H), 1.56 (3 H, S, 2-CCH3')' 1.24 (3 H,s, 2-CCH3").

Conditions used for G. C. Analysis (not quantiative):

G.C. Apparatus: Pye Series, 104 Chromatograph, Pye Unicam.

Column:

Gases:

10% Apiezon on Chomosorb W, AW/DMCS, 100/200 mesh.

Air: 11; Hydrogen: 17.

Carrier Gas: Nitrogen.

Detector: FID.

Chart Recorder: Omniscribe Recorder, Houston Instruments.

Program: Oven temperature: 1aO QC;

Retention times of standard solutions:

6-cyano-2,2-dimethyl-2H-1-benzopyran: 4.5 minutes;

6-cyano-2,2-dimethyl-2H-1-benzopyran oxide: Decomposition noted.

2 peaks: 6 and 7.75 minutes.

129

Page 151: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Oxidation of 6-cyano-2,2-dimethyl-2H-1-benzopyran

CN CN, o

• o

To 6-cyano-2,2-dimethyl-2H-1-benzopyran (0.150 g, 0.810 mmole) at 0 cC,

was added a solution of dimethyldioxirane in acetone (0.088 M, 11.04 ml,

0.972 mmole). This was left to stir at 0 cC for 10 hours after which time the

reaction appeared to have gone to completion by tic analysis (silica with 70 :

30 light petroleum [b.p. 40 - 60 cC] : diethyl ether); alkene R, 0.28; epoxide R,

0.63. Concentration of the reaction mixture in vacuo yielded a colourless oil

which was purified by flash chromatography (silica with 85 : 15 light

petroleum [ b.p. 40 - 60 cC] : diethyl ether) to yield a colourless solid; (0.072

g, 89%); m.p. 50.5 - 53.5 cC; Vmax (film) 2983,2227 (CN), 1616, 1494, 1279

1162 cm-1; OH (250 MHz, CDCI3) 7.65 (1 H, d, J2.1 Hz, 5-H), 7.52 (1 H, d,d, J

8.6, 2.1 Hz, 7-H), 6.86 (1 H, d, J8.5 Hz, 8-H), 3.91 (1 H, d, J4.3 Hz, 4-H),'

3.54 (1 H, d, J4.3 Hz, 3-H), 1.60 (3 H, S, 2-CCH3')' 1.30 (3 H, S, 2-CCH3"); Oc

(62.5 MHz, CDCI3) 133.3 (8a-C), 134.3 (NC-C=C), 133.7 (NC-C=C'), 121.0

(4a-C), 118.9 (8-C), 118.6 (NC-C), 104.2 (NC), 74.6 (OCMe2), 62.2 (4-C),

49.8 (3-C), 25.4 (a-i3), 22.9 (CH3).

Conditions used for G. C. Analysis (not quantiative):

G.C. Apparatus: Pye Series, 104 Chromatograph, Pye Unicam.

Column: 10% Apiezon on Chomosorb W, AW/DMCS, 100/200 mesh.

Gases: Air: 11; Hydrogen: 17. Carrier Gas: Nitrogen.

Detector: FID.

Chart Recorder: Omniscribe Recorder, Houston Instruments.

Program: Oven temperature: 180 cC;

Retention times of standard solutions:

6-cyano-2,2-dimethyl-2H-1-benzopyran: 4 minutes;

6-cyano-2,2-dimethyl-2H-1-benzopyran oxide: Decomposition noted.

2 peaks: 5.75 and 7 minutes.

130

Page 152: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Oxidation of Metoclopramide

0 Et 0 Et 0 I Cl N~N'Et C N~~

H • H H2N OMe H2N OMe

To metoclopramide (O.1S0 g, O.SOO mmole) was added a solution of the

dioxirane derivative of cyclohexanone in cyclohexanone (0.09 M, S.S6 ml,

O.SOO mmole). this was left to stir over night. After this time a further aliquot of

the dioxirane derivative of cyclohexanone (5.S6 ml, O.SOO mmole) was added

and the solution left to stir at room temperature for 6.S hours. A further aliquot

(S.S6 ml, O.S mmole) was then added and the solution left to stir overnight, at

room temperature. Any remaining oxidant was then destroyed by the addition

of sodium sulfite. After stirring for 1 minute, solvent was removed in vacuo to

yield ari orange residue. This was purified by flash chromatography (alkaline

alumina with dichloromethane to elute the residual cyclohexanone, increasing

to 1 % MeOH in dichloromethane to elute the starting material, then S%

MeOH in dichloromethane to isolate the product). Tic analysis (neutral

alumina with S% MeOH in dichloromethane): metoclopramide: Rr 0.63 and

product: Rr 0.36.

