252
Applications of Metal Triflates and Assisted Acids as Catalysts for Organic Transformations By Mike Sbonelo Sibiya Thesis submitted in fulfilment of the requirements for the degree Philosophiae Doctor in Chemistry in the Faculty of Science of the University of Johannesburg Promoter: Prof. D.B.G. Williams Co-supervisor: Dr. P.S. Van Heerden September 2012

Applications of metal triflates and assisted acids as

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Applications of metal triflates and assisted acids as

Applications of Metal Triflates and Assisted Acids as Catalysts for Organic

Transformations By

Mike Sbonelo Sibiya

Thesis submitted in fulfilment of the requirements for the degree

Philosophiae Doctor in

Chemistry in the

Faculty of Science of the

University of Johannesburg

Promoter: Prof. D.B.G. Williams

Co-supervisor: Dr. P.S. Van Heerden

September 2012

Page 2: Applications of metal triflates and assisted acids as

i

Acknowledgements

Thanks be to God for all the grace and strength and enabling me to execute this work.

First and foremost I would like to thank Professor D. Bradley G. Williams for his incredible

support, guidance and supervision. Thank you for the enthusiasm showed during our

consultation sessions. The time you took from your busy schedule to read the manuscript is

much appreciated. The skills and knowledge I have acquired during my studies have made me

a better scientist and researcher through your guidance.

I would like to thank Dr. Pieter S. Van Heerden, who has been my manager for quite a long time

at Sasol Technology R&D. Your support, guidance and co-supervision of the project will always

be appreciated. Thank you for introducing me to research and patiently mentoring me in the

organisation up to this stage.

I would also like to thank the following people:

Prof. Mike Green (Newcastle University) for all the support and encouragement while still at

Sasol Technology R&D.

Dr. Cathy Dwyer, for patiently allowing me to take some time and complete this work.

Mr. Petrus Makgoba and Mr Raymond Hunt for all the support and also for understanding if I

had to take leave to execute this manuscript.

Dr. Andile Mzinyati and Mr. Nceba Magqi for taking some time to proof-read and review the

document.

Dr. Mokgolela Mogorosi, Dr. Chris Maumela and Dr. Kenny Tenza for all the support and

scientific discussions on the project.

Dr. Sibusiso Mtongana and Dr. Rita Jordaan for assisting with the NMR analysis at Sasol

Technology Research and Development.

Dr. Megan Shaw, your assistance with NMR analysis at the University of Johannesburg is much

appreciated.

Dr. Michelle Lawton for assisting with the pH measurement instruments.

I would like to thank all my colleagues at Sasol Technology R&D who have sacrificed their time

and effort, to mention, but a few: Mrs. Mariam Abrahams, Dr. Norah Maithufi, Mr. Molladi

Page 3: Applications of metal triflates and assisted acids as

ii

Maseloane, Dr. Dave Morgan, Dr. Hendrik van der Westhuizen, Dr. Peter Bungu, Dr Irene

Kamara, Mr. John Smith and Dr. Esti van Ryneveld for support and encouragement.

I would also like to pass a word of thank you to the Acid and Base Catalysis team at Sasol

Technology R&D for all the support (Bongani Mdakane, Zibuyile Mncwabe and Dr. Daphne

Radebe).

To my family

I dedicate this work to my late Mother (Mrs. Minha Khonzaphi Sibiya), I will always remember

her as a supportive, loving and courageous person. “Noma ucashele amehlo ethu, kodwa

izinhliziyo ziyakuzwa nokho” siyabonga ngakho konke owasenzela kona, lala uphumule manje.

I would like to pass a word of thank you my father (Mr. Elphas M. Sibiya), my brothers

(Manzolwandle Tu, Phumlane Tile, Siyabonga Ndoda, Siyabonga Gede and Lethukuthula,

Mfana) bafethu nina niyikho konke kimi, amathemba ami ahlezi kunina kukhona konke. To my

sisters (Nokwazi Pita, Londi Lo, Philile Disco, maZakwe Sibiya, maNene Sibiya kanye no

Noxolo, Xolo) ngiyabonga kakhulu maGumede sengathi uThixo angangigcinela nina, niqhubeke

nokwenza okwamukelekile ngaso sonke isikhathi.

Page 4: Applications of metal triflates and assisted acids as

iii

Synopsis

The research contained in this thesis was aimed at the applications of Lewis acids (metal triflate

salts in particular) and Brønsted acids as catalysts for various organic synthesis reactions. The

ultimate objective was to prepare combinations of the Lewis and Brønsted acids to form

assisted acids. The assisted acids yield to the formation of highly acidic assisted acids which

exhibit high activity as compared to the individual Lewis and Brønsted acids. A detailed

literature study was undertaken, with emphasis on the applications of metal triflate salts as

catalysts for various organic reactions and the applications of assisted acids.

The study was motivated by the fact that metal triflate Lewis acids are thermally stable, non

corrosive and water tolerant catalysts, hence can be used industrially to replace the corrosive,

moisture sensitive acids as catalysts. However, metal triflates have not yet been recognised and

utilised in the chemical industry. On the other hand, the active Brønsted acids such as triflic

acid, H2SO4 etc. are corrosive, which restricts the type of construction material to hastelloy.

However, the assisted acids composed of less corrosive Brønsted acids and metal triflate Lewis

acid is desirable to address the corrosion and safety challenges.

The metal triflate salts and Brønsted acids were evaluated as catalysts for etherification

reactions of alcohols and olefins, Friedel-Crafts alkylation reactions phenolic substrates with

isobutylene. The study showed that some dependence of the charge density to the activity, i.e.

metal triflate salts such as Al(OTf)3, Zr(OTf)4 and Sc(OTf)3 with relatively high charge density

were more effective in catalysing the reactions than those with relatively smaller charge density

such as lanthanides, which were virtually active. The activity of Brønsted acids showed a clear

dependence on the acid strength pKa, with H3PO4 giving the least activity.

The assisted acids formed via a combination of metal triflate salts with mineral Brønsted acids

showed a significant enhancement of the reaction rates as compared to the individual acids.

This set of new combined acids was proven to be excellent catalysts for the etherification

reactions, Friedel-Crafts alkylation reactions and also for the synthesis of biologically active

compounds called chromans. The assisted acids as well as Al(OTf)3, and Zr(OTf)4 could be

recycled at least four times without significant loss of activity. The study also showed that

assisted acids could be recycled for both etherification and Friedel-Crafts reactions.

