122
PREPARATION OF HIGHER ALPHA OLEFIN OLIGOMERS by CHANTELLE CRAUSE Dissertation submitted in fulfilment of the requirements for the degree MASTER OF SCIENCE in CHEMISTRY in the FACULTY OF SCIENCE at the RAND AFRIKAANS UNIVERSITY Supervisor: PROF. F.R. VAN HEERDEN DECEMBER 1996

Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

PREPARATION OF HIGHER ALPHA OLEFIN OLIGOMERS

by

CHANTELLE CRAUSE

Dissertation

submitted in fulfilment of the requirements

for the degree

MASTER OF SCIENCE

in

CHEMISTRY

in the

FACULTY OF SCIENCE

at the

RAND AFRIKAANS UNIVERSITY

Supervisor: PROF. F.R. VAN HEERDEN

DECEMBER 1996

Page 2: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

TABLE OF CONTENTS

Synopsis iv

Ops omming vi

List of Schemes viii

List of Figures

List of Tables r xi

Abbreviations xii

Chapter 1

The oligomerization of higher alpha olefins : a literature study

1.1 Alpha olefin beneficiation: the key to value-adding and economic growth 1

1.1.1 The South African perspective 1

1.1.2 Focus on alpha olefins 2

1.1.3 Alpha olefin oligomers as base stocks for synthetic lubricants 3

1.1.4 Use and functionalization of poly(alpha olefins) 4

1.2 General principles of oligomerization 5

1.3 Olefin oligomerization catalysts 6

1.3.1 Friedel-Crafts catalysts 6

Polyalphaolefins 6

Cationic acid catalysed oligomerization 7

BF3-catalysed oligomerization 8

Mechanistic considerations 10

Correlation of paraffin structure with physical

properties 16

Reaction variables 18

1.3.2 Free radical catalysts 19

1.3.3 Transition metal catalysts 20

Heterogeneous Ziegler-Natta catalysts 23

Homogeneous Ziegler-Natta catalysts: Metallocenes 26

(a) Nature and formation of the active species 27

Page 3: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

11

1.4 Conclusion

Kinetics and mechanism of alpha olefin

polymerization 29

Stereoregular polymers from chiral metallocene

catalysts 32

Reaction control and product distribution 35

Future developments 36

(iii) Other important transition metal catalysts 36

Supported reduced transition metal oxide catalysts 36

Homogeneous nickel catalysts 37

38

Chapter 2

Discussion

2.1 Aim of research 39

2.2 Synthesis of model oligomers 40

2.2.1 Synthetic target and strategy 41

2.2.2 Synthesis of 1,2-epoxypentane 42

(i) Direct epoxidation 43

Magnesium monoperphthalate 44

Urea hydrogen peroxide 45

Dimethyldioxirane 45

(ii) Indirect epoxidation 46

Iodohydrin 47

Bromohydrin 48

1,2-Epoxypentane 49

2.2.3 Synthesis of the dimer analogue 50

2.2.4 Synthesis of the trimer analogue 57

(i) Bromination of the C 10 alcohol 57

(ii) Synthesis of the C15 alcohol 59

Formation of Grignard reagent 59

Carbon-carbon bond formation 61

(iii) Synthesis of a model trimer 62

2.2.5 Attempted synthesis of a model tetramer 63

Page 4: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

111

2.3 Catalyst survey 65

2.3.1 Cationic oligomerization 65

2.3.2 Radical oligomerization 67

2.3.3 Metallocene catalysts 69

Metal 76

Metallocene trichloride 77

Substituted cyclopentadienyl rings 78

Temperature 79

Purity of the alpha olefin monomer 80

2.3.4 Comparison of different catalysts 81

Comparison of viscosity data 81

General comparison 81

2.4 Conclusion 82

Chapter 3

Experimental

3.1 General 84

3.2 Standard chemical methods 86

3.3 Preparation of model oligomer 88

3.4 Oligomerization reactions 95

References 102

Acknowledgments 109

Page 5: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

SYNOPSIS

The economic growth of a country can be stimulated by value-adding to basic resources e.g.,

chemical intermediates. Alpha olefins are products of the Sasol Fischer-Tropsch process and

oligomerization is one of the most significant beneficiation processes available for these

alpha olefins.

Alpha olefin oligomers obtained from the cationic oligomerization with BF3/ROH, have

been used mainly as base stocks in synthetic lubricants. These oligomers consist of a

mixture of isomeric products that complicates the investigation and rational modification of

this and other closely related oligomerization processes. A comparison of these isomeric

mixtures to alkane standards could assist in the study of these processes, but almost no

standards are available. This project was aimed, in part, at establishing a methodology for

the synthesis of 1-pentene model oligomers. A model branched 1-pentene dimer and trimer

as well as a precursor to the analogous tetramer were prepared. The syntheses of other

model oligomers in conjunction with the comparison of the properties of model oligomers

with different branching patterns will certainly be of value.

Some of the main classes of catalysts were evaluated for the oligomerization of higher alpha

olefins and especially of 1-pentene. The BF3/n-BuOH cationic oligomerization of 1-pentene

yielded dimers, trimer, tetramers and pentamers consisting of a large number of isomers

and the product distribution peaked at the trimer. This is in agreement with previous

results for other monomers. Radical oligomerization of 1-pentene with organic peroXide

initiators proceeded in low yield. Dimers and trimers consisting of a mixture of isomers

were mainly formed.

Metallocene catalysts offer a convenient route to single component, structurally well

characterized alpha olefin oligomers. A variety of functional group transformations and

consequent applications are possible for these oligomers. The general nature of the

oligomerization of 1-pentene, 1-hexene and 1-octene with metallocene/methylaluminoxane

catalysts was investigated. The influence of different metals, the number of

cyclopentadienyl groups and substitution on the cyclopentadienyl groups in the

metallocene moiety, the reaction temperature and the purity of the monomer were studied.

By employing the appropriate metallocene in combination with methylaluminoxane, higher

Page 6: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

V

alpha olefin dimers and trimers could be formed in high yield at ambient temperatures.

Cp2ZrC12 was the most active of the metallocene catalysts and single isomers were formed

with high selectivity.

The metallocene-based oligomerization of higher alpha olefins offers great scope for further

research and promises a new era in olefin oligomerization.

Page 7: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

OPSOMMING

Die ekonomiese groei van 'n land kan bevorder word deur waarde-toevoeging tot basiese

hulpbronne, byvoorbeeld chemiese grondstowwe. Alfa-olefiene word in die Sasol Fischer-

Tropsch proses vervaardig en oligomerisasie is een van die belangrikste prosesse vir

waarde-toevoeging tot hierdie alfa-olefiene.

Alfa-olefienoligomere verkry deur die kationiese oligomerisasie met BF3/ROH, word

hoofsaaklik as basisolies vir sintetiese smeermiddels aangewend. Hierdie oligomere bevat

'n mengsel van isomeriese produkte wat die ondersoek en rasionele modifikasie van hierdie

en verwante oligomerisasieprosesse bemoeilik. Die vergelyking van hierdie isomeriese

mengsels met alkaanstandaarde kan waardevol wees in die studie van hierdie prosesse,

maar byna geen standaarde is beskikbaar nie. Hierdie projek was gedeeltelik daarop gemik

om 'n metodologie daar te stel vir die sintese van 1-penteen modeloligomere. 'n Vertakte

model 1-penteendimeer en - trimeer asook 'n voorloper tot die analoe tetrameer is berei. Die

sintese van ander modeloligomere, tesame met die vergelyking van die eienskappe van

modeloligomere met verskillende vertakkingspatrone, sal van verdere waarde wees.

Sommige van die hoofgroepe katalisatore is geevalueer in die oligomerisasie van hoer alfa-

olefiene en spesifiek vir 1-penteen. Dimere, trimere, tetramere en pentamere, bestaande uit

'n groot aantal isomere, is gevorm in die BF3/n-BuOH oligomerisasie van 1-penteen en die

produk verspreiding het gepiek by die trimeer. Dit is in ooreenstemming met resultate

verkry vir ander monomere. 'n Lae opbrengs is verkry in die radikaal oligomerisasie van 1-

penteen met organiese peroksiede as inisieerders. Hoofsaaklik dimere en trimere bestaande

uit 'n aantal isomere, is gevorm.

Metalloseenkatalisatore bied 'n eenvoudige metode om enkel-komponent, struktureel goed-

gekarakteriseerde alfa-olefien oligomere te berei. 'n Verskeidenheid funksionele groep

transformasies en gevolglike toepassings is moontlik vir hierdie oligomere. Die algemene

eienskappe van die oligomerisasiereaksies van 1-penteen, 1-hekseen en 1-okteen met

metalloseen/metielaluminoksaan katalisatore is ondersoek. Die invloed van verskillende

metale, die aantal siklopentadienielgroepe en substitusie op die siklopentadienielgroepe in

die metalloseeneenheid asook die reaksie temperatuur en suiwerheid van die monomeer is

bestudeer. Hoer alfa-olefiendimere en - trimere is gevorm by kamertemperatuur deur die

Page 8: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

vii

gepaste metalloseen saam met metielaluminoksaan as katalisator to gebruik. Cp2ZrC12 was

die aktiefste metalloseenkatalisator en enkele isomere is gevorm met 'n hoe selektiwiteit.

Die metalloseengebaseerde oligomerisasie van hoer alfa-olefiene bied ruimte vir verdere

navorsing en lei 'n nuwe era in olefienoligomerisasie in.

Page 9: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

vii'

LIST OF SCHEMES

Chapter 1

The oligomerization of higher alpha olefins : a literature study

Scheme 1.1. Kinetics of cationic chain growth oligomerization. 8

Scheme 1.2. Formation of catalyst complex. 9

Scheme 1.3. Conventional cationic mechanism. 11

Scheme 1.4. Shubkin's mechanism. 12

Scheme 1.5. Oligomerization of 1-butene and cis-2-butene. 15

Scheme 1.6. First step in oligomer formation. 16

Scheme 1.7. Free radical polymerization of vinyl monomers. 20

Scheme 1.8. Simplified hydride mechanism. 21

Scheme 1.9. Formation of active heterogeneous Ziegler-Natta catalyst. 24

Scheme 1.10. Monometallic Ziegler-Natta mechanism. 25

Scheme 1.11. Formation of active metallocene/ MAO species. 27

Scheme 1.12. Formation of active MAO-free catalysts. 29

Scheme 1.13. Basic steps in polymerization of alpha olefins using

metallocene-based catalysts. 30

Scheme 1.14. Illustration of agostic interactions. 31

Scheme 1.15. An example of stereocontrol in metallocene/ MAO polymerization. 34

Chapter 2

Discussion

Scheme 2.1. General route to tetraalkylmethanes. 41

Scheme 2.2. Retrosynthetic strategy. 43

Scheme 2.3. MMPP epoxidation of 1-pentene. 44

Scheme 2.4. UHP epoxidation of 1-pentene. 45

Scheme 2.5. Dimethyldioxirane epoxidation of 1-pentene. 46

Scheme 2.6. Formation of iodohydrin. 47

Scheme 2.7. Formation of bromohydrin. 48

Scheme 2.8. Illustration of the preparation of an epoxide from a halohydrin. 49

Page 10: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

ix

Scheme 2.9. Preparation of 1,2-epoxypentane. 50

Scheme 2.10. Formation of C10 alcohol. 53

Scheme 2.11. Formation of two stereogenic centres. 56

Scheme 2.12. Bromination of 2-heptanol. 58

Scheme 2.13. Bromination of alcohol 44. 58

Scheme 2.14. Generation of Clo Grignard reagent. 59

Scheme 2.15. Preparation of C10 iodide. 60

Scheme 2.16. Formation of C15 alcohol. 61

Scheme 2.17. Preparation of model trimer. 63

Scheme 2.18. Decomposition of a secondary organic peroxide. 69

Scheme 2.19. Formation of methylated oligomers. 73

Scheme 2.20. Formation of normal dimers and trimers. 74

Page 11: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

LIST OF FIGURES

Chapter 1

The oligomerization of higher alpha olefins: a literature study

Figure 1.1. Possible functionalization and application of single

component alpha olefin oligomers. 5

Figure 1.2. Gas chromatogram of typical oligomer. 9

Figure 1.3. Gas chromatogram of 1-decene trimer. 10

Figure 1.4. Bent metallocene. 26

Figure 1.5. Relationship of pro-catalyst symmetry to polypropylene

stereostructure. 33

Figure 1.6. Two types of supported chromium catalysts used in ethylene

polymerization. 37

Chapter 2

Discussion

Figure 2.1. 1-Pentene model trimer. 42

Figure 2.2. Gas chromatogram of two alcohol produdts. 54

Figure 2.3. Gas chromatogram of 1-pentene trimer. 66

Figure 2.4. Structure of oligomers. 75

Page 12: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

xi

LIST OF TABLES

Chapter 2

Discussion

Table 2.1. Types of organocopper reagents. 51

Table 2.2. Summary of model C-C bond formation reactions. 52

Table 2.3. Influence of different organometallic nucleophiles on the addition

of a Grignard reagent to an epoxide. 55

Table 2.4. Attempted functional group transformations. 64

Table 2.5. Product distribution of the BF3/n-BuOH oligomerization of 1-pentene. 66

Table 2.6. Radical oligomerization of 1-pentene with organic peroxide initiators. 68

Table 2.7. Oligomerization of alpha olefins with titanocene dichloride/MAO. 71

Table 2.8. Oligomerization of alpha olefins with Cp2ZrC12/ MAO. 77

Table 2.9. Oligomerization of alpha olefins with CpTiC13/ MAO. 78

Table 2.10. Room temperature oligomerization of alpha olefins with

Cp2ZrC12/ MAO. 80

Table 2.11. Viscosity/temperature relationship of 1-pentene trimer fractions 81

Page 13: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

xii

ABBREVIATIONS

by boiling point

Bu butyl

CI chemical ionization

Cp cyclopentadienyl

El electron ionization

fg functional group

GC gas chromatograph

MAO methylaluminoxane

MCPBA meta-chloroperoxybenzoic acid

Me methyl

MMPP magnesium monoperphthalate

MS mass spectrometer

Ms mesylate

NMR nuclear magnetic resonance

PAO polyalphaolefin

Pr propyl

psi pounds per square inch

rf radio frequency

rt room temperature

SHOP Shell Higher Olefin Process

THE tetrahydrofuran

Ts tosylate

UHP urea hydrogen peroxide

Page 14: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

CHAPTER 1

THE OLIGOMERIZATION OF HIGHER ALPHA OLEFINS:

A LITERATURE . STUDY

1.1 Alpha olefin beneficiation: the key to value-adding and economic

growth

1.1.1 The South African perspective

The value-adding to manufactured goods is one of the pivotal factors governing the

economic growth and competitiveness of a country.. South Africa is blessed with abundant

natural resources. Gold, platinum, diamonds and coal are some examples of South Africa's

mineral wealth. Coal is of particular importance to the local chemical industry since it can

be converted into liquid hydrocarbons. The industrial process utilizing coal derived

mixtures of carbon monoxide and hydrogen to produce hydrocarbons, is based on the work

of Fischer and Tropsch. 1 South Africa, starting in the 1950's, developed an updated

Fischer-Tropsch synthetic fuel oil plant (Sasol) using improved engineering and technology.

These synfuels and synchemical plants are a lucrative means of adding substantial value to

more than 40-million tons of bituminous coal. 2

The Sasol synfuel plants were originally designed to produce mainly fuels, ethylene and

unavoidable byproducts such as ammonia. The Sasol/ Lurgi gasification process uses coal,

water and oxygen to generate synthesis gas (syngas: CO + H2) and a range of hydrocarbon

chemicals. The Fischer-Tropsch high-temperature synthesis reaction employs syngas to

create liquid fuels and a broad spectrum of useful olefins, alcohols and ketones. Market

growths in conjunction with the development of new separation and purification

technologies has realized the opportunity to redirect synfuel feedstocks to petrochemical

feedstocks. The 1994 Sasol alpha olefin plant, producing comonomer quality 1-hexene and

1-pentene, is a prime example of a profitable value-adding expansion.

Page 15: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

2

Approximately two million tons of olefins (C2 - Cio range) are produced annually in the

Fischer-Tropsch process, of which about a third is currently recovered for the petrochemical

market. The balance is still being converted into liquid fuels. The new alpha olefin plant has

the potential to yield 450 000 tons per year of high quality 1-pentene and 1-hexene. This

beneficiation project forms part of Sasol's strategy to add greater value to its treasure chest

of petrochemicals. Alpha olefins represent a major niche-market opportunity for Sasol, viz.:

the 1-hexene market has been in tight supply, while the 1-pentene comonomer was not

previously available in commercial quantities. The feasibility of the production of additional

linear alpha olefins in the C2 to Cu range, is under consideration. 3,4

Sasol boasts a competitive advantage in petrochemicals, which stems from its unique one-

step process of producing fuels and petrochemicals without the larger capital investment

and higher operating costs of naphtha crackers. Seen in the light of this competitive edge,

further attention should be focused on in-house alpha olefin beneficiation.

1.1.2 Focus on alpha olefins

"Alpha olefins" is a generic term referring to linear alkenes with the double bond in the first

or alpha position. These olefins are important feedstocks in the petrochemical industry for

they are versatile chemical intermediates. Higher linear alpha olefins (C6 - Go) can be

transformed into biodegradable surfactants, new types of polymers, lubricants and several

industrially useful chemicals.

The widespread commercial usage of alpha olefins can be attributed to their ready

availability, low cost, high reactivity and ease of transformation into a range of valuable

products. Higher linear olefins have been available in commercial quantities since the

1960's. The processes used for the preparation of alpha olefins comprise: thermal and

catalytic cracking of paraffins, oligomerization of ethylene, dehydrogenation of paraffins,

dimerization and metathesis of olefins, dehydration of alcohols and electrolysis of C3 - C30

straight-chain carboxylic acids. The first process is used for commercial production of C2 - C5

olefins and the second for the large scale manufacturing of C9 - C30 alpha olefins.

Page 16: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

3 •

The high reactivity of the alpha olefins can be explained in terms of the basic structure of

alkenes. The carbon-carbon double bond is best described by using an sp 2 hybridized

orbital model which predicts a strong c bond and a weaker it bond. The overlapping 7C-TC

bond behaves as a high-electron-density site and can be accessible to electrophilic reagents.

In alpha olefins the double bond is in the terminal position and is more available than

internal alkenes that is more sterically hindered. Therefore, alpha olefins are generally more

reactive than internal olefins. Additionally, functional group modifications often result in

products that are also linear, and this may be advantageous to the industrial user, e.g. the

linear surfactants obtained from alpha olefins are biodegradable. 4,5

1.1.3 Alpha olefin oligomers as base stocks for synthetic lubricants

Oligomerization of alpha olefins to produce feedstocks for synthetic lubricants is one of an

array of processes available for the beneficiation of alpha olefins. In the most fundamental

sense, a lubricant may be defined as a substance that has the ability to reduce friction

between two solid surfaces rubbed against each other. Animal fats, vegetable oils as well as

mineral oils have been used throughout history to achieve this effect. For more than a

millennium petroleum products have been employed in a primitive fashion and to this day,

petroleum based lubricants play a dominant role in the world lubricant market.

The first synthetic hydrocarbon oils were produced in 1877 by Charles Friedel and James

Mason Crafts. Synthetic lubricants are not obtained directly from crude oil and may be

described as blends of base stocks made from low-molecular weight (chiefly petrochemical)

building blocks. These base stocks consist of a limited number of components with uniform

molecular structures and, therefore, well defined properties that can be tailored to specific

applications.

Initial attempts to commercialize synthetic hydrocarbon oils failed due to insufficient

demand. However, this situation is changing. Technological advances in modern machines

and engines have been placing a greater pressure on the performance of existing functional

fluids. Today's lubricants are facing increasingly stringent demands. Some of these

challenges are cost effectiveness, lessened dependance on crude oil reserves,

biodegradability, specialized performance requirements and operation under more severe

Page 17: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

4

conditions. These limitations led to the development of synthetic lubricants such as

polyalphaolefins, organic esters, silicones and polyglycols. Although the replacement of

mineral oil by synthetic lubricants seems farfetched, it should be noted that current

estimates predict that we may run out of crude oil in the next century.

Hydrogenated polyalphaolefins form a class of high-performance fluids considered to be the

most versatile of the synthetic lubricants. The term polyalphaolefin (PAO) is commonly

used to designate such fluids derived from the catalytic oligomerization of linear alpha

olefins having six or more (usually ten) carbon atom s . Owing to inherent and highly sought

after characteristics of PAO's these fluids are rapidly gaining ground as high-performance

lubricants. PAO's are currently being used as base stocks in, for example, automatic

crankcase lubricants, hydraulic oils and turbine oils. Synthetic lubricants manufactured

from alpha olefins can improve motor fuel efficiency and prolong oil lifetime. 6,7

1.1.4 Use and functionalization of poly(alpha olefins)

Alpha olefin oligomers and their derivatives are central to modem life, not only as synthetic

lubricants but in a variety of end-uses. Attention will henceforth be focused on the general

oligomerization process, use and functionalization of oligomers and the possible catalyst

systems active in this reaction.

Isomeric mixtures of olefin oligomers accessible via cationic catalysts like the BF3/ROH

system are commonly used in lubricant applications. Metallocene catalysts now offer a

convenient route to single component, structurally well characterized alpha olefin

oligomers. The possible functionalization and potential application of these oligomers are

depicted in Figure 1.1.8

Page 18: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

additives for lacquers, / paper and leather industry

fuel additives

OH NH2

5

R OH R'..,. .OH -I. adhesives (Co-)Polymer

lubricants ointment white oil R' CH3

--PI.. fragrances

Figure 1.1. Possible functionalization and application of single component alpha olefin

oligomers. 8

1.2 General principles of oligomerization

According to the IUPAC nomenclature, oligomers 9 are molecules containing a few

constitutional units repetitively linked to each other. No specific molecular weight or degree

of oligomerization is defined. It states, however, that the physical properties of oligomers

vary with the addition or removal of one or several units. Polyolefin chain formation,

depending on a number of n reacting molecules, can in most cases be named as follows:

dimerization (n = 2), oligomerization (2 < n < 100) and polymerization (n > 100). 5

Alpha olefins may be oligomerized thermally or catalytically; the latter having industrial

application. A variety of oligomerization catalysts are known and reactions can be classified

accordingly as acid-, base-, radical- or metal catalysed.

The typical catalytic cycle of these catalysts consists of initiation, propagation and

termination steps. The relative propagation (r e) and termination (1-.1 rates determine the

molecular weight of the obtained product. If r p > rt many propagation steps take place

before termination and a high molecular weight polymer is formed, when rt > r p dimers are

produced, and finally when r t rp oligomers are obtained.5

Page 19: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

6

The main oligomerization catalysts will be discussed in the following section, with specific

reference to higher alpha olefin monomers, reaction mechanisms and the factors governing

the product distribution.

1.3 Olefin oligomerization catalysts

1.3.1 Friedel-Crafts catalysts -

As early as 1877 Charles Friedel and James Mason Crafts found that anhydrous aluminium

chloride could be used to catalyse the polymerization of unsaturated hydrocarbons. During

the ensuing years aluminium chloride was shown to be a most versatile catalyst effecting an

almost unlimited number of reactions. Today any substitution, isomerization, elimination,

cracking, polymerization, or addition reaction that takes place under the catalytic effect of

Lewis acid-type acidic halides (with or without cocatalysts) or proton acids, are regarded as

Friedel-Crafts-type reactions. 10

i) Polyalphaolefins

Friedel-Crafts type catalysts are currently the only ones used commercially for producing

PAO base stocks for synthetic lubricants. A few aspects concerning polyalphaolefins should

be highlighted and expanded. Polyalphaolefins are the largest volume synthetic lubricant

base stock worldwide and are considered to have the largest potential for growth. The

combination of their relative low cost and desirable properties make the PAO's the preferred

base stock for a wide variety of synthetic lubricant applications. Among these favorable

properties are:

A wide operational temperature range

High viscosity index

Thermal stability

Oxidative stability

Hydroly tic stabili ty

• Shear stability

Low corrosivity

Page 20: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

7

Compatibility with mineral oils

Compatibility with various materials of construction

Low toxicity

Manufacturing flexibility that allows the "tailoring" of products to specific end-use

application requirements.

However, PAO's lack the ability to swell the elastomers used in the seals of many lubricant

systems, and they have limited ability to dissolve the additives needed to enhance critical

properties. As a result, they are often used in combination with petroleum-based oils

and/or esters (primarily dibasic acid esters) to overcome these deficiencies. The synthetic

lubricants compete with petroleum lubricants on a cost-competitive basis, but some key

environmental and quality policy changes could affect future growth positively. 6,7

ii) Cationic acid catalysed oligomerization

Lewis - and Bnansted acids, that is, the common Friedel-Crafts catalysts, may effect the

oligomerization of alkenes. 11 Cationic oligomerization is distinguished by the nature of the

carrier ions on the growing chains: in the process of growth, an ion pair is formed in which

the chains carry carbocations. Initiation occurs by electrophilic attack of the initiators on the

monomers. A general scheme for cationic oligomerization is presented in Scheme 1.1. The

coinitiator is presented as A, the initiator as RH, and the monomer as M.