The product was isolated as a yellow solid (0.115 g, 73%). This was

recrystallised from dichloromethane to give a colourless solid (O.OS g, 32%);

bH (400 MHz, CDCI3, DMSO) 8.88 (1 H, m, CON,"", 8.02 (1 H, s, CICC,"",

6.34 (1 H, s, hCCOMe), 4.74 (2 H, s, NH2 ), 3.93 - 3.89 (2 H, q, J6.3 Hz,

CH2NHCO), 3.88 (3 H, s, OC~), 3.39 (2 H, t, J6.3 Hz, CH.!NOEt2), 3.28 (4 H,

q, J7.3 Hz, CH2CH3), 1.33 (6 H, t, J7.3 Hz, CH2CH3 ); be (62.S MHz, CDCI3,

DMSO) 170.0 (C=O), 162.9 (COMe), 152.2 (CNH2), 137.6 (Cl-iCCI), 116.7

(CC=O), 116.0 (CCI), 102.7 (Q-INH2), 67.8 (Q-i2NO), 6S.6 (Cl-i2NH), 61.0

(OQ-i3), 39.S (Cl-i2CH3), 13.8 (CH2 Q-13); m/z (FAB [Positive-ion], glycerol) 316

131

Page 153: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

(MH+, 51.6%), 227 (ArCONHEt, 58), 185 (Gly., 79), 184 (ArCO+, 33), 93

(Gly., 100),86 (19),73 (40), 57 (20).

132

Page 154: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

References and Notes

Page 155: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

References and Notes

1. Saeyer, A V. and Villiger, V., Ber., 1899, 32, 3625.

2. Doering, W. von E. and Dorfman, E., J Am. Chem. Soc., 1953, 75, 5595.

3. Rozen, S., Bareket, Y. and Kol, M., Tetrahedron, 1993, 49(36),8169.

4. Tal bott, R.1. and Thompson, P.G., US Patent 3 632 606, 1972.

5. Montgomery, R.E., J. Am. Chem. Soc., 1974, 96, 7820.

6. Edwards, J.O., Pater, S.H., Curci, R. and Di Furia, F., Photochem. Photobio/., 1979, 30, 63.

7. Curci, R., Fiorentino, M., Troisi, L., Edwards, J.O. and Pater, R.H., J Org. Chem., 1980, 45, 4758.

8. Gallopo, AR. and Edwards, J.O., J Org. Chem., 1981, 46, 1684.

9. Cicala, G., Curci, R., Fiorentino, M. and Laricchiuta, 0., J Org. Chem., 1982, 47, 2679.

10. Curci, R., Fiorentino, M. and Serio, M.R., J. Chem. Soc., Chem. Commun., 1984, 3, 155.

11. Mello, R., Fiorentino, M., Sciacovelli, O. and Curci, R., J Org. Chem., 1988,53,3890.

12. Armstrong, A., Clarke, P.A. and Wood, A, J Chem. Soc., Chem. Comm., 1996,849.

13. Murray, R.W. and Jeyaraman, R., J Org. Chem., 1985, 50,2847.

14. Curci, R., Mello, R., Fiorentino, M. and Sciacovelli, 0., J Org. Chem., 1988, 53, 3891.

15. Ferrer, M., Sanchez-Saeza, F., Casas, J. and Messeguer, A., Tetrahedron Lett., 1994, 35(18), 2981.

16. Adam, w., Sialas, J. and Hadjiarapoglou, L., Chem. Ber., 1991, 124, 2377.

17. Murray, R.W., Singh, M. and Jeyaraman, R., J Am. Chem. Soc., 1992, 114,1346.

133

Page 156: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

1.8. Russo, A. and DesMarteau, 0.0., Angew. Chem. Int. Ed Engl, 1993, 32 (6),905.

19 .. Burger, H., Weinrath, P., Arguello, GA, Julicher, B., Willner, H., DesMarteau, D.D. and Russo, A., J MolecularSpectroscopy, 1994, 168, 607.

20. Jones, C.w., Sankey, J.P., Sanderson, W.R., Rocca, M.C. and Wilson, S.L., J Chem. Research (S), 1994, 114.

21. Kirshleld, A., Muthusamy, S. and Sander, w., Angew. Chem. Int. Ed Engl, 1994, 33(21), 2212.

22. Reviews 01 Dioxirane Chemistry: a) Murray, R.w., Chem. Rev., 1989, 89, 1187; b) Adam, w., Curci, R. and Edwards, J.O., Acc. Chem. Res., 1989, 22, 205; c) Adam, W. and Hadjiarapoglou, L., Topics in Current Chemistry, 1993, 164,45; d) Curci, R., Advances in Oxygenated Processes, 1990, 2, 1; e) Curci, R., Dinoi, A. and Rubino, M.F., Pure and Applied Chem., 1995, 67 (5),811.