Page 5: Applications of metal triflates and assisted acids as

iv

Abbreviations

H0 Hammett acidity function

HSAB Hard and soft acids and bases

PA Proton affinity

LAB Linear alkyl benzene

LA Lewis acid

Ln Lanthanide

ScCO2 Supercritical carbon dioxide

BMIM 1-butyl-3-methylimidazolium

EMIM 1-ethyl-3-methylimidazolium

TFA Triflic acid

NTF2 Triflimide

MTBE Methyl tertiary butyl ether

BHT Butylated hydroxyl toluene

S-ZrO2 Sulfated zirconia

TPA Tungstophosphoric acid

DPE Diphenylether

TAME Tertiary-amyl methyl ether

TAEE Tertiary-amyl ethyl ether

2M2B 2-methyl-2-butene

2M1B 2-methyl-1-butene

DM1B 2,3-dimethyl-1-butene

DM2B 2,3-dimethyl-2-butene

GC Gas chromatography

NMR Nuclear magnetic resonance

Page 6: Applications of metal triflates and assisted acids as

v

BLA Brønsted acid-assisted Lewis acid

LLA Lewis acid-assisted Lewis acid

LBA Lewis acid-assisted Brønsted acid

BBA Brønsted acid-assisted Brønsted acid

FID Flame ionisation detector

TOF Turn over frequency

SPA Solid phosphoric acid

2M2M3B 2-methoxy-2-methyl-3-butene

2,4DMB 2,4-dimethoxybutane

1M3M2B 1-methoxy-3-methyl-2-butene

MCP Methylenecyclopentane

UOP Inc. Universal Oil Products Incorporated

PC para-Cresol

MC meta-Cresol

DBPC 2,6-di-tert-butyl-para-cresol

MBPC Mono-butylated-para-cresol

DBMC 2,4-di-tert-butyl-meta-cresol

MBMC Mono-butylated-meta-cresol

HOTf Trifluoromethanesulfonic acid

GC-MS Gas Chromatography-Mass Spectroscopy

Bp Boiling point

Mp Melting point

HIV Human immunodeficiency virus

Page 7: Applications of metal triflates and assisted acids as

vi

Table of contents

CHAPTER 1: Literature Review

1.1 Introduction 1

1.1 Liquid phase Brønsted acid catalysts 3

1.2 Lewis acid catalysis: A focus on metal triflate salts 4

1.3 General applications of metal triflate as catalysts 6

1.3.1 Mukaiyama Reaction 6

1.3.2 Diels Alder Reactions 8

1.3.3 General Friedel-Crafts alkylation reactions 17

1.3.4 Friedel-Crafts alkylation of phenolic compounds 22

1.3.4.1 Friedel-Crafts alkylation of cresols 22

1.3.4.2 Friedel-Crafts alkylation of anisole 32

1.3.4.3 Friedel-Crafts alkylation of diphenylether (DPE) 35

1.4 Etherification of alcohols and olefins 38

1.5 Condensation of phenols with dienes 45

1.6 The concept of combined acid systems 49

1.7 Summary 50

CHAPTER 2: Hydroalkoxylation of alcohols

2.1 Introduction 52

2.2 Screening of Lewis acids 55

Page 8: Applications of metal triflates and assisted acids as

vii

2.3 Screening of Brønsted acids 56

2.4 Optimisation of reaction conditions 58

2.4.1 The influence of varying the alcohol to olefin mole ratio. 58

2.4.2 The influence of changing catalyst concentration 59

2.4.3 The influence of changing the reaction temperature 60

2.5 Recycling of Al(OTf)3 and Zr(OTf)4 61

2.6 The evaluation of Lewis assisted Brønsted acids 63

2.7 The influence of varying La(OTf)3/H3PO4 concentration 68

2.8 The influence of varying temperature on La(OTf)3/H3PO4 catalysed reactions 69

2.9 Effect of varying the catalyst composition 71

2.10 Recycling of La(OTf)3/H3PO4 assisted acids 72

2.11 Solid phosphoric acid (SPA) catalysis 73

2.12 Comparison of activity of Amberlyst 15 and La(OTf)3/H3PO4 75

2.13 The reactivity of other olefins with methanol 76

2.13.1 The reactivity of isoprene with methanol 76

2.13.2 Etherification of styrene with methanol 79

2.13.3 Etherifcation of methylenecyclopentane (MCP) with methanol 81

2.14 Etherification of tertiary olefins in the presence of other olefins 87

2.15 Etherification of other alcohols with 2-methyl-2-butene 89

2.16 Conclusions 91

CHAPTER 3: Friedel-Crafts alkylation

Page 9: Applications of metal triflates and assisted acids as

viii

3.1 Introduction 93

3.2 Alkylation of cresol 95

3.3 Evaluation of metal triflate Lewis acids as catalysts for the butylation of cresol 100

3.3.1 Catalyst recycling studies 108

3.3.2 The influence of catalyst concentration 110

3.3.3 The influence of varying stirring speed 112

3.4 Screening of Brønsted acids 114

3.5 Evaluation of assisted acids 115

3.6 Evaluation of La(OTf)3/SPA assisted acid 120

3.7 Friedel-Crafts alkylation of anisole 124

3.7.1 Evaluation of individual Lewis and Brønsted acids 126

3.7.2 The evaluation of assisted acids as catalysts for the butylation of anisole 130

3.8 Alkylation of diphenylether (DPE) 133

3.9 Conclusions 140

CHAPTER 4: Synthesis of chromans

4.1 Introduction 142

4.2 Reaction of phenol with 2-methyl-1,3-butadiene (isoprene) 143

4.3 Reaction of phenols with 2,3-dimethyl-1,3-butadiene 149

4.4 Conclusions 153

Page 10: Applications of metal triflates and assisted acids as

ix

CHAPTER 5: Experimental protocol

5.1 General protocol for preparation of ethers 154

5.2 General protocol for alkylation of phenolic compounds 162

5.3 General protocol for synthesis of chromans 168

REFRENCES 181

APPENDIX 193

Page 11: Applications of metal triflates and assisted acids as

x

List of schemes

Scheme 1.1: Yb(OTf)3 catalysed Mukaiyama reactions 6

Scheme 1.2: Lanthanide triflate catalysed aldol reactions of silyl enolates with aldehydes

and acetals 7

Scheme 1.3: Michael addition reaction catalysed by Yb(OTf)3 8

Scheme 1.4: Diels-Alder reaction of methyl vinyl ketone with isoprene in the presence of

Sc(OTf)3 8

Scheme 1.5: Sc(OTf)3 catalysed Diels-Alder reaction in aqueous medium 10

Scheme 1.6: Diels-Alder reaction in water catalysed by polymer supported Sc(OTf)3 10

Scheme 1.7: Ionic liquids used in the investigation 11

Scheme 1.8: Diels-Alder reactions catalysed by ionic liquids and metal triflates 11

Scheme 1.9: Diels-Alder reactions catalysed by Yb(OTf)3 at 0˚C 13

Scheme 1.10: Reaction catalysed by Yb(OTf)3-H2O under high pressure. 13

Scheme 1.11: Intramolecular Diels-Alder reaction of oxazole-olefin 14

Scheme 1.12: Preparation of chiral rare-earth M(OTf)3 catalyst 15

Scheme 1.13: Diels-Alder reaction catalysed by “chiral Yb triflate.” 15

Scheme 1.14: Preparation of tetrahydropyridines in the presence of Sc(OTf)3 17

Scheme 1.15: Friedel-Crafts alkylation of benzene with benzyl alcohol. 17

Scheme 1.16: Friedel-Crafts alkylation of toluene with alkyl halides 19

Scheme 1.17: Friedel-Crafts alkylation of furan with α-chloro-α-(ethylthio)acetate 20

Scheme 1.18: Friedel-Crafts alkylation in ionic liquids 20

Scheme 1.19: Alkylation of indoles with aldehydes and ketones using Dy(OTf)3 21

Page 12: Applications of metal triflates and assisted acids as

xi

Scheme 1.20: Dehydration of tert-butanol and butylation of p-cresol 23

Scheme 1.21: Alkylation of p-cresol with tert-butanol in the presence of TPA/ZrO2 26

Scheme 1.22: tert-Butylation of phenol 27

Scheme 1.23: Butylation of cresols with tert-Butanol using K10-Fe-120 at 80 °C 28

Scheme 1.24: Butylation of p-cresol with MTBE 29

Scheme 1.25: Alkylation of m-cresol with isopropanol 30

Scheme 1.26: Butylation of anisole with isobutylene 33

Scheme 1.27: Etherification reaction of methanol and isoamylene 38

Scheme 1.28: Zeolites catalysed etherification of methanol with isobutylene 42

Scheme 1.29: Etherification of isopropanol with isobutylnene 43

Scheme 1.30: Mechanism for synthesis of 2,2-dimethylchromane 45

Scheme 1.31: Products from the reaction of phenol and isoprene 46

Scheme 1.32: Production of chromans and chromenes catalysed by Amberlyt 15 46

Scheme 1.33: [(acac)2Mo(SbF6)2]-catalysed preparation of chroman 47

Scheme 1.34: Products for the reaction of isoprene with phenol in the presence of

Al(OPh)3. 47

Scheme 1.35: Ag(OTf) catalysed reaction of isoprene with 4-methoxy phenol 48

Scheme 2.1: Etherification of alcohols with 2M2B 53

Scheme 2.2: Lewis acid catalysed etherification 54

Scheme 2.3: Another possible mechanism for a Lewis acid catalysed etherification 54

Scheme 2.4: Hydration reaction of 2M2B 56

Page 13: Applications of metal triflates and assisted acids as

xii

Scheme 2.5: Brønsted acid catalysed etherification reaction of methanol and 2M2B 57

Scheme 2.6: Addition of HCl to 2M2B 58

Scheme 2.7: Proposed reaction of Brønsted and Lewis acid 65

Scheme 2.8: Other olefins used for etherification with methanol 76

Scheme 2.9: Products distribution during etherification of isoprene 77

Scheme 2.10: Reaction of styrene with methanol 79

Scheme 2.11: Etherification and isomerisation of methylenecyclopentane 81

Scheme 2.12: Etherification of ethanol with 2M2B 89

Scheme 3.1: Alkylation of phenol and o-cresol in the presence of sulfonic acids 96

Scheme 3.2: Aryl sulfonic acid catalysts 97

Scheme 3.3: Rearrangement of 2-tert-butylphenol to 4-tert-butylphenol 97

Scheme 3.4: De-alkylation 2,4-di-tert-butylphenol 98

Scheme 3.5: The acid catalysed Friedel-Crafts alkylation of p-cresol 99

Scheme 3.6: Alkylation of m-cresol with isobutylene 100

Scheme 3.7: Dimerisation of isobutylene under acidic conditions 101

Scheme 3.8: Formation of a complex via combination of La(OTf)3/H3PO4 116

Scheme 3.9: Alkylation of anisole with isobutylene 124

Scheme 3.10: Generation of isobutylene from tert-BuOH and MTBE 125

Scheme 3.11: Alkylation of DPE with 1-decene 135

Scheme 4.1: Addition of 1,3-diene to a phenol derivative. 142

Page 14: Applications of metal triflates and assisted acids as

xiii

Scheme 4.2: Proposed mechanism for AgOTf catalysed synthesis of chromans. 143

Scheme 4.3: Preparation of 2,2,6-trimethylchroman catalysed by La(OTf)3/H3PO4 144

Scheme 4.4: Reaction of isoprene with hydroquinone in the presence of La(OTf)3/H3PO4 146

Scheme 4.5: Products obtained from the reaction of phenol and isoprene. 146

Scheme 4.6: Brønsted acid catalysed preparation of chroman. 147

Scheme 4.6: Brønsted acid catalysed preparation of chroman. 151

Page 15: Applications of metal triflates and assisted acids as

xiv

List of figures

Figure 2.1: Influence of varying the methanol to olefin mole ratio using 1.5 mol% Al(OTf)3 59

Figure 2.2: The influence of changing the catalyst concentration 60

Figure 2.3: Influence of changing the reaction temperature 61

Figure 2.4: Recycling of Al(OTf)3 during etherification reactions. 62

Figure 2.5: Recycling of Zr(OTf)4 during etherification reactions. 63

Figure 2.6: Etherification reactions catalysed by La(OTf)3/H3PO4 assisted acid 64

Figure 2.7: Activity of metal triflates/H3PO4 acid combinations 66

Figure 2.8: Comparison of metal triflate/mineral acids activity. 67

Figure 2.9: Comparison of metal triflates/ organic acids for etherification 68

Figure 2.10: Effect of increasing La(OTf)3/H3PO4 concentration 69

Figure 2.11: Effect of changing the reaction temperature 70

Figure 2.12: The effect of varying the catalyst composition 72

Figure 2.13: Recycling of La(OTf)3/H3PO4 73

Figure 2.14: The activity of La(OTf)3/SPA during etherification 74

Figure 2.15: Comparison of La(OTf)3/H3PO4 and Amberlyst 15 for etherification reactions 75