Generally, cationic oligomerization depends on the initial formation of positive and negative

ions in an organic environment. Initiating systems include: Lewis acid/Bronsted acid

combinations, Brensted acids by themselves, stable cationic salts, some organometallic

compounds, and some cation forming substance's. Propagation takes place through

successive addition of monomeric units to the charged or reactive groups of the propagating

chains. Terminations are the result of unimolecular reactions, or transfers to other

molecules (such as monomers, solvents or impurities like moisture). They can also result

from quenching by deliberate addition of reactive terminating species.'2

Page 21: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

8

Cationic Initiation Process :

A + RH K

[H+] [AR- ]

[HM+][AR- ]

[HMnM+][Ak-

Mn+1 + [HM+][AR-

Mn+1 [HI [AR-

Rate of initiation :

Ri = KKi[A][RH][M]

Rate of propagation :

Rp = KpRHM+) (AR - )] [M]

Rate of termination :

Rt = Kt [ (HMnM+)(ARli

where Kt is a combination of

Ktr, Kst and Kr

Ki [H+][AR- ] + M

Cationic Propagation Process :

[HMn+] [AR- ] + M Kp

Cationic Termination :

By transfer :

+ M [HMnM+] [AR- Kt _

By spontaneous termination

[HMnM+] [AR] Kst

By rearrangement of the kinetic chain :

[HMnM+][AR- ] Kr HMnMAR

Scheme 1.1. Kinetics of cationic chain growth oligomerization. 12

iii) BF3-catalysed oligomerization of alkenes

Cationic catalysts, such as AlCb, have been reported for the oligomerization of alkenes as

early as 1931. 13 In 1951 a patent was issued to Gulf oil company describing the use of AlCb

for the oligomerization of 1-octene. 14 Brennan (Mobil Oil) . showed in 1968 that

oligomerization of alpha olefins could be carried out in a reproducible fashion with boron

trifluoride in the presence of a protic cocatalyst. 15 Commercial PAO's are made from 1-

decene. The low viscosity products are made using BF3 catalysts, while the higher viscosity

PAO's are produced with catalysts like a trialkylaluminium or alkylaluminiumhalide in

conjunction with a halogen or halide source. 716

BF3 is the catalyst of choice for the most common grades of PAO's. While incapable of

catalyzing the oligomerization of alpha olefins on its own, BF3 is active when used in

conjunction with a proton-donating cocatalyst (typically a low molecular weight alcohol

Page 22: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

9

such as n-butanol). BF3 complexes with proton-donating compounds by one of the

following general reactions:

BF3 (g) + ROH (1) --> BF3.ROH (1)

or

BF3 (g) + 2 ROH (1) ---> BF3.2ROH (1)

Scheme 1.2. Formation of catalyst complex!

Since BF3 is normally used in excess, the BF3 and alcohol cocatalyst react in a 1:1 molar ratio

to form the catalyst complex BF 3.ROH. The BF3.ROH system is unique for two reasons: the

oligomer peaks at the trimer (Figure 1.2) and the products exhibit unexpectedly good

low-temperature viscosity characteristics due to the presence of excess branching in the

molecular skeleton.

TA ImEF,S

TC7.7AmEA S

CMEPS

?ENT AmEqS

HEXAMEAS

10' 1Sr 2d 23. T al. fain)

Figure 1.2. Gas chromatogram of typical oligomer.16

Figure 1.2 is a high resolution GC trace of a typical oligomerization product resulting from

the reaction of 1-decene in the presence of the BF 3.ROH complex. Figure 1.3 shows clearly

that the single oligomer fractions are composed of a very large number of branched isomers.

Page 23: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

10

These isomer mixtures form even under tightly controlled reaction conditions.

Notwithstanding, this has the advantage, especially in lubricant applications, of lowering

the melting point of the trimer mixture relative to that of pure compounds of comparable

molecular weight and structure. 17

lOr 12. 14' le 18" 20' 22' 24' 19" 19'

Ti.. (ain)

Figure 1.3. Gas chromatogram of 1-decene trimer.' 6

a) Mechanistic considerations

1H NMR studies of the hydrogenated oligomers indicated about one additional methyl

group compared to that predicted on the basis of the conventional cationic mechanism

(Scheme 1.3). As yet no mechanism that explains this branching has been proven

conclusively in view of the complexity of the products. Various studies of the hydrogenated

and distilled oligomer fractions indicate anomalies with the theoretically expected results.

Skeletal rearrangement at the dimer stage, 18 rearrangement of the monomer 19 as well as

positional isomerization of the monomer 20 have been proposed in order to explain these

discrepancies.

2522a

Page 24: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

- H-F

H-F Isomers

11

H+

- H+

H+

- H+

H+

- H+

R— CH-2 CH= CH2

MONOMER

CH3 R— CH-2 CH— CH= CH—CH-2 R

(+ isomers)

DIMER

CH3 R— CH— CI H— CH—C H— CH-2 R 2 2

CH I I

(+ isomers) CH

H2

R- CH2 CH—CH3

R—CH-2 CH=CH2

CH I + R— CH-2 CH— CH-2 2 CH—CH— R

R— CH-2 CH= CH2

CH3 R— CH-2 CH— CH-2 CH— CH-2 R

CH2 +CH

CH2 R R

TRIMER

Scheme 1.3. Conventional cationic mechanism.

Experimental work of Shubkin et al.18 indicated that the conventional mechanism did not

explain the composition of the oligomerization products adequately. Firstly, the unique

product distribution could not be clarified: at lower temperature (30 °C) mainly trimer was

formed, while the product peaked at the dimer at elevated temperature (85 °C). Arguments

based on steric crowding were unconvincing. Secondly, the physical properties of the

oligomerization products differed from the expected: lower viscosities and significantly

lower pour points than would have been predicted from the expected molecular structures,

were found. Further investigation revealed that on average two-thirds of the oligomers had

one more methyl group (in other words one more branch) than expected.

The observed nature of the BF3-catalysed oligomerization reaction, as well as the physical

and chemical properties of the products, led Shubkin 18 to suggest a mechanism that includes

skeletal rearrangement (Scheme 1.4). According to Shubkin, rearrangement occurs only at

the dimer stage, proceeding via a protonated cyclopropyl intermediate.

Page 25: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

12

R—CH=CH—CH3 3

H+ Double bond migration

H-F R—CH-2 CH— CH3 2

+ 1

CH H \ 3/

C'■ 2 CH2

R—CH—

4

2 R

R—CH-2 CH=CH2 -

1

CH I 3

R—CH2 CH

Cx, / —CH2—R

H+ CH2 6

1 CH3

R—CH-2 C—CH—CH-2 R

7 CH3 +

H+1[ -H+

ICH3

R—CH=C—CH—CH—R I 2 CH3

+

C H3 I

8

CH3 R—CH-2 CH—CH=CH—CH-2

5

CH3

R—CH-2 C—CH 2—CH—R 1 \

CH2 CH3 + CH

I 9 CH2 I R

+ 1

H+ Double bond R migration - H+

H\14+

—— R—CH2 C= C —CHR I 2

CH3 Tetramer

11

CH3 R—CH2 —C—CH—CH—R

1 \ 2

CH CH3 CH 10

CH2 R

Scheme 1.4. Shubkin's mechanism's

Page 26: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

13

Protonation of the olefin (1 ---> 2) is followed by nucleophilic attack by another olefin

molecule to give a protonated dimeric species (4). This species may then deprotonate to

form dimeric product (5) and/or isomerize by double bond migration. Shubkin states,

however, that there is a unique feature about intermediate 4 which is the key to this

reaction: there is a labile, tertiary hydrogen p to the carbonium ion. Skeletal rearrangement

through a protonated cyclopropyl intermediate (6) to give the more stable tertiary

carbocation 7 is thermodynamically favored. Intermediate 7 may deprotonate to 8 and

undergo isomerization by double-bond migration. But the inherent stability of a tertiary

carbocation makes it available for attack by another olefin to produce the trimeric

carbocation 9. This is a secondary carbocation and therefore less stable than 7; furthermore,

it cannot rearrange the way that 4 does. The carbon atoms p to the carbocation either

contain no hydrogen atom or only secondary hydrogen atoms which are not particularly

labile. Thus deprotonation to 10, the trimeric olefin, is favored over reaction with an

additional alpha olefin to give 11, the tetramer.

Onopchenko and coworkers 19 studied the oligomerization of 4-methyl-1-pentene by the BF3

catalyst system to test Shubkin's proposed mechanism. 1H NMR of the dimer and trimer

product fractions showed 5.0 and 6.8 methyl groups, respectively, instead of the expected 6

and 8 predicted by Shubkin's mechanism. According to Onopchenko, this indicated that: (a)

no methyl group rearrangement occurred at the dimer stage or, (b) that the initial secondary

4-methylpentyl carbocation had rearranged at the monomer stage and was behaving as the

tertiary carbocation derived from the vinylidene alkenes.

In order to provide evidence that methyl group migrations can occur at the monomer stage

(and not only at the dimer stage as claimed by . Shubkin et al.18), the composition of

unreacted decenes were determined under varying conditions. Onopchenko et al. 19

confirmed the results of Brennan: 17 At high 1-decene conversion (90%), the unreacted alkene

contained only 60% of 1-decene, 35% of internal alkenes and 5% of methyl nonenes.

Conversely, at low 1-decene conversion (30%), less than 0.5% of methylnonenes was

detected. This data suggested that the BF3 complex is a mild isomerization catalyst, but an

effective oligomerization catalyst.

The complexity of the oligomers formed from 1-decene necessitated the examination of a

simpler system. The structures of the dimers from the oligomerization of 1-butene and

Page 27: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

14

cis-2-butene using the BF3/mannitol complex were investigated (theoretically only 18

isomers are possible for C8F112paraffins).

The data obtained showed that all of the isomers had either three, four or five methyl

groups. The similarity of the product isomers from the two butenes suggested that the

same intermediate had to be involved in both cases. The common intermediate is the sec-

butyl carbocation. The only significant difference in the oligomerization products from the

two butenes was the formation of substantially more 3-methylheptane from 1-butene and

3,4-dimethylhexane from 2-butene.

Onopchenko et al.19 assumed that two combinations (sec-butyl 12 and tert-butyl 13 in Scheme

1.5) were involved in the dimerization of the two butenes. This assumption followed from

the postulated mechanism for the Koch reaction, 21 where a protonated cyclopropane is the

key-intermediate in the formation of tertiary alkyl carboxylic acids. Onopchenko and

coworkers used this reaction to predict the nature of the major isomers.

Following the protonation of either 1- or 2-butene, the secondary butyl carbocation

rearranges to the tertiary butyl carbocation, and both then add to an isomerized mixture of

alkenes. The intermediate ions can then undergo deprotonation to afford the corresponding

olefinic dimers and their isomers. Eventually six paraffins (14 - 19) would result from each

butene, but these accounted for less than 70% of the experimentally determined number of

isomers. Eight additional isomers were present and this suggested that some of the

expected intermediates had rearranged further and thus the model used by Onopchenko et

al. was only an approximation. The total percentage of isomers 16 - 19, nevertheless, could

be used to estimate the extent of methyl group migration at the monomer stage - it was in

the range of 14 - 22% for the two butenes considered.

Driscoll and Linkletter 20 used NMR to analyse 1-decene dirner fractions. Their conclusion

was that the identified structures were formed by the attack of secondary carbocations on

secondary olefin. Further studies, incorporating 1-decene selectively enriched in position 2,

suggested that in most cases oligomerizations took place internally and not at the terminal

double bond.

Page 28: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

15

C—C—C=C -

- C—C=C—C rearrangement C—c—C

12 13

C C C I I I

C—C—C—C C—C—C—C C—C—C—C C—C—C C—C—C C—C—C + + + + + +

1 C=C—C—C 1C—C=C—C 1C —LC 1C=C—C—C 1C—C=C—C 1C—LC

C—C—C—C

C

+C

C

CI

1- H+ 1- H+ H+

H+ 1- H+

C C C

I I I c—c—c—c c—c—c—c c—c—c—c c—C—C C—C—C C—C—C

CI I I I I I

II c—c C C C—C

c II II II I I II

C C C—C C C C—C I I I I I I c c C C C C

1 i C c

H2 1 H2 "1 H2 I H2 1 H2 1 H2

C C C

I I I C—C—C—C C—C—C—C C—C—C—C C—C—C C—C—C C—C—C

C C I I I

C I

C C—C C—C

15 c1 1

c1

18 c1 i

14 C 16 c—c 17 19 c—c

C I I I I I c c C c C

C C

3-methyl- 3,4-dimethyl- 2,4-dimethyl- 2,2-dimethyl- 2,2,3-trimethyl- 2,2,4-trimethyl-

heptane hexane hexane hexane pentane pentane Total:

1-C4= : (34.4%) (8.6%) (14.3%) (2.2%) (1.8%) (3.8%) 64.1%

(3.0%) 69.7%

Scheme 1.5. Oligomerization of 1-butene and cis-2-butene."

c c c I I I

c—c—c—c C—C—C—C C—C—C C—C—C C—C—C I I I I

C C CI

I C—C C—C

-EC C—C+ I

+C I

+C I

C—C + I

cl

cI

CI

C I

CI

2-C4= : (14.2%) (41.2%) (8.7%) (1.5%) (1.1%)

Page 29: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

16

Thus, double bond isomerization took place appreciably faster than oligomerization and the

first step in the mechanism can be depicted as follows:

H3c(042)7&1=cH2

H3C(CH2)mCH=CH(CH2)nCH3 H2C=CH(CH2)6h-I2CH3

Scheme 1.6. First step in oligomer formation 2°

This step embodies the formation of the dimer through the reaction of a secondary

carbocation with a primary or secondary olefin. According to Driscoll and Linkletter 2° linear

olefins such as decene, in all probability have limited thermodynamic preference for the

internal olefin isomers and as a result, carbocation formation will proceed at random along

the chain. The attack on the primary olefin will be preferred sterically, but not statistically.

This led Driscoll and Linkletter to claim that random bonding, primarily by secondary

carbon atoms, would take place to afford a large number of isomers containing the labelled

carbon next to the methyl group. This proposed mechanism explains the presence of more

methyl groups than predicted by the conventional cationic mechanism (attack of a

secondary carbocation on a primary olefin)..

b) Correlation of paraffin structure with physical properties

Before the synthesis of novel fluids with improved physical and thermal properties can be

considered, some understanding of the structure-property relationship is required.

Cupples, Kresge, Onopchenko and Pelligrini 22 related the structure of paraffins to physical

properties such as viscosity, pour point and in some cases thermal stability. The pour point

of a lubricant may be described as the temperature just before the fluid solidifies during

cooling, in other words it is the freezing point.

As the carbon number of the feed decreased, the low temperature viscosity increased; thus

the viscosity-temperature relationship worsened. Likewise, as more branching was

introduced into the molecule, the viscosity of the oligomers deteriorated. Introduction of a

short chain (methyl) in the 2-position of n-alkanes practically had no effect on the viscosity,

but moving the methyl group toward the centre of the alkane backbone showed some

Page 30: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

17

deviation. The introduction of a single methyl branch into a linear chain lowered the

melting point significantly, but the effect was most pronounced when the branch was near

the centre of the chain. Lengthening the branch to C4 or longer reduced the melting point

even further. When considering the low temperature performance of fluids, including both

viscosity and pour point, it should be noted that the structural requirements for good low

temperature viscosity and low pour point are conflicting. The low degree of branching of

the linear alpha olefin oligomers which imparts their superior viscosity properties is

precisely what will cause these oligomers to be the first to fail a given pour point. In short,

the preferred structure would appear to be a molecule having linear characteristics but with

sufficient branching to provide the needed pour points.

Thermal stability seemed to be favored by less branched, more linear structures; as the

number of tertiary hydrogens increased, the thermal stability decreased. This, of course, is

consistent with the data of Benson, 23 who reported that the tertiary carbon-hydrogen bonds

have the lowest bond dissociation energies and are therefore considered the weak points to

thermal stability.

Brennanl7 found that the following theoretical molecular structure, in which the

concentration of atoms is very close to the centre of the chain of atoms, is associated with

superior low temperature fluidity:

R1

R— C— R2

R3

where R and R2 are nearly identical in size, R is generally a hydrogen but may be an alkyl

group, and R3 can vary from a methyl group to a group larger than R and R 2.. The carbon

atoms comprising R, R2 and R3 may be arranged in a chain or a ring. Compounds with

carbon numbers up to 36 in which Ri is a hydrogen and R, R2 and R3 are alkyl groups and

nearly identical in size, exhibited very good wide temperature range fluidity. Long chain

linear alpha olefins should yield trimer fractions structurally similar to this ideal and thus

the trimerization of long-chain olefins is a one-step synthesis of a superior lubricant

molecule.17

Page 31: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

18

c) Reaction variables

Even though the mechanism of the BF3.ROH catalysed oligomerization remains to be fully

elucidated, researchers have learned how to advantageously control the composition of the

final PAO products in order to tailor the oligomer distribution to fit the requirements of

specialized end-use applications. 16 This customizing may be done by manipulation of the

reaction variables, which include: chain length of olefin raw material, temperature, reaction

time and pressure, catalyst concentration, cocatalyst type and concentration, cocatalyst feed

rate, olefin feed rate, reaction quench and recovery procedures, hydrogenation catalyst and

conditions, and distillation. 6

It is now clear why BF3/decene was adopted as the basis of manufacture by every major

producer of low viscosity PAO's. It is, however, normally not possible to manufacture

higher viscosity products using BF3.ROH technology, but other catalyst systems are known

that can give the desired products. These include alkylaluminium compounds in

conjunction with TiC4 or allcylhalides, peroxides and A1Cb catalysts. 6,16

The function of the cocatalyst is obviously complex. Shubkin and Kerkemeyer 16 showed

empirically that the choice of cocatalyst and the concentration at which it is used plays an

important role in determining the composition of the product PAO. Priola et al.24 observed

that an increased temperature decreased the molecular weight of the oligomers as well as

increased the presence of fragmentation products. Shubkin et al. 16 also established that

higher temperatures increased the branching in the oligomers as well as promoted some

depolymerization. Relative to the reaction time, both the composition and the isomeric

distribution of the product change. This represents a useful opportunity for tailor making

PAO's with special physical properties. 6

In addition to controlling the relative distribution of oligomers by manipulation of the

reaction parameters, major alterations can be made to the product properties by the choice

of starting olefin to better satisfy the requirements for a particular end-use application.

Fluids prepared from short-chain monomers tend to have low pour points and moderately

low viscosity indices, whereas fluids prepared from long chain monomers have moderately

low pour points and high viscosities. 22 A final consideration is the position of the double

bond in the starting olefin. Shubkin et al. 16 established that increased levels of internal

Page 32: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

19

olefins in the starting mixture resulted in slower reactions and more dimer in the product.

In addition the low temperature properties of the hinter were influenced: the pour point

decreased and the viscosity increased.

The cationic oligomerization of unrefined C 10 cuts obtained from the Fischer-Tropsch

process have also been researched. 28,26 The raw Cio Fischer-Tropsch products are suitable

for oligomerization and the included alcohols serve as promoters. Oxygenated compounds,

aromatics as well as branched, internal and cyclic olefins were found to be more

problematic. The presence of ketones led to early termination of the oligomerization

reaction and as a result the oligomer distribution was shifted to the lower end.

Additionally, the aromatics caused a lowering of the viscosity indices. Preliminary

purification of the feed enabled better reaction control and product quality could be

improved.

Much of the recent research activity is centred on the development of either new catalysts

that would alleviate the environmental problems associated with BF3, or alternate feedstocks

that are less expensive than 1-decene. 7

1.3.2 Free radical catalysts

The use of organic peroxides as free-radical initiators for polymerization became important

during World War II as a result of the increased demand for synthetic rubber and plastics. 27

It has been reported in a Mobil European patent application that high viscosity PAO's may

be produced by dimerizing lower oligomers with peroxides. 28 In reality this is not currently

economically feasible, since stoichiometric quantities of di-tert-butyl peroxide are used.

The steps in free radical polymerization are illustrated below. The formation of initiating

radicals is the rate-determining step in the initiation reaction. Numerous reactions lead to

the formation of free radicals: thermal decompositions with azo and peroxy groups are

common sources, but "redox" reactions, light induced decompositions and ionizing

radiation can also be used to form initiating radicals. 12 These initiating radicals must be

energetic enough to react with the vinyl compounds. Once the initiating radical is formed,

there is competition between the addition to the monomer and all other secondary reactions.

After the initiating radical has diffused into the proximity of the monomer, the capture of

Page 33: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

20

the free radical by the monomer is a straightforward addition reaction and completes the

step of initiation. By comparison to other steps in the polymerization, initiation is slow and

requires high activation energy. The rate of propagation depends upon both the reactivity

of the monomer and the growing radical chain. Steric hindrance, polar effects and

resonance are important factors in this reaction. The termination processes in free-radical

polymerization are: combination, disproportionation or chain transfer.

Initiation :

R—R --I,- 2 R".

R* + CH2= CHX R—CHT CHX •

Propagation :

R— CH-2 CHX " + n CH =---- CHX ---0- R— (—CH-2 CHX—)n— CH-2 CHX • 2

Termination :

By combination

R— (— CH— CHX--)n—CHT CHX • + • XHC— CHT(—XHC— CH2)n—R

R— (—CHTCHX—)n— CHTCHX— XHC— CHT (— XHC— CHT-) n — R

By disproportionation

R— (— CHT CHX — CHT CHX • + • XHC— CHT (— XHC— CHHn—R

R— (— CHT CHX—)n— CHT CH2X + XHC:= CH— (— XHC— CHT-)n— R

By transfer

R— (— CHT CHX--)n— CHT CHX + R'H

-111' R— (— CH2— CHX —)n — CHT CH 2X + R''

Scheme 1.7. Free radical polymerization of vinyl monomers. 12

1.3.3 Transition metal catalysts

In the last two decades, there has been as surge of interest in the chemistry and application

of catalysis by certain transition metal complexes. Polymerization catalysts have been the

focus and examples of industrial application abound. The oligomerization of alkenes by

Page 34: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

H

LnIvi—R' LnM—C—C—R

21

transition metal catalysts is however of considerable academic and industrial interest for the

synthesis of linear and branched higher mono-olefins. The oligomerization of ethylene and

propylene are the most well known, e.g. in the formation of higher linear alpha olefins in the

Shell Higher Olefin Process (SHOP). 29,30

A number of catalysts are known for the oligomerization of higher alpha olefins. Catalysts

similar to those used for ethylene and propylene oligomerization and polymerization can be

used for higher alpha olefin oligomerization. It is assumed that the polymerization

mechanisms of these catalysts can be applied to the corresponding oligomerizations of

higher linear alpha olefins. Among the most popular catalysts are titanium, zirconium and

nickel complexes; homogeneous and heterogeneous Lewis and Bronsted acids; and

inorganic oxides. 5 The reactivity of the olefinic substrates can be correlated to their relative

coordination abilities. Olefin reactivity decreases in the following order : ethylene >>

propylene > 1-butene > 1-pentene > 1-hexene. 31,32

In general, two different reaction mechanisms are considered relevant: 29,30 a hydride

mechanism and a mechanism based on metallacyclic intermediates. Keim 29 states that the

`hydride insertion-alkyl migration mechanism is the most generally accepted one, but proof

is still questionable. It is, however, the accepted one for the majority of oligomerization

reactions. Considerable attention has also been attracted by the catalytic sequence involving

metallacyclopentane intermediates. A simplified scheme for the metal-hydride mechanism

is portrayed in Scheme 1.8.

Kp

propagation n C=C—R

Scheme 1.8 Simplified hydride mechanism.29,3°

Page 35: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

22

The first step comprises the insertion of an olefin into a metal-hydrogen bond. Further

insertion of olefins into the metal-alkyl bond leads to growth products (propagation step).

The elimination of the thus formed oligomers occurs by a n-hydrogen elimination returning

to Li,M-H which re-enters in the catalytic cycle.

Analyzing the sequence given in Scheme 1.8, the relative propagation and elimination rates

again determine whether polymerization, oligomerization or dimerization will be observed.

Which of these reactions prevail is dependent on the nature of the metal and its oxidation

state, the nature of the ligand and the reaction parameters. These parameters critically

influence the electronic structure of the complex.

According to Keim 29 the following generalizations can be made: Group VIII transition

metals favor 13-hydrogen transfer (elimination) . and thus olefin dimerization and

oligomerization are observed preferentially. In contrast, with Group IV to VI transition

metals propagation prevails and predominately polymerization catalysts are found in this

class. With the same metal, chain propagation decreases with increasing oxidation state.

Furthermore, the metal character can be altered by adding ligands. Donor ligands tend to

increase and acceptor ligands to decrease the chain propagation.

There should be distinguished between the two possible modes of addition of a metal-

carbon or metal-hydrogen bond to an unsymmetrically substituted olefin, namely the

Markovnikov and anti-Markovnikov addition. The direction of addition determines

whether linear or branched products are formed. Higher alpha olefins can form a greater

number of isomers, as a result the selectivity of oligomerization is lower. Highly selective

formation of certain isomers can be achieved by the selection of appropriate ligands.

Homogeneous, transition metal based catalysis has received attention over the past number

of years and a variety of industrial processes eXiSt. 33,34,35 It is broadly assumed that the true

catalysts consist of a metal surrounded by a ligand field. The ligands are essential in

controlling the activity, selectivity and catalyst life. 36 The theoretical interest in

homogeneous catalysts, as opposed to heterogeneous catalysts, stems from the unique

insight into the mechanism of catalytic action. The nature and surfaces of solids are not well

known or understood, but the nature of molecular and ionic species existing in solution has

been relatively well established.

Page 36: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

23

For catalyst preparation one of three routes are generally followed. Transition metal salts in

a higher oxidation state are reduced to the catalyst, low valent transition metal salts are

oxidized to the catalyst or organometallic complexes are used as catalyst precursors. 36

Organometallic complex catalyst precursors are favored, because they tend to be highly

active, specific and selective thus minimizing unwanted reactions. 37

i) Heterogeneous Ziegler-Nana catalysts

The Ziegler-Natta catalysts received their initial attention when Ziegler showed that some

transition metal halides, upon reaction with aluminium alkyls, can initiate the

polymerization of ethylene. 38 Simultaneously, Natta demonstrated that similar catalysts can

polymerize various other olefins like propylene, butylene and higher alpha olefins. 39 Higher

molecular weight linear polymers formed as well, and more importantly, stereospecific

ones. These reactions require much lower pressures than do free-radical polymerizations

and form the foundation of today's polyalkene industry. Despite the consequent

development of a wide range of highly reactive and selective catalysts, the complex

multisited nature of heterogeneous Ziegler-Natta catalysts has hindered their detailed

characterization and rational modification. As a result, fundamental questions concerning

the mechanism of monomer addition and the origin of stereochemical control in

heterogeneous Ziegler-Natta catalysts remain unanswered.