23. Baumstark, A.L. and McCloskey, C.J., Tetrahedron Let!., 1987, 28, 3311.

24. Baumstark, A.L. and Vasquez, P.C., J Org. Chem., 1988, 54,3437.

25. Bach, R.D., Owensby, A.L., Andres, J.L. and Schlegel, H.B., J. Am. Chem. Soc., 1991, 113, 7031.

26. Denmark, S.E., Forbes, D.C., Hays, D.S., DePue, J.S. and Wilde, R.G., J Org. Chem, 1995, 60(5), 1391.

27. Curci, R., D'Accolti, L., Fiorentino, M. and Rosa, A., Tetrahedron Left., 1995, 36 (32), 5831.

28. Brown, D.S., Marples, BA, Smith, P. and Walton, L., Tetrahedron, 1995, 51 (12), 3587.

29. Crandall, J.K., Batal, D.J., Sebesta, D.P. and Lin, F., J Org. Chem., 1991,56,1153.

30. Crandall, J.K., Batal, D.J., Lin, F., Reix, T., Nadol, G.S. and Ng, RA, Tetrahedron Left., 1992, 48, 1427; Crandall, J.K. and Rambo, E., J Org. Chem., 1990, 55, 5929.

134

Page 157: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

31. All "Method A,,32 reactions were carried out with a solution of the dioxirane in its parent ketone, unless otherwise noted. Any additional solvent has been recorded. Only when the number of equivalents of dioxirane added, was important to the result, has this quantity been specified.

32. Method A: isolated; Method B: ill situ generated.

33. Bovicelli, P. and Lupattelli, P. and Mincione,E., J. Org. Chem., 1994, 59, 4304.

34. Iyer, R.S., Coles, B.F., Raney, K.D., Thier, R., Guengerich, F.P. and Harris, T.M., J. Am. Chem. Soc., 1994, 116 (4), 1603. NB. Experimental details given in: Baertschi, S.W, Raney, K.D., Stone, M.P. and Harris, T.M., J. Am. Chem. Soc., 1988, 110,7929.

35. Nan, F., Chen, X., Xiong, Z., Li, T. and Li, Y.,Synth. Comm., 1994, 24 (16),2319.

36. Adam, Wand Sauter, M., Chem. Ser., 1993, 126,2697.

37. Adam, W, Peters, K. and Sauter, M., Synthesis, 1994, 111.

38. Adam, Wand Sauter, M., Tetrahedron, 1994, 50 (28),8393.

39. Adam, W, Kab, G. and Sauter, M., Chem. Ser., 1994, 127, 433.

40. Adam, W, Sauter, M. and ZOnkler, C., Chem. Ser., 1994, 127, 1115.

41. Adam, W, Hadjiarapoglou, L., Peters, K. and Sauter, M., J. Am. Chem. Soc., 1993, 115(19),8603.

42. Adam, W, Hadjiarapoglou, L., Peters, K. and Sauter, M., Angew Chem. Int. Ed. Eng/., 1993, 32 (5), 735.

43. Zhang, X. and Foote, C.S., J. Am. Chem. Soc., 1993,115(19),8867.

44. Adam, W, Halasz, J., Lavai, A., Nemes, C., Patonay, T. and T6th,G., Uebigs Ann. Chem., 1994, 795.

45. Kurihara, M., Ito; S., Tsutsumi, N. and Miyata, N., Tetrahedron Lett., 1994, 35 (10), 1577.

46. Beasdale, C., Golding, B.T. and Kitson, S.L., Sioorganic and Medicinal Chemistry Letters, 1994, 4 (2),291.

47. Swindell, C.S. and Ch and er, M.C., Tetrahedron Left., 1994, 35 (33), 6001.

135

Page 158: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

48. Armstrong. A.. Barsanti. P.A. and Clarke. P.A.. Tetrahedron Left .. 1994. 35 (33).6155.

49. Reisch. J. and Schiwek. K .. Uebigs Ann. Chem.. 1994. 317.

50. Adam. w.. ?rechtl. F .. Richter. M.J. and Smerz. A.K .. Tetrahedron Left .. 1993. 34 (52).8427.

51. Crandall. J.K. and Rambo. E .. Tetrahedron Left .. 1994. 35 (10). 1489.

52. Crandall.J.K. and Reix. T .. Tetrahedron Left .. 1994. 35 (16).2513.

53. Allemann. S. and Vogel. P .. Tetrahedron. 1994. 50 (8). 2469.

54. Choo. H.-Y .. P .. Bull. Korean Chem. Soc .. 1994. 15 (2). 104.

55. Baumstark, A.L. and Harden. Jr.. D.B .. J. Org. Chem.. 1993,58 (26). 7615.

56. Adger. B.J .. Barrett. C .. Brennan. J .. McGuigan. P .. McKervey. A. and Tarbit. B .. J. Chem. Soc. Chem. Comm.. 1993. 1220; Adger. B.J .. Barrett. C .. Brennan. J .. McKervey. A. and Murray. R.W.. J. Chem. Soc. Chem. Comm.. 1991. 1553.