Figure 2.16: Etherification of isoprene with methanol in the presence of Zr(OTf)4 78

Figure 2.17: Reaction of isoprene with methanol catalysed by La(OTf)4/H3PO4 79

Figure 2.18: Results for etherification of styrene with methanol 80

Figure 2.19: Etherification of MCP with methanol in presence of Al(OTf)3 83

Figure 2.20: Products distribution afforded by various acids 84

Page 16: Applications of metal triflates and assisted acids as

xv

Figure 2.21: Etherification of MCP with methanol catalysed by La(OTf)3/H3PO4 85

Figure 2.22: Etherification MCP with methanol catalysed by La(OTf)3/H2SO4 86

Figure 2.23: GC trace for selective etherification of 2M2B 88

Figure 2.24: Reactivity of methanol with 2M2B in the presence of 1-hexene and

cyclohexene 89

Figure 2.25: The activity of various acids during etherification of ethanol and 2M2B 90

Figure 2.26: Reaction catalysed by assisted acids 91

Figure 3.1: Butylation of m/p-cresol in the presence of Zr(OTf)4 (0.05wt%) at 70 °C 104

Figure 3.2: Summary of reaction rates for various assisted acids 124

Figure 3.3: Alkylation of anisole (560 mmol) with 1.5 bar isobutylene catalysed by 1.46

mmol Zr(OTf)4 at 70 ˚C, and stirred at 1000 rpm. 126

Figure 3.4: Alkylation of anisole (560 mmol) with 1.5 bar isobutylene catalysed by

Zr(OTf)4 (1.46 mmol ) at 100 ˚C, and stirred at 1000 rpm. 127

Figure 3.5: Alkylation of anisole (560 mmol) with 1.5 bar isobutylene catalysed by HOTf

(0.2 mol%) at 100 ˚C, and stirred at 1000 rpm. 129

Figure 3.6: The activity of 2.0 mol% H3PO4 during alkylation of anisole. 130

Figure 3.7: Alkylation of anisole (560 mmol) with 1.5 bar isobutylene catalysed by

La(OTf)3/H3PO4 (0.8 mol%) at 100 ˚C, and stirred at 1000 rpm 131

Figure 3.8: Alkylation of anisole (560 mmol) with 1.5 bar isobutylene catalysed by

Al(OTf)3/H3PO4 (0.8 mol%; 0.8 mol%) at 100 ˚C, and stirred at 1000 rpm 132

Figure 3.9: Alkylation of anisole (560.01mmol) with 1.5 bar isobutylene catalysed by

Gd(OTf)3/H3PO4 (0.8 wt%; 0.2 mol%) at 100 ˚C, and stirred at 1000 rpm 133

Page 17: Applications of metal triflates and assisted acids as

xvi

Figure 5.1: Reactor set-up used during etherification of alcohols and olefins 155

Figure 5.2: Experimental set-up for the purification of ethers 157

Figure 5.3: The experimental set-up for the butylation of phenolic compounds 162

Figure 5.4: Kugelrhor Distillation setup used during purification of chromanes 169

Page 18: Applications of metal triflates and assisted acids as

xvii

List of tables

Table 1.1: The super acids with their Hammet constants 2

Table 1.2: Diels-Alder reaction of 1 or 4 with 5 12

Table 1.3: Metal triflates catalysed intramolecular reaction of oxazole olefin 14

Table 1.4: Enantioselective Diels-Alder reactions catalysed by chiral rare-earth metal

triflates 15

Table 1.5: Lewis acid-catalysed benzylation with benzyl alcohola 18

Table 1.6: Dy(OTf)3 catalysed alkylation of indoles with aldehydes and ketones 21

Table 1.7: The products afforded by Nafion® resin/silica and Amberlyst 15. 25

Table 1.8: Butylation of phenol with tert-butanol using K10 catalysts at 80 °C 27

Table 1.9: Activity of various acids for butylation of p-cresol with MTBE 29

Table 1.10: Comparison of various solid acid catalysts during etherification of DM1B with

methanol. 41

Table 2.1: Reaction of methanol with 2M2B and correlation to ionic radiia 55

Table 2.2: Catalytic activity of Brønsted acids during etherification reaction 57

Table 2.3: Turn over frequencies over a period of 15 minutes. 70

Table 2.4: Comparison of activity of various catalysts for etherification of MCP and

methanol 87

Table 3.1: Results for butylation of p-cresol in the presence of Zr(OTf)4 103

Page 19: Applications of metal triflates and assisted acids as

xviii

Table 3.2: Butylation of m/p-cresol in the presence of Al(OTf)3 (0.05 wt%), at 70 ˚C 105

Table 3.3: Butylation of m/p-cresol in the presence of 0.05 wt% Sc(OTf)3, at 70 ˚C 106

Table 3.4: Comparison of activity of the triflate salts of Zr, Al and Sc. 107

Table 3.5: Recycling of Al(OTf)3 (0.09 wt%) during butylation of m/p-cresol at 70 °C,

stirring at 1200 rpm and using 1.1 bar of isobutylene 109

Table 3.6: The influence of catalyst concentration during butylation of m/p-cresol in the

presence of Al(OTf)3 at 70 °C, 1.1 bar isobutylene pressure and stirring at

1200 rpm 110

Table 3.7: Influence of stirring speed on butylation of m/p-cresol in the presence of 0.09

wt% Al(OTf)3 at 70 °C 113

Table 3.8: Butylation of the m/p-cresol in the presence of H2SO4 114

Table 3.9: Butylation of m/p-cresol in the presence of [La(OTf)3/H3PO4] 117

Table 3.10: Butylation of m/p-cresol in the presence of [Gd(OTf)3/H3PO4] 118

Table 3.11: Butylation of m/p-cresol catalysed by Al(OTf)3/H3PO4 119

Table 3.12: Butylation of m/p-cresol in the presence of La(OTf)3/SPA. 120

Table 3.13: Comparison of cresols conversions and product selectivities afforded by

La(OTf)3/H3PO4 and La(OTf)3/SPA after 120 minutes

121

Table 3.14: The results obtained during recycling of La(OTf)3/SPA. 123

Table 3.15: Comparison of the acid strengths of H3PO4 and HOTf 128

Table 3.16: Properties of the products obtained during alkylation of diphenylether 137

Table 3.17: The variation of reaction parameters during butylation of DPE (0.36mol) in

the presence of Zr(OTf)4 138

Table 3.18: The results obtained during evaluation of other catalysts for butylation of

DPE (0.57 mol%). 139

Page 20: Applications of metal triflates and assisted acids as

xix

Table 3.19: Results obtained during the evaluation of assisted acids for the butylation of

DPE 140

Table 4.1: Reaction of various phenols (2.1 mmol) with isoprene (7.0 mmol) in the

presence of assisted acid La(OTf)3/H3PO4 148

Table 4.2: Reaction of substituted phenols with 2,3-dimethyl-1,3-butadiene. 152

Table 5.1: Reagents used during etherification reaction of olefins and alcohols 156

Table 5.2: GC operating conditions during analysis of the etherification reaction mixture. 158

Table 5.3: Reagents used during alkylation or butylation of phenolic compounds 163

Table 5.4: Analysis conditions of the butylated compounds 164

Page 21: Applications of metal triflates and assisted acids as

1

CHAPTER 1

Literature Review This chapter entails the literature review on Lewis and Brønsted acid catalysis with high

emphasis on Friedel-Crafts alkylation reaction of aromatic compounds, etherification of olefins

and alcohols and condensation reactions of phenols with dienes.

1.1 Introduction The concept of acidity was established by S.A. Arrhenius 1 in 1903. This Nobel Prize winner

described an acid as any hydrogen containing species able to release protons and base as any

species able to release hydroxide ions. This concept was modified by J.M. Brønsted2 and T.M.

Lowry3 independently in 1923. According to these pioneers, a base is any species that is able to

combine with protons. In this view, equilibrium exchange of a proton from an acid to a base,

which could also be a solvent, generates a conjugate base and conjugate acid. In the same

year (1923), G. N. Lewis4 proposed a different approach. He described an acid as any species

that can accept an electron pair (octet rule of electron configuration), thus forming a dative or

coordinative bond. Conversely, a base is any species bearing a non-bonding electron pair which

can be donated to form a dative bond.

The acid strength in water is determined by the acid dissociation constant (Ka), which is

promoted by the solvation of the protons by water molecules forming hydronium ions (H3O+).

The extent of dissociation of pure anhydrous acids is far lower than for the corresponding water

solutions, i.e. the dissociation constants (Ka) for sulfuric acid and hydrofluoric acids in water are

> 102 and 10-3, respectively, but they decrease to 10-4 and 10-10 respectively in the absence of

water.5 The acidity of the highly concentrated or anhydrous acids is described according to the

Hammett acidity function (H0), proposed by Hammett in the 1930s.6

H0 = pKBH++ log [B]/[BH+]

Where B is a basic indicator and pKBH+ is the pKa of its conjugate acid. The obtained value of H0

is almost independent of the indicator base B. The H0 values for anhydrous H2SO4 and HF are -

12 and -15, respectively. One of the strongest acids known is SbF6/HSO3F called “magic’ acid,

which is a composed of Lewis acid component (SbF6) and Brønsted acid component (HSO3F).