0

Ziegler-Natta catalysts are products from reactions of metal alkyls or hydrides of Group I to

III of the Periodic Table with metals, salts or complexes of Group IV to VIII. The catalysts

that originally gained commercial importance were combinations of TiC14 or TiC13 with R 3A1

or R2A1C1. These so-called first generation heterogeneous Ziegler-Natta catalysts exhibited

rather low activity, water-sensitivity and were susceptible to catalyst poisoning. High-

activity-high-yield (second generation) catalysts were developed incorporating trivalent

alkoxy chlorides of transition metals with certain halogen-free organoaluminium

compounds. The third generation comprises supported Ziegler-Natta catalysts prepared by

reacting titanium compounds with hydroxylated magnesium derivatives, magnesium

alkoxides, or magnesium organometallics followed by activation with aluminium alkyls.11,12

Page 37: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

24

The Ziegler-Natta catalysts can be subdivided into two groups: (1) heterogeneous insoluble

catalysts, and (2) homogeneous or soluble, ones. Soon after the initial discoveries of Ziegler

and Natta, efforts were made to devise homogeneous models of the heterogeneous catalysts

that would prove more amenable to study. In 1957 Nattao and Breslow 41 reported that

Cp2TiC12 could be activated for alkene polymerization by Et 3A1 or Et2A1C1.

Several mechanisms have been proposed to explain the action of heterogeneous Ziegler-

Natta catalysts. All agree that the polymerization takes place at localized active sites on the

catalyst surfaces by coordinated anionic mechanisms. The organometallic component is

believed to activate the site on the surface by alkylating the transition metal. Some

controversy still exists about the exact mechanism of catalytic action, whether it is

monometallic or bimetallic. The majority of opinion leans to the former. In addition, it is

well accepted that the monomer insertion into the polymer chain is between the transition

metal atom and the terminal carbon of the growing polymeric chain. 42

The role of the metal alkyl in forming the active catalyst is illustrated in Scheme 1.9. After

complexation with TiC14 (or in general, with the transition metal compound), the transition-

metal alkyl is generated through halogen-alkyl exchange. This complex is then reduced in a

subsequent intricate reduction step including dealkylation to yield low-valence titanium.

These reduced metal species are believed to constitute the active catalytic species. 11

TiC14 + AlEt3 TiC14.AlEt3

TiC14.AlEt3 EtTiC13.AlEt2C1

2 EtTiC13.A1Et2C1 2 TiC13.A1Et2C1 C2H4 C2H6

Scheme 1.9 Formation of active heterogeneous Ziegler-Natta catalyst. 11

All evidence to date indicates that chain growth occurs through repeated four-centred

insertion reactions of the monomer into the transition metal-carbon cy-bond. It is still not

established whether the basic metal alkyls serve only to produce the active centres and have

no additional function. This would imply that the mechanism is monometallic. If the active

centres must be stabilized by coordination with base metals, then the mechanism would be

bimetallic.

Page 38: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

25

An example of the monometallic mechanism is one proposed by Cossee and Arlman

(Scheme 1.10). 43 This mechanism assumes that the reaction occurs at a transition metal ion

on the surface layer of the metal trichloride (or perhaps dichloride) lattice. Here, the halide

is replaced by an alkyl (R) group. The adjacent chloride site is vacant and accommodates

the incoming monomer molecule. The newly formed transition metal-alkyl bond becomes

the active centre and a new vacant site form in place of the previous transition metal alkyl

bond.

R' a ss

cl al I

R R' Cl 1 I ...-• CH—R

+ CH=CH —110- Cl—Ti--H 2 al 1 CH2 CI

CH—R

R'

a '

,CH, -

al I Cl

four-centre intermediate

R'

CH

/CH— R /

al2

Cl

Scheme 1.10 Monometallic Ziegler-Natta mechanism. 11,12

Terminations in coordination polymers usually occur by chain transfer either by an internal

hydride transfer or by transfer to a monomer. Hydrogen or protonic acids also terminates

the reaction and is often used to control the molecular weights of the products. 12

Ziegler catalysts yield outstanding wide-temperature range hydrocarbon fluids when used

in the oligomerization of alpha olefins. Compared to similar cationic catalysed reaction

products, the Ziegler derived materials exhibit higher oxidative and thermal stability as well

as a slightly higher viscosity index. The pour point is lower due to the presence of highly

linear chain segments in the oligomer. The high quality of these Ziegler prepared fluids can

be ascribed to the little or no isomerization that takes place during oligomerization. The BF3

cationic oligomerization still holds some advantages: while BF3 uses none, the Ziegler

process uses an unspecified amount of solvent; the Ziegler oligomerization takes much

longer; and the conversion for the BF3 process as well as the yield of the trimer are higher.17

Page 39: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

26

ii) Homogeneous Ziegler-Natta catalysts: Metallocenes

Forty years after the discoveries of Ziegler and Natta we are witnessing the evolution of new

generations of catalysts and polyolefin materials which originate from studies on

homogeneous, metallocene-based polymerization catalysts. Research on

metallocene-catalysed olefin polymerization derived much of its impetus from the desire to

model the reaction mechanisms of the -heterogeneous polymerization catalysts.

The first soluble catalysts (Cp2TiC12 activated by Et3A1 or Et2A1C1) exhibited lower activities

in ethylene polymerization than their heterogeneous counterparts and were inactive for

propylene. 40,41 This situation changed dramatically in the early 1980's when Sinn and

Kaminsky44• 5 discovered that partially hydrolyzed Me3A1 activated Group 4 metallocenes

catalysed the polymerization of ethylene and other alpha olefins.

Active homogeneous olefin polymerization catalysts are obtained on mixing bent

metallocenes Cp'2MX2 (Cp' = general cyclopentadienyl ligand; M = Zr, Hf, Ti; X = Cl, Me),

with excess methylaluminoxane (MAO) in inert solvents such as toluene (Figure 1.4). Such

uniform site catalysts generally contain a single kind of active site and offer potential

advantages over traditional multi-site Ziegler-Natta polymerization catalysts. The most

important is the ability to control polymer structure and properties by variation of the

catalyst structure, allowing production of a wide range of new polymers, unavailable using

the classic catalysts. 46

M x

where M = Hf, Zr, or Ti

and X= Me or Cl

Figure 1.4 Bent metallocene.

Page 40: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

27

a) Nature and formation of the active species

Initial metallocene research was focused on establishing the catalytically active species of the

metallocene/MAO system. Both a cationic monometallic species and a neutral bimetallic

species have been proposed. In 1960, Shilov 47 proposed that the active species was a 14-

electron cationic metal alkyl of the form [Cp2MR] 4- .• The isolation of the cationic insertion

product from the reaction of Cp2TiC12 -/MeA1C12 with an alkyne by Eich48 supported Shilov's

proposal. This result suggested that the metal alkyl cation [Cp2TiMe]+ was the initially

formed catalytically active species for this alkene polymerization system. In addition, the

catalytic nature of discrete do metallocenes (n = 0, 1) substantiated the existence of

a cationic active site. Strong evidence for the involvement of cationic intermediates came in

1986 when Jordan49 isolated and characterized the cationic complex

[Cp2ZrMe(THF)]+[BPh4] , which polymerized ethylene in the absence of a cocatalyst.

Controlled hydrolysis of Me3A1 produces a mixture of methylaluminoxanes (MAO) thought

to have the general formula [Al(Me)O]n (5 < n , 20). 50 MAO plays at least three roles in

polymerization: as a alkylating reagent for the generation of transition metal-alkyl adducts;

as a Lewis acid for anion abstraction from the complex generating an electrophilic species;

and as a scavenger for the removal of impurities, particularly water in the olefin and

solvent.46

Highly active alkene polymerization catalysts are formed when a large excess of MAO is

mixed with Group 4 metallocenes (Scheme 1.11). It is thought that the excess of MAO is

required to shift an equilibrium involving a metal alkyl (20) and MAO to the active cationic

species (21). 51 Initially a fast ligand exchange takes place between Cp2MC12 and MAO to

generate primarily the monomethyl complex Cp2M(Me)Cl; excess MAO leads to

Cp2M(Me)2.52

---(- Cp2MC12 + jid—C) --IP- Cp2M(Me)C1 + MAO' [Cp2MMer[(MA0 1)Clr CH3 n

20 21.

Scheme 1.11 Formation of active metallocene/MAO species.

Page 41: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

28

It is generally assumed that some of the Al centres in MAO have an exceptionally high

propensity to abstract a CH3 ion from Cp2M(Me)2 and to sequester it in a weakly

coordinating ion CHs—MAO-. High catalyst activity suggests that this MAO derived anion

is labile and easily displaced by the monomer prior to insertion. This weak association gives

way in the presence of substituted olefins, to olefin-separated ion pairs

[Cp2MR(olefin)]+CH3—MAO-, the presumed prerequisite for olefin insertion into the M-R

bond (where R is the growing polymeric chain). This hypothesis - that the unusually low

coordinating capability of the anion AT in [Cp2M(Me)]+A - is crucial for catalytic activity 53 -

led to the discovery of a series of highly active cationic metallocene catalysts for the

polymerization of propylene and higher alpha olefins.

The key to the synthesis and activity of such active cationic metallocenes is the utilization of

robust and poorly coordinating anions showing good solubility in inert aromatic solvents.

Early attempts to use the traditional 'non-coordinating anions', such as BF3 and PF6, were

unsuccessful as fluorine transfer to the cation led to deactivation. 54 Even for large, weakly

coordinating anions such as (C6H5)413 - and C2B9Hu- fairly strong interactions have been

observed with cationic (alkyl)zirconocene species and as a result propylene was

polymerized only at low rates, if at all. 55'56'57 A breakthrough in this regard was the

introduction of perfluorinated tetraphenylborate as counterion by Hlatsky and Turner 58 and

by Marks et al.56 These was the first well-defined metallocene catalysts capable of

polymerizing propylene and higher olefins at high rates without addition of further

activators. Similar activities for propylene polymerization were subsequently observed with

other base-free or weakly stabilized (alkyl)metallocene cations. Tetraphenyl borates, alkyl

triphenylborates and, to a lesser extent, carborane anions have found wide use in cationic

metallocene chemistry. 54,55,56,59,60

The formation of these highly active, base-free cationic metallocene alkyls is represented in

Scheme 1.12. Usually an ion pair such as [Cp'2ZrMe]f(C6F5)13 - (where Cp' is some

substituted Cp or indenyl ligand) is formed in which residual coordinative contacts between

the cationic Zr centre and its counterion exists. This contact appears to resemble those yet

unidentified interactions that stabilize an (alkyl)metallocene cation in contact with a

CH3—MAO- counterion. In both cases, a fast ligand exchange occurs between the cationic

and anionic complex moieties; most importantly, bo. th types of contact appear to be weak

Page 42: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

HIGH ACTIVITY

Me /

Zr (C6F5)3

` Me

ZrMe2 B(C6F5)3

29

enough to allow an alpha olefin to displace the anion from its coordination site at the metal

centre. 59,61

ZrMe2 + [Bu3NFIRB(C6F5)4]

- Bu3N

Zr+ - B(C6F5)4

`Me

ZrMe2 + [Ph3C][B(C6F5)4]

MODERATE ACTIVITY

Scheme 1.12 Formation of active MAO free catalysts. 46

Of even greater simplicity are catalysts based on an (alkyl)metallocene complex containing a

Scm, Yifi or a trivalent lanthanide centre. Both Ballard 62 and Watson63 found that neutral

complexes of the type Cp2MEHR acted as single component catalysts for the oligomerization

of alpha olefins. 64,65 These catalysts are however, more difficult to prepare than the Group 4

metallocene systems.

b) Kinetics and mechanism of alpha olefin polymerization

The availability of MAO-free polymerization catalysts has allowed the study of the olefin

insertion and termination (chain transfer) steps in the catalytic cycle. Important aspects of

Page 43: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

30

alpha olefin insertion are the regioselectivity (1,2-insertion is normally found for simple

olefins), the stereospecificity and the mode of stabilization of the insertion product

(presumed to be via weak secondary bonding interactions). In Scheme 1.13 the basic steps

believed to be involved in polymerization using metallocene-based catalysts are shown.

n R

Scheme 1.13. Basic steps in polymerization of alpha olefins using metallocene-based

catalysts.66

The reactive complex, [Cp2MR(olefin)]+, is firstly formed by the displacement of an anion

CH3—MAO- from its [Cp2MR]+ counterion by an olefin molecule. Coordination of the

alkene followed by migratory insertion generates a coordinatively unsaturated species (22)

that can repeat this propagation sequence. For alpha olefins the insertion step is typically

highly regioselective, with greater than 98% of the monomer units undergoing primarily

1,2-cis insertion. 67

Kinetic isotope studies by Bercaw 68 and Brintzinger, 69 as well as molecular modeling"

suggested that the insertion process can be accompanied by an agostic interaction between

the metal centre and one of the a-hydrogen atoms of the growing polymer chain. According

to Morokuma et al.71 this transition state for alkene insertion into [M]—R+ exhibits an

interaction of an a-H atom of the growing chain with the electron-poor metal, such that the

electron pair in the a-C-H a-bond is partly donated to the metal; this is the so-called

Page 44: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

31

a-agostic interaction (Scheme 1.14). The stabilization of the alkyl product of insertion by an

p-agostic interaction (of a (3-C-H bond with the metal) is also suggested by Morokuma and

coworkers, although others support either a y-agostic interaction" or no agostic interaction

at all.72

+ M —CH-2 R'

CH2 2

M CHR

CH2

1 + ,,CH

2 R'

/CSR

CH2

,,H

' M•

„CHR \

H2C

transition state

p-agostic interaction a-agostic interaction

Scheme 1.14 Illustration of agostic interactions."

Several chain transfer processes have been observed in these homogeneous systems

(Scheme 1.13). In the absence of a chain transfer agent, intramolecular (3-hydrogen

elimination produces a vinylidene-terminated polymer (23). Intramolecular p-alkyl

elimination forms an allyl terminated polymer (24). 13-Alkyl elimination is commonly

observed with catalysts possessing pentamethylcyclopentadienyl (Cp*) ligands, or when the

resulting polymer chain has no [3-hydrogens and no chain transfer agents are present.

Hydrogen is an effective chain transfer agent in these systems, which produces polymers

with saturated end groups (25). Alkyl aluminium compounds are also chain transfer

agents, where transmetallation produces an aluminium-terminated polymer (26).66

Page 45: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

32

c) Stereoregular polymers from chiral metallocene catalysts

An advantage of well-defined homogeneous catalysts over their heterogeneous equivalents

is that the ligands of soluble metal complexes can be systematically modified to study the

effect of catalyst structure on the stereoselectivity of the reaction. The symmetry of the

metallocene, as well as the nature of the cocatalyst and the polymerization temperature

determine the polymer stereostructure.

Using ethylene bridged ligands, Wild and Brintzinger 73 obtained ethylenebis(indenyl)- and

ethylenebis(tetrahydroindenyl) titanium complexes, rac-(en)(ind)2TiC12 (27) and rac-

(en)(thind)2TiC12 (28), and their zirconium analogs. The conformationally constrained

indenyl and tetrahydroindenyl give these complexes chiral structures which could be

expected to be retained even under catalysis conditions. When activated with MAO, these

ansa-metallocenes were indeed found by Ewen 67 and also Kaminsky and Ktilper, 74 to

polymerize propylene and other alpha olefins to give highly isotactic polymers.

27

28

A range of polypropylene (PP) stereostructures can be formed by using chiral

metallocene/ MAO systems. In addition to atactic PP (random orientation of methyl

groups), isotactic PP (methyl groups all lie on one side of the outstretched polymer chain)

and syndiotactic PP (alternating methyl groups), other stereoforms of PP have been

attained. In hemiisotactic PP, every second methyl group is orientated in the same

direction, whereas the methyl groups in between adopt a random orientation. Stereoblock

isotactic PP contains isotactic blocks of alternating orientation. Another form of stereoblock

PP contains isotactic blocks alternating with atactic blocks. These stereostructures and the

pro-catalysts used are illustrated in Figure 1.5.46

Page 46: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

(C2v ) F. F

Atactic PP

33

O O

(C2)

Syndiotactic PP

Heirdisotactic PP

Isotactic PP

(:).)1\

TiPh2 (C2v)

Stereoblock isotactic PP

(Ci)

Stereoblock isotactic-atactic PP

Figure 1.5 Relationship of pro-catalyst symmetry to polypropylene stereostructure.46

Page 47: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

)

CH3

,\ CH3

R 01 Me

M Me • H I

CH3 Isotactic polypropylene

,\ CH3

favored

H

C

R

C

34

Two mechanisms regulate stereospecific poly(alpha olefin) formation, depending on the

catalyst symmetry: enantiomorphic-site and chain-end stereochemical control. Efficient

control of the chiral catalyst over the enantiofacial orientation of the entering olefin is

contingent on the presence of a metal-alkyl segment with at least two C atoms. This

segment appears to act as a sort of lever in transmitting the effects of the P-substituents on

the orientation of the prochiral alpha olefin. Corradini et al.75 proposed that repulsive

interactions force an olefin into that .enantiofacial approach to the metal-alkyl unit which

places the olefin substituent trans to the p-C atom of the metal-bound alkyl chain. The

metal-alkyl chain, in turn, was suggested to favor an orientation in which its C(a) - C((3)

segment is in the most open sector of the chiral ansa-metallocene ligand framework (Scheme

1.15).

14 valence electron active species

ri indicates vacant coordination site

Scheme 1.15. An example of stereocontrol in metallocene/MAO polymerization.76

Page 48: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

35

Leclerc and Brintzinger77 rationalized the stereocontrol in terms of a-agostic interactions. Of

the two alternative Zr-C(a)-C((3) orientations resulting from binding one of the a-H atoms to

the metal centre, one orientation is sterically inaccessible since it would cause the growing

chain to collide with a 13-substituent of the chiral ligand framework; the other one is

sterically unencumbered. The a-agostic interactions control the stability of alternative

transition states for inserting an olefin into a metal-alkyl bond. The a-agostic model thus

describes the re or si orientation of an inserting olefin as being controlled by the required

placement of the p-c atom of the growing chain in the more open of the two possible

positions at a rigid three-membered Zr-H(a)-C(a) ring. 78

d) Reaction control and product distribution

An important aspect in which homogeneous olefin polymerization by metallocene catalysts

differs from heterogeneous Ziegler-Natta catalysts is the narrow molecular weight

distribution of homogeneously produced polymers. Molecular weight is strongly

influenced by the metallocene catalyst and reaction conditions employed.

A number of generalizations can be made concerning metallocene/ MAO catalysts. Catalyst

activities increase with increase in the Al : Metal ratio. As mentioned previously, large

excesses of MAO (Al : Metal > 500) is usually required for acceptable activity. A decrease in

the Al : Metal ratio and an increase in temperature result in lower average molecular weight

and (for stereospecific catalysts) decreased stereoregularity. Furthermore, the order of

activity is generally Zr > Hf > Ti. In addition alternative, potentially cheaper aluminoxanes,

such as (-EtA10-),, or (-i-BuA10-)r, and other alkylaluminium cocatalysts (e.g. the mixture of

A1Me3 and A1Me2F) show inferior activity."

In short, by raising the reaction temperature, lowering the alkene concentration and

decreasing the Al : Metal ratio, oligomers may be synthesized.

Page 49: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

36

e) Future developments

The most remarkable aspect of metallocene-based catalysts appears to be that they can

produce an unprecedented variety of polyolefins. The range of stereo- and regioregularities,

molecular weights, copolymers and cyclopolymers available with these catalysts exceeds

that accessible with classical heterogeneous Ziegler-Natta catalysts and opens new

possibilities for producing tailored poryolefin materials. 78

The extent to which metallocene catalysts will replace traditional catalysts will depend

greatly on the relative costs of making the products. The greatest challenges involved in

utilizing this new catalyst technology in an economically feasible process are reduction of

the high cost of the cocatalyst, control of polymer morphology and the adaption of the

catalysts for use in existing polymerization plants. 79 The lack of morphological control and

reactor fouling can be addressed in practical application of metallocene catalysts by their

pre-adsorption on solid supports such as alumina or silica gels. Such heterogenized

catalysts can then readily be used in existing Ziegler-Natta plants, for instance in

solvent-free slurry or gas-phase reaction systems. Supported catalysts may permit the use

of less MAO whilst still offering facile control of polymer properties by ligand variation. 46,79

iii) Other important transition metal catalysts

a) Supported reduced transition metal oxide catalysts

Besides Ziegler-Natta catalysts, one-component, that is, metal-alkyl-free transition metal

catalyst systems are also highly active in alkene polymerization. The most important of

these are the oxide-supported catalysts such as the original Phillips chromium oxide on

silica or silica-alumina. 80

Such catalysts are prepared from Ti02, V205, Nb205, Ta2O5, Cr03, Cr203, Mo03, W03, NiO

and CoO. Supports include charcoal, silica, alumina, zinc oxide or alumina-silica. An

important characteristic of these catalysts is their specificity. The Phillips catalyst, for

instance, although highly efficient for ethylene polymerization, is ineffective in the

Page 50: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

37

polymerization of propylene. Moreover, higher temperatures and pressures are required

for catalyst activity than do the typical Ziegler-Natta catalysts. 11

Diverse chromium-based catalysts are now used in the Phillips polymerization process.

Broadly, there are two families of supported chromium catalysts: organochromium

compounds, and catalysts based on chromium oxide. Suitable organochromium

compounds include the 7c-bonded "sandwich" compounds, bis(arene)-chromium(0) and the

chromocene derivatives, allyl-chromium(II) and - chromium(III), and various alkyls of Cr(II)

and Cr(IV) such as the (trimethylsilyl)methyl or neopentyl. Compounds varying in valence

from 0 to 4 have all produced good activity once attached to a carrier. The two types of

supported chromium catalysts that are commonly employed are depicted in Figure 1.6. 81

0 0 .• Cr

* Cr —ligand I ■ ol

0 0

(a) (b)

Figure 1.6. Two types of supported chromium catalysts used for ethylene

polymerization. In one type (a), Cr is converted to the hexavalent oxide; the

other (b) employs a reactive organochromium compound. 81

b) Homogeneous nickel catalysts

The use of complex nickel catalysts for the homogeneous dimerization and oligomerization

of ethylene has by far attracted the greatest interest among the Group VIII metals.

Unfortunately only a few cases are known in which well-defined complexes have been

applied.

Tsutsui reported that dicyclopentadienylnickel was a homogeneous catalyst for the

dimerization of ethylene. 82 Barnett83 discovered that [Ni(1e-05H7)Cp] was a highly selective,

homogeneous catalyst for the conversion of ethylene to 1-butene and n-hexenes.

Page 51: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

38

Keim84 found that complexes of type (29) were very active catalysts for the oligomerization

of ethylene. Additionally, complex 30 converted ethylene to alpha olefins possessing 99%

linearity and >95% alpha olefin content. The geometric distribution of olefins could be

tailored by pressure and addition of Ph3P. By using phosphines other than Ph3P the catalyst

could be altered towards the synthesis of linear polyethylene. 85 Complexes similar to (30)

may be formed on reacting bis(1,5-cyclooctadiene)nickel(0) with Ph2PCH2CO2H, thiolactic

acid, o-mercaptobenzoic acid or phosphorous ylides. All these systems have been patented

by Shell and they are effective in the SHOP reaction.2 9'86

H Ph

Ph ?= gNi; Ph .7 .7 Ph \3113

29 30

R1 :..._. R2 = CF3

R1 = CF3; R 2 = CH3

1.4 Conclusion

Catalytic oligomerization 5 of olefins has attracted a lot of interest because oligomers are

useful intermediates for speciality chemicals. Whereas catalytic systems for ethylene

oligomerization are well established, processes for the oligomerization of alpha olefins by

organometallic catalysts were rare before the discovery of metallocenes. Here, metallocene

catalysis is promising, because easily accessible metallocenes like Cp2ZrC12 or alkyl-

substituted derivatives can be used as very effective catalysts for oligomerization with high

regioselectivity and also high specificity in the chain-termination reaction. 87,88

However, there are limited examples of the oligomerization of higher alpha olefins with

metallocene catalysts. Furthermore, at this stage, a detailed investigation into the general

nature of the oligomerization of higher alpha olefins with these catalysts is lacking. In the

light of the mentioned advantages, a systematic study of these reactions is certainly

warranted in order to broaden the knowledge base in this field.

Page 52: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

CHAPTER 2

DISCUSSION

2.1 Aim of research

Linear alpha olefins are versatile and reactive intermediates for the petrochemical industry.

Alpha olefin derivatives are used in surfactants, synthetic lubricants, plastics, seals in space

shuttles and for the extraction of crude oil form the ground, to name but a few. The use of

these olefins as comonomers in the polymerization of ethylene and as feedstocks for the

production of polyalphaolefins are showing the highest growth.

The oligomerization of ethylene is the primary source of linear alpha olefins today.

However, this process restricts the manufacturer to a relatively inflexible, even-numbered

distribution of alpha olefins. The Fischer-Tropsch synthesis may be tailored to produce

liquid fuel with a larger light alkene content. 89,90 The favorable surface composition of the

catalyst suppresses secondary transformations (alkene hydrogenation and isomerization),

thus ensuring selective alpha olefin formation. 91 These alpha olefins can then be recovered

from the raw synthetic fuel streams.