57. Abad. J.-L .. Casas. J .. Sanchez-Baeza. F. and Messeguer. A.. J. Org. Chem., 1993. 58 (15).3991.

58. Nemes. C .. Levai. A .. Patonay. T .. T6th. G .. Boros. S .. Halasz. Adam. W. and Golsch. D .. J. Org. Chem.. 1994. 59 (4).900.

59. Patonay. T .. T6th. G. and Adam. w.. Tetrahedron Left.. 1993.34 (32). 5055

60. Curci. R .. Detomasa. A.. Prencipe. T. and Carpenter. G.B .. J. Am. Chem. Soc.. 1994. 116(18).8112.

61. Lupattelli, P .. Saladino. R. and Mincione. E .. Tetrahedron Left .. 1993. 34 (39).6313.

62. Meyer. C. and Spiteller. G .. Uebigs Ann. Chem.. 1993. 17.

63. Lluch. A.-M .. Sanchez- Baeza. F .. Messeguer. A.. Fusco. C. and Curci. R .. Tetrahedron. 1993. 49 (28).6299.

64. Marzabadi, C.H. and Spilling. C.D .. J. Org. Chem.. 1993. 58 (14).3761.

136

Page 159: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

6S. Crimmins, M.T., Jung, D.K. and Gray, J.l., J. Am. Chem. Soc., 1993, 115 (8),3146.

66. Myers, A.G. and Dragovieh, P.S., J. Am. Chem. Soc., 1993, 115 (15), 7021.

67. Liang, X., Kingston, D.G.I., Lin, C.M. and Hamel, E., Tetrahedron Left., 1995, 36 (17), 2901.

68. Murray, R.W and Singh, M., J. Org. Chem., 1993, 58(19), S076.

69. Timmers, C.M., van der Marel, GA and van Boom, J.H., Recl des Trav. Chim. desPays-Bas, 1993, 112(11),609.

70. Link, J.T., Danishefsky, S.J. and Sehulte, G., Tetrahedron Left., 1994, 35 (49),9131.

71. Elmore, S. Wand Paquette, LA, J. Org. Chem., 1995, 60 (4), 889.

72. Adam, W, Halasz, J., Jambor, Z., Levai, A., Nemes, C., Patonay, T. and T6th, G., J. Chem. Soc., Perkin Trans. 1, 1996, 39S.

73. Murray, R.W, Jeyaraman, R. and Mohan,L., J. Am. Chem. Soc., 1986, 108,2470; Melio,R., Cassidei, L., Fiorentino, M., Fuseo, C., HOmmer, W., Jager, V. and Curci, R., J. Am. Chem. Soc., 1991, 113, 220S.

74. Reiser, 0., Angew. Chem. Int. Ed. Eng/., 1994, 33 (1),69.

7S. Kuek, D., Sehuster, A., Fuseo, C., Fiorentino, M. and Curei, R., J. Am. Chem. Soc., 1994, 116, 237S.

76. Bovieelii, P., Lupattelii, P., Fraeassi, D. and Mineione, E., Tetrahedron Lett., 1994, 35 (6), 93S.

77. Bovieelii, P., Lupattelli, P., Fiorini, V. and Mineione, E., Tetrahedron Lett., 1993, 34 (38),6103.

78. Sehwarz,S., Ring, S., Weber,G., TeiehmOlier, G., Palme, H.-J., Pfeiffer, C., Undeutseh, B., Erhart, B. and Grawe, D., Tetrahedron, 1994, 50 (36), 10709.

79. Dixon, J.T., Holzapfel, C.W. and van Heerden, F.R., Synth. Comm., 1993,23 (2), 13S.

80. Bisseret, P. and Rohmer, M., Tetrahedron Left., 1993,34 (1),1131.

81. Curci, R., D'Aeeolti, L., Detomaso, A., Fuseo, C., Takeuehi, K., Ohga, Y., Eaton, PE. and Vip, Y.C., Tetrahedron Left., 1993, 34 (28),6299.