Page 22: Applications of metal triflates and assisted acids as

2

The Hammett number of this complex is between -23/-26, and the 1:1 complex of SbF6/HF gives

H0 = -28. These highly acidic liquids are called super acids, a term first used by Conant in 1972

to describe very strong acids.7 This very term was also used by Gillespie to indicate acids that

are stronger than 100% sulfuric acid the 1960’s.8 This field of superacid catalysis has been

widely explored by Olah and co-workers. 9 A list of super acids is reported in Table 1.1. The

carborane acids of the formula HCB11H11-xXx, where X = Cl, Br or I, have been developed by

Reed and co-workers.10 These acid systems are strong yet gentle solid acids and are able to

protonate extremely weakly basic substrates such as olefins and aromatics.

Table 1.1: The super acids with their Hammet constants11

Acid Name Molecular formula H0

Hydrogen fluoride/Antimony fluoride HF/SbF6 -28

Antimony fluoride (“magic acid”) SbF6/HSO3F -23/-26

Hydrogen fluoride HF (anhydrous) -15

Fluorinated sulfuric acid HSO3F -15

Disulfuric acid H2S2O7 -15

Hydrogen chloride/aluminium chloride HCl/AlCl3 -15/-14

Hydrogen fluoride/trifluoroborane HF/BF3 -15/-14

Water / trifluoroborane H2O/BF3 -15/-14

Triflic acid CF3SO3H -14.1

Sulfuric acid H2SO4 (100%) -11.9

In 1963, Pearson12 and co-workers introduced the concept of hard and soft acid and bases

(HSAB). According to this theory, hard acids are defined as highly positively charged, small

sized and not easily polarisable electron acceptors: hard acids prefer to associate with hard

Page 23: Applications of metal triflates and assisted acids as

3

bases. Hard bases are substances that hold their electron pair tightly due to high

electronegativity, low polarisability and difficulty of oxidation of their donor atoms. Furthermore,

soft acids prefer to associate with soft bases. According to this Pearson theory, the proton is a

hard acid and metal cations may have different hardness properties. The acid and base

properties may be evaluated experimentally or estimated by theory. The enthalpy of

deprotonation of the conjugate acid in the gas phase leads to an acidity or basicity scale on

proton affinity (PA):

PA = -ΔHprotonation

1.1 Liquid phase Brønsted acid catalysts

A wide range of liquid acids has been evaluated as catalysts for hydrocarbon conversion both in

the industry and academic research projects. The advantages associated with liquid acids

include their activity and selectivity under mild conditions as compared to solid acids. The

drawback associated with liquid acids is separation of catalyst from the products, loss of

catalysts and environmental factors. A common liquid acid used in hydrocarbon chemistry is

sulfuric acid (H2SO4). H2SO4 is a strong diprotic acid with H0 ranging between -3.5 and -11.9,

depending on the extent of dilution. There are several industrial processes that are still based

on H2SO4 e.g. Friedel-Crafts alkylation reactions, hydration of olefins and oligomerisation of

reactions.

Sulfur trioxide (SO3) is soluble in H2SO4, resulting in the formation of disulfuric acid (H2S2O7)

also known as “oleum”. Oleum is a liquid super acid with H0 = -15. This super acid is mostly

used for sulfonation reactions. However, the main drawback associated with usage of strong

sulfuric acid is related to the difficulty of its regeneration, purification and concentration.

Furthermore, H2SO4 is capable of oxidising paraffins which results in the formation of SO2 and

alkyl sulfates. An additional disadvantage associated with H2SO4 includes its corrosive nature,

which imposes usage of specialised material for construction of reactors and distillation towers.

It is also essential that spent catalyst is carefully disposed of in a safe manner.

Hydrofluoric acid (HF) is a super acid with H0 = -15 in pure liquid form and H0 = -11 in the

presence of small amounts of water. In water solution HF is a weak acid [Ka = 2-7 x10-4]. Its

acidity increases as a function of an increase in concentration, due to increase in stabilisation of

F- anion when its surroundings become more ionic. The acidity of HF is further increased by

combination with a Lewis acid such as SbF5, and this acid combination HF/SbF5 is the strongest

known acid system with H0 = -28. Literature suggests that the acidity is increases by the

Page 24: Applications of metal triflates and assisted acids as

4

formation of (H2F)+ ions [H2F+(HF)n] and of solvated ions such as (Sb2F11)- and Sb3F15..13 HF has

been widely used as a catalyst for the isobutene or isobutylene alkylation process.14 HF has

also been reportedly used for benzene alkylation processes such as the synthesis of linear alkyl

benzene (LAB), and cumene.15 During the synthesis of LAB, the feed contains approximately

79% HF. The reaction temperature ranges between 0 and 10 °C at ambient pressure with a

large excess of benzene (4-10 mol benzene/olefin).16 The advantage associated with HF over

H2SO4 is in the ease of catalyst separation and purification by means of distillation, due to the

high volatility of HF, and hence the loss of this substance is reduced. The main drawback

associated with HF is related to its toxicity together with its volatility, which promote the

formation of toxic aerosol clouds, thereby posing a high safety hazard. However, the use of

suppressants such as nitrogen-donor bases as additives limits the shortcoming to some extent

e.g. pyridinium poly-(hydrogen fluoride), also known as Olah’s reagent.17 It is indicated in the

literature that amine poly-(hydrogen fluoride) complexes are active catalysts and are associated

with lowered HF vapour pressure.18 The UOP process for the production of cumene via

alkylation of benzene with propene shows that the suppression additives can mitigate the risk of

HF release by 90%.19

The aluminium trichloride Lewis acid, was discovered in the 19th century by C. Friedel and J.M.

Crafts and applied specifically to alkylation reactions. The solid has a melting point of 193 ˚C

and can form the dimer complex Al2Cl6. It also produces other low temperature complexes with

other metal chlorides and gives rise to liquid complexes with hydrocarbons20 and ionic liquid

precursors.21 In solution this Lewis acid produces ionic liquid species such as AlCl4-, Al2Cl7-. Al in

the monomeric complex is coordinatively unsaturated and is only saturated in the polymeric

anions by weak nucleophile such as the Cl- anions.

1.2 Lewis acid catalysis: A focus on metal triflate salts

The Lewis acid (LA) catalysts are of great importance and interest in organic chemistry,

especially during carbon-carbon bond formation and because of their unique activities and

selectivities under mild reaction conditions.22 A variety of Lewis acid-catalysed reactions have

been investigated and these acids have been used during synthesis of natural and unnatural

products. The major disadvantages associated with traditional Lewis acids such as AlCl3, BF3,

TiCl4 etc. is that more than stoichiometric amounts of the acid are required in most cases.

Furthermore, these Lewis acids are moisture sensitive and easily decompose in the presence of

even small amounts of water. But, most importantly, these Lewis acids cannot be easily

Page 25: Applications of metal triflates and assisted acids as

5

recovered and reused, if at all, in a subsequent reaction. Recent discoveries indicate that

trifluoromethanesulfonate, or triflate (CF3SO3-), complexes are an alternative to the traditional

metal halide Lewis acids for catalysing a variety of organic transformations. In 1991, the first

report was issued on water tolerant Lewis acids, relating to lanthanide triflates [Ln(OTf)3]. 23 The

most characteristic nature of Ln(OTf)3 complexes is that they are stable and function as Lewis

acids in water. Not only Ln(OTf)3, where Ln=La, Ce, Pr, Nd, Pm, Sm, Eu, Gd, Tb, Dy, Ho,Er,

Tm, Yb and Lu) but also scandium (Sc) and Ytterbium (Y) were shown to be water compatible

Lewis acids and these complexes have been regarded as new set of Lewis acids since 1991.

Many organic transformations can be catalysed by these metal triflates in aqueous medium and

only catalytic amounts are required to complete the reaction in most cases. Moreover, metal

triflates can be easily recovered when the reaction is completed and reused in the subsequent

reaction without any loss of activity.24 The triflates are still active in the presence of many Lewis

bases containing hetero-atoms such as nitrogen, oxygen, phosphorus and sulfur. Metal triflate

Lewis acids are mostly used in catalytic amounts, while the conventional Lewis acids are used

in stoichiometric quantities. Thus, the use of conventional acids requires treatment of the

residues at the end of the reaction, which may induce major environmental problems and

financial cost. Metal triflates are regarded as environmental friendly Lewis acids. The efficacy of

metal triflate-based Lewis acid as catalysts may vary from one reaction to another. Recently,

studies were carried out to measure the relative acidity of Lewis acids based on their

competitive ligand dissociation from the M(OTf)3(L4) complexes, (L= hexamethylphosphoramide,

triethylphosphine oxide or trimethyl phosphate) using tandem mass spectroscopy.25 The results

showed a good correlation between the acidity and extraordinary catalytic activity of Sc(III) and

Yb(III) in the Lewis acid catalysed reactions. The high Lewis acidity of Sc(III) and Yb(III) may be

associated with their small ionic radii, which implies high charge density, and hence high

catalytic activity.26

The metal triflate Lewis acids are prepared by heating metal oxides or chlorides in an aqueous

trifluoromethane sulfonic acid (TfOH).27 These metal triflate-based Lewis acids can also be

prepared by reacting aqueous silver triflate with metal halides.28 The metal triflate molecule

normally contains nine water molecules, and anhydrous samples can be prepared by drying at

higher temperatures under high vacuum.29

Page 26: Applications of metal triflates and assisted acids as

6

1.3 General applications of metal triflate as catalysts

1.3.1 Mukaiyama Reaction

Lewis acid catalysed carbon-carbon bond forming reactions have been of great interest in

organic synthesis due to the selectivity achievable and for the mild reaction conditions used.30

The Mukaiyama reaction is one of the important C-C bond forming reactions. It involves the

reaction of silyl enol ethers with aldehydes or ketones in the presence of a Lewis acid such as

TiCl4.31 In the past, TiCl4 was found to be the best Lewis acid as catalyst for the reaction;

however, this Lewis acid requires strictly anhydrous conditions.