Sasol isolates alpha olefins from coal-derived fuel (with a value comparable to ethylene) and

can target only those alpha olefins that are the most profitable. In addition odd-numbered

alpha olefins, like 1-pentene, can be produced. These odd-numbered alpha olefins make a

host of new products accessible to the chemical industry. Hitherto, the plastics industry

could not shift from the C4 to C5 alpha olefins in polymerization or use 1-pentene as a

comonomer in the production of polyethylene due to insufficient access to 1-pentene.

Furthermore, 1-pentene is a new raw material that can be funtionalized for use in variety of

applications.

The aim of this project was to study the oligomerization of higher alpha olefins and,

specifically that of 1-pentene. A dual approach was followed, namely (i) the synthesis of a

model 1-pentene oligomer and (ii) an evaluation of the different oligomerization catalysts.

Page 53: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

40

2.2 Synthesis of model oligomers

As mentioned previously, a large number of isomers form during the cationic

oligomerization of alpha olefins, especially if BF3/ROH is used. In order to investigate and

rationally modify this and other closely related oligomerization processes, it was necessary

to obtain information on the structures and the degree of branching of the different isomers

present in the product mixture. Unfortunately, there is no unambiguous method to predict

the structures of the major isomers in the product mixture, or to ensure that peaks

appearing in the gas chromatogram are single isomers rather than mixtures of closely

related isomers. Furthermore, there is only a limited number, if any, of the corresponding

alkane standards commercially available. Since it is impractical to synthesize all possible

isomers, some model is usually applied to predict the major isomers. The best that can be

done at present, is to determine the average "hypothetical" structure of the product. It

should be kept in mind that the properties of such a compound might not correspond to

those of the mixture of isomers. 18 One of the aims of this project was to synthesize such a

model 1-pentene oligomer. A literature search revealed that the synthesis of these model

oligomers and related compounds have almost never been attempted.

Guppies and coworkers 22 prepared tetrasubstituted methanes having three approximately

equal alkyl branches, and a methyl branch in order to determine the effect of branching on

viscosity and thermal stability. The synthesis of 11-methyl-11-octylheneicosane was

attempted and it was found that the direct route via the addition of an organometallic

reagent to a tertiary alkyl halide did not proceed in reasonable yield. A general, indirect

route to tetraalkylmethanes in high yield was established (Scheme 2.1). This procedure

involved the condensation of ethyl cyanoacetate with a ketone to afford ethyl

alkylidenecyanoacetate, followed by the 1,4-addition of a Grignard reagent to give

a-cyano-B,[3,13-trialkylproprionate. Alkaline hydrolysis and decarboxylation yields

1343,13-trialkylproprionitrile. The addition of another Grignard reagent affords a ketone and

finally the reduction of this ketone to the alcohol, dehydration and hydrogenation produce

an alkane.

Page 54: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

41

-

Et0OCCI1CN +R1_8_R2 CN

C= (._ R2/ 'COOEt

R3MgX I-1,30+

R3

Ri_LCH2CN

R1

2

R4MgX H30+

a) Alkaline hydrolysis

4111,) Cu, 150 °C, (-04)

R3 CN

I 'Coot

R3 0 R3 R3 KHSO4

L

R1 -0 _(::H2cH(014)R4 R1-—CH=CHR4 + isomers

R1

2 R1

2 L2

Ni, H2

R3 -- C1--120-12 R4

R2

Scheme 2.1. General route to tetraalkylmethanes. 22

2.2.1 Synthetic target and strategy

The synthesis of the following range of hypothetical 1-pentene oligorners was attempted.

31

32

33

Dimer

Trimer Tetramer

Page 55: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

42

It can be seen that the trimer (32) will correspond to the previously mentioned ideal

structure for a lubricant molecule. 16 This states that in 32 (Scheme 2.1) R and R2 should be

nearly identical in size. R 1 is generally a hydrogen and R3 can vary from a methyl group to

larger than R and R2. Due to the low carbon number of this trimer, C15, it can be foreseen

that the physical properties of this model will not be ideal for application as an automotive

crankcase lubricant although it might be useful in other applications.

R1

R2

R3

32

R1 = H

R = C5

R2 = C5—Me

R3 = C3 ('\/)

Figure 2.1. 1-Pentene model trimer.

The study of this model oligomer is, however, valuable in order to understand and analise

higher alpha olefin oligomerization reactions. The general retrosynthetic strategy to attain

the synthetic targets is shown in Scheme 2.2. 1-Bromopentane (34), 1,2 -epoxypentane (35)

and 2-bromopentane (36) are employed as starting materials in this proposed synthesis.

2.2.2 Synthesis of 1,2-epoxypentane

Oxiranes are common intermediates in organic synthesis. The facile nucleophilic opening of

oxiranes provides a convenient route for forming C-C es-bonds and in most cases it is

possible to anticipate the stereochemistry and regiochemistry of the ring-opening. Oxiranes

play an important role as 1,2-functionalized synthons.92

Page 56: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

29

43

28

C-C bond formation

III

O

X = leaving group e.g. halide, OTs, 0

%

FG transformation

34

FG transformation

O

III

FG transformation Metal

FG transformation

OH

C-C bond formation

0

C-C bond formation

OH

O

OH OH

0

Metal

3.5

FG transformation

B r

36

Scheme 2.2. Retrosynthetic strategy.

i) Direct epoxidation

The direct epoxidation of an olefinic double bond is a simple route to oxiranes. This reaction

is achieved most conveniently by employing ni-chloroperbenzoic acid (MCPBA). Pure

MCPBA is, however, both shock-sensitive and potentially explosive, and consequently strict

Page 57: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

44

regulations govern transport and supply. Although it would be the oxidation agent of

choice, the demise of MCPBA necessitated the use of safer alternatives in the large-scale

epoxidation of 1-pentene.

a) Magnesium monoperphthalate

1.0

Magnesium monoperphthalate (MMPP) is by comparison a non-shock-sensitive and non-

deflagrating peroxygen product and is thus much safer to use. It can be used to carry out a

variety of oxidation reactions. The epoxidations of cylcohexene and 2-methylbut-2-ene in 85

and 98% yield, respectively, are illustrative. 93,94

Attempted oxidation of 1-pentene with a stoichiometric amount of MMPP in aqueous

i-PrOH afforded 1,2-epoxypentane in low yield (estimated by GC). A variety of conditions

were employed, including different reaction temperatures (0 °C, 25 °C, 40 °C) and extended

reaction times, but the yield could not be improved beyond 30% (Scheme 2.3).

MMPP, i-PrOH/ H 20

C3' 35 37

bp 30 °C

40 °C, 16 h, 30 %

Scheme 2.3. MMPP epoxidation of 1-pentene.

This can presumably be attributed to the low boiling point of 1-pentene (bp 30 °C). Most

MMPP epoxidations are carried out at temperatures higher than 30 °C. 93,94 This would

imply that, depending on the reactivity of the alkene substrate, most reactions require some

activation energy to proceed satisfactorily. The epoxidation of a monosubstituted alkene,

which is generally less reactive, requires the use of stronger oxidation agents and more

severe conditions. 95 Low yields can be due to the high volatility of 37 and consequent

losses, even at room temperature. 1-Pentene's low boiling point and accompanying high

volatility proved to be a challenge in all consequent reactions with this substrate.

Page 58: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

45

Urea hydrogen peroxide

The hydrogen bonded adduct, urea hydrogen peroxide (UHP), is formed when urea is

crystallized from aqueous hydrogen peroxide. 96

NH2 ---- HO—OH

38

UHP in combination with either acetic anhydride or trifluoroacetic anhydride can be used to

effect a wide range of peroxidations. Which anhydride is used, depends on the

nucleophilicity of the alkene: terminal alkenes give good yields using the trifluoroacetic

anhydride procedure. 94,97

UHP - Na2HPO4 0 „..-%

'.. (CF3CO)20 37 35

40 °C, 18 h, 25 %

by 30 °C

Scheme 2.4. UHP epoxidation of 1-pentene.

Attempted epoxidation of 37 by peroxytrifluoroacetic acid generated from the anhydride

and 38 in the presence of a phosphate buffer, proceeded in a low yield (Scheme 2.4). This

can again be attributed to the low boiling point of 37, since standard UHP epoxidations are

carried out in refluxing methylene chloride (bp 40 °C).

Dimethyldioxirane

Murray et al.98 have shown that solutions of dimethyldioxirane (39) in acetone (0.1M) can

be obtained through low temperature distillation of 2KHSO5.KHSO4.K2SO4 and acetone

reaction mixtures. A solution of 39 in dry acetone reacts readily with disubstituted alkenes

such as cis-3-hexene and trans-3-hexene to furnish the corresponding epoxides in nearly

quantitative yield. 99

39

Page 59: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

46

Murray's procedure for making and isolating 39 is not always practical for large-scale

synthesis. Edwards and Curci, 180 as well as Murray98 have reported a method for forming

an excess of 39 in situ under phase-transfer conditions. Both the improved in situ generated

and isolated dimethyldioxirane were used to epoxidise 1-pentene in moderate yield (Scheme

2.5).

o—o

X (2 equivalents)

acetone

RT, 2 h, 40 % 37 35

2KHSO5.KHSO4.K2SO4 (3.2 eq)

acetone n-Bu4NHSO4, CH 2C12

NaHPO4 (aq) buffer

pH– 8, 0 °C, 50 %

37

35

Scheme 2.5. Dimethyldioxirane epoxidation of 1-pentene.

The consequent isolation of 35 proved to be more problematic. Both atmospheric and

vacuum distillation was not successful in separating the epoxide from the solvent, acetone.

This epoxide 35 was found to be unstable in extended atmospheric distillation. Separation

by vacuum distillation at a lower temperature was unsuccessful because the boiling points

of the epoxide and acetone were in the same range.

ii) Indirect epoxidation

Although several other oxidation reagents are generally used for the epoxidation of

alkenes,92 it was decided to abandon the direct epoxidation route. The indirect route via a

halohydrin is a viable method for producing epoxides which formally uses the reaction with

a halogen to effect the oxidation.

1-Pentene was transformed into the iodohydrin (Scheme 2.6) or bromohydrin (Scheme 2.7)

derivatives through the in situ generation of the hypohalous acids (HOX) from H5106 or

NaBrO3 in the presence of a suitable reducing agent.101

Page 60: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

a) Iodohydrin

47

H5106

OH

I 37

NaHS03

0 °C to rt, 4 h, 85 % 40

Scheme 2.6. Formation of iodohydrin.

Since the reaction is exothermic, the reducing agent (NaHSO3 as a 1M solution) was added

dropwise to a mixture of H5106 and 37 in CH3CN/H20 at 0 °C in order to minimize losses of

1-pentene. After the addition was completed the reaction mixture was allowed to warm to

room temperature and to proceed for another 4 hours, after which work-up afforded a

yellow-brown liquid in 85% yield. The liquid was analysed by GC and GC/MS which

indicated the presence of two products in a 1 : 2 ratio.

The MS data of all organic compounds were determined by using a quadrupole ion trap

mass spectrometer as detector for a gas chromatograph. This spectrometer employs three

electrodes - two end-cap electrodes that are normally at ground potential and between them

a ring electrode to which a radio frequency (rf) voltage, often in the megaHertz range is

applied - to generate a quadrupole electric field. 102

The two products had identical MS data. The EI spectra showed a weak ion at m/z 213

which corresponds to hydride abstraction from 40. This phenomenon (the presence of a

M+ - 1 peak) was observed in the mass spectra of all'of the organic compounds. The loss of

water to yield C5H9I+ with m/z 196, of an iodide ion (m/z 127) to furnish an ion with m/z

87 and of both to form an ion with m/z 69 were observed. The following fragments were

also present: C4H7+ (m/z 55), CH31+ (m/z 142), C3F17+ (m/z 43) and C2H401+ (m/z 171). The

1H and 13C NMR spectra also indicated the presence of two compounds. The main

compound could be identified as the desired 1-iodo-2-pentanol. The 1 H-NMR spectrum of

the mixture showed the expected resonances of one methyl group (5 0.82, t, J 6.8 Hz), 4 alkyl

protons as two overlapping multiplets at 6 1.27 - 1.44, the hydroxylic proton as a broad

singlet (6 2.33) and the proton on the hydroxylic carbon as a 1-proton multiplet (6 3.40 -

3.43). The iodohydrin moiety was recognized by a characteristic AB system indicative of the

diastereotopic protons lb-H (8 3.12, jib,2 6.6 Hz, Jas 10.2 Hz) and la-H (5 3.36, Jia,2 3.7 Hz, JAB

Page 61: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

48

10.2 Hz). The 13C NMR of the mixture also confirmed the presence of 40 as the main

compound in the mixture. The resonances of the methyl carbon (6 13.20, C-5), two alkyl

carbons (6 17.83, C-4; 6 35.21, C-3), iodo-substituted carbon (6 20.09, C-1) and hydroxylic

carbon (5 68.11, C-2) were present.

OH

41

It can be argued that the second compound is a regioisomer (41) or conformational isomer of

40, but no firm evidence exists. Although the literature 101 states that the iodohydroxylation

of terminal alkenes occurs with high regioselectivity in a Markovnikov fashion to afford the

1-iodo-2-hydroxy derivative, we consider the alternative, i.e. the formation of

conformational isomers as unlikely.

b) Bromohydrin

OH NaBrO3, pH — 1

Br NaHSO3

35

0°C to rt, 4 h, 80 % 42

Scheme 2.7. Formation of bromohydrin.

A similar procedure to the iodohydrin formation was used to generate 42 in 80% yield.

NaBrO3 could be used in an acidic (pH = 1) medium in the presence of NaHSO3 to form the

hypobromous acid. The GC and GC/MS analyses indicated the presence of three products

with identical MS data in a 1 : 2 : 4 ratio. Both 1 H and 13C NMR also showed the presence of

additional products with closely related structures.

By following the same reasoning as before it can be postulated that these products are

regioisomers. Again this could not be proven without any doubt. The analytical data did

confirm that the major component in the mixture was 1-bromo-2-pentanol.

The molecular ion was absent in the EI-MS and CI-MS spectra of 42. The loss of water from

42 gave two prominent peaks at m/z 148 and m/z 150 corresponding to fragments

Page 62: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

49

containing the two bromine isotopes; the loss of a bromide ion from this ion (m/z 69)

furnished an intense peak. Other weak ion peaks could be attributed to the fragmentation

of 42. The 1H NMR spectrum indicated the presence of a methyl group (6 0.90, t, J 6 Hz)

and 4 alkyl protons as two overlapping multiplets (6 1.34 - 1.57). The hydroxylic proton was

represented by a broad singlet 6 2.20 (confirmed by D20 exchange) and the proton on the

hydroxylic carbon as a multiplet 6 3.74 - 3.76 (confirmed by proton decoupling). The

diastereotopic halide protons were again characterized by an AB spectrum: lb-H (6 3.38, J1b,2

7.1 Hz, JAB 10.3 Hz) and 1a-H (6 3.53, Jia,23.3 Hz, JAB 10.3 Hz). 13C NMR clearly identified 42

as the main product in the mixture. The data confirmed the presence of the methyl carbon

(6 13.80, C-5), two alkyl carbons (6 18.73 and 6 37.07, C-4 and C-3), bromo-substituted carbon

(6 40.52, C - 1) and the hydroxylic carbon (6 70.72, C - 2).

c) 1,2-E poxypentane

Epoxides can be formed from halohydrins by treatment with a base to generate the alkoxide.

The alkoxide should be anti to the adjacent halide in order to allow the substitution of the

halide to give the epoxide via a SN2 process (Scheme 2.8).

OH

„X R

B-

Scheme 2.8. Illustration o f the preparation of an epoxide from a halohydrin.

The epoxide (35) was prepared without any solvent by adding the halohydrin dropwise to a

large excess of finely powdered NaOH (Scheme 2.9). The reaction was exothermic and 35

could be distilled under vacuum as it formed; this method circumvented the problem of

separating the epoxide from a solvent. The epoxide isolated in this manner was found to be

pure (> 95%) by GC analysis. The boiling point of this compound was determined as

67 - 69 °C. The total yield for the synthesis of the epoxide from 1-pentene is approximately

65% which is higher than the yields obtained in the direct route.

Page 63: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

50

OH Base

40

35

OH

Br Base

42 35

0 80%

O 75%

Scheme 2.9. Preparation of 1,2-epoxypentane.

The EI-MS of 35 afforded a weak peak at m/z 85 which corresponds to hydride abstraction

from 35. Further fragments observed were the loss of H2O (m/z 68) and a CH3 fragment

(m/z 71). The 1H NMR of the epoxide in CDCb indicated the presence of two compounds

of which one could be the epoxide. The GC analysis of the product in CDC13 showed that a

new product formed: two peaks were observed instead of one. It was thought that the other

originated from the addition of residual DC1 (HC1) in the CD03 to the epoxide. Due to the

large extent of peak overlap the spectrum could not be assigned with certainty and the

NMR analyses were repeated in deuteurated pyridine in which only one product was

observed. The 1H NMR confirmed the presence of the methyl group (6 0.94, tt, J 7.5 and 0.9

Hz), two alkyl groups as a 4-proton multiplet (6 1.42 - 1.49) and the secondary epoxide

proton (6 2.80 - 2.84, m). The diastereotopic protons lb-H (6 2.38, Jlb,2 2.77 Hz, JAB 5.4 Hz)

and la-H (6 2.94, ,la,2 4.2 Hz, JAB 5.4 Hz) on the terminal carbon atom again yielded an AB

spectrum. The 13C NMR spectrum showed the characteristic resonances of C-5 (6 14.03), C-4

and C-3 (6 19.63 and 6 34.90) and the two epoxide carbons C-1 and C-2 (6 46.21 and 6 51.54).

2.2.3 Synthesis of the dimer analogue

The reaction of an epoxide with a carbon nucleophile generates a new carbon-carbon

6-bond, and is a versatile protocol for the preparation of a variety of alcohols provided that

the ring opening takes place regioselectively. Owing to the polarity and the strain of the

three-membered ring, epoxides undergo reaction with a number of organometallic

reagents.92

Page 64: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

51

A summary of organocopper reagent types is given in Table 2.1. 103 Addition reactions of

these species are sometimes curtailed owing to the competing reactions arising from the

Lewis acidity or basicity of the organometallics used. 104,105 The carbon nucleophiles of choice

are organocuprates, because Grignard reagents sometimes result in concomitant formation

of side products arising from the substitution by a halide ion.104,105,106 Higher order mixed

organocuprates of the general formula R2Cu(CN)Li are superior to R2CuLi. 107 Although the

nucleophilic opening of oxiranes by Grignard reagents has been limited in scope due to side

reactions, it has been shown that the reaction proceeds smoothly in the presence of copper

catalysts. 108

Table 2.1 Types of organocopper reagents 103

Type Preparation

Copper-catalysed Grignard reagents

Mono-organocopper reagents

(organocopper reagents)

Homocuprate reagents

(cuprate or Gilman reagents)

Heterocuprate reagents

Higher order homocuprates, e.g.

Rn,+„CurnLin (m + n > 2)

Higher order heterocuprates,

e.g. higher order cyanocuprates

RMgX (or RLi) + < 25 mol% CuX

RM + CuX —> RCu

2 RM + CuX —> R2CuM

RM + CuZ --> RCu(Z)M

3 RLi + 2 CuI —> R3Cu

5 RLi + 3 CuI -3 R5Cu3Li2

2 RM + CuCN —> R2Cu(CN)M2

Species Comments

R = alkyl, aryl, vinyl, etc.

CuX is usually CuI or CuBr•SMe2

CuZ is usually CuSPh or CuO-t-Bu

RM is usually RLi or RMgX

Ether or THE used as solvent

Reactions usually carried out at - 78 ° C to 0 ° C

under N2 or Ar

Solubilizing ligands [e.g. Me2S, (EtO)3P, Bu3P] are

often added

Addition of Lewis acids (e.g. BF3, A103) or

Me3SiX can enhance activity

Copper-catalysed Grignard reagents are still of great utility owing to their ease of

preparation and the normal requirement of only a modest excess of organometallic reagent.

They are considered as the first choice of reagent for nucleophilic addition and

stoichiometric reagents is only investigated if problems are encountered.103

Page 65: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

52

Copper-catalysed Grignard openings of epoxides are commonly used reactions. 109 Some

model C-C bond formation reactions were attempted using this methodology; the results are

summarized in Table 2.2.

Table 2.2 Summary of model C-C bond formation reactions

Grignard reagent a

Epoxide Reaction Product c Yield

conditions b

MgBr

MgBr

0 °C, 2h

/KT 10% CuId -30 °C to rt, 5.5 he

Mixture of more than 10 products

Mixture of 4 products

OH 79%

OH 10% CuCN f

99% MgBr -30 °C to rt, 5.5 he

O g 10% CuCNIf 30 °C to rt, 5.5 he

OH

95%

Grignard reagents were prepared by standard procedurello from 1-bromobutane as a 0.8

mol/dm3 solution in dry diethyl ether.

Reactions were carried out under positive N2 pressure in flame dried glassware using dry,

oxygen-free THF as solvent.

Product mixtures were analysed by GC and GC/MS and purified by Kugelrohr distillation prior

to NMR analysis. Analytical data confirmed the structure and constitution of the alcohols.

Cul was prepared from CuSO4 and NaI.

The Cu(I) salt (10 mol %) and the Grignard reagent (1 equivalent) were stirred at -30 °C for 30

minutes. An equimolar amount of epoxide in THF was added dropwise to the reaction mixture

at -30 °C and the mixture was kept at this temperature for 1 hour. The reaction mixture was then

allowed to warm to room temperature over a period of 4 hours.

Obtained commercially.

Prepared from 1-iodo-2-pentanol and determined by GC analysis to be 95% pure.

Page 66: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

53

The results confirmed the expected side reactions for the addition of a Grignard reagent to

an epoxide. The facile addition of the Grignard reagent in the presence of catalytic Cu(I)

salts to the oxirane moiety is in accord with the literature. 108 A high quality of the Cu(I) salt

is of utmost importance to the success of the subsequent reaction - this is also reflected in the

model studies. The prepared Cul can contain impurities such as iodine and the

accompanying product of disproportionation Cu(II). 109 The higher yield with CuCN can

possibly be attributed to the higher purity of the commercially obtained salt.

Considering the results obtained with the model studies, the formation of the dimer carbon

skeleton was attempted using copper-catalysed Grignard reagent ring opening of

1,2-epoxypentane (Scheme 2.10). The Grignard reagent was easily generated by standard

procedurello from 2-bromopentane as a 0.8 mol/dm 3 solution in diethylether. Two closely

related products (with identical MS data) were detected in a 1 : 1 ratio by GC and GC/ MS

(Figure 2.2).

OH MgBr

10% CuCN

-30 °C to rt, 5.5 h, 70%

43 35 44

Scheme 2.10. Formation of Cio alcohol.

Initially it was envisaged that these products formed as a result of a loss of regioselectivity

thus forming the following byproduct:

OH

45

Page 67: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

100z

TOT -

500 550 8.33 9.16

Figure 2.2. Gas chromatogram of two alcohol products.

It is known that the nature of the organometallic reagent can have a marked effect on the

regioselectivity of the reaction. 103,109 Different organometallic reagents were generated and

their addition to 35 studied in order to observe their influence on the product distribution.

These results are summarised in Table 2.3 and it was clear that the product distribution was

not influenced by using different nucleophilic species. Thus, the possibility that the

products were regioisomers was unlikely. It was proposed that the two products were

diastereomers resulting from the generation of two stereogenic centres in 44. It should be

noted at this stage that although the presence of four stereoisomers can be expected,

enantiomeric pairs could not be observed by the analytical techniques employed.

Consequently although enantiomers certainly form only the formation of diastereomers are

discussed.

54

44

Page 68: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

55

Table 2.3 Influence of different organometallic nucleophiles on the following general

reaction:

OH Metal

44

35

Organometallic

reagenta

Mg Br

Reaction

conditions b

No additive

Product

distribution c

Mixture of more than 10 products

Yield

Mg Br

/\) 10% CuCN 1 : 1 70 %

30% CuCN 1 : 1 69 %

60% CuCN 1 : 1 69 %

Cu(CN)MgBr

/\) No additive 1 : 1 72%

Cu(CN)(MgBr)2

)2

No additive 1 : 1 75%

The organometallic reagents were prepared according to standard procedure.lmim

The reaction mixture was kept at -30 °C while 35 dissolved in THE was added dropwise. The

reaction mixture was kept at this temperature for 1 hour and then allowed to warm to room

temperature over 4 hours.

Product distribution was determined by GC/MS (approximate ratios are given).

These stereogenic centres form as a result of the stereospecific trans opening of the epoxide

by two possible nucleophilic approaches (Scheme 2.11)

Page 69: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

H IMe

(c),

H Me

56

Scheme 2.11. Formation of two stereogenic centres.

The mixture was purified by Kugelrohr distillation (70 °C, 0.2 mrnHg) and an inseparable

mixture of the two diastereomers in a 1 : 1 ratio was obtained. These two products had

identical MS spectra. The EI-MS indicated that the molecular ion of 44 was very unstable

and fragmentation occured to a large extent. This is characteristic of long chain alkyl

compounds (especially branched ones) and the chemical ionization technique was

subsequently used in order to observe the molecular ion. The CI-MS of 44 afforded a very

weak molecular ion at m/z 158 which is in agreement with a molecular formula of C121 -1220.