137

Page 160: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

82. D'Accolti, L., Detomaso, A, Fusco, C., Rosa, A. and Curci, R., J. Org. Chem., 1993, 58 (14),3600.

83. Bovicelli, P., Lupattelli, P. and Sanetti, A, Tetrahedron Left., 1995, 36 (17),3031

84. Csuk, R. and D6rr, P., Tetrahedron, 1994, 50 (33),9983.

85. Bovicelli, P., Lupattelli, P., Sanetti, A. and Mincione, E., Tetrahedron Left., 1994, 35 (45),8477.

86. Kovac, F. and Baumstark, AL., Tetrahedron Left., 1994, 35 (47), 8751.

87. Teager, D.S. and Murray, Jr., R.K., J. Org. Chem., 1993, 58 (20), 5548.

88. Winkler, J.D, Bhattacharya, S.K., Liotta, F., Batey, RA, Hefferman, G.D., Cladingboel, D.E. and Kelly, R.C., Tetrahedron Left., 1995, 36 (13), 2211.

89. Asensio, G., Gonzalez-Nunez, M.E., Bernardini, C.B., Mello, R. and Adam, A, J. Am. Chem. Soc., 1993, 115 (16), 7250; Asensio, G.A., Gonzalez-Nunez, M.E. and Mello, R .. , Anales de Quimica, 1993, 89, 125.

90. Curci, R., D'Accolti, L., Dinoi, A., Fusco, C. and Rosa, A, Tetrahedron Left., 1996, 37(1), 115.

91. Asensio, G., Mello, R., Gonzalez-Nunez, M.E., Castellano, G. and Corral, J., Angew. Chem. Int. Ed Engl., 1996, 35 (2),217 .

. "92. Murray, R.W, Jeyaraman, R. and Mohan, L., Tetrahedron Left., 1986, 27, 2335.

93. Zabrowski, D.L., Moorman, AE. and Beck, K.R., Jr., Tetrahedron Left., 1988,29,4501.

94. Eaton, P.A. and Wicks, G.E., J Org. Chem., 1988, 53, 5353.

95. Murray, R.W, Jeyaraman, R. and Pilay, M.K., J Org. Chem., 1987, 52, 746.

96. Murray, R.W and Jeyaraman, R., J Org. Chem., 1985, 50,2847; Adam, W., Chan, Y.-Y., Cremer, D., Guass, J., Scheutzow, D. and Schindler, M., J Org. Chem., 1987, 52, 2800.

97. Adam, W, Curci, R., and Mello, R., Angew. Chem., 1990, 102,916.

138

Page 161: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

98. Claudia C., Mincione, E., Saladino, R. and Nicoletti, R., Tetrahedron, 1994,50(10),3259.

99. Sakanishi, K, Kato, Y., Mizukoshi, E. and Shimizu, K, Tetrahedron Left., 1994, 35 (27), 4789.

100. Neset, S.M., Benneche, T. and Undheim, K., Acta Chemica Scandinavica, 1993, 47, 1141.

101. Munson, M.C. and Barany, G., J. Am. Chem. Soc., 1993, 115(22), 10195.

102. Adam, w., Hadjiarapoglou, L., Mielke, K and Treiber, A, Tetrahedron Lett., 1994, 35 (31),5625.

103. Hamilton, R., McKervey, M.A., Rafferty, M.D. and Walker, B.J., J. Chem. Soc. Chem. Comm., 1994, 37.

104. Saba, A., Synthetic Communications, 1994, 24 (5), 695.

105. Olah, G.A., Liao, Q., Lee, C.-S. and Prakash, GKS., Synlett, 1993, 427.

106. Murakata, M., Imai, M., Tamura, M. and Hoshino, 0., Tetrahedron: Asymmetry, 1994, 5(10), 2019.

107. Altamura, A, Curci, R. and Edwards, J.O., J Org. Chem., 1993, 58 (25), 7289.

108. Nelsen, S.F., Scamehorn, R.G., De Felippis, J. and Wang, Y., J Org. Chem., 1993, 58(7),1657.

109. Darkins, P., McCarthy, N., McKervey, M.A. and Ye, T., J. Chem. Soc. Chem. Comm., 1993, 1222.

110. Hu, J. and Miller, M.J., JOrg. Chem., 1994, 59(17),4858.

111. Ferrer, M., Sanchez-Baeza, F., Messeguer, A, Diez, A and Rubiralta, M., J. Chem. Soc. Chem. Comm., 1995,293.

112. Webb, K.S., Tetrahedron Left., 1994, 35 (21),3457.

113. Bisseret, P., Seemann, M. and Rohmer, M., Tetrahedron Lett., 1994, 35 (17),2687.

114. Danelon, G.O., Mata, E.G. and M ascaretti , OA, Tetrahedron Lett., 1993, 34 (49),7877. NB. Correction given in: Danelon, G.O., Mata, E.G. and M ascaretti , OA, Tetrahedron Left., 1994, 35 (15),2254.