The use of lanthanide triflate salts as catalysts for the Mukaiyama reaction was first reported by

Kobayashi et al.32 Interestingly, the reactions were performed in the presence of water (1:1

mixture of water and THF) and it was found that silylenol ethers react readily with aldehydes in

the presence of Yb(OTf)3 (Scheme 1.1). Catalytic amounts (1.0 mol%) of Yb(OTf)3 were used

for the reactions, with yields ranging from 77 to 94%, depending on the silyl enol ether used.

Other aldehydes and ketones such as benzaldehyde, salicylaldehyde, 2-

pyridinecarboxyaldehyde etc. can also be used in the Mukaiyama reaction.

1.1 1.2 1.3

1.1 1.4 1.5 1.6

1.1 1.7 1.8

OSiMe

Ph

O

Ph OH

O

H H

OSiMe O

OHYb(OTf)3

92%

O

SEt

OSiMe

Yb(OTf)3 O

SEt

OH

H

OSiMe

OH

O

OSiMe

83% overall

Yb(OTf)3

94%

O

H H

O

H H

Scheme 1.1: Yb(OTf)3 catalysed Mukaiyama reactions

Page 27: Applications of metal triflates and assisted acids as

7

The optimisation studies carried out in the presence of benzaldehyde showed that Yb(OTf)3,

Lu(OTf)3 and Gd(OTf)3 gave the highest yields, and a water to THF ratio of 1:4 was found to be

the most effective solvent mixture. In his study, Kobayashi mentioned that salicylaldehyde is

incompatible with Lewis acids because of the free hydroxyl group, and 2-

pyridinecarboxyaldehyde is also difficult to use with Lewis acids because the nitrogen

coordinates with the metal and resulting in catalyst deactivation.33 The application of Ln(OTf)3 as

catalysts is not only restricted to aqueous solutions. Kobayashi employed organic solvents to

perform the lanthanide triflate catalysed aldol reaction of silyl enolates with aldehydes and

acetals (Scheme 1.2).

1.9

1.9

1.10 1.11

1.12 1.13

PhCOHOSiMe3

OMe

Yb(OTf)3Ph

OH O

OMe

95%

PhCOHOSiMe3

SEt

Yb(OTf)3Ph

OH O

SEt

100%syn:anti 35:65

Scheme 1.2: Lanthanide triflate catalysed aldol reactions of silyl enolates with aldehydes and acetals

It is important to note that these reactions are heterogeneous when performed in organic

solvents without any water present, and the yields are still generally good (75 to 95%).

The lanthanide triflates also catalyse the Michael addition reaction, where the conjugate addition

of silyl enolates to unsaturated ketones occurs to give 1,5-dicarbonyl compounds (Scheme 1.3).

Page 28: Applications of metal triflates and assisted acids as

8

O

Ph Ph

OSiMe3

OMePh

O Ph O

OMeYb(OTf)3 H+

1.14 1.10 1.15

Scheme 1.3: Michael addition reaction catalysed by Yb(OTf)3.

Scandium triflate is reported to be the best catalyst for Michael addition reactions, giving 1,5-

dicarbonyl compounds in 80 to 88% yield with no 1,2-addition products.34 These reactions are

performed at room temperature in the presence of dichloromethane as a solvent, followed by

acid work-up. Furthermore, it is reported that the lanthanides evaluated in the study were

reusable. Also noteworthy is the fact that reactions proceeded even in the presence of 1.0 mol%

of catalyst.35

1.3.2 Diels Alder Reactions

The Diels-Alder reaction forms part of the most important transformations in organic chemistry.

This reaction also leads to the formation a carbon-carbon bond. Furthermore, this reaction

allows the formation of a six-membered cycloadducts of biological importance with fine control

over their steroselectivities. This reaction can be performed in the absence of a catalyst at

elevated temperatures.36 However, some organic compounds are thermally unstable, thus

reactions have to be carried out at mild conditions (temperature and pressure). The reaction of

methyl vinyl ketone with dienes in the presence of Sc(OTf)3 with dichloromethane as a solvent,

followed by water workup gave Diels-Alder products with high selectivity to a desired endo

products (Scheme 1.4).37

1.16 1.17 1.18

OO

Sc(OTf)3

CH2Cl2

91%

Scheme 1.4: Diels-Alder reaction of methyl vinyl ketone with isoprene in the presence of Sc(OTf)3

Page 29: Applications of metal triflates and assisted acids as

9

The Diels-Alder reaction between cyclopentadiene and butyl acrylate in the presence of

Sc(OTf)3 and supercritical carbon dioxide (scCO2) as reaction medium was reported by Rayner

et al.38 During their study, it was noted that the reaction performed in the presence of scCO2

was complete within 15 hours at 50 °C and 6.5 mol% catalyst concentration, while the

uncatalysed reaction was only 10% complete over a period of 24 hours under the same set of

conditions. Furthermore, the reaction afforded high selectivity to the endo product. In the

conventional solvents such as toluene, the reaction yielded 10:1 endo:exo stereoselectivity;

changing to a solvent such as chloroform slightly increased the stereoselectivity to 11:1. On the

other hand, the reaction performed in scCO2 at the same temperature with variation of pressure

allowed the optimisation of selectivity with a maximum of 24:1 endo:exo, which is a significant

improvement of stereoselectivity. The optimisation study also undertaken by Rayner and co-

workers,39 showed that, as the density of the reaction medium was increased (by increasing

pressure), the stereoselectivity also increased and reached a maximum (attained above the

critical conditions of the reaction medium) and then begins to decline. Scandium

perfluoroalkanesulfonate was also found to be an efficient catalyst for Diels-Alder reactions.40

Rayner has further showed that better results are obtained when bulky amines are used during

preparation of the catalyst. For example, a yield of 82% with 85/15 endo/exo adducts and

70%ee to the endo adduct was obtained when diisopropylethylamine was used. In comparison,

though, Et3N gave slightly better yield, but the selectivity was worse (87% yield, endo/exo =

76/24 and 33% ee). Furthermore, they have shown that even better results are obtained when

bulky amines are combined with molecular sieves 4Å (cis-1,2,6-trimethylpiperidine as the

additive, 91% yield, endo/exo=86/14, endo=90%ee), and the enantiomeric excess was further

improved when the reaction was performed at 0 °C.

The metal triflates salts are known to be more stable in the presence of water compared to

metal chloride counterparts.41 Consequently, Rideout et al.42 have reported the use of Sc(OTf)3

as catalyst for Diels-Alder reaction of naphthaquinone with cyclopentadiene in aqueous medium

to give high yields with high selectivity to the desired endo adduct (Scheme 1.5). Interestingly, it

was highlighted that, when this reaction is performed in an organic solvent (CH2Cl2), low yields

are obtained, although the endo adduct selectivity remains high.

Page 30: Applications of metal triflates and assisted acids as

10

O

O

CH2Cl2, 83%, endo/exo (100/0)THF/H2O (9:1): 93%, endo/exo (100/0)

O

O

H

H10 mol% Sc(OTf)3

1.19 1.20 1.21

Scheme 1.5: Sc(OTf)3 catalysed Diels-Alder reaction in aqueous medium

It has also been reported that polymer-supported Sc(OTf)3 has high activity in promoting the

Diels-Alder reaction in aqueous medium (Scheme 1.6).43

O

ON

O

H2O, rt, 12h O

O

CON

cat. (1.6 mol%)

cat.=polymer-supported Sc(OTf)3

(endo/exo = 92/8)99.9%

1.22 1.20 1.23

Scheme 1.6: Diels-Alder reaction in water catalysed by polymer supported Sc(OTf)3

The application of ionic liquids as solvents and catalysts for Diels-Alder reactions has also been

reported in the literature.44 Aggarwal et al.45 have reported the reaction of cyclopentadiene with

methyl acrylate in the presence of ionic liquids. The highest yield (93%) was obtained when

using 1-butyl-3-methylimidazolium hexafluorophosphate [BMIM][PF6] ionic liquid with endo:exo

= 4.8. Consequently, the investigation of the combined effect of metal triflates with ionic liquids

in promoting the Diels-Alder reactions was carried out by Kumar et al.46 The ionic liquids used in

their study are shown in Scheme 1.7 below.

Page 31: Applications of metal triflates and assisted acids as

11

NN(CH2)n X-

n=1 [EMIM]=1-ethyl-3-methylimidazolium

n=3 [BMIM]=1-ethyl-3-methylimidazolium

n=1 [EMIM]=1-ethyl-3-methylimidazolium

X=BF4, PF6, LactateTFA, NTf2

Scheme 1.7: Ionic liquids used in the investigation

The results obtained by Kumar and co-workers have shown that usage of catalytic amounts (2.0

mol%) of metal triflates (La(OTf)3 and Sc(OTf)3) concurrently with ionic liquids to promote Diels

Alder reactions shown in Scheme 1.8 significantly accelerate the reaction rate, as compared to

the reactions performed in ionic liquids in the absence of metal triflate Lewis acids.

O

O

O

O

O

O

O

O

R

O

ON

O

O

R

ON

O

O

R

ON

O

1.20 1.191.21

1.241.25

1.261.27

1.28

1.19

1.20

Scheme 1.8: Diels-Alder reactions catalysed by ionic liquids and metal triflates

Some of the results obtained in the duration of their studies are shown in Table 1.2. Other

examples of Diels-Alder reactions (Scheme 1.8) catalysed by various ionic liquids combined

Page 32: Applications of metal triflates and assisted acids as

12

with metal triflate are also given in the article, which also serves to support the notion of

combined ionic liquid-metal triflate catalysis.