Further fragmentation included the loss of a water molecule (m/z 140) and the subsequent

typical fragmentation of a saturated hydrocarbon, i.e. a series of m/z differing by 14 amu

(corresponding to a difference of CH2). The 1 H NMR spectrum of the mixture indicated the

presence of 9 protons as three overlapping triplets (5 0.83 - 0.94), 10 alkyl protons as a

multiplet centred at 5 1.36 overlapping with a 1-proton multiplet centred at 5 1.67, the

hydroxylic proton (5 2.5) and the hydrogen on the alcoholic carbon (5 3.65 - 3.71, m). The

13C NMR spectrum corroborated the presence of two similar compounds; almost all peaks

were duplicated and the chemical shift difference was more pronounced in proximity to the

stereogenic centres. The DEPT technique was used to assist in the assignment of the carbon

atoms. The analysis confirmed the presence of three methyl carbons [5 14.11 (2 C), 14.31 and

14.36, C-1 and C-9; 5 19.26 and 20.37, C-7'], five CH2 carbons, all duplicated (5 18.37 and

18.85, C-2; 5 19.95 and 20.03, C-8; 5 38.99 and 40.02, C-7; 5 40.22 and 40.59, C-3; 5 45.07 and

45.36, C-5) and two CH carbons also in duplicate (5 28.97 and 29.41, C-6; 5 69.42 and 69.81,

C-4).

Page 70: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

57

It was assumed that since there were no anomalies in the analytical data two diastereomers

of the expected product 44 did form and no skeletal rearrangement took place.

2.2.4 Formation of the trimer analogue

i) Bromination of the Cio alcohol

Alkyl bromides are useful precursors for organometallic compounds and can be formed by

the substitution of a hydroxy group. HBr is generally used but for secondary alcohols a SN1

and SN2 mechanism can operate simultaneously and competitively. The intermediate in the

SN1 pathway has the tendency to isomerize to a more stable cation which gives rise to

isomeric products.

Promotion of an SN2 displacement mechanism, and hence greater regioselectivity may be

effected by using PBr3 or PBrs. The use of phosphorus reagents, however, avoids

rearrangement only to some extent in secondary alcohol substrates. Other regioselective

brominating reagents are chlorotrimethylsilane/lithium bromide and hexamethyldisilane/

pyridinium perbromide. 111 Reaction with Ph3P and Br2 or CBr4 as well as the Mitsunobu

procedure (Ph3P, diethyl azodicarboxylate, LiBr) 112 allows the formation of alkyl bromides

under mild conditions.

A model secondary alcohol, 2-heptanol (prepared in the study of C-C bond formation

reactions, Table 2.2) was brominated using PPh3/CBr 4 (Scheme 2.12). Two equivalents of

PPh3 were added portionwise to a mixture of one equivalent of the alcohol and two

equivalents of CBr4 at 0 °C in CH3CN. The reaction mixture was then refluxed for 30

minutes and the resultant product after work-up was analysed with GC/MS. The alcohol

45 was converted quantitatively to the corresponding alkyl bromide 46, but separation from

the reaction byproducts such as phosphines and bromomethanes proved difficult.

Page 71: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

OH

58

45

Ph3P / CBr4

CH3CN

0 °C to reflux, 1 h 46

Scheme 2.12. Bromination of 2-heptanol.

The branched alcohol 44 was also brominated by using this procedure in order to ascertain if

this product could be isolated (Scheme 2.13). The purification was unsuccessful.

Alcohol 44 was subsequently brominated with PBr3 using standard procedure. 113,114,115 PBr3

(0.33 equivalents) was added dropwise to the neat alcohol (44) while the temperature was

kept between -5 °C and 5 °C. The reaction was allowed to warm to room temperature and

to proceed for eight hours, after which work-up and purification by Kugelrohr distillation

afforded the product mixture as a clear liquid in 80% yield.

OH Ph3P / CBr4

CH3CN

0 °C to reflux, 1 h 44

47

Scheme 2.13. Bromination of alcohol 44.

Both GC and GC/MS indicated that carbocation rearrangement did take place and the

product contained 10% of an isomeric alkylbromide (lowering the yield of 47 to

approximately 72%). This isomer formed in the same proportion under a variety of

controlled reaction conditions (e.g. under nitrogen, rigorous drying and purification of

reagents, in dry diethyl ether at -30 °C, the removal of excess HBr by passing a N2 stream

through the solution and shorter reaction times). Furthermore, this byproduct could not be

separated from the desired product and it was decided to use the mixture in further

reactions.

The analysis also indicated that as expected, two diastereomers of 47 were present in a 1: 1

ratio. The CI-MS (and El-MS) data of all three compounds were the same. The molecular

ion (m/z 221) was absent, but the ion corresponding to the loss of HBr (m/z 141) and the

Page 72: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

59

typical alkane fragmentation (repetitive loss of 14 mass units) were present. 1H NMR

indicated the resonances of 47 as the main product: the three methyl groups overlapped as a

9-proton multiplet 6 0.83 - 0.93, the secondary and tertiary protons overlapped as a 11-

proton multiplet 6 1.17 - 1.92 and the bromide proton was represented by a 1-proton

multiplet centred at 6 4.09. The 13C NMR spectrum confirmed the presence of three

compounds of which two were diastereomers of 47. The DEPT technique was used to assist

in the assignment of the peaks of these diastereomers; almost all peaks were duplicated.

The resonances of three methyl carbons (6 13.47, 6 14.25 and 14.30, C-9; 6 18.67 and

19.73, C-7'), the five methylene carbons (6 19.93, 6 20.70 and 20.79, C-8; 6 38.13 and

39.66, C-7; 6 41.21 and 42.08, C-3; 6 46.51 and 46.99, C-5) and two CH carbons (6 30.85, C6; 6

56.50 and 56.89, C-4) were found. The peaks representing the third compound were in the

same range as the peaks of the diastereomers thus confirming the theory that it was a

structural isomer.

ii) The synthesis of the C15 alcohol

a) Grignard formation

Br

47

Mg

MgBr

diethyl ether

48

Scheme 2.14. Generation of the Cio Grignard reagent.

Attempted synthesis of the secondary Grignard reagent 48 by standard procedure 11° in

diethyl ether failed. No active organometallic species formed, even under tightly controlled

reaction conditions (e.g. under a nitrogen atmosphere, rigorous drying and purification of

alkyl bromide and solvent), extended reflux (up to 7 days), use of a higher boiling solvent

(THF, dlisopropyl ether) and a large excess of magnesium. The alkyl halide 47 might be

unreactive because it is a secondary alcohol and the branching and long chain nature may

cause steric hindrance.

Page 73: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

60

It is known116 that simple organic halides are sometimes rather unreactive in the preparation

of their Grignard reagents. Modifications have been developed to enable the formation of

these difficultly accessible Grignard reagents.117 ,116 These includes the use of higher boiling

and more strongly coordinating solvents, 118 catalysis with transition metal salts / 119

physicall 2o and chemical121 methods to produce highly activated magnesium. Among these

modifications, chemically activated magnesium has been widely used.

The addition of catalytic amounts of iodine to the reaction mixture, the use of stoichiometric

iodine-activated magnesium (Mg/ I2), finely divided Mg (formed by stirring Mg turnings

under N2 for one day) 122 and magnesium activated by acidic wash did not result in the

formation of the Grignard reagent.

If the R group remains the same the reactivity of organic halides in the formation of

Grignard reagents is: RI > RBr > RC1. 1111 Consequently the bromide 47 was converted into

the alkyl iodide 49 by refluxing with 1.25 equivalents of NaI in acetone (Scheme 2.15). This

iodide was also unreactive towards magnesium.

Br Nal

acetone

44 reflux, 6h, 80% 46

Scheme 2.15. Preparation of Clo iodide.

The entrainment procedurel 23 was successful in generating the Grignard reagent 48 in

approximately 60% yield. In this procedure all the "inert" halide 47 was added to

magnesium (> 2 equivalents) in dry diethyl ether and the resulting suspension refluxed.

The entrainment reagent ethylene dibromide (stoichiometric amount) dissolved in dry

diethyl ether was then added dropwise. The auxiliary halide keeps the magnesium clean

and forms ethylene and magnesium dibromide, thus activating the magnesium for reaction

with the inert halide without introducing a second Grignard rPagent.

Page 74: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

49 35

MgBr

10% CuCN

-30 °C to rt, 5.5 h, 70%

61

b) Carbon-carbon bond formation

The nucleophilic ring opening of 35, using the same methodology as before proceeded

smoothly to afford 50 as a clear oil in 70% yield (Scheme 2.16). Three stereogenic centres are

present in alcohol 50 and thus a total of eight stereoisomers can be expected of which the

four diastereomers were observed.

OH

Scheme 2.16. Formation of C15 alcohol.

The product was purified by Kugelrohr distillation (130 °C, 0.15 mm Hg) and analysed by

GC/MS. The chromatogram confirmed the presence of the diastereomers as well as some

structural isomers of 50. The constitutional isomer of 47 could also have formed a Grignard

reagent and reacted with 35 to yield another isomer of 50. These isomers represented less

than 5% of the total reaction mixture. The different isomers could not be separated and thus

alcohol 50 was not purified any further.

The molecular ion of 50 (m/z 228) was absent in the CI-MS and the loss of water (m/z 210)

gave rise to a weak ion peak. This can be attributed to the highly branched nature of 50

which causes the molecular ion to fragment easily. The expected fragmentation to form

clusters of peaks 14 mass units apart was observed. Other fragments within each cluster

corresponding to additional losses of one or two hydrogen atoms were observed. Cleavage

to yield secondary carbocations is responsible for the relatively high intensities of a C101 -119+

(m/z 139) and a C5F111+ (m/z 71) fragment. The 1 H NMR spectrum verified the presence of

4 methyl groups resonating as overlapping triplets 8 0.78 - 0.89, 8 methylene groups

resonating as a multiplet centred at 8 1.23 overlapping with 2 CH protons (5 1.63, m). The

resonances of a hydroxylic proton (6 2.33, s) and the proton on the alcoholic carbon (6 4.52,

m) were also observed. The 13C NMR spectrum established the presence of the four

Page 75: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

62

diastereomers of 50, some peaks were quadrupled. The DEPT technique was used to assist

in the assignment of the carbon atoms. The resonances of 3 methyl carbons as exceptionally

broad peaks were found (6 14.13, 14.40 and 14.48; C-1, C-9' and C-11). The broadening of

the peaks are due to the presence of diasteromeric carbons resonating at nearly identical

chemical shifts. The four diastereomer peaks of the methyl carbon on the stereogenic centre

(8 19.81, 19.91, 20.00, 20.35; C-9"), the eight groups of methylene carbons C-2 (6 20.07, 20.13

and 20.21), C-10 (8 21.09, 21.17 and 21,24), C-8' (6 32.56, 32.71, 32.97 and 33.00), C-7' (8 33.21,

33.39, 33.42 and 33.66), C-9 (6 37.88, 38.36, 38.40 and 38.62), C-7 (6 38.64, 38.87 and 38.98),

C-5 (6 39.66, 39.78 and 39.87), C-3 (6 40.11, 40.15, 40.27 and 40.41) and the three CH carbons

C-8 (6 30.14, 30.22, 30.27 and 30.34), C-6 (5 36.10, 36.22, 36.29 and 36.34) and C-4 (6 67.60,

broad) were observed.

iii) Synthesis of a model trimer

Although the C15 model trimer could be formed from alcohol 50 by the elimination of the

alcohol group followed by reduction, this was viewed as a cumbersome route. It was

envisaged that alkane 32 could be formed by the displacement of a secondary halide by an

organometallic species (Scheme 2.17).

Metal

47

32

Scheme 2.17. Preparation of model trimer.

This C-C bond formation was attempted by using a Grignard reagent as the organometallic

species. This Grignard reagent was prepared by standard procedureno from

1-bromopentane as a 0.8 mol.dm- 3 solution in diethyl ether. The copper(I) catalysed

nucleophilic displacement of the halide did not proceed at all. Neither did reaction with the

higher order organocuprate nor catalysis of this reaction with a Lewis acid (BF3.Eb0).

Page 76: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

63

It is known that most higher order cyanocuprates are very effective in the displacement of

secondary bromides. 124 The role of the gegenion in the displacement reaction has been

studied. 125 The replacement of the organolithium with an organomagnesium reagent

usually dramatically decreases the rate of the reaction.

An organolithium reagent was formed by standard procedure 126 from the primary alkyl

halide and lithium pellets. From this organolithium a higher order cyanocuprate was

formed and reacted with 47 at -50 °C. The reaction was allowed to warm to 0 °C and

proceed for 6 hours after which work-up afforded a clear oil in 50% yield.

The GC/ MS confirmed the formation of two diastereomers of 32 in a 1: 1: ratio as the main

products. A byproduct representing 5% of the reaction mixture was observed. The CI-MS

furnished an extremely weak ion peak at m/z 211 which corresponds to hydride abstraction

from 32. Furthermore, the characteristic alkane fragmentation i.e. a series of m/z differing

by 14 mass units was observed. The 1H NMR spectrum confirmed that no functional groups

were present, only alkane peaks were detected. The resonances of the main classes of

protons were found, namely, the four methyl groups (6 0.83 - 0.89), nine methylene groups

(centred at 8 1.22) overlapping with the two CH groups centred at 5 1.29. The 13C NMR data

corroborated to some extent that two diastereomers were present; only some of the peaks

were duplicated. The presence of the following alkane carbons were established using the

DEPT technique: 4 methyl carbons [6 14.06 (3 C) and 14.49; C-11, C-1, C-9 and C-5']; 9

methylene carbons (6 22.67 - 36.08, some peaks overlapped); and two CH carbons [6 32.45, (2

C), C-4; 6 37.15 and 37.36, C-6].

2.2.5 Attempted synthesis of model tetramer

In order to synthesise the model tetramer (33) by the same protocol as before, the alcohol

group in 50 had to be converted into a good leaving group. This C15 alcolhol proved to be

extremely unreactive and prone to elimination as exemplified in Table 2.4.

The unreactivity of this secondary alcohol lies in part in the long chain branched nature of

the alkyl substituents. This functional group is sterically hindered and hence a low

Page 77: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

64

reactivity was observed. Conversely, the 13-hydrogen is more accessible and elimination to

form an alkene was favored. In other words, the reaction proceeding via a SN1 pathway

yields an unstable carbocation while the reaction via a SN2 pathway is sterically hindered.

The functionalization of the substrate and the attempted synthesis of the tetramer using this

route was abandoned.

Table 2.4: Attempted functional group transformation

General reaction :

OH

50

X

Reaction conditions

Result (leaving group)

Br PBr3 (0.33 eq) 0 °C to rt, 16 h No reaction 80 °C, 2 h No reaction 80 °C, 16 h No reaction

PBr3 (3 eq) 140 °C, 12 h Elimination

I

Mitsunobu conditions PPh3, imdazole, 12 100 °C,16 h Elimination rt, 1 h Elimination

OTs TsC1 (1.5 eq), pyridine 0 °C, 1h

TsC1 (1.5 eq), NaH (1.2 eq)

No reaction

Tosylate formed, product unstable and eliminated

Page 78: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

65

2.3 Catalyst survey

Although the oligomerization of other alpha olefins has been studied, there is limited

knowledge concerning the oligomerization of 1-pentene with common catalysts.

Furthermore, there are almost no systematic studies on the oligomerization of higher alpha

olefins that enables the comparison of different catalysts and alpha olefin activity. In this

project, some of the main oligomerization catalysts were evaluated for the dimerization and

trimerization of higher alpha olefins and especially of 1-pentene.

2.3.1 Cationic oligomerization

The first and most common of these catalysts is the BF 3/ROH cationic oligomerization

system. BF3/n-BuOH was employed to produce 1-pentene oligomers and this reaction was

used as standard to compare the different catalysts.

The reaction was carried out in a reproducible fashion by bubbling BF 3(gas) rapidly through

a mixture of 1-pentene (2.7 equivalents), heptane (3 equivalents) and n-BuOH (0.05

equivalents) for 5 minutes, the gas flow was then reduced to such a rate that the pressure in

the system remained approximately constant. The temperature of the reaction mixture rised

initially to ± 40 °C after which the reaction was allowed to proceed for another 30 minutes

and then quenched with aqueous NaOH. Work-up afforded a light brown oil in high yield.

This oil was hydrogenated over 1% Pd/C (10% m/m) with H2 (40 - 50 bar) for 2 hours. The

degree of unsaturation was determined by standard procedure 127 and it was found that the

product was completely hydrogenated. Both this product and the unsaturated oil were

analyzed by GC and GC/ MS. Dimers through to pentamers were formed and the general

distribution is summarized in Table 2.5. The product distribution peaked at the trimer

which is in accord with results obtained for other feedstocks. At higher temperatures the

distribution shifted to the higher oligomers but still peaked at the trimer.

Page 79: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

66

Table 2.5. Product distribution of BF3/n-BuOH oligomerization of 1-pentene.

Oligomer Cu C15 C20 C25

Oh 15 47 23 15

The GC/MS analyses indicated that, as expected, individual fractions consisted of a large

number of isomers (Figure 2.3) - even more than can be explained in terms of the classic

cationic mechanism. It has been found that product quality not only depends on oligorner

distribution, but also on the specific structures present in the various oligomers. 26 Due to

the formation of stable secondary ions, the mass spectra of branched hydrocarbons should

include intense peaks that deviate from that expected for straight-chain hydrocarbons. This

phenomenon provides a convenient method to determine the branching pattern of the

isomeric mixtures formed in the BF3/n-BuOH oligomerization of 1-pentene. It was,

however, found that although each oligomer had a unique fragmentation pattern, it was

impossible to assign specific structures unambigUously to the isomers. The GC/MS

analyses did indicate that, as expected, branched oligomers were formed, even though the

different isomers could not successfully be distinguished, because of the complexity of the

mixture.

100v.

TOT-

600 000 1000 1200 1400 10.00 13.33 16.66 19.99 23.33

Figure 2.3. Gas chrornatrogram of 1-pentene trimer.

Page 80: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

67

The trimer (C15) was isolated by high vacuum distillation (0.20 mmHg): dimer 80 - 85 °C,

trimer 85 - 115°C, and tetramer 125 - 145 °C. The isolated, hydrogenated trimer fraction was

used in viscosity studies of 1-pentene oligomers, which will be discussed later.

2.3.2 Radical oligomerization

Almost all organic peroxides are photo- and thermally sensitive because of the facile

cleavage of the weak oxygen-oxygen bond; AH = -125.6 to -184.2 kJ/mo1. 128 The initially

generated species, the free radicals, are reactive intermediates with very short lifetimes, e.g.

t112 = 10-3 S.129 As a consequence organic peroxides are used as initiators for many different

free-radical reactions, especially in commercial polymerizations.

The activity of different organic peroxide radical initiators in the oligomerization of 1-

pentene was investigated. These include: dibenzoyl peroxide, one of the first peroxides

used in the industry; 26 di-tert-butyl peroxide, patented by Mobil for the oligomerization of

lower oligomers from other processes; 27 and dicumyl peroxide, reported in a French patent

application for the oligomerization of C9 - Cu alpha olefins. 13°

Unless otherwise stated, the reactions were performed under pressure in an autoclave with

the catalyst, monomer and a small amount of heptane (as internal standard); initiation

occurring by thermal homolysis. The results are summarized in Table 2.6. GC and GC/ MS

analyses of the reaction mixtures were complicated by the presence of catalyst

decomposition products. Most of these decomposition products had retention times and

peak intensities similar to those of the oligomers. The estimation of the product distribution

and yields are thus only rough approximations.

Page 81: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

68

Table 2.6. Radical oligomerization of 1-pentene with organic peroxide initiators.

Initiator Reaction conditions. Resultsb•

Conversion % C10 C15 C20

Dibenzoyl peroxide

37 °C; Reflux, UV light 24 h; 5 mol% catalyst

- - - -

100 °C; 40 psi 24 h; 2 mol% catalyst

8 5 3 -

150 °C; 80 psi 24 h; 5 mol% catalyst

10 4 5 1

150 °C; 140 psi 24 h; 10 mol% catalyst

10 4 5 1

150 °C; 170 psi 24 h; 15 mol% catalyst

10 4 5 1

Dicumyl peroxide

100 °C; 40 psi 4 h; 2 mol% catalyst

- - -

100 °C; 40 psi 24 h; 2 mol% catalyst

5 1 4 -

150 °C; 80 psi 24 h; 2 mol% catalyst

5 1 4 -

150 °C; 80 psi 24 h; 5 mol% catalyst

15 5 10 -

150 °C; 86 psi 24 h; 10 mol% catalyst

25 10 15 -

150 °C; 96 psi 24 h; 15 mol% catalyst

25 10 15 -

Di-t-butyl peroxide

100 °C; 50 psi 24 h; 2 mol% catalyst

- - -

150 °C; 90 psi 24 h; 5 mol% catalyst

30 5 20 5

150 °C; 90 psi 24 h; 10 mol% catalyst

40 10 30 -

150 °C; 90 psi 24 h; 15 mol% catalyst

50 15 35

1-Pentene : Heptane in a 3 : 1 ratio.

Maximum pressures are quoted (after temperature equilibration, about 1 hour).

Reaction mixtures were allowed to cool down to room temperature and analyzed with

GC/MS.

Conversions and product composition were calculated by comparison with the peak area of

the internal standard and unconverted monomer.

Primary and secondary dialkyl peroxides undergo competing electrocyclic reactions upon

thermal decomposition (Scheme 2.18) and are more susceptible to radical-induced

decomposition. 131,132 Consequently di-tert-alkyl peroxides are used as initiators in most

applications. In the absence of reactive substrates, tertiary dialkyl peroxides yield primarily

Page 82: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

69

t-alcohols, ketones, hydrocarbons and minor amounts of ethers, epoxides and carbon

monoxide. Decomposition products from primary and secondary dialkyl peroxides include

aldehydes, ketones, alcohols, hydrogen, hydrocarbons, carbon monoxide and carbon

dioxide.

H H 0 , 1 I .. • II

R— C-0-0=CR,

R"CR' + H2 I R" R

I "

solvent

Scheme 2.18. Decomposition of a secondary organic peroxide.

Conversions with each of these catalysts were low (< 30%), but it seemed that di-t-butyl

peroxide was the better initiator. Although each oligomer consisted of a number of isomers,

fewer isomers formed than in the cationic oligomerization. An increase in catalyst

concentration increased the conversion. A certain temperature was necessary in order to

initiate radical formation, for these catalysts 150 °C was sufficient.

2.3.3 Metallocene catalysts

Metallocenes in the presence of methylaluminoxane are receiving a lot of attention, both

academically and industrially as a new generation of homogeneous, single-site Ziegler-

Natta catalysts. The focus of this investigation was the oligomerization of alpha olefins,

because higher C5 - C20 alpha olefins are of increasing importance as a source for many

industrially important chemicals. Although various catalysts have been described for the

oligomerization of alpha olefins, little attention has been paid to the application of

metallocene/ MAO systems in this process. These catalysts should advantageously combine

high catalyst activity with the potential to control properties such as molar mass, molar

mass distribution and stereoregularity through rational ligand design at the transition metal

centre. Depending on the chain termination, the type of end group may also be controlled.

In this preliminary study, the potential of metallocene/MAO catalysts to fine-tune the

oligomer characteristics, was investigated.

Page 83: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

51

Ti

70

As a starting point in our study, the oligomerization of 1-pentene, 1-hexene and 1-octene by

titanocene dichloride (51) in the presence of MAO was examined. As mentioned previously

a decreased Al / Metal ratio, increased temperature and lowered alkene concentration

should result in the formation of oligomers. Oligomerizations were carried out in 100 ml

flamed-out Schlenk-flasks under argon at an oil bath temperature of 40 °C, unless stated

otherwise. The MAO and metallocene (Al : Metal in a ratio of 150: 1 as opposed to a ratio of

greater than 500 required for polymerization) were preactivated for a short period before the

monomer was added. After the specified reaction time, the reaction mixtures were allowed

to cool to room temperature and the reactions were quenched with a 10% HC1/ Me0H

solution. After work-up, the isolated oligomers were analyzed with GC/MS. The results

are compiled in Table 2.7.

Initially the yields of the reaction were not reproducible, especially when 1-pentene was

used as starting material. As before, this might be attributed to the high volatility and low

boiling point of 1-pentene. Losses of monomer at the reaction temperature of 40 °C could

influence conversion and product distribution. To ensure reproducibility, the

oligomerization reactions were repeated until at least two runs were consistent.

The data indicated maximum conversions of 37% and 48% for 1-pentene and 1-hexene

respectively after 24 - 48 hours. Longer reaction times increased the selectivity for higher

oligomers. 1-Octene showed decreasing conversion with longer reaction times and this

could be due to the formation of higher oligomers or polymers which could not be observed

by GC analyses. In strong contrast to the cationic oligomerization where a multitude of

isomers formed, each oligomer fraction consisted of one compound only (>95%). A small

amount of products was formed with 15 mass units more than the expected oligomer, e.g. a

C11 compound where a C10 was expected.

Page 84: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

71

Table 2.7. Oligomerization of alpha olefins with titanocene dichloride/MAO.'

Alpha

olefin"

Reaction Conditions

Results`

Conver-

siond TONe Yields % f

(Selectivity % g) Dimer Dimer Trimer Trimer Tetra-

+ Me + Me mer 1-pentene 50 °C, 2h 0 - - - - - - - (800 eq.) 10 ml toluene

1-pentene 50 °C, 24h 0.30% 2 - - 0.08 - 0.22 (800 eq.) 10 ml toluene (25) (75)

1-pentene 40 °C, 24h 33% 124 14 1 18 - - (375 eq.) (43) (2) (55)

1-pentene 40 °C, 48h 37% 139 16 1 20 - - (375 eq.) (44) (2) (54)

1-pentene 40 °C, 72h 37% 139 9 1.5 20 1.5 5 (375 eq.) (25) (4) (54) (4) (13) 1-hexene 50 °C, 24h 39% 146 36.6 2 0.4 (375 eq.) (94) (5) (1) 1-hexene 40 °C, 24h 48% 180 46.6 1 - 0.4 - (375 eq.) (97) (2) (1) 1-hexene 40 °C, 48h 48% 180 43 2 3 - - (375 eq.) (89) (4) (7) 1-hexene 40 °C, 72h 64% 240 53 1.3 7.7 12 - (375 eq.) (83) (2) (12) (3) 1-octene 40 °C, 24h 50% 250 50 - - - (375 eq.) (100) 1-octene 40 °C, 48h 35% 131 34 - 1 - - (375 eq.) (97) (3) 1-octene 40 °C, 72h 16% 60 16 - - - (375 eq.) (100)

Cp2TiC12 (1 eq.); MAO (150 eq.).