139

Page 162: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

115. Glass, R.S. and Liu, Y., Tetrahedron Left., 1994, 35 (23),3887.

116. Lu, S., Pearce, E.M. and Kwei, TK, Macromolecules, 1993, 26 (14), 3514.

117. Webb, K.S., Tetrahedron Left., 1994, 35 (21),3457.

118. Derbesy, G. and Harpp, D.N., J. Org. Chem., 1995, 60, 1044.

119. Adam, W, van Barneveld, C. and Golsch, D., Tetrahedron, 1996, 52 (1), 2377.

120. Adam, W Hadjiarapoglou, L. and Wang, X., Tetrahedron Left., 1989, 30, 6497; Adam, W, Hadjiarapoglou, L., Jager, V., Seidel, B. and Wang, X., Chem. Ber., 1991, 124, 2377.

121. Adam, Wand Prechtl, F., Chem. Ber., 1991, 124,2369.

122. Lluch, A.-M., Jordi, L., Sanchez- Baeza, F., Ricart, S., Camps, F., Messeguer, A., Moreto, J.M., Tetrahedron Left., 1992, 33 (21),3021.

123. Adam, Wand Prechtl, F., Chem. Ber., 1994, 121, 667.

124. Adam, W, Muller, M. and Prechtl, F., J. Org. Chem., 1994, 59, 2358.

125. Perez-Encabo, A., Perrio, S., Slawin, A.M., Thomas, S., Wierzchleyski, A.T. and Williams, D.J., J. Chem. Soc. Chem. Comm., 1993, 1059.

126. Griffiths, S.L., Marcos, C.F., Perrio, S., Saberi, S.P., Thomas, S.E., Tustin, G.J. and Wierzchleyski, A.T., Pure and App/. Chem., 1994, 66 (1), 1565.

127. Christian, P.WN., GiI, R., Mufliz-Fernandez, K., Thomas, S.E. and Wierzchleyski, A.T., J. Chem. Soc., Chem. Commun., 1994, 1569.

128. Lluch, A-M., Sanchez-Baeza, F., and Messeguer, M., Anales de Quimica, 1993, 89, 133.

129. Adam, W, Schuhmann, R.M., J. Org. Chem., 1996, 61,874.

130. Lluch, A.-M., Gilbert, M., Sanchez-Baeza, F. and Messeguer, A., Tetrahedron, 1996, 52 (11),3973.

131. Murray, R.W and Gu, D., J. Chem. Soc., Perkin Trans.2, 1993, 2203.

132. Murray, R.W and Gu, D., J. Chem. Soc., Perkin Trans. 2, 1994,451.

140

Page 163: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

133. Murray, R.W, Singh, M., Williams, B.L. and Moncrieff, H.M., Tetrahedron Left., 1995, 36 (14), 2437.

134. Murray, R.W, Pillay, M.K. and Jeyaraman, R., J. Org. Chem., 1988, 53, 3007; Mello, R., Ciminale, F., Fiorentino, M., Fusco, C., Prencipe, T. and Curci, R., Tetrahedron Left., 1990, 31, 6097; Adam, W, Bottle, S.E. and Mello, R., J Chem. Soc. Chem. Comm., 1991, 771; Adam, W, Asensio, G., Curci, R., Gonzalez-Nufiez, M.E. and Mello, R., J Am. Chem. Soc., 1992, 114,8345.

135. Adam, Wand Golsch, D., Angew. Chem. Int. Ed. Eng., 1993, 32(5), 737.

136. Adam, W, Haas, Wand Sieker, G., J. Am. Chem. Soc., 1984, 106, 5020; Adam, W, DOrr, H., Haas, Wand Lohray, B.B., Angew. Chem. Int. Ed. Eng., 1986, 25, 101; Adam, Wand Lohray, B.B., Angew. Chem. Int. Ed. Eng., 1986, 25, 188; Adam, W, Haas, Wand Lohray, B.B., J Am. Chem. Soc., 1991, 113,6202; Adam, W, Golsch, D. and Gorth, F.C., Chemistry-A European Journal. 1996, 2(3), 255.

137. Adam, Wand Golsch, D., Chem. Ser., 1994, 127, 1111.

138. Baumstark, A.L. and Vasquez, P.C., J. Org. Chem., 1988, 53,3437; Murray, R.W, Jeyaraman, R. and Pillay, M.K., J. Org. Chem., 1987, 52, 746.