Table 1.2: Diels-Alder reaction of 1 or 4 with 5

Entry Diene Catalyst Ionic liquid Time Yield (%)

1 20 - [BMIM][BF4] 0.5 h 77

2 20 - [BMIM][PF6] 0.5 h 74

3 20 - [EMIM][TFA] 0.5 h 78

4 20 - [EMIM][NTF2] 0.5 h 79

5 24 - [BMIM][BF4] 6 h 42

6 24 - [BMIM][BF6] 6 h 48

7 24 - [BMIM][TFA] 6 h 49

8 24 - [BMIM][NTF2] 6 h 50

9 20 aLa(OTf)3 [BMIM][BF4] 10 min 82

10 20 aSc(OTf)3 [BMIM][BF4] 10 min 84

11 24 bLa(OTf)3 [BMIM][BF4] 2 h >99

12 24 bSc(OTf)3 [BMIM][BF4] 2 h >99

a 1.0 mol% catalyst, b 2.0 mol% catalyst

Kobayashi and co workers47 have reported the use of ytterbium triflates as catalysts for Diels-

Alder reaction at temperatures as low as 0 ˚C (Scheme 1.9). It is noteworthy that catalytic

amounts of the acid were required to promote the reaction and they were able to recover and

reuse the catalyst.

Page 33: Applications of metal triflates and assisted acids as

13

CHO

O

O

O

Yb(OTf)3 (20 mol%)

CH2Cl2, 0 deg.C

Yb(OTf)3 (20 mol%)

CH2Cl2, 0 deg.C

Yb(OTf)3 (10 mol%)

CH2Cl2, 0 deg.C

O

O

86% (exo/endo=10/90)

93% (exo/endo=0/100)

CHO

77% (exo/endo=93/7)

H

O

1.29 1.20

1.20

1.20

1.30

1.311.32

1.19 1.21

Scheme 1.9: Diels-Alder reactions catalysed by Yb(OTf)3 at 0˚C

The Diels-Alder reaction of various electron deficient dienophiles with 1,3-cyclohexadiene under

high pressure (13 kbar) were effectively catalysed by Yb(OTf)3, (Scheme 1.10).

1.33 1.34 1.35

R1 R2

O

13 kbar, CH2Cl2

Yb(OTf)3.H2O (10 mol%)

COR2

R1

60-94%60-95% ee

Scheme 1.10: Reaction catalysed by Yb(OTf)3-H2O under high pressure.

The intramolecular Diels-Alder reaction of oxazole-olefins (Scheme 1.11) are catalysed by Sc(OTf)3, Yb(OTf)3 and most effectively by Cu(OTf)2, yielding pyridine derivatives, as shown in Table 1.3, as reported by Ohba et al.48

Page 34: Applications of metal triflates and assisted acids as

14

R1

R2R2O

NO

M(OTf)3

o-dichlorobenzene150 oC

M= Sc, Cu, Yb

O

R2R2

NR1

1.36 1.37

Scheme 1.11: Intramolecular Diels-Alder reaction of oxazole-olefin

Table 1.3: Metal triflates catalysed intramolecular reaction of oxazole olefin

R1 R2 Catalyst (mol%) Time (h) Yield (%)

H Me - 24 21

Yb(OTf)3 (10) 2 37

Sc(OTf)3 (10) 3 40

Cu(OTf)2 (10) 3 48

Cu(OTf)2 (2)a 1 55

H H - 24 6

Sc(OTf)3 (10) 8 15

Cu(OTf)2 (10) 8 24

CO2Me Me Cu(OTf)2 (2)a 0.5 95

Enantioselective Lewis acid-promoted Diels-Alder reactions have been reported in the

literature.49,50,51 A chiral Yb(OTf)3–derived catalyst was reported by Kobayashi and co-workers

for the Diels-Alder reaction.52 Chiral catalysts are prepared via a reaction of (R)-(+)-binaphthol

with metal triflates, especially those of Yb and Sc and in the presence of dichloromethane and

an amine (Scheme 1.12).

Page 35: Applications of metal triflates and assisted acids as

15

M(OTf)3OH

OH

2.4 eq. amine

0 deg oC, 30 min.

Chiral M(OTf)3

(1.2 equiv.)

M=Sc, Yb

OH

OH

M(OTf)3

N

N1.38

1.39

Scheme 1.12: Preparation of chiral rare-earth M(OTf)3 catalysts.

In their study, Kobayashi and co-workers used the Diels-Alder reaction as a model reaction for

testing the activity of their chiral Yb(OTf)3 catalyst (Scheme 1.13). The use of additives to the

Diels-Alder reactions was of importance in stabilising the catalysts and also helped in controlling

the enantioselectivities. The usage of 3-acetyl-1,3-oxazolidin-2-one as an additive enhanced the

selectivity to the endo-adduct to 93% ee, with (2S,3R) configuration, and the usage of 3-

phenylacetylacetone as an additive gave endo-adduct in 81% ee with (2R,3S) configuration

(Table 1.4).

1.22 1.20 1.23

1.40 1.41

O

ON

O "Chiral M-triflate"(20 mol%)

MS 4A, CH2Cl2, 23 oCO

O

CON

ON

O O O O

Ph

Additive:

Scheme 1.13: Diels-Alder reaction catalysed by “chiral Yb triflate.”

Page 36: Applications of metal triflates and assisted acids as

16

Table 1.4: Enantioselective Diels-Alder reactions catalysed by chiral rare-earth metal triflates.

Catalyst (mol%) Additive Yield (%) endo/exo ee (%), configuration

Yb (20) none 77 89/11 95 (2S,3R)

Yb (20) 40 77 89/11 93 (2S,3R)

Yb (20) 41 83 93/7 81 (2S,3R)

Sc (20) none 94 89/11 92 (2S,3R)

Sc (10) none 84 86/14 96 (2S,3R)

Sc (3) none 83 87/13 92 (2S,3R)

Aza-Diels-Alder reactions (conversion of imines and dienes into tetrahydropyridines, where the

nitrogen atom can be part of the diene or dienophile involved during the Diels Alder reaction)

promoted by metal triflates are also reported in the literature.53 These reactions allow easy

excess to nitrogen containing compounds. Sc(OTf)3 was found by Kobayashi and co-workers to

be efficient as a catalyst for the latter reactions, Scheme 1.14. 54,55 Other metal triflate Lewis

acids such as Yb(OTf)3,56 In(OTf)357 and chiral metal triflate58 salts are also reported to promote

these types of reactions. It is therefore clear that metal triflate salts play a major role as Lewis

acid catalysts for Diels-Alder reactions.

Page 37: Applications of metal triflates and assisted acids as

17

1.42 1.431.20

N

R

Sc(OTf)3 (10 mol%)

CH3CN, rtN

H

H

Ph

R

H

tetrahydroquinolines

Scheme 1.14: Preparation of tetrahydropyridines in the presence of Sc(OTf)3

1.3.3 General Friedel-Crafts alkylation reactions

Friedel-Crafts alkylation is among the most fundamental reactions in the field of synthetic

organic chemistry, and leads to the synthesis of various important aromatic compounds via

formation of C-C bonds. This reaction is widely utilised in both the laboratory and at the

commercial industrial scale for the production of fine chemical and valuable synthetic

intermediates. Anhydrous AlCl3 has maintained its wide use as catalyst for the process ever

since it was introduced by Friedel and Crafts in 1877.59 Since then, a number of Lewis and

Brønsted acids were introduced for the process. This section of the literature survey will focus

on the use of metal triflate Lewis acids as catalysts for Friedel-Crafts alkylation reactions.

It has been reported in the literature that lanthanide triflates and Sc-triflate Lewis acids were

found to be efficient catalysts for Friedel-Crafts reactions.60,61 The benzylation of aromatic

compounds with various benzyl alcohols (Scheme 1.15) using scandium triflate as catalyst was

disclosed by Fukuzawa and co-workers.62 In their study, it was shown that different metal

triflates promote alkylation of benzene (56.1mmol) with benzyl alcohol (1.0 mmol) differently

(Table 1.5).

Page 38: Applications of metal triflates and assisted acids as

18

HO

MX3 H2OM=Yb, Sc, Nd, Sm, Y

X=Cl, OTf1.44 1.45 1.46

Scheme 1.15: Friedel-Crafts alkylation of benzene with benzyl alcohol.

Table 1.5: Lewis acid-catalysed benzylation with benzyl alcohola

MX3 Yield (%)

YbCl3 Trace

ScCl3 Trace

Nd(OTf)3 quantitative

Sm(OTf)3 quantitative

Yb(OTf)3 24

Sc(OTf)3 91

Y(OTf)3 trace

a Reaction conditions: MX3 = 0.1 mmol; 115-120 °C

Their study has shown that metal halides were less active in the alkylation reaction, which could

be due to weak Lewis acidity of the salts. On the other hand, metal triflates of Nd, Sm and Sc

efficiently catalysed the reaction, and the triflates salts of Yb and Y gave low yields. Fukuzawa

and co-workers have further showed that Sc(OTf)3 could catalyse Friedel-Crafts reactions using

various alkylating agents such as allylic alcohols, arenecarbaldehydes and arenecarbaldehyde

acetals with alkylated benzenes in the presence of 1,3-propane diol, giving high yields. It was

also shown that the catalyst is recyclable with a slight loss of activity with the yield from the

original run being 91%, and that of the two subsequent runs being 84%.

Page 39: Applications of metal triflates and assisted acids as

19

In an independent study by Fujiwara and co-workers,63 it was demonstrated that Sc(OTf)3 could

be recycled during benzylation of aromatic compounds using benzyl halides as alkylating

agents. Olah disclosed a patent which describes the usage triflates of boron, aluminium and

gallium as efficient catalysts for Friedel-Crafts alkylation reactions.64 The metal triflates were

prepared via a reaction of triflic acid with the corresponding metal halides.