Alpha olefins were dried over activated molecular sieves prior to use. •

Yields and seletivities were calculated from the GC/MS peak areas of each oligomer by

comparison with an internal standard.

Conversion —

mass of productx 100%

mass of monomer

TON = turnover number — mol of converted monomer

mol of catalyst

1) mol of compound X

Yield of compound X = x 100% mol of monomer

yield of compound g) Selectivity for compound X — x 100%

total yield

Page 85: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

72

The reaction mechanism for the formation of these methylated products with the titanocene

dichloride/ MAO catalyst is depicted in Scheme 2.19. Firstly an excess of MAO generates

the active methylated metallocene complex 52. The alkene coordinates with the transition

metal followed by the insertion of the olefin into the metal-alkyl bond to form complex 53.

As noted in the literature study, 1,2-cis olefin insertion is mainly observed for metallocene

catalysts. The second olefin also coordinates with the transition metal and inserts into the

growing metal-alkyl chain to give complex 54. This complex (54) can then either form the

dimer 56 and a hydridotitaniumcomplex 55 by 13-hydride elimination or react with a third

olefin to yield trimer 57. This termination reaction is the most commonly observed for the

metallocene catalysts and may occur via an intermediate hydride complex or by direct

hydride transfer to the monomer. 46

Thus this dimer (and trimer) contains one more methyl group than expected due to methyl

insertion from a Ti-Me starting group. This was also observed by Janiak and coworkers in

the dimerization of propylene and 1-hexene by Cp2ZrC12 and MA0. 133 This effect is,

however, only important in the dimerization step; in subsequent catalytic cycles a Cp2Ti+-H

species formed by 13-H elimination is the only discerning chain transfer process.

The complex 55 is the main catalyst present in the reaction mixture and forms the normal

dimers and trimers as shown in Scheme 2.20. As before, propagation occurs by insertion of

the monomer double bond into the growing chain and termination occurs by 13-hydrogen

transfer.

The 1-pentene dimer (C1oH20) and the trimer (C15F130), 1-hexene dimer (C12H24), and 1-octene

dimer (C16H32) was isolated and purified by Kugelrohr distillation. The structures of these

products (Figure 2.4) were determined by 1 H and 13C NMR. GC analyses of the product

indicated that only one compound was present in each fraction.

Page 86: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

73

Me

Ti

+ [—A1-0—] n Me

CI

52

+Ti? \ Me

MAO

51

+Tf Me

54

0

+Ti

1,2-cis-insertion

13-hydride elimination

52

1,2-cis-insertion

1,2-cis-insertion

O +Ti

(3-hydride elimination R R R 55 +

Me

57

Scheme 2.19. Formation of methylated oligorners.

Page 87: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

0

+Tr"

H

74

R\ 55 +Tf

H

O •■•.,..R

1,2-cis-insertion

1,2-cis-insertion

O +Ti

n-hydride elimination of)

59

60 R = C3H7

61 R = C4H9

62 R = C6H13

1,2-cis-insertion i e •- • ■,

O +Ti

3-hydride elimination 55 +

63 R = C3H7

Scheme 2.20. Formation of normal dimers and trimers.

Page 88: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

75

60

63

1 -pentene dimer

1 -pentene trimer

61

62

1 -hexene dimer

1-octene dimer

Figure 2.4. Structure of oligomers.

The EI/MS of the 1-pentene dimer 60 (C1oH20) gave a molecular ion at m/z 140 and an ion at

m/z 139 corresponding to hydride abstraction. Further fragmentation was the typical

clusters of peaks spaced 14 mass units apart. The 1H NMR spectrum indicated the

resonances of two methyl groups as triplets (6 0.87, J 6 Hz; 6 0.88, J 7.5 Hz) and 8 methylene

protons as overlapping multiplets 6 1.26 - 6 1.47. The resonances of the four methylene

protons on the carbons alpha to the quaternary carbon overlapped in a broadened triplet

(6 1.98, J 7.8 Hz) and the vinylidene protons resonated as a singlet at 6 4.68. Ten carbon

peaks were distinguished in the 13C NMR spectrum of 60. The multiplicities of the peaks

were confirmed by the proton-coupled 13C NMR spectrum: two methyl carbons as quartets

(6 13.8 and 14.0), six methylene carbons as triplets (6 20.8 - 38.1), the terminal vinylidene

carbon (6 108.5, t) and the quaternary alkene carbon (6150.2, s) were observed.

The structures of the other dimers were elucidated in analogous manner, the resonances of

the expected number of methyl and methylene groups were found and each oligomer

exhibited the characteristic vinylidene end group resonances. The diagnostic resonances of

the CH group, the three methyl groups, the nine methylene groups and the vinylidene

group of the trimer 63 were observed.

The 1-pentene tetramer was also isolated but NMR analyses proved to be complicated due

to the large extent of peak overlap of the resonances of the methylene carbons in the 13C

NMR spectrum. The presence of the vynilidene end group and the four methyl groups

Page 89: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

76

were, however, established. A molecular ion with m/z 280 corresponding to a constitution

of C2oH40, was observed in the CI-MS data.

The metallocene component of the catalyst and the reaction conditions were subsequently

modified to study the influence on the oligomerization reactions.

i) Metal

It was noted in Chapter 1 that in general zirconocene derivatives are more active

polymerization catalyst than their titanocene counterparts. The activity of zirconocene

dichloride 64 as a catalyst for the oligomerization of alpha olefins was investigated and the

results are given in Table 2.8. The general reaction conditions are similar to those quoted in

the previous table.

64

Oligomerization reactions using Cp2ZrC12/MAO as catalyst was more reproducible than

reactions using Cp2TiC12/ MAO. In general, conversions were higher, only marginally so for

1-pentene, but this catalyst was clearly superior for 1-hexene and 1-octene. Longer reaction

times increased conversions slightly, except for 1-octene, as was observed previously.

Dimers were predominantly formed without significant portions of the higher oligomers.

These dimers consisted of single compounds that formed with >95% selectivity.

Page 90: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

77

Table 2.8. Oligomerization of alpha olefins with Cp2ZrC12/ MAO.

Alpha olefin

Reaction Conditions

Results

Conver- sion

TON Yields % (Selectivity %)

Dimer Dimer + Me

Trimer Trimer + Me

Tetra- mer

1-pentene 40 °C, 24h 37 139 35.5 - 1.5 - - (375 eq.) .. (96) (4)

1-pentene 40 °C, 48h 36 135 34.6 - 0.4 - - (375 eq.) (96) (4)

1-pentene 40 °C, 72h 42 158 40.7 - 1.3 - - (375 eq.) (97) (3) 1-hexene 40 °C, 24h 76 285 76 - - - (375 eq.) (100) 1-hexene 40 °C, 48h 78 293 77.2 - 0.8 - - (375 eq.) (99) (1) 1-hexene 40 °C, 72h 83 311 82.2 - 0.8 - - (375 eq.) (99) (1) 1-octene 40 °C, 24h 89 334 89 - - - (375 eq.) (100) 1-octene 40 °C, 48h 56 210 54 - 2 - - (375 eq.) (96) (4) 1-octene 40 °C, 72h 41 154 40 - 1 - - (375 eq.) (98) (2)

ii) Metallocene trichloride

The relationship of the electron density at the metal centre and the oligomerization activity

was investigated. The substitution of a cyclopentadienyl ligand with a chloride should

increase the positive charge on the metal due to the greater electron-withdrawing properties

of the halide. This in turn should strengthen the bonding between the metal and other

ligands, particularly the bond to the olefin which will then become less reactive. Titanocene

trichloride (65) was used as catalyst precursor. The results of the oligomerizations with

CpTiCl3/MAO are compiled in Table 2.9.

\T,

/Cl

/ a a

65

Page 91: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

78

Table 2.9. Oligomerization of alpha olefins with CpTiC13/ MAO. a

Alpha olefin

Reaction Conditions

Results

Conver- sion

TON Yields % ' (Selectivity %)

Diener Dimer + Me

Trimer Trimer + Me

Tetra- mer

1-pentene (375 eq.)

40 °C, 24h 5 19 ..

- - 5 (100)

- -

1-pentene (375 eq.)

40 °C, 48h 8 30 0.41 (5.1)

0.07 (0.9)

7.52 (94)

- -

1-pentene (375 eq.)

40 °C, 72h 9 34 - - 7.83 (87)

0.63 (7)

0.54 (6)

1-hexene (375 eq.)

40 °C, 24h 14 53 11.6 (83)

- 2.4 (17)

-

1-hexene (375 eq.)

40 °C, 48h 14 53 13.9 (99)

0.1 (1)

- -

1-hexene (375 eq.)

40 °C, 72h 6 23 3 (50)

0.6 (10)

2.4 (40)

- -

1-octene (375 eq.)

40 °C, 24h 5 19 4.8 (96)

- 0.2 (4)

- -

1-octene (375 eq.)

40 °C, 48h 8 30 8 (100)

- - -

1-octene (375 eq.)

40 °C, 72h 6 23 5.9 (99)

- 0.1 (1)

- -

a) All reaction conditions and calculations are analogous to those quoted in the previous tables.

As expected, the conversions were considerably lower than for the dichloride catalyst

species. A higher selectivity was observed for the higher oligomers. Furthermore, the

individual oligomer fractions and in particular the dimer fraction of 1-pentene and 1-hexene

consisted of more than five products. Steric effects may decrease the selectivity: a number of

insertion or termination reactions may take place, because there is less steric hindrance

around the metal centre.

iii) Substituted cyclopentadienyl rings

Tetramethyl substituted cyclopentadienyl ligands were used to correlate substitution on the

ligands at the transition metal centre with alpha olefin oligornerization activity. The

metallocene 66 was prepared from TiC14 and [CpMe4]1.1+. The EI-MS of this compound

furnished a molecular ion at m/z 360 corresponding to a constitution of (C9H13)2TiC12 and

titanium isotope peaks were observed. Both the 1H and 13C NMR spectra revealed the

Page 92: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

79

expected equivalence of the ring ligands. The 1H NMR resonances of two different sets of

methyl groups as singlets (6 1.89 and 6 2.03) and the two equivalent ring protons (6 5.97)

were found. The 13C spectrum was in accord with this: two sets of methyl carbons (8 13.02

and 6 14.81) and three different sets of ring carbons (8 113.09, 8 125.16 and 6 135.91) were

observed.

Attempted oligomerization reactions with 66 in the presence of MAO (similar ratios of

catalyst, cocatalyst and monomer than before) were unsuccessful. Neither 1-pentene, 1-

hexene or 1-octene were oligomerized, even after a 100 hours at 40 °C. This can be ascribed

to the steric bulk of the ligand around the metal centre. Although polymerization of

ethylene and co-oligomerization with alpha olefins have been observed with substituted

cyclopentadienyl ligands, 1M it seems as if the steric hindrance is detrimental to the

oligomerization of higher alpha olefins. In several studies 135'136 it has been established that

highly substituted cyclopentadienyl ligands for example Cr (Cp*

pentamethylcyclopentadienyl) exhibit the lowest activities in polymerization reactions.

iv) Temperature

Ambient reaction temperatures would be economically favored in industrial applications.

The oligomerization activity of zirconocene dichloride at room temperature was studied

since it was the most active metallocene catalyst at 40 °C (Table 2.10). Monomer conversions

comparable to the higher temperature oligomerizations were observed after 24 hours at

room temperature. After this period the yields decreased, probably due to polymerization

Page 93: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

80

and the formation of higher oligomers. Dimers consisting of only one product were isolated

and insignificant amounts of higher oligomers formed.

Table 2.10. Room temperature oligomerization of alpha olefins with Cp2ZrC12/ MAO.

Alpha olefin

Reaction Conditions

Results ' Conver-

sion - TON Yields %

(Selectivity %) Dimer Dimer

+ Me Trimer Trimer

+ Me Tetra- mer

1-pentene rt, 24h 36 135 35 - 1 - - (375 eq.) (96) (4)

1-pentene rt, 48h 9 34 7.7 - 1.3 - - (375 eq.) (86) (14)

1-pentene rt, 72h 6 23 4.5 - 1.5 - - (375 eq.) (75) (25) 1-hexene rt, 24h 90 338 90 - - - (375 eq.) (100) 1-hexene rt, 48h 33 124 33 - - - - (375 eq.) (100) 1-hexene rt, 72h 39 146 32 - 7 - - (375 eq.) (83) (17) 1-octene rt, 24h 86 323 86 - - - (375 eq.) (100) 1-octene rt, 48h 12 45 12 - - - - (375 eq.) (100) 1-octene rt, 72h 12 45 12 - - - - (375 eq.) (100)

While this study was in progress Bergman and Christoffers 137 also published that

Cp2ZrC12/ MAO catalysed the oligomerization of C4 - C7 alpha olefins at 25 °C to form

dimers in a 80 - 90% yield.

v) Purity of the alpha-olefin monomer

1-Hexene containing 2 - 3% oxygenate impurities were used as monomer to examine the

sensitivity of metallocene/ MAO catalysts to impurities. Oligomerizations with both

Cp2TiC12/ MAO and Cp2ZrC12/ MAO yielded conversions and product distributions almost

identical to those obtained with pure 1-hexene. This illustrated one of the proposed

functions of methylaiuminoxane as cocatalyst, namely as a scavenger of impurities.

Page 94: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

81

2.3.4 Comparison of different catalysts

Comparison of viscosity data

The viscosity of the hydrogenated trimer fractions from the BF3/n-BuOH and

Cp2TiC12/MAO oligomerization of 1-pentene was determined at different temperatures by

using an Ostwald viscometer. 138,139 These results are tabulated below:

Table 2.11. Viscosity/temperature relationship of 1-pentene trimer fractions.

Temperature Viscosity

(°C) •(cStokes)

BF3/n-BuOH Cp2TiC12/MAO

30 22 63.1

35 19.1 44.6

40 17.3 38.3

The viscosity of the single component trimer was higher than the viscosity of the

corresponding mixture of isomers formed with the BF3/n-BuOH system. This may be due

to the higher degree of regularity that can prevent the molecules from sliding over each

other with ease.

General comparison

The BF3/n-BuOH system yielded a 1-pentene oligomer distribution that peaked at the trimer

and consisted of a large number of isomers. Although the monomer conversion was high,

the trimer was not formed with a very high selectivity. The radical oligomerizations of 1-

pentene gave low conversions and a smaller number of isomers under the selected reaction

„ditto k_viLns.

Page 95: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

82

Of the three classes of catalysts surveyed, metallocenes clearly offers new avenues to explore

in oligomerization. By employing the appropriate metallocene in the presence of

methylaluminoxane, higher alpha olefin dimers and trimers can be formed in a high yield at

ambient temperatures. Furthermore, a single oligomer can be formed with high selectivity

as opposed to a number of isomers. Of the metallocene catalysts investigated Cp2ZrC12 was

the most active. It should further be possible to tailor reaction conditions in order to form

dimers or trimers exclusively.

The choice of which catalyst to use and develop will depend on the intended application of

the oligomer. This will determine whether a single component oligomer is required or

whether a mixture of isomers is advantageous. At this stage the BF3/ROH catalyst system

is more economical and forms trimers in a higher yield than the metallocene catalysts. On

the other hand, the metallocene can form dimers with a 100% selectivity and in addition

oligomers consisting of only one component.

In the production of synthetic lubricants the formation of highly branched longer alpha

olefin (Cs to C12) oligomers by the BF 3/ROH catalyst is favorable. The oligomers of shorter

alpha olefins (C4 to C7), however, exhibits too much branching and thus the metallocene-

based catalysts present the opportunity to form less branched oligomers for application in

this field.

2.4 Conclusion

The synthesis of model oligomers represented a unique opportunity to relate specific

molecular structures with certain properties. Firstly, the analytical data of branched alkanes

are somewhat ambiguous. The mass spectra of these compounds give an indication of the

branching in the molecule, but molecules with more than one branch fragment easily,

therefore complicating structural determination. Comparison with the GC/MS data of

standard branched alkanes, particularly those structurally similar to alpha olefin oligomers,

should be a valuable tool in the structure determination of these compounds.

Furthermore, it was found possible to differentiate between diastereomers in the GC/MS

analyses of these types of compounds. The implication is that the composition of

Page 96: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

83

oligomerization reaction mixtures may be much less complex than appears at first. The

presence of these diastereomeric compounds with identical mass spectra should always be

considered in the analyses of complex mixtures.

A route to a model 1-pentene dimer and trimer, and possibly a model tetramer, was

developed. By using this methodology model 1-pentene oligomers can be prepared if

certain factors are kept in mind: the unreactivity of substrates, ease of cationic isomerization

and the separation of reaction products. The synthesis of other model oligomers with

different branching patterns would certainly be of value.

Metallocene-based oligomerization of higher alpha olefins presents opportunities for further

research. The reduction of the MAO/metallocene ratio is of great industrial importance and

so is the use of cationic metallocene catalysts. In addition, the use of bridged ligands will

facilitate steric control and thus the syntheses of optically active oligomers.

Metallocene catalysts have started a revolution in the polymer industry and also promise a

new era in the oligomerization of higher alpha olefins.

Page 97: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

84

CHAPTER 3

EXPERIMENTAL

3.1 General

Apparatus and standard procedure

All air- and moisture sensitive reactions were performed under an inert gas (N2 or Ar)

atmosphere with dry solvents in flamed-out glassware. Room temperature (rt) refers to ca.

20 - 25 °C

Solvents and reagents

Solvents were purified and distilled prior to use. Ether, tetrahydrofuran, pentane, hexane,

benzene and toluene were dried over sodium wire, with benzophenone as indicator.

Chloroform and dichloromethane were dried over phosphorous pentoxide. Alpha olefins

(1-pentene, 1-hexene and 1-octene) were left over activated molecular sieves for 24 hours,

distilled and stored over freshly activated molecular sieves. 1-Pentene was obtained from

Sasol with >99% purity. 1-Hexene with 97 - 98% purity (2 - 3% oxygenates) was also

obtained from Sasol and higher purity 1-hexene was purchased commercially. All other

reagents were purchased from commercial suppliers and used without further purification

unless stated otherwise.

Instrumentation

1) Gas chromatography

A Varian 3400 gas chromatograph equipped with a hydrogen flame ionization detector and

a variable-split glass capillary injector was used in this study. The carrier gas was nitrogen

with a linear flow rate of 2 ml.min -1 . The data system used in conjunction with the Varian

3400 was a Microsep Datamodule M741. A DB1 fused-silica capillary column

Page 98: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

85

(30m x 0.25 mm i.d.) coated with 100% methyl groups, was used. Samples were injected

into the gas chromatograph (20/1 split, 250 °C), the oven was kept at 50 °C for 5 minutes,

programmed to 250 °C at 10 °C.min- 1 and then held isothermally for 10 minutes. The

detector was operated at 300 °C. Retention times (R e) are given in minutes.

Gas chromatography/mass spectrometry

A Varian Saturn 3 GC/MS was used to collect GC, El-MS and CI-MS data of all organic

compounds (mass values m/z 40 •- 600). A DB5-MS fused-silicon capillary column

(30m x 0.25 mm i.d.) coated with 5% phenyl and 95% methyl groups was used. Helium

was the carrier gas with a flow rate of 1 ml.min -1 . Samples were injected into the gas

chromatograph (10/1 split, 260 °C) with an appropriate oven program; retention times (Re)

are given in minutes. The oven was held at 40 °C for 5 minutes and programmed to 250 °C

at 10 °C.min-1, 10 minutes isotherm, unless stated otherwise. The GC/MS transfer line was

maintained at 260 °C with a filament emission of 12 1.t.A for both EI and CI, an ionizing

voltage of 50 eV and a source temperature of 170 °C. • Chemical ionization was accomplished

using methane as reagent gas.

NMR spectroscopy

1H- and 13C NMR spectra were recorded by means of a Varian VXR 200 (200 MHz) or Varian

Gemini 2000 (300 MHz) spectrometer. Tetramethylsilane and deuterochloroform were used

as internal standards for 1H- and 13C-spectra, respectively. Unless otherwise stated, the

spectra were recorded in CDCb. Chemical shifts are given in parts per million (5) downfield

from the signal of tetramethylsilane.

The following abbreviations are used throughout:

s = singlet ct = quartet

d = doublet

m = multiplet

t = triplet

br = broadened

Page 99: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

86

4) Mass spectroscopy

Mass spectra of organometallic catalysts were recorded on a Finnigan-MAT 8200 mass

spectrometer (70 eV).

3.2 Standard chemical methods

Standard Grignard solutions

Magnesium (1.05 eq.) was placed in a flamed-out flask under N2, together with dry diethyl

ether (5 eq.), a crystal of iodine and a third of the alkyl halide (0.33 eq.). The reaction started

refluxing at room temperature and the remainder of the alkyl halide (0.66 eq.) in diethyl

ether was added at such a rate that the ether refluxed smoothly. The reaction was refluxed

for 30 minutes, cooled to room temperature and then standardized.

Titration of organometallic reagents

The organometallic reagent was added dropwise to a solution of 2,2'-bipyridyl (10 mg) in

dry THE (1 ml) under N2 until a color change was observed (dark red). Another 1.00 ml

aliquot portion of the organometallic reagent was added to the solution after which the

resulting solution was titrated with dry s-BuOH until the color of the charge transfer

complex was discharged (dark red —> milky yellow). This method was then repeated by

adding 1.00 ml of the organometallic followed by titration with s-BuOH. The molarity was

determined from the following equations :

col (RM) Molarity (RM) =

ml (RM) used

mmol (RM) = mmol (s - BuOH)

density (s - BuOH) x ml (s - BuOH)

MW (s - BuOH)

where RM = RMgX or RLi

At least three consistent measurements were made to ensure accuracy.

Page 100: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

87

Gilman test for free RLi or RMgX

An equal volume of a 1% benzene solution of Michler's ketone

[4,4'-bis(dimethylamino)benzophenone] was added to an ethereal solution of the

organometallic (e.g. 2.5 ml sample) and the mixture was allowed to warm to room

temperature. Water (1.5 ml) was then added and the mixture was stirred vigorously for 5

minutes. A 0.2% solution of 12 in glacial acetic acid was added dropwise and the

development of a characteristic greenish-blue aqueous layer confirmed the presence of

RMgX or RLi in the original solution.

Addition of a Grignard reagent to an epoxide in the presence of catalytic Cu(I) salts

The ethereal solution of the Grignard reagent (1 eq.) was added dropwise to a stirred slurry

of CuCN (0.1 eq.) in THF (1 m1/0.1 g CuCN) under N2 at -30 °C and the mixture was stirred

at this temperature for 30 minutes. The reaction mixture was then cooled to - 50 °C, the

epoxide (1 eq.) in THF was added dropwise and the reaction mixture was allowed to warm

to room temperature over a period of 5 hours. The reaction was quenched with a

NH4C1/ NH3 solution (pH 8 - 9) and filtered through Celite ® . The filtrate was extracted with

ether (3x) and the combined organic layers washed with a saturated solution of NaCl (aq)

and dried over MgSO4. After filtration the solvent was removed in vacuo and the crude

product usually purified by distillation.

Hydrogenation of oligomers

The oligomers were hydrogenated over 1% Pd/C (10% m/m) with H2 (40 - 50 bar) for 2

hours at room temperature. The resulting liquid and catalyst were separated by filtration

through a sintered glass funnel. The degree of unsaturation was determined by halogen

addition.

Determination of the degree of unsaturation

A portion of a 0.3 N bromine solution (10 - 15 ml) was placed in a 500 ml flask fitted with a

ground glass opening. The flask was closed immediately and cooled in ice-water. The

sample (0.2 - 0.3 g = VV) was added, the closed flask shaken vigorously and placed in a cold

Page 101: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

88

bath for 2 minutes. 20 ml of a 10% KI solution and 75 ml of H2O was added and the iodine

titrated with a 0.1 N thiosulfate solution (standardized) in the presence of a starch indicator.

A blank control was titrated at the same time. After the first titration, 5 ml of a 2% K103

solution was added and the iodine liberated was again titrated with thiosulfate.

b — a — 2c Bromine number = x C x 7.99

a = nil of thiosulfate needed for titrating the sample;

b = ml of thiosulfate needed in the control;

c = ml of thiosulfate needed after adding iodate to the titrated

solution;

W= sample weight in g;

C = concentration of thiosulfate solution.

3.3 Preparation of model oligomers

Epoxidations of 1-pentene with MMPP, 93m UHP94,97 and dirriethyldioxirane 98,100 were carried

out according to standard procedure.

1-Iodo-2-pentanol (40)

1 M NaHSO3 (96 ml) was added dropwise over a period of about 4 hours to a solution of 1-

pentene (40 mmol, 4.4 ml) and HI04.2H20 (48 mmol, 10.94 g) in CH3CN (80 ml) and H2O (24

ml) at 0 °C and the mixture was stirred at room temperature for another 8 hours. • The

resulting solution was extracted with diethyl ether (150 ml x 3) and the combined organic

layers washed with aqueous Na2SO3 and dried over MgSO 4. After filtration, the solvent was

evaporated in vacuo to leave the crude product as a yellow-brown liquid consisting of two

compounds (by GC) which was used without further purification (7.28 g, 85%). These

products exhibited identical EI-MS data as determined by GC-MS.