139. Clennan, E.L. and Yang, K., J Org. Chem., 1993, 58(16). 4504.

140. Ballistreri, F.P., Tomaselli, GA, Toscano, R.M., Bonchio, M., Conte, V. and Di Furia, F., Tetrahedron Left., 1994, 35(43), 8041.

141. a) Singh, M., J Org. Chem., 1992,57, 4263; Adam, W, Curci, R., Gonzalez-Nufiez, M.E. and Mello, R., J Am. Chem. Soc., 1991, 113, 7654. b) Curci, R., Dinoi, A., Fusco, C. and Lillo, MA, Tetrahedron Left., 1996, 37 (2),249; Minisci, F., Zhao, L., Fontana, F. and Bravo, A., Tetrahedron Left., 1995, 36, 1697.

142. Hull, LA and Budhai, L., Tetrahedron Left., 1993, 34 (32), 5039.

143. Adam, W, Chan, Y., Cremer D., Guass, J., Scheutzow, D.C. and Schindler, M., J Org. Chem., 1987, 52, 2800; Singh, M. and Murray, R.W, J Org. Chem., 1992, 57, 4263.

144. Brunnelle, WH., Chem. Rev., 1991,91 (3), 335.

145. Bach, R.D., Andres, J.L., Owensby, A.L., Schlegel, H.B. and Mc Douall, J.J.W, J Am. Chem. Soc., 1992, 114 (18),7202.

141

Page 164: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

146. Murray, A.W., Kong, W. and Aajadhyaksha, S.N., J Org. Chem., 1993, 58(2),315.

147. Kopecky, KA., Xie, Y. and Molina, J., Can. J Chem., 1993, 71,272.

148. lesce, M.A., Cermola, F., Graziano, M.L. and Scarpati, R., J Chem. Soc., Perkin Trans. 1, 1994, 147.

149. Bach, A.D., Andres, J.L., Su, M.-D. and McDouall, J.J.w., J Am. Chem. Soc., 1993,115(13),5768.

150. Takahshi, J., Tsuchiya, J. and Kawasaki, K, Chem. Phys. Left., 1994, 222,319.

151. Cremer, D., Guass, J., Kraka, E., Stanton, J.F. and Bartlett, R.J., Chem. Phys. Left., 1993, 209 (5,6),547.

152. a) Marples, B.A., Muxworthy, J.P. and Baggaley, K.H., Tetrahedron Letters, 1991, 32 (4), 533; b) Marples, B.A., Muxworthy, J.P. and Baggaley, KH., J Chem. Research (S), 1992,28; c) Marples, B.A., Muxworthy, J.P. and 8aggaley, KH., Synlett, 1992, 8, 646.

153. Curci, A., Fiorentino, M., Troisi, L., J Org. Chem., 1980, 45, 4758.

154. Muxworthy, J.P., Synthetic and Mechanistic Aspects of Dioxirane Chemistry, PhD Thesis Loughborough University 01 Technology, 1992.

155. Davis, F.A. and Weismiller, M.C., J Org. Chem., 1990, 55, 3715; Davis, F.A., Sheppard, A.C., Chen, B.-C. and Haque, M.S., JAm. Chem. Soc., 1990, 112, 6679; Davis, F.A. and Kumar, A, Tetrahedron Left., 1991, 32,7671.

156. Meyer, A and Jaouen, G., J Chem. Soc., Chem. Commun., 1974, 787; Jaouen, G. and Meyer, A, J Am. Chem. Soc., 1975, 97 (16),4667.

157. Genet, J.P., Plister, X., Aatovelomanana-Vidal, V., Pinel, C. and Laffitte, J.A., Tetrahedron Left., 1994, 35, 4559.

158. Mander, L.N. and Sethi, P., Tetrahedron Left., 1983, 24, 5425; Aldrichimica Acta, 1987, 20 (2),53.

159. Stork, G., Brizzolara, A, Landesman, H., Szmuszkovica, J. and Terrel, A., J Am. Chem. Soc., 1963, 85, 207.

160. House, H.O., In Modern Synthetic Reactions, 2nd edition, w.A.

142

Page 165: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

Benjamin Inc., Menlo Park, 1972, pp. 734-816 .

. 161. Westlake, P.J., Radical Ring Expansion of Benzocyclic Carbonyl Compounds, PhD Thesis, Loughborough University of Technology, 1992.

162. Titley, A.F., J Am. Chem. Soc., 1982, 2571; Chatterjee, A, Bannerjee, O. and Bannerjee, S., Tetrahedron Left., 1965,3851.