MX3 3CF3SO3H 3HX

M = B, Al, Ga; X =Cl, Br

M(CF3SO3H)3

The evaluation of the metal triflates was carried out using the Friedel-Crafts alkylation reaction

of toluene with various alkyl halides in a weakly coordinating solvent (dichloromethane) and

strongly coordinating solvent (nitromethane), Scheme 1.16. In Olah’s study, it was observed that

B(OTf)3 was soluble in both solvents, whereas Ga(OTf)3 and Al(OTf)3 were only soluble in

nitromethane, hence were studied as heterogeneous catalysts in CH2Cl2.

R=Me, Eth, i-Pr, t-Bu

R XR

R

R

M(OTf)3

Solvent, 25 deg.C

M=B, Al, Ga; X=Cl, F

Solvent = CH2Cl2, CH2NO2

1.47 1.48 1.49 1.50

Scheme 1.16: Friedel-Crafts alkylation of toluene with alkyl halides

The latter reactions were performed at 25 °C, using toluene/alkyl halide/ catalyst mole ratio of

12:2:1. The yields of methylation of 14-41%, ethylation 21-53%, iso-propylation = 29-60% and

tert-butylation = 30-78% were obtained. In the case of B(OTf)3, dimethylated and diethylated

products (3-8%) were obtained, with 3-6% di-alkylated products obtained during propylation and

butylation. Ga(OTf)3 gave 1-2% diisopropylated and dibutylated products and Al(OTf)3 gave no

apparent dialkylated products. It is known in the literature that the reaction conditions such as

Page 40: Applications of metal triflates and assisted acids as

20

temperature, catalyst amount and, most importantly, solvents have a large effect on the

orientation of products formed.65

The Friedel-Crafts alkylation of aromatic compounds such as anisole, methylnaphthalene, furan,

etc. with α-chloro-α-(ethylthio) acetate in the presence of Yb(OTf)3 was reported.66 The catalyst

showed high activity and selectivity, even during alkylation of furan, affording only

monoalkylated products in high yields (Scheme 1.17), while ZnCl2 afforded mono and

dialkylated products.

O Yb(OTf)3 (5 mol%)

Cl-CH(SEt)CO2Et, CH3NO2, 6h

OCO2Et

SEt

81%

1.50 1.51

Scheme 1.17: Friedel-Crafts alkylation of furan with α-chloro-α-(ethylthio)acetate

The alkylation of aromatic compounds with olefins in the presence of Yb(OTf)3 immobilised in

ionic liquid was also reported.67 (Scheme 1.18).

1.52 1.53 1.54

Yb(OTf)3 (20 mol%)

ionic liquid20 deg.C, 12h

R R

R=H, OH, OMe

NNBuSbF6

-ionic liquid:

Scheme 1.18: Friedel-Crafts alkylation in ionic liquids

During the Friedel-Crafts alkylation of indole with aromatic and aliphatic aldehydes and ketones

to form bisindolyl-methanes (Scheme 1.19), Dy(OTf)3 was found to be most effective among the

lanthanide triflate salts in aqueous medium and the results (Table 1.6) showed high yields when

aldehydes are used as compared to ketones.68 Wang has also reported the use of Dy(OTf)3 and

Y(OTf)3as the best catalyst for the reaction of indole with imines. 69 Other lanthanide triflates (La,

Page 41: Applications of metal triflates and assisted acids as

21

Nd, Eu, Gd and Er) and lanthanide chlorides (Nd, Dy and Y) also catalysed the reaction, while

no reaction occurred in the absence of catalyst.

1.55 1.561.57

R2

O

R3N

R1

HEtOH/H2O12-36 h

(10 mol%)Dy(OTf)3

N

R2 R3R1

N

R1

H H

Scheme 1.19: Alkylation of indoles with aldehydes and ketones using Dy(OTf)3

Table 1.6: Dy(OTf)3 catalysed alkylation of indoles with aldehydes and ketones

R1 R2 R3 Yield %

H Ph H 95

H p-MeOC6H4 H 98

H p-ClC6H4 H 99

H n-Pent H 84

MeO n-Pent H 81

H Me Me 76

H Ph Me 77

The lanthanide triflate salts of La, Nd, Dy, Yb and Sc were used as catalysts for alkylation of

electrophillic arenes with ethylglyoxylate. The highest yields up to 90% were obtained with

Yb(OTf)3. The metal halide Lewis acids did not afford any of the desired products.70 It was also

reported that the Friedel-Crafts reaction of phenolic compounds with α-imino esters was

effectively catalysed by Sc(OTf)3 and high yields of the products were obtained.71 The use of

Hf(OTf)3 as catalyst for Friedel-Crafts alkylation of benzene with benzylchloride in the presence

of lithium perchlorate-nitromethane solvent mixture was reported by Kobayashi et al.72 The

Page 42: Applications of metal triflates and assisted acids as

22

monoalkylated and dialkylated products were obtained in 39% and 29% respectively, and

thereafter the reaction parameters were optimised to get high yields.

1.3.4 Friedel-Crafts alkylation of phenolic compounds

Alkylphenols are very important compounds due to their diverse applications. For instance,

butylated hydroxyl toluene (BHT) in is an important industrial antioxidant, 4-methyl-2,6-ditert-

butylphenol is used in motor and aviation fuel, insulating oils, natural and synthetic rubber

stabiliser. 6-Isopropyl-3-methylphenol (thymol) leads to the formation of menthol when

hydrogenated.73 Menthol is used as raw material for the production of antiseptic, anaesthetic,

antibacterial and antifungal preparations, flavours, fragrances and preservatives.74 Dibutyl m-

cresol is reported to be one of the best softeners for certain types of synthetic rubber,75 while

thymol is also used as an antioxidant.76,77 Alkylphenols are manufactured via addition of the alkyl

group to the corresponding phenols. The alkylating reagents include alcohols, alkylchlorides,

olefins and tertiary ethers such as methyl tertiary butyl ether (MTBE). Olefins are more

preferred due to their low cost and ease of handling. There are numerous homogenous and

heterogeneous acidic catalyst systems that have been reported in the literature to promote

Friedel-Crafts alkylation reactions of phenolic compounds. The advantages associated with

heterogeneous or solid acids for the process are well known and include ease of catalyst

recovery. This section of this literature survey will focus on catalysts that have been cited in the

literature to date for alkylation of phenolic compounds.

1.3.4.1 Friedel-Crafts alkylation of cresols

Weinrich and co-workers have reported the use of concentrated sulphuric acid for butylation of

p-cresol with isobutylene to form 2-tert-butyl-4-methylphenol (mono-butylated p-cresol) and 2,6-

di-tert-butyl-4-methylphenol (di-alkylated p-cresol) or BHT.78 In their study, the dehydration of

tert-butanol to form isobutylene in the presence of activated alumina was carried out at 625 °C

(step1 of Scheme 1.20) and thereafter the butylation reaction was performed at 60-75 °C and

using 5% concentrated H2SO4 (step 2 and 3 of Scheme 1.20). While m- and p-cresols form one

monobutylated cresols, o-cresol formed two mono-butylated products i.e., 4-tert-butyl-o-cresol

and 6-tert-butyl-o-cresols, and on their second butylation only one 4,6-di-tert-butyl-2-

methylphenol is formed.

Page 43: Applications of metal triflates and assisted acids as

23

OHalumina

625 deg.C

OH OH

H2SO4

OH OH

Step1

Step2

Step3

H2O

H2SO4

1.58 1.59

1.59

1.60 1.61

1.59

1.61 1.62

Scheme 1.20: Dehydration of tert-butanol and butylation of p-cresol

It has also been stated that oligomerisation of isobutylene (side reaction) to form diisobutylene

and triisobutylene was also evident. The amount of isobutylene oligomers is dependent on the

reaction time and that, when the reaction is carried out in liquid phase, more oligomers are

obtained compared to a melt phase reaction. It was evident that the butylation reaction

proceeds faster than the oligomerisation reaction. The butylation reactions are exothermic and,

when the heat generation drops, the reaction is almost complete. It is evident that H2SO4 has

been widely used for the alkylation of cresol in the past. For example, Koenig79 used an excess

of H2SO4 for the alkylation of cresol with olefins, while Niedel and Natelson80 used equimolar

amounts of the acid for alkylation below 0°C.

Some alkylated phenols, such as m- and p-cresols, 2,4-, 2,5- and 2,6-xylenols have similar

boiling points and it is therefore difficult to separate these compounds by distillation. In a study

by Sharma and co-workers,81 the alkylation of a mixture of m/p-cresol with α-methyl styrene was

carried out in the presence of a homogeneous para-toluenesulfonic acid (pTSA) and with a

heterogeneous ion exchange resin (Amberlyst 15) at 60-100 °C, and cumene was used as a

solvent. It was observed that the alkylation of cresols resulted in the formation of di-alkylated

cresol products. On prolonged stirring at reaction temperature, dealkylation occurred giving

Page 44: Applications of metal triflates and assisted acids as

24

back the cresol. It was also observed that at higher temperature (100 °C) Amberlyst 15 formed a

substantial amount of phenylindan, while pTSA promoted dialkylation of p-cresol. There was no

significant difference in the rate of conversion of both m- and p-cresols between heterogeneous

and homogeneous catalysts. At 33 mol% conversion of α-methyl styrene, the ratio of p-cumyl-

m-cresol to o-cumyl-p-cresol was found to be 77:33. This mixture was successfully separated

via dissociation extraction (the mixture was treated with 50% w/w cumene and 20% aqueous

NaOH) and 95% pure p-cumyl-m-cresol was obtained. Cracking of the alkylated isomers (p-

cumyl-m-cresol: o-cumyl-p-cresol; 95:5) under mildly acidic conditions and atmospheric

pressure gave back 98%of m-cresol and 2% of p-cresol at 50% cracking level, suggesting that

p-cumyl-m-cresol underwent cracking at faster rate than o-cumyl-p-cresol.