Rt (GC) 15.06 (50%) and 16.39 (50%)

EI-MS:

m/z 213 (M+ - 1, 0.4%), 196 (41), 142 (39), 127 (16), 87 (100), 69 (38), 55 (47), 45 (68),

43 (74), 41 (84)

Page 102: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

89

NMR (CDC13, 200 MHz):

814 0.82 (3 H, t, J4,5 6.8 Hz, 5-Me), 1.27 - 1.44 (4 H, m, 3-CH2, 4-CH2),

2.33 (1 H, br s, OH), 3.12 (1 H, dd, J1,2 6.6 Hz, JAB 10.2 Hz, lb-H),

3.36 (1 H, dd, J1,2 3.7 Hz, JAS 10.2 Hz, la-H), 3.40 - 3.43 (1 H, m, 2-H);

8c 13.20 (C-5), 7.83 (C-4), 20.09 (C-1), 35.21 (C-3), 68.11 (C-2).

1-Bromo-2-pentanol (42)

NaBrO3 (48 mmol, 7.24 g) was dissolved in water (24 ml) and the solution was adjusted to a

pH of 1 with 2 M H2SO4. The resulting solution was cooled to 0 °C and 1-pentene (40 mmol,

4.4 ml) and CH3CN (80 ml) were added; 1 M NaHSO 3 (96 ml) was added dropwise over a

period of 4 hours and the reaction was stirred at room temperature for a further 8 hours.

The reaction mixture was extracted with diethyl ether (150 ml x 3), the combined organic

layers washed with aqueous NaSO3 and dried over MgSO4. After filtration the solvent was

evaporated in vacuo to leave the crude product as a dark brown liquid consisting of 3

products (by GC) which was used without further purification (5.35 g, 80%). These

products exhibited identical EI-MS and CI-MS data as determined by GC/MS.

Rt (GC) 13.9 (14%), 14.3 (29%) and 15.6 (57%).

CI-MS:

m/z 167 (M+ + 1, absent), 166 (M+ - 1, absent), 150 (18%) 148 (20), 81 (1.4) 79 (1.6),

69 (100)

NMR (CDC13, 200 MHz):

8H 0.90 (3 H, t, J1,2 6 Hz, 5-Me), 1.34 - 1.57 (4 H,,m, 3-CH2, 4-CH2),

2.20 (1 H, s, disappears on deuteration, OH),

3.38 (1 H, dd, J1,2 7.1 Hz, JAB 10.3 Hz, lb-H),

3.53 (1 H, dd, J1,2 3.3 Hz, JAB 10.3 Hz, la-H),

3.74- 3.76 (1 H, m, verified by spin-spin decoupling, 2-H);

8c 13.30 (C-5), 18.73 (C-4), 37.07 (C-3), 40.52 (C-1), 70.72 (C-2).

1,2-Epoxypentane (35)

The reaction was carried out in vacuum distillation apparatus fitted with a liquid nitrogen

trap to facilitate collection of the product. The neat halohydrin 40 (0.1 mol, 24.40 g) or 42

Page 103: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

90

(0.1 mol, 16.71 g) was added dropwise to stirred, finely powdered NaOH (1 mol, 40 g) under

a vacuum of ± 30 mm Hg. The reaction was exothermic (temperatures of up to 60 °C was

noted) and the epoxide 35 was distilled and trapped as it formed. The resultant clear liquid

(6.46 - 6.89 g, 75 - 80%) was dried over MgSO4, filtered and stored under N2 over activated

molecular sieves. GC analysis indicated that 35 was isolated with > 97% purity and no

further purification was undertaken.

Rt (GC) 6.9

CI-MS:

m/z 85 (M+ - 1, 2%), 71 (100), 68 (16)

NMR (pyridine-ds, 300 MHz):

5H 0.94 (3 H, tt, J3,5 0.9 Hz, 14,5 7.5 Hz, 5-Me),

1.42 - 1.49 (4 H, m, 3-CH2, 4-CH 2),

2.38 (1 H, dd, J1,2 2.7 Hz, JAB 5.4 Hz, lb-H),

2.64 (1 H, dd, J1,2 4.2 Hz, JAB 5.4 Hz, la-H), 2.80 - 2.84 (1 H, m, 2-H);

Sc 14.03 (C-5), 19.63 (C-4), 34.90 (C-3), 46.21 (C-1), 51.54 (C-2).

Butylmagnesium bromide

Butylmagnesium bromide was prepared by standard procedure from Mg (0.105 mol, 2.55 g),

a crystal of iodine, 1-bromobutane (0.1 mol, 10.5 ml) and dry diethyl ether (100m1).

[BuMgBr] = 0.80 mol.dm-3

2-Heptanol (45)

2-Heptanol (1.15 g, 99%) was prepared as a dark yellow liquid by standard addition of

ethereal butylmagnesium bromide (0.01 mol, 12.5 ml) to propylene oxide (0.01 mol, 0.70 ml)

in the presence of CuCN (0.001 mol, 0.09 g) with THE (2 ml) as solvent.

Rt (GC): 12.25

EI-MS:

m/z 116 (M+ - 1, 14%), 98 (95), 97 (100), 83 (76), 70 (43), 69 (34)

NMR (CDC13, 200 MHz):

SH 0.83 (3 H, br t,16,7 7.3 Hz, 7-Me), 1.11 (3 H, d, 11,2 6.18 Hz, 1-Me),

Page 104: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

91

1.23 - 1.36 (8 H, m, 4 x CH2), 2.17 (1 H, br s, OH), 3.71 (1 H, m, 2-H).

4-Nonanol

4-Nonanol (1.37 g, 95%) was prepared as a light yellow liquid by standard addition of

butylmagnesium bromide (0.01 mol, 12.5 ml) to 1,2-epoxypentane (0.01 mol, 0.86 g) in the

presence of CuCN (0.001 mol, 0.09 g) with THF (2 ml) as solvent.

Rt (GC): 16.5

EI-MS:

m/z 143 (M+ - 1, 7%), 126 (9), 125 (15), 111 (8), 101 (22), 97 (14), 85 (55), 83 (100), 71 (51),

69 (26)

NMR (CDC13, 200 MHz):

51-1 0.86 and 0.90 (6 H, 2 x t, 1-Me, 9-Me), 1.20 - 1.42 (12 H, m, 6 x CH2),

2.82 (1 H, s, OH), 3.72 (1 H, m, 4-H);

5c 13.99 (C-9), 14.06 (C-1), 18.78 (C-2), 22.61 (C-8), 25.28 (C-6), 31.89 (C-7), 37.29 (C-5),

39.49 (C-3), 6 72.12 (C-4).

2-Pentylmagnesium bromide (43)

Grignard reagent 43 was prepared by standard procedure from Mg (0.525 mol, 12.76 g), a

crystal of iodine, 2-bromopentane (0.5 mol, 63 ml) and dry diethyl ether (500 ml).

[2-Pentylmagnesium bromide] = 0.75 mol.dm- 3

6-Methyl-4-nonanol (44)

a) By copper(I) catalysed Grignard addition to epoxide 35

Alcohol 44 (22.38 g, 70%) was prepared by standard addition of Grignard reagent 43

(0.20 mol, 270 ml) to epoxide 35 (0.20 mol, 17.23 g) in the presence of CuCN (0.02 mo1,1.79 g)

in THF (40 ml). The product was purified by Kugelxohr distillation (70 °C, 0.2 mm Hg) and

a mixture of 2 diastereomers (1 : 1) was isolated. The EI-MS and CI-MS data of these two

compounds as determined by GC/ MS were identical.

Page 105: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

92

By addition of a lower order cuprate of 43 to epoxide 35

CuCN (0.01 mol, 0.9 g) was suspended in 50 ml of dry THF, cooled to -78 °C and Grignard

reagent 43 (0.02 mol, 26.65 ml) was added. The mixture was stirred at this temperature until

a negative Gilman test resulted (± 15 minutes) and then epoxide 35 (0.01 mol, 0.86 g) in THF

(5 ml) was added dropwise over 30 minutes. Stirring was continued at -78 °C for 1 hour and

then aqueous NH4C1/NH3 (pH 8 - .9) was added. Similar work-up as before afforded

alcohol 44 as a mixture of two diastereomers (1 : 1) in a yield of 72% (1.14 g).

By addition of a higher order cyanocuprate of 43 to epoxide 35

CuCN (0.01 mol, 0.9 g) was suspended in 50 ml of dry THF, cooled to -78 °C and Grignard

reagent 43 (0.02 mol, 26.65 ml) was added. The reaction was stirred at this temperature until

a negative Gilman test resulted (± 15 minutes). The mixture was then allowed to warm until

all of the CuCN was dissolved (ca. -10 °C), thereby forming the higher order cuprate. The

reaction was again cooled to -78 °C and the epoxide'35 (0.01 mol, 0.86 g) in THF (5 ml) was

added dropwise over 30 minutes. Stirring was continued at -78 °C for 1 hour and the same

work-up as before gave the 1 : 1 mixture of diastereomers of 44 (1.2 g, 75%).

Rt (GC): 17.48 and 17.66

EI-MS:

m/z 158 (M+, 0.8%), 157 (4), 156 (9), 140 (11), 139 (17), 138 (15), 115 (18), 114 (14), 98 (24),

97 (99), 96 (36), 85 (100), 84 (31), 83 (34), 70 (75), 69 (24), 68 (26)

NMR (CDC13, 200 MHz):

OFT 0.83 - 0.94 (9 H, m, 3 x Me), 1.36 (10 H, m, 5 x CH2), 1.67 (1 H, m, 6-H),

2.5 (1 H, br s, OH), 3.65 - 3.71 (1 H, m, 4-H);

8c 14.11 (C-1), 14.31 and 14.36 (C-9), 19.26 and 20.37 (C-7'), 18.37 and 18.85 (C-2),

19.95 and 20.03 (C-8), 28.97 and 29.41 (C-6), 38.99 and 40.02 (C-7),

40.22 and 40.59 (C-3), 45.07 and 45.36 (C-5), 69.42 and 69.81 (C-4).

Multiplicities were confirmed using a DEPT experiment.

Page 106: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

93

4-Bromo-6-methylnonane (47)

PBr3 (0.033 mol, 0.32 ml) was added dropwise to the neat alcohol 44 (0.01 mol, 1.60 g) kept

between -5 °C and +5 °C. After the addition was completed, the reaction mixture was

allowed to warm to room temperature and to proceed for eight hours. The reaction was

shaken with one-third its volume of concentrated sulfuric acid and the acid layer was drawn

off and discarded. The residual organic phase was mixed with an equal volume of 50%

Me0H and aqueous ammonia was added with intermittent shaking until the solution was

alkaline to phenolphtalein. The organic layer was separated and washed once with an equal

volume of 50% Me0H, then dissolved in diethyl ether, dried over MgSO4, filtered and the

solvent removed in vacuo. The crude product was purified by Kugelrohr distillation (120 °C,

0.2 mm Hg) to yield a mixture of three compounds as a clear oil (1.77 g, 80%). These

compounds exhibited identical EI-MS data as determined by GC/ MS.

Rt (GC/ MS): 14.64 (45%), 14.71 (45%) and 15.16 (10%)

EI-MS:

m/z 222 (M+ + 1, absent), 220 (M+ - 1, absent), 140 (8%), 98 (18), 97 (25), 96 (9), 85 (100),

84 (22), 83 (18), 71 (79), 70 (19), 69 (24)

NMR (CDC13, 200 MHz):

8H 0.83 - 0.93 (9 H, m, 3 x Me), 1.17 - 1.92 (11 H,. m, 5 x CH2, 6-H), 4.09 (1 H, m, 4-H);

8c 13.47 (C-1), 14.25 and 14.30 (C-9), 18.67 and 19.73 (C-7'), 19.93 (C-2),

20.70 and 20.79 (C-8), 30.85 (C-6), 38.13 and 39.66 (C-7), 41.21 and 42.08 (C-3),

46.51 and 46.99 (C-5), 56.50 and 56.89 (C-4). Multiplicities were confirmed using a

DEPT experiment.

6-Methylnonyl-4-magnesium bromide (48)

Alkyl bromide 47 (0.03 mol, 6.63 g) dissolved in dry ether (30 ml) was added to magnesium

turnings (0.07 mol, 1.69 g) under N2 and maintained at a gentle reflux. 1,2-Dibromoethane

(0.03 mol; 2.16 ml) in dry ether (30 ml) was added dropwise over a period of 10 hours. The

mixture was then refluxed for another 2 hours, allowed to cool to room temperature and

standardized.

[6-Methylnonyl-4-magnesium bromide] = 0.28 mol.dm-3

Page 107: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

94

8-Methyl-6-propyl-4-undecanol (50)

Alcohol 50 (2.40 g, 70%) was prepared as a viscous yellow oil by standard addition of

Grignard reagent 49 (0.015 mol, 53.6 ml) to epoxide 35 (0.015 mol, 1.30 g) in the presence of

CuCN (0.0015 mol, 0.13 g) in THE (3 ml). The product was purified by vacuum distillation

(130 °C, 0.15 mm Hg) and a mixture of 5 products was isolated. These products exhibited

identical EI-MS data as determined by. GC/MS.

Rt (GC/MS): 11.63 (24%), 11.65 (24%), 11.71 (24%), 11.74 (24%)and 12.01 (4%)

EI-MS:

m/z 228 (M+, absent), 209 (0.8%), 140 (5), 139 (16), 138 (6), 111 (7), 110 (8), 98 (15), 97 (34),

96 (54), 85 (48), 84 (30), 83 (17), 71 (90), 70 (46), 69 (25), 57 (94), 56 (27), 55 (39),

43 (100), 42 (19), 41 (74)

NMR (CDC13, 200 MHz):

8H 0.78 - 0.89 (12 H, m, 4 x Me), 1.23 (16 H, m, 8 x CH 2), 1.63 (2 H, m, 6-H, 8-H),

2.33 (1 H, s, OH), 4.52 (1 H, m, 4-H);

Sc 14.13, 14.40 and 14.48 (br, C-1, C-9' and C-11), 19.81, 19.91, 20.00 and 20.35 (C-9"),

20.07, 20.13 and 20.21 (C-2), 21.09, 21.17 and 21.24 (C-10),

30.14, 30.22, 30.27 and 30.34 32.56, 32.71, 32.97 and 33.00 (C-8'),

33.21, 33.39, 33.42 and 33.66 (C-7'), 36.10, 36.22, 36.29 and 36.34 (C-6),

37.88, 38.36, 38.40 and 38.62 38.64, 38.87 and 38.98 (C-7),

39.66, 39.78 and 39.87 (C-5), 40.11, 40.15, 40.27 and 40.41 (C-3), 67.60 (br, C-4).

Multiplicities confirmed using a DEPT experiment.

Pentyllithium

Lithium shot (0.25 mol, 1.74 g) was placed in a flame-dried flask under an Ar atmosphere

and washed with dry hexane (5 x 8 ml) followed by addition of 6.5 ml hexane. The slurry

was cooled to -10 °C and 1-bromopentane (0.10 mol, 12.40 ml) dissolved in 12.5 ml hexane

was added slowly over a period of 40 minutes while the temperature was maintained at

-10 °C. After the addition was completed the reaction was warmed to room temperature

and stirring was continued for 2 hours. The slurry became light purple in color and

afforded a clear solution after filtration under. Ar via a cannula through a sintered glass

funnel.

Page 108: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

95

[C5H11Li] = 3.20 mol.dm-3

4-Methyl-6-propylundecane (32)

CuCN (0.01 ml, 0.90g) was stirred in dry THF (25 ml) and cooled to -78 °C. Pentyllithium

(0.02 mol, 3.2 ml) was added and the reaction was stirred until a negative Gilman test

resulted (± 15 minutes). The reaction mixture was allowed to warm to -10 °C to effect

complete solution of the CuCN. The reaction mixture was cooled to -78 °C and the alkyl

bromide 47 (0.01 mol, 2.2 g) in THF (5 ml) was added dropwise, after which the reaction was

allowed to proceed for 1 hour and then quenched with aqueous NH4C1/ NH3 (pH 8 - 9). The

resulting suspension was filtered through Celite ®, the filtrate extracted with ether (20 ml x 3)

and washed with Na2SO3. The solution was dried over MgSO4, filtered and the solvent

removed in vacuo to yield alkane 32 as a clear oil (1.06 g, 50%). Three products with

identical EI-MS data were distinguished by GC/MS.

Rt (GC/MS): 17.50 (48%), 17.58 (48%) and 17.98 (4%)

EI-MS:

m/z 211 (M+ - 1, 0.4%), 210 (0.6), 209 (8), 167 (6), 153 (2), 139 (9), 125 (10), 111 (16), 97 (29),

85 (80), 71 (100)

NMR (CDC13, 300 MHz):

514 0.83 - 0.89 (12 H, m, 4 x Me), 1.22 (18 H, m, 9 x CH2), 1.29 (2 H, m, 4-H, 8-H);

5c 14.06 (3 C, C-1, C-9; and C-11), 14.49 (C-5'), 19.77 (C-2), 22.68 (2 C, C-10, C-8'),

26.31 and 26.62 (C-8), 29.66 and 29.80 (C-9), 31.93 (C-7), 32.36 (C-7'), 32.45 (br, C-4),

33.61 and 33.66 (C-3), 36.08 (C-5), 37.15 and 37.36 (C-6). Multiplicities confirmed

using a DEPT experiment.

3.4 Oligomerization reactions

BF3/n-BuOH oligomerization of 1-pentene

1-Pentene (0.34 mol, 37 ml), heptane (45 ml) and n-BuOH (0.027 mol, 0.25 ml) were placed in

a three-necked round-bottomed flask fitted with a thermometer, gas inlet, reflux condenser

and two-way stopcock to a gas outlet or to a manometer. BF3 (g) was rapidly bubbled

Page 109: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

96

through the solution for 5 minutes. The gas flow was adjusted to maintain a constant

pressure in the closed system. The temperature rised to 37 °C, after 30 minutes the gas flow

was stopped and the reaction quenched by addition of 25% NaOH (aq). The organic layer

was diluted with diethyl ether, washed with H2O (30 ml x 3), dried over MgSO4 and filtered.

The solvents and unreacted 1-pentene were removed in vacuo to yield the oligomeric

products as a yellow-brown oil (total conversion > 90%).

Bp (0.2 mm Hg): Dimer (80 - 85 °C), trimer (85 - 115 °C), tetramer (125 - 145 °C)

Rt (GC): Dimer (11.7 - 13.1, 15%), trimer (23.2 - 28.0, 47%), tetramer (32.8 - 37.4, 23%),

pentamer (40.6 - 46.8, 15%). The GC oven was programmed from 50 °C to 250 °C at

a rate of 5 °C.min- 1, 20 minutes isotherm.

EI-MS:

The complete resolution of any single product was impossible due to the large number of

isomers in each oligomer fraction. The MS data at the maximum of the broad peaks

representing the different oligomers are quoted.

Dimer

m/z 140 (M+, 1.6%), 139 (2.4), 124 (1.6), 123 (2.4), 112 (11), 111 (5), 110 (2.4), 98 (16),

97 (100), 96 (12), 85 (38), 84 (41), 83 (74), 71 (68), 70 (91), 69 (68)

Trimer

m/z 210 (M+, 1.6%), 209 (15), 153 (12), 140 (15), 139 (25), 126 (6), 125 (19), 112 (13),

111 (35), 110 (10), 98 (17), 97 (56), 96 (17), 85 (16), 84 (23), 83 (81), 82 (18), 81 (20),

71 (21), 70 (34), 69 (100), 68 (16), 67 (28)

Tetramer

m/z 280 (M+, 10%), 266 (4), 252 (4), 237 (10), 224 (8), 210 (8), 196 (10), 181 (17), 168 (4),

154 (26), 140 (35), 125 (43), 111 (72), 96 (44), 83 (68), 69 (100)

Hydrogenation of the oligomers obtained from 1 -pentene with BF3/n -BuOH as catalyst

The 1-pentene oligomers (10 g) obtained with BF3/ri-BuOH as catalyst were hydrogenated

over Pd/C (0.1 g) according to standard procedure. The bromine number of the product

was determined to be ca. 0 and thus complete hydrogenation took place.

Rt (GC): Dimer (11.6 - 13.96), trimer (23.8 - 28.2), tetramer (31.6 - 39.2),

Page 110: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

97

pentamer (39.9 - 48.0). The GC oven was programmed from 50 °C to 250 °C at a

rate of 5 °C.min- 1, 20 minutes isotherm.

Hydrogenation the 1 -pentene trimer obtained with of BF3/n -BuOH as catalyst

The 1-pentene trimer (20 g) obtained with BF3/n-BuOH as catalysts was hydrogenated over

Pd/C (0.2 g) according to standard procedure. The bromine number of the product was

determined to be ca. 0 and thus complete hydrogenation took place.

Rt (GC/MS): 15.1 - 23.5

Viscosity (Ostwald viscometer, cStokes): 30 °C (22), 35 °C (19.1), 40 °C (17.3)

Radical oligomerization of 1 -pentene with organic peroxide initiators

The autoclave was charged with 1-pentene (3 eq.), heptane (1 eq.) and the specified mol

percentage of the organic peroxide. The system was closed, set to the required temperature

and reaction time and pressure was measured from temperature equilibration. Afterwards

the reaction vessel was allowed to cool to room temperature and analyzed without further

work-up.

Benzoyl peroxide initiated oligomerization of 1-pentene

Reagents :

1-Pentene (0.14 mol, 15 ml)

Heptane (10 ml)

Benzoyl peroxide 2% (0.003 mol, 0.94 g); 5% (0.007 mol, 2.35 g), 10% (0.014 mol,

4.69 g), 15% (0.020 mol, 7.04 g).

Dicumyl peroxide initiated oligomerization of 1-pentene

Reagents :

1-Pentene (0.14 mol, 15 ml)

Heptane (10 ml)

Page 111: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

98

Dicumyl peroxide 2% (0.003 mol, 0.74 g); 5% (0.007 mol, 1.85 g), 10% (0.014 mol,

3.7 g), 15% (0.020 mol, 5.55 g).

c) tert-Butyl peroxide initiated oligomerization of 1-pentene

Reagents :

1-Pentene (0.14 mol, 15 ml)

Heptane (10 ml)

Di-tert-butyl peroxide 2% (0.003 mol, 0.6 ml); 5% (0.007 mol, 1.45 ml), 10% (0.014 mol,

2.9 ml), 15% (0.020 mol, 4.35 ml).

Bis(tetramethylcyclopentadienyl)titanium(IV) dichloride (66)

Tetramethylcyclopentadienyl (0.01 mol, 1.22 g) was dissolved in dry THF (25 ml), cooled to

-78 °C and n-BuLi (1.47 mol.dm-3, 0.01 mol, 6.80 ml) was added dropwise over a period of 30

minutes. The solution was stirred for another 30 minutes at this temperature and then

allowed to warm to room temperature to form a light yellow suspension of the anion. Dry

THF (20 ml) was carefully added to TiC14 (0.005 mol, 0.55 ml) under an Ar atmosphere in

order to generate the THF adduct. The anion was slowly added to this mixture via a

cannula and the resulting solution stirred for 1 hour. The solvent was evaporated and the

metallocene recrystallized from chloroform/pentane as deep red crystals (1.80 g, 50 %).

EI-MS

m/z 362 (M+, 10 %) , 360 (15), 326 325 (5), 324 (13), 241 (16), 239 (24), 205 (5), 204 (7),

203 (12), 202 (5), 201 199 121 (100), 105 (25), 93 (12), 91 (16), 79 (11), 77 (9)

NMR (CDCb, 300 MHz):

SH 1.89 (12 H, s, 4 x Me), 2.03 (12 H, s, 4 x Me), 5.97 (2 H, s, 2 x CH);

Sc 13.02 (4 C, 2-Me, 5-Me), 14.81 (4 C, 3-Me, 4-Me), 113.09 (2 C, C-1),

125.16 (4 C, C-2, C-5), 135.91 (4 C, C-3, C-4).

Page 112: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

99

Alpha olefin oligomerization with the metallocene/MAO catalyst system

Methylaluminoxane was obtained as a 10% solution in toluene. Although it was thought

that the MAO could decompose with time, the concentration of the MAO was not estimated

and the molar mass of the [Al-Me-0] monomer was used in calculations.

The metallocene (0.05 mmol) catalyst was preactivated with the methylaluminoxane

solution (0.0075 mol, 0.5 ml) for 15 minutes at room temperature in a 100 ml flamed-out

Schlenk flask under an Ar atmosphere. The alpha olefin was added (0.019 mol), the flask

tightly sealed and the reaction mixture stirred in an oil bath at 40 °C. After the specified

reaction time, the reaction mixture was allowed to cool to room temperature and quenched

carefully with a 10% HC1/ Me0H solution. The resulting solution was extracted with

toluene (15 ml x 3), washed with water and dried over MgSO4. After filtration, the solvent

was removed in vacuo to yield the oligomeric products as a colored oil (depending on the

color of the metallocene).

Metallocenes

Cp2TiC12 (12.5 mg)

Cp2ZrC12 (13.5 mg)

CpTiC13 (11 mg)

(C9H13)2TiC12 (17.9 mg)

Alpha olefins

1-Pentene (1.33g, 2.1 ml)

1-Hexene (1.60 g, 2.4 ml)

1-0c tene (2.13 g, 3 ml)

2-Propyl-1-heptene (60)

The 1-pentene dimer 60 was prepared by standard oligomerization of 1-pentene with

Cp2TiC12/ MAO.

Page 113: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

100

Itt (GC): 10.22

EI-MS:

m/z 140 (M+, 8%), 139 (32), 137 (3), 111 (15), 97 (41), 96 (100), 95 (42), 85 (42), 84 (28),

83 (44), 69 (40)

NMR (CDC13, 300 MHz):

51-1 0.87 (3 H, t, 16,7 6 Hz, 7-Me), 0.88 (3 H, t, J4•5 , 7.5 Hz, 5'-Me),

1.26 - 1.47 (8 H, m, 4 x CH2), 1.98 (4 H, br t, J3',4• 13,4 7.8 Hz, 3-CH2, 3'-CH2),

4.68 (2 H, s, 1-CH2)

5c 13.8 (q, C-7), 14.0 (q, C-5'), 20.84 (t, C-4'), 22.55 (t, C-6), 27.45 (t, C-4), 31.64 (t, C-5),

35.96 (t, C-3), 38.18 (t, C-3'), 108.5 (t, C-1), 150.2 (s, C-2).