163. Stork, G., Brizzolara, A, Landesman, H., Szmuszkovica, J. and Terrel, R., J Am. Chem. Soc., 1963, 85, 207.

164. White, WH. and Weingarten, H., J Org. Chem., 1967,32, 213.

165. Banks, R.E., Murtagh, V. and Tsiliopoulos, J Fluorine Chemistry, 1991, 52, 389.

166. Williams, D.H. and Fleming, I., Spectroscopic Methods in Organic Chemistry, 4th Ed., McGraw-Hill Book Company, Maidenhead, 1989.

167. Carried out by Or. D.S. Brown of this Department. Positional and temperature parameters, bond lengths, and angles have been deposited with the Cambridge Crystallographic Data Centre, Lensfield Road., Cambridge, CB2 1 EW

168. (Review) Meskens, F.A.J., Synthesis, 1981, 7, 501.

169. Corey, E.J. and Suggs, J.w., Tetrahedron Left., 1975, 44, 3775.

170. Verboom, W, Visser, G.W and Reinhoudt, D.N., Synthesis, 1981, 807.

171. Tsunoda, T., Suzuki, M. and Noyori, R., Tetrahedron Left., 1980, 21, 1357; El Gihani, M. and Heaney, H., Synlett, 1993,433.

172. Moretti, I. and Torre, G., Synthesis, 1970, 141.

173. Wadsworth, WS. and Emmons, W.O., Organic Syntheses, 1969, 45, 44.

174. Wittig, G. and Schollkopf, U., Chem. Ber., 1954, 87, 1318.

175. Metz, G., Synthesis, 1976, 614; Novak, J. and Salemink, C.A., J Chem. Soc., Perkin Trans. 1,1982,2403.

176. Olah, G.A., Gupta, B.G.B., Narang, S.C. and Malhotra, R., J Org. Chem., 1979, 44 (24),4272.

177. Frew, AJ. and Proctor, J Chem. Soc., Perkin Trans. 1,1980,1245.

143

Page 166: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and

178. Pandit. U.K., van der Vlugt, FA, van Dalen, AC., Joustra, A.H. and Schenk, H., Tetrahedron Left., 1969, 42, 3693.

179. The EDTA.Na2 solution was composed of 4 9 EDT A.Na2 in 2500 ml of distilled water. The EDT A Na2 was present to sequester any metal ions present in the reaction.

180. The acetone was distilled from KMn04.

181. The phosphate buffer solution was made up using 20 9 [85%] H3P04 in 2000 ml of EDTA.Na2 solution (ct. Ref. 179), made to pH 7.2 with NaOH (-29 g).

182. Dichloromethane was distilled from P205 before use.

183. The KOH solution was made up of 10% KOH in EDTANa2 solution (ct. Ref. 179).

184. Frost, C., Dioxirane Chemistry - Novel Oxidations, Final Year Research Project, Loughborough University, 1991.

185. The acetic acid I acetone solution was made from glacial acetic acid and acetone solution (ct. Ref. 180) in a ratio of 3 : 2 by volume.

186. The potassium iodide solution was made up t01 0% wlv in distilled water.

187. The sodium thiosulfate solution was made to 0.01182 N in distilled water.

188. The phosphate buffer was made from KH2P04 (1.179 g) and Na2HP04 (4.302 g) in distilled water (1000 ml). This gave a pH of 7.5 at 5 ac.

189. The phosphate buffer (pH 7.4) was made from NaH2P04 (174 g) and Na2HP04 (56 g) made to 150 ml in distilled water. NB. Addition of Oxone® (1 g) to this buffer (5 ml) gave a pH of 6.9.

190. Previous syntheses of ~-methylstyrene oxide (optical rotation of enantiomers measured): Irie, R., Noda, K., Ito, Y., Matsumoto, N. and Katsuki, T., Tetrahedron.·Asymmetry, 1991, 2(7), 481; Conte, V., Di Furia, F., Modena, G., Sbampato, G. and Valle, G., Tetrahedron: Asymmetry, 1991, 2(4), 257; Castedo, L., Castro, J.L. and Riguera, R., Tetrahedron Left., 1984, 25(11),1205; Witkop, B. and Foltz, C.M., J Am. Chem. Soc., 1957, 79, 197.

191. Nahm, S. and Weinreb, S., Tetrahedron Lett., 1981, 22(39),3815.

192. Adams, R. and Ulrich, L.H., J Am. Chem. Soc., 1920, 42, 599; Aldrich Technical Information Bulletin Form Number A 74.

144

Page 167: Epoxidation using dioxiranes · 2019. 9. 7. · monoperoxysulfuric acid. This theory later lost out to a competing mechanism, primarily as a result of an 180 study by Doering and