2,4-Propy1-1-nonene (63)

The 1-pentene trimer 63 was prepared by standard oligomerization of 1-pentene with

Cp2TiC12/ MAO.

Rt (GC): 16.83

EI-MS:

m/z 210 (M+, 0.4%), 209 (2.4), 208 (2.8), 165 (14), 138 (13), 123 (26), 110 (36), 97 (70),

96 (61), 84 (81), 83 (100), 82 (63), 70 (44), 69 (62), 68 (42)

NMR (CDC13, 300 MHz):

OH 0.80 - 0.91 (9 H, m, 3 x Me), 1.14 - 1.30 (14 H, m, 7 x CH2), 1.39 - 1.47 (1 H, m, 441),

1.91 - 1.96 (4 H, m, 3-CH2, 3'-CH2), 4.65 - 4.71 (2 H, 2 x d, Jcis 15 Hz, 1-CH2);

8c 13.84 (q, C-9), 14.06 (q, C-7"), 14.43 (q, C-5'), 19.64 (t, C-4'), 20.77 (t, C-6"),

22.67 (t, C-8), 26.15 (t, C-6), 32.30 (t, C-7), 33.35 (t, C-5), 34.98 (d, C-4), 35.80 (t, C-5"),

37.81 (t ,C-3'), 40.27 (t, C-3), 110.20 (t, C-1), 148.91 (s, C-2).

Viscosity (Ostwald viscometer, cStokes): 30 °C (63.1), 35 °C (44.6), 40 °C (38.3)

2-Butyl-1-octene (61)

The 1-hexene dimer 61 was prepared by standard oligomerization of 1-hexene with

Cp2ZrC12/ MAO.

Rt (GC): 14.16

Page 114: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

101

EI-MS:

m/z 168 (M+, 12%), 167 (28), 166 (86), 124 (24), 110 (100), 93 (57), 81 (71), 68 (44), 67 (45)

NMR (CDC13, 300 MHz):

5H 0.90 (3 H, t, 17,86.6 Hz, 8-Me), 0.92 (3 H, t, J5',6' 6.6 Hz, 6'-Me),

1.20 - 1.46 (12 H, m, 6 x CH2), 2.01 (4 H, br t, '74 J3,4 7.6 Hz, 3-CH2, 3'-CH2),

4.70 (2 H, 2 x d, 1 Hz, 1-CH2);

5c 13.98 (q, C-8), 14.08 (q, C -6'),-22.55 (t, C-7), 22.70 (t, C-5'), 27.88 (t, C-4),

29.19 (t, C-4'), 30.12 (t, C-5), 31.86 (t ,C-6), 35.84 (t, C-3), 36.16 (t, C-3'),

108.38 (t, C-1), 150.31 (s, C-2).

2-Hexyl-1-decene (62)

The 1-octene dimer 62 was prepared by standard oligomerization of 1-octene with

Cp2ZrC12/ MAO.

Rt (GC): 19.54

EI-MS:

m/z 224 (M+, 14%), 223 (12), 222 190, 193 (2), 169 (4), 154 (6), 138 (48), 125 (16), 111 (100),

97 (42), 83 (59), 69 (66)

NMR (CDC13, 300 MHz):

8H 0.85 - 0.89 (6 H, m, 2 x Me), 1.26 - 1.42 (20 H, m, 10 x CH2),

1.98 (4 H, br t, J3,4 7.8 Hz, 3-CH2, 3'-CH2), 4.67 (2 H, 2 x d, J ets 1.2 Hz, 1-CH2);

5c 14.03 (q, 2 C, C-10, C-8'), 22.61 (t, C-9), 22.64 (t, C-7'), 27.77 (t, C-4), 27.80 (t, C-4'),

29.10 (t, C-7), 29.28 (t, C-6), 29.44 (t, C-5'), 29.50 (t, C-5), 31.77 (t, C-8), 31.88 (t, C-6'),

36.07 (t, 2 C, C-3, C-3'), 108.35 (t, C-1), 150.52 (s, C-2).

Page 115: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

102

REFERENCES

F. Fischer and H. Tropsch, Ger. Pat. 484377/1929; Brennstoff-Chem., 1923, 4, 276;

Brennstoff-Chem., 1926, 7, 97; Chem. Ber., 1926, 59, 830, 832, 923.

W. Theron, Sasol News, May 1994.

M. Waddacor, Chemical Processing SA, 1994, 1 (10), 2.

G.R. Lappin and J.D. Sauer, Alpha Olefins Application Handbook, Marcel Dekker, Inc.,

New York, 1989, p.353.

J. Skupenska, Chem. Rev., 1991, 91, 613.

R.L. Shubkin, Synthetic Lubricants and High Performance Functional Fluids, Marcel

Dekker, Inc., New York, 1993, p.1.

R.E. Bolan, Synthetic Lubricant Base Stocks, SRI International, Report No. 125A, Sept.

1989.

K.-D. Hungenberg, J. Kerth, F. Langhauser, H.J. Muller and P. Muller, Angezv.

Makromol. Chem., 1995, 227, 159.

IUPAC Commision on Macromolecular Nomenclature, Pure Appl. Chem., 1979, 40, 479.

G.A. Olah and D. Meimar, in Kirk-Othmer Encyclopedia of Chemical Technology, eds.

M. Grayson and D Eckroth, John Wiley & Sons, New York, Third Edition, vol. 11,

p.269.

G. Olah and A. Molnar, Hydrocarbon Chemistry, John Wiley and Sons, Inc., New York,

1995.

A. Ravve, Principles of Polymer Chemistry, Plenum Press, New York, 1995.

F.W. Sullivan, Jr., V. Vorhees, A.W. Neeley and R.W. Shankland, Ind. Eng. Chem., 1931,

23, 604.

C.W. Montgomery, W.I. Gilbert and R.E. Kline, USP 2559984/1951, to Gulf Oil

Company.

J.A. Brennan, USP 3383291/1968, to Mobil Oil.

R.L. Shubkin, M.E. Kerkemeyer, D.K. Walters and J.V. Bullen, in Mechanical

Engineering, 1990, 80, 125.

J.A. Brennan, Ind. Eng. Chem. Prod. Res. Dev., 1980, 19 , 2.

R.L. Shubkin, ivi.S. Baylerian and A.R. Maier, Ind. Eng. Chem. Prod. Res. Dev., 1980, 19,

15.

Page 116: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

103

A. Onopchenko, B.L. Cupples and A.N. Kresge, Ind. Eng. Chem. Res. Dev., 1983, 22,

182.

G.L. Driscoll and S.J.G. Linkletter, AFWAL-TR-85066, Synthesis of Synthetic

Hydrocarbons via Alpha Olefins, Air Force Wright Aeronautical Laboratories,

May 1985.

H. Koch, Brennstoff-Chem, 1955, 36, 321.

B.L. Cupples, A.M. Kresge, A.Pnopchenko, J.P. Pellegrini, Jr., AFWAL-TR-81-4109,

Synthesis of Synthetic Hydrocarbons via Alpha Olefins, Gulf Research and

Development Company, Sept. 1981.

S.W. Benson, J. Chem. Educ., 1965, 42, 502.

A. Priola, G. Gozzelino and F. Ferrero, Polymer Bulletin, 1985, 13, 245.

F.H. Dressler and S. Vermaire, Makromol. Chem., Makromol. Symp., 1988, 13/14, 271.

A. Ranwell, Die Oligomerisasie van Sasol Alfa-Olefienfraksies, M.Sc. Thesis, Rand

Afrikaans University, 1994.

C.S. Sheppard and O.L. Mageli, in Kirk-Othmer Encyclopedia of Chemical Technology, eds.

M. Grayson and D Eckroth, John Wiley & Sons, New York, Third Edition, vol. 17, p.

27.

H. Ashijan, EP Application 88312436.4/1989, to Mobil Oil.

W. Keim, A. Behr and M. Roper, in Comprehensive Organometallic Chemistry, ed.

G. Wilkinson, Pergamon Press, Oxford, 1982, vol. 8, p. 372.

W. Kaminsky and R. Kranolowsky, in Inorganic Reactions and Methods, ed. A.D.

Norman, VCH Publishers, New York, 1991, vol. 16, p. 278.

W. Kelm, E. Hoffmann, R. Lodewick, M. Pauckert, G. Schmitt, J. Fleischhauer and U.

Meier, J. Mol. Catal., 1979, 6, 79.

V.Y. Kissin, J. Polym. Sci. Part A, Polym. Chem., 1989, 27, 605, 623.

G.W. Parshall and S.D. Ittel, Homogeneous Catalysts, Wiley-Interscience, New York,

1992.

W. Keim, A. Montreux and F. Petit, Industrial Applications of Homogeneous Catalysts,

Reidel, Dordrecht, 1988, p. 335.

W. Keim, Fundamental Research in Homogeneous Catalysts, Plenum, New York, 1984, 4,

p. 131.

W. Keim, New J. Chem., 1994, 18, 93.

G.A. Mills and J.A. Cusumano, Kirk-Othmer Encyclopedia of Chemical Technology, eds.

M. Grayson and D. Eckroth, John Wiley & Sons, New York, Third Edition, vol. 5, p. 17.

Page 117: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

104

K. Ziegler, E. Holzkamp, H. Breil and H. Martin, Angew. Chem., 1955, 67, 426.

G. Natta, P. Corradini, F. Danusso, E. Mantica, M. Moraglio, P. Pino and G. Mazzanti,

J. Am. Chem. Soc., 1955, 77, 1708.

G. Natta, P. Pino, G. Mazzanti and U. Gianni, J. Am. Chem. Soc., 1957, 79, 2975.

D.S. Breslow and N.R. Newburg, J. Am. Chem. Soc., 1957, 79, 5072.

B. Rieger, X. Mu, D.T. Mallin, M.D. Raush and J.C.W. Chen, Macromolecules, 1990, 23,

383; N.H. Cheng and J.A. Ewen, Macromol. Chem., 1989, 190, 1931; A. Grassi, A.

Zambelli, L. Resconi, E. Albizzti and R. Mazzocchi, Macromolecules, 1988, 21, 617; Y.V.

Kissin, T.E. Nowlin and R.I. Mink, Macromolecules, 1993, 26, 2151.

E.J. Arlman, J. Catal., 1964, 3, 89; P. Cossee, J. Catal., 1964, 3, 80; E.J. Allman and P.

Cossee, J. Catal., 1964, 3, 99; E.J. Allman, J. Catal., 1966, 5, 178.

H. Sinn and W. Kaminsky, Adv. Organomet. Chem., 1980, 18, 99.

H. Sinn, W. Kaminsky, H.J. Vollmer and R. Woldt, Angew. Chem., Int. Ed. Engl., 1980,

19, 390.

A.D. Horton, Trends in Polym. Sci., 1994, 2, 158.

A.K. Zetirova and A.E. Shilov, Proc. Acad. Sci. USSR, 1961, 136, 77.

J.J. Eisch, A.M. Piotrowski, S.K. Brownstein, E.J. Gabe and F.L. Lee, J. Am. Chem. Soc.,

1985, 107, 7219.

R.F. Jordan, C.S. Bajgur, R. Willet and B. Scott, J. Am. Chem. Soc., 1986, 108, 7410.

M.R. Mason, J. M. Smith, S.G. Bott and A.R. Barron, J. Am. Chem. Soc., 1993, 115, 4971.

D. Fischer and R. Miilhaupt, J. Organomet. Chem., 1991, 417, C7.

W. Kaminsky and R. Steiger, Polyhedron, 1988, 7, 2375; W. Kaminsky, A. Bark and R.

Steiger, J. Mol. Catal., 1992, 74, 109.

A.R. Siedle and R.A. Newmark, J. Organomet. Chem., 1995, 497, 119.

R.F. Jordan, Adv. Organomet. Chem., 1991, 32, 325.

G.G. Hlatsky, H.W. Turner and R.R. Eckman, J. Am. Chem. Soc., 1989, 111, 2728;

Organometallics, 1992, 11, 1413.

X. Yang, C.L. Stern and T.J. Marks, Organometallics, 1991, 10, 840.

R.F. Jordan, R.E. LaPointe, P.K. Bradley and N. Baenziger, Organometallics, 1989, 8,

2892; R.F. Jordan, R.E. LaPointe, P.K. Bradley and D.F. Taylor, New. J. Chem., 1990, 14,

505.

G.G. Hlatsky, D.J. Upton and H.W. Turner, USP Appl. 459921/1990 (Chem. Abstr.

1991, 115, 256897v).

X. Yang, C.L. Stern and T.J. Marks, J. Am. Chem. Soc., 1991, 113, 3623.

Page 118: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

105

J.A. Ewen, Makromol. Chem., Macromol. Symp., 1991, 48/49, 253; J.A. Ewen and M.J.

Elder, Makromol. Chem., Macromol. Symp., 1993, 66, 179; A.D. Horton and J.H.G. Frijns,

Angew. Chem., Int. Ed. Engl., 1991, 30, 1152; A.D. Horton and A.G. Orpen,

Organometallics, 1991, 10, 3910.

X. Yang, C.L. Stern and T.J. Marks, J. Am. Chem. Soc., 1994, 116, 10015; Angew. Chem.,

Int. Ed. Engl., 1992, 31, 1375.

D.G.H. Ballard, A. Courtis, J. Helton, J. McMeeking and R. Pearce, J. Chem. Soc., Chem.

Commun., 1987, 994.

P.L. Watson, J. Am. Chem. Soc., 1982, 104, 337; P.L Watson and D.C. Roe, J. Am. Chem.

Soc., 1982, 104, 6471; P.L. Watson and G.W. Parshall, Acc. Chem Res., 1985, 18, 51.

G. Parkin, E. Bunel, B.J. Burger, M.S. Trimmer, A. Van Asselt and J.E. Bercaw, J. Mol.

Cat., 1987, 41, 21; B.J. Burger, M.E. Thompson, W.D. Cotter and J.E. Bercaw, J. Am.

Chem. Soc., 1990, 112, 1566.

G. Jeske, H. Lauke, H. Mauermann, P.N. Swepston, H. Schumann and T.J. Marks, J.

Am. Chem. Soc., 1985, 107, 8091.

G.W. Coates and R.M. Waymouth, in Comprehensive Organometallic Chemistry II, eds.

G. Wilkinson, E.W. Abel and F.G.A. Stone, Pergamon Press, Oxford, 1995, vol. 12,

p. 1193.

H.N. Cheng and J.A. Ewen, Makromol. Chem., 1989, 190, 1931; J.A. Ewen, J. Am. Chem.

Soc., 1984, 106, 6355.

W.E Piers and J.E. Bercaw, J. Am. Chem. Soc., 1990, 112, 9406.

H. Krauledat and H.H. Brintzinger, Angew. Chem., Int. Ed. Engl., 1990, 29, 1412.

M.H. Prosenc, C. Janiak and H.H. Brintzinger, Organometallics, 1992, 11, 4036.

H. Kawamura-Kuribayashi, N. Koga and K. Morokuma, J. Am. Chem. Soc., 1992, 114,

8687.

L.A. Castonguay and A.K. Rapp& J. Am. Chem. Soc., 1992, 114, 5832.

J.A. Smith, J. Von Seyerl, G. Huttner and H.H. Brintzinger, J. Organomet. Chem., 1979,

173, 175.

W. Kaminsky, K. Killper, H.H. Brintzinger and F.R.W.P. Wild, Angew. Chem., Int. Ed.

Engl., 1985, 24, 507.

75, P. Corradini, V. Busicr, Guerra, 4-1 Transition Metals and Organometallics as

Catalysts for Olefin Polymerization, eds. W. Kaminsky and H. Sinn, Springer, Berlin,

1988, p337; L. Cavallo, G. Guerra, L. Olivia, M. Vacatello and P. Corradini, Polym.

Page 119: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

106

Commun., 1989, 30, 16; V. Venditto, G. Guerra, P. Corradini and R. Fusco, Polymer,

1990, 31, 530; L. Cavallo, P. Corradini, G. Guerra and M. Vacatello, Polymer, 1991, 32,

1329.

M. Bochman, Organometallics 2, Oxford University Press, Oxford, 1994, p57-58.

M. Leclerc and H.H. Brintzinger, J. Am. Chem. Soc., 1995, 117, 1651.

H.H. Brintzinger, D. Fischer, R. Millhuapt, B. Rieger and R.M. Waymouth, Angew.

Chem., Int. Ed. Engl., 1995, 34, 1143.

A.M. Thayer, Chem. Eng. News, 1995, 73(38), 15.

J.P. Hogan and R.L. Banks, USP 2825721/1958; A. Clark, J.P. Hogan, R.L. Banks and

W.C. Lanning, Ind. Eng. Chem., 1956, 48, 1152.

M.P. McDaniel, Ind. Eng. Chem. Res., 1988, 27, 1559.

M. Tsutsui and T. Koyano, J. Polym. Sci. (A), 1967, 1, 682.

J.D. McClure and K.W. Barnett, J. Organomet. Chem., 1974, 80, 385.

W. Keim, F.H. Kowaldt, R. Goddard and C. Kruger, Angew. Chem., Int. Ed. Engl., 1978,

17, 466.

W. Keim and F.H. Kowaldt, Erdoel Kohle, Erdgas, Petrolchem., Brennst. Chem.,

Compendium, 1978, 78/79, 453.

E.R. Freitas and C.R. Gum, Chem. Eng. Progr., 1979, 75, 73.

T. Mise, A. Kageyama, S. Miya and H. Yamazaki, Chem. Lett., 1991, 1525.

L. Resconi, F. Piemontesi, G. Franciscono, L. Abis and T. Fiorani, J. Am. Chem. Soc.,

1992, 114, 1025.

J. Haggin, Chem. Eng. News, 1981, 59, 22.

H. Kolbel and M. Ralek, Catal. Rev.-Sci. Eng., 1980, 21, 225.

S. Soled, E. Iglesia and R.A. Fiato, Catal. Lett., 1990/91, 7, 271.

A.S. Rao, S.K. Paknikar and J.G. Kirtane, Tetrahedron, 1983, 39, 2323.

P. Brougham, M.S. Cooper, D.A. Cummerson, H. Heaney and N. Thompson,

Synthesis, 1987, 1015.

H. Heaney, Aldrichimica Acta, 1993, 26, 35. .

S.N. Lewis, in Oxidation, ed. R.L. Augustine, Marcel Dekker, New York, 1969, vol. 1,

p. 214.

C.-S. Lu, E.W. Hughes and P.A. Gigue.re„I. Am. Chem. Coc., 1941, 63, 1507.

M.S. Cooper, H. Heaney, A.J. Newbold and W.R. Sanderson, Synlett., 1990, 533.

R.W. Murray and R. Jeyaraman, J. Org. Chem., 1985, 50, 2847.

A.L. Baumstark and P.C. Vasquez, J. Org. Chem., 1988, 53, 3437.

Page 120: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

107

J.O. Edwards, R.H. Pater, R. Curci, L. Troisi and M. Fiorentino, J. Org. Chem., 1982, 47,

2670.

H. Masuda, K. Takase, M. Nishio, A. Hasegawa, Y. Nishiyama and Y. Ishii, J. Org.

Chem., 1994, 59, 5550.

R.G. Cooks, G.L. Glish, S.A. McLuckley and R.E. Kaiser, Chem. Eng. News, 1991, 69, 26.

R.J.K. Taylor, Organocopper Reagents A Practical Approach, Oxford University Press,

Oxford, 1994.

R.J. Anderson, J. Am. Chem Soc., 1970, 92, 4978.

R.W. Herr and C.R. Johnson, J. Am. Chem Soc., 1970, 92, 4979.

C.R. Johnson, R.W. Herr and D.M. Wieland, J. Org. Chem., 1973, 38, 4263.

B.H. Lipshutz, J.A. Kozlowski and R.S. Wilhelm, J. Am. Chem. Soc., 1982, 104, 2301.

C. Huynh, F. Derguini-Boumechal and G. Linstrumelle, Tetrahedron Lett., 1979, 1503.

B.H. Lipshutz, in Organometallics in Synthesis, ed. M. Schlosser, John Wiley & Sons,

New York, 1994, p. 283.

M.S. Kharasch and 0. Reinmuth, Grignard Reactions of Nonmetallic Substances,

Prentice-Hall, Inc., New York, 1954.

G.A. Olah, B.G. Balaram Gupta, R. Malhotra and S.C. Narang, J. Org. Chem., 1980, 45,

1638.

H. Loibner and E. Zbiral, Helv. Chico. Acta, 1976, 59, 2100; 0. Mitsunobu, Synthesis,

1981, 1; S. Manna and J.R. Falck, Synth. Commun., 1985, 15, 663.

A.I. Vogel, J. Chem. Soc., 1943, 636.

R.B. Reynolds and H. Adkins, J. Am. Chem. Soc., 1929, 51, 279.

C.-M. Hsueh and C.S. Marvel, J. Am. Chem. Soc., 1928, 50, 855.

Y.-H. Lai, Synthesis, 1981, 585.

A. FUrstrier, Angew. Chem., Int. Ed. Eng., 1993, 32, 164.

H.E. Ramsden, J. Org. Chem., 1957, 22, 1202; H.E. Ramsden, J. Org. Chem., 1957, 22,

1602; C.S. Marvel and R.G. Woolford, J. Org. Chem., 1958, 23, 1658;

W.L. Respess and C. Tamborski, J. Organomet. Chem., 1969, 18, 263; H. Felkin and G.

Swierczewski, Tetrahedron, 1975, 31, 2735.

K.J. Klabunde, H.F. Efner, L. Satek and W. Donley, J. Organomet. Chem., 1974, 71, 309.

D.E. Pearson, D. Cowan and J.D. Beckler„I. 01-g. Chem., 1959, 24, 504; R.D. Rieke and

P.M. Hudnall, J. Am. Chem. Soc., 1972, 94, 7178; R.D. Rieke and S.E. Bales, J. Chem. Soc.,

Chem. Commun., 1973, 879; R.D. Rieke and S.E. Bales, J. Am. Chem. Soc., 1974, 96, 1775;

E.C. Ashby, S.H. Yu and R.G. Beach, J. Am. Chem. Soc., 1970, 92, 433; S.H. Yu and E.C.

Page 121: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

108

Ashby, J. Org. Chem., 1971, 26, 2123; M. Gomberg and W.E. Backmann, J. Am. Chem.

Soc., 1927, 49, 236; R. Robinson and S.N. Slater, J. Chem. Soc., 1941, 376.

K.V. Baker, J.M. Brown, N. Hughes, A.J. Skarnulis and A. Sexton, J. Org. Chem., 1991,

56, 698.

D.E. Perason, D. Cowan and J.D. Beckler, J. Org.Chem., 1959, 24, 504.

B.H. Lipshutz, R.S. Wilhelm and J.A. Kozlowski, Tetrahedron, 1984, 40, 5005.

B.H. Lipshutz, R.S. Wilhelm, J.A. Kozlowski and D. Parker, J. Org. Chem., 1984, 49,

3928.

M. Schlosser, in Organometallics in Synthesis, ed. M. Schlosser, John Wiley & Sons, New

York, 1994, p.104.

K. Muller, Functional Group Determination of Olefinic and Acetylenic Unsaturation,

Academic Press, London, 1975, p. 74.

R. Hiatt, in Organic Peroxides, ed. D. Swem, John Wiley & Sons, New York, 1970, vol. 2,

chapter 1.

D. Griller and K.U. Ingold, Acc. Chem. Res., 1976, 9, 13.

P. Bedague, G. Leleu and G. De Gaudemaris, FR 1549182/1968 (Chem. Abstr. 1969, 71,

127010z).

R. Hiatt and P.M. Rahimi, Int. J. Chem. Kinet.,1978, 10, 185.

F. Tang and E.S. Huyser, J. Org. Chem., 1978, 43, 1016.

C. Janiak, K.C.H. Lange and P. Marquardt, Macromol. Rapid Commun., 1995, 16, 643.

P.C. Miihring and N.J. Coville, J. Organomet. Chem., 1994, 91, 1.

C.W. Chien and B.P. Wang, J. Polym. Sci. A, Polym. Chem., 1990, 28, 15.

J.C.W. Chien and A. Razavi, J. Polym. Sci. A, Polym. Chem., 1988, 26, 2369.

J. Christoffers and R.G. Bergman, J. Am. Chem. Soc., 1996, 118, 4715.

O.F. Steinbach, C.V. King, Experiments in Physical Chemistry, American Book

Company, New York, 1950, p. 55.

E.R. Booser, in Kirk-Othmer Encyclopedia of Chemical Technology, eds. M. Grayson and

D. Eckroth, John Wiley & Sons, New York, Third Edition, vol. 14, p. 472.

Page 122: Preparation of higher alpha olefin oligomers · Formation of Grignard reagent 59 Carbon-carbon bond formation 61 (iii) Synthesis of a model trimer 62 2.2.5 Attempted synthesis of

109

ACKNOWLEDGMENTS

A work of this magnitude could not have been accomplished alone. I would like to extend

my heartfelt gratitude to :

My supervisor, Prof. Fanie van Heerden, for her guidance and counsel over the past two

years.

Hennie Swanepoel, for his understanding and encouragement as well as for helping with

the typesetting of this thesis.

My parents, for their unconditional love and support through all the years.

My colleagues at RAU for practical suggestions, interesting discussions and providing a

congenial atmosphere to work in.

Dr. Hannes Malan at Sastech R&D for expert advise.

SASOL and RAU for financial assistance.