The kinetics for tert-butylation of p-cresol with isobutylene in the presence of an ion exchange

resin (Amberlyst 15) was carried out by Santacesaria and co-workers.82 The kinetics of all the

possible reactions were considered in their study, i.e. mono-alkylation and successive di-

alkylation of cresol, dimerisation and trimerisation of isobutylene. A second order rate law was

found to be suitable to describe the behaviour of all the reactions involved.

Santacesaria et al.83 have also used Amberlyst 15 as a catalyst for alkylation of p-cresol with

isobutylene, which was selected as a model reaction when performing comparative studies

between a well-stirred slurry reactor and a spray tower loop reactors. The quantitative

conversion of p-cresol was obtained. They have also found that the reaction occurs in two

steps, the first step being mono-alkylation of p-cresol to give 2-tert-butyl-p-cresol and the

subsequently di-alkylation takes place, giving 2,6-di-tert-butyl-p-cresol. The side reaction where

dimerisation and trimerisation of isobutylene also occur was evident. In their study it was found

that spray tower loop reactor has low performances at lower temperatures (60 °C). At 72 °C the

performances are comparable. According to Santacesaria, the explanation to the differing

behaviour of the two reactors is that at low temperature the spray nozzle is not efficient enough

to produce very small droplets for the high viscosity liquid.

Sulfated zirconia (S-ZrO2) is a super acid with a Hammett acidity (H0) value of -16.04.84 The

latter acid was reported by Yadav and co-workers as an efficient catalyst for the alkylation of p-

cresol with isobutylene.85 It was found that the S-ZrO2 exhibits more surface area and high

selectivity towards alkylated products as compared to Amberlyst 15. Yadav has also reported

the use of sulfated zirconia (20% w/w dodecatungstophosphoric acid supported on K10 clay and

ZnCl2/K10) as catalyst for alkylation of p-cresol with cyclohexene.86 The catalyst was highly

selective to 1-cyclohexyloxy-4-methylbenzene (O-alkylated product).The C-alkylated product (4-

Page 45: Applications of metal triflates and assisted acids as

25

cyclohexyl-4-methyl phenol) was also formed. The production of O-alkylated product is favoured

by lower temperatures (80 °C) and C-alkylated product by higher temperatures. The

comparative study of the activity of Nafion® resin/silica composite and Amberlyst 15 as catalysts

for alkylation of p-cresol with isobutylene was undertaken by Harmer and coworkers. They have

shown that a Nafion® resin/silica composite is more effective (Table 1.7). The latter catalyst also

has the advantages of being used at high temperatures, while the Amberlyst 15 is thermally

unstable.

Table 1.7: The products afforded by Nafion® resin/silica and Amberlyst 15.

Catalyst Conversion (%)

Selectivity

(%)

Alkylation rate

(mM/g cat H)

Ether Alkylates

13% Nafion®/SiO2 82.6 0.6 99.4 581

Nafion®-NR50 19.5 28.2 71.8 54.8

Amberlyst15 62.4 14.5 85.5 171

The use of 12-tungstophosphoric acid supported on zirconia as a catalyst for butylation of p-

cresol with tert-butanol (Scheme 1.21) was studied by Halligudi and co-workers.87 The reaction

gave three alkylation products (2-tert-butyl-p-cresol, 2,6-di-tert-butyl-p-cresol and an ether. The

ether is formed when p-cresol undergoes alkylation on the oxygen (O-alkylation), water is

formed as by-product of this side reaction.

Page 46: Applications of metal triflates and assisted acids as

26

Ether

OH

catalystOOHOH

H2OOH

1.62

1.58

1.61 1.62 1.63

Scheme 1.21: Alkylation of p-cresol with tert-butanol in the presence of TPA/ZrO2

Halligudi and co-workers optimised the reaction conditions by focussing on the effect

temperature, space velocity and molar ratio of reactants. Their study showed that a catalyst with

15% loading of TPA on ZrO2 calcined at 750 °C was the most effective for butylation reactions.

The study also showed that the optimum conditions were: temperature of 150°C, with tert-

butanol to p-cresol mole ratio of 3:1 and liquid hourly space velocity of 4 h-1. Under the optimised

conditions, conversion of p-cresol was found to be 61 mol%, with 81.4% selectivity to 2-tert-

butyl-p-cresol, 18.1% to 2,6-di-tert-butyl-p-cresol and 0.5% tert-butyl-p-tolyl ether (ether). The

catalyst showed loss of activity in terms of cresol conversion to be 6%, with catalyst time on

stream of 100 hours.

Halligudi et al.88 has performed a similar study, where the reaction of p-cresol with tert-butanol

was carried out in the presence of tungsten oxide (WO3) supported on zirconia (WO3/ZrO2). In

this study, different loadings of WO3 (3-30 wt%) were prepared and calcined at 800 °C. The

catalyst with 15% WO3/ZrO2 was found to be most effective. At the optimum conditions

(temperature of 130°C, tert-butanol to p-cresol mole ratio of 3:1 and flow rate of 10 ml.h-1) the p-

cresol conversion of 69.8% with 92.4% selectivity to 2-tert-butyl-p-cresol, 6.3% to 2,6-di-tert-

butyl-p-cresol and 1.3% to tert-butyl-p-tolyl ether (ether) were obtained. In the course of their

study, the reactions catalysed by sulphated zirconia, USY- and Hβ-zeolites as well as K-10

montmorillonite at optimum conditions were performed and these catalysts (USY-, K-10 ans Hβ-

gave low conversion (25%) of p-cresol.

The tert-butylation of cresols (including phenol) with tert-butanol in the presence of FeCl3-

modified montmorillonite K-10 as catalyst was performed by Samat and co-workers.89 During

Page 47: Applications of metal triflates and assisted acids as

27

butylation reactions, high excess of phenol was used, the products formed are shown in

Scheme 1.22.

1.64 1.65 1.66 1.67 1.68

OH OHOHOHO

tert-BuOH

K10-Fe catalyst

Scheme 1.22: tert-Butylation of phenol

The phenol butylation reactions were performed at 80 °C and 100% butylation was obtained

with all the K10 catalysts. The Fe-modified K10 catalysts exhibited high activity, while the

unmodified catalyst gave low rates of reaction (Table 1.8). The O-alkylated product 1.65 did not

form in these reactions, and the authors suggest that selectivity to O-alkylation is dependent on

the acid strength of the catalyst. In general, C-alkylation requires stronger acid sites while the O-

alkylation requires weaker acid sites

Table 1.8: Butylation of phenol with tert-butanol using K10 catalysts at 80 °C

Catalyst Time (h)c Selectivity (%)

1.66 1.67 1.68

K10-120 6 35.4 62 2.6

K10-Fe-O120a 1 34.8 62.2 3.0

K10-Fe-A120b 0.5 30.5 66.8 2.7

aCatalyst prepared in acetonitrile, bCatalyst prepared in water, cTime for 100% butylation

In their study, it was found that K10-Fe-120 was the most effective catalyst for butylation of

phenol, hence the latter catalyst was evaluated for butylation of cresols with tert-butanol

(Scheme 1.23). The quantitative conversion of cresols was observed within 30 min. in all

Page 48: Applications of metal triflates and assisted acids as

28

instances. Dibutylation of m- and p-cresol was evident, while only mono-alkylated products were

obtained in the case of o-cresol. Interestingly, this is the first report where 3-tert-butyl-p-cresol is

observed, and presumably arises due to the presence of the activating p-CH3 group. They have

also shown that the catalysts could be recycled three times with no significant loss of activity

and selectivity.

OH

OH

OH

OHOHOH

OHOHOH

OHOHOH

(17%) (83%) (0%)

(86.7%) (11.5%) (1.8%)

(11.6%) (86.7%) (1.7%)

1.69 1.70 1.71 1.72

1.73 1.74 1.75 1.76

1.60 1.61 1.61 1.62

Scheme 1.23: Butylation of cresols with tert-Butanol using K10-Fe-120 at 80 °C.

Yadav et al.90 have reported the butylation of p-cresol with MTBE in the presence of zirconia

based-mesoporous superacid catalyst (UDCaT-1),91 Scheme 1.24. UDCaT-1 is prepared by via

combination of hexagonal mesoporous silica and sulfated modified zeolite. This reaction occurs

in two steps. In the first step, cracking of MTBE to form isobutylene and methanol takes place.

In the second step, the addition isobutylene to p-cresol forming 2-tert-butyl-p-cresol, followed by

addition of another molecule to eventually 2,6-di-tert-butyl-p-cresol. According to Yadav,

isobutylene formed in situ is quantitatively consumed and none of the isobutylene is in the gas

phase, hence no oligomerisation side reaction occurs. Furthermore, it was noticed that no O-

alkylated products were formed above 100 °C.

Page 49: Applications of metal triflates and assisted acids as

29

OH OHOH

OH+

CH3OH

H+ H+

1.58 1.59

1.60 1.61 1.62

Scheme 1.24: Butylation of p-cresol with MTBE

All the catalysts evaluated in their study were selective to 2-tert-butyl-p-cresol (

Table 1.9).. However, UDCaT-1 gave the highest conversion. In order to avoid excessive

formation of isobutylene, the conversion of p-cresol was limited to 50% using 1:1 mole ratio of p-

cresol to MTBE, stirring at 700 rpm and the reaction performed at 100 °C to avoid the

oligomerisation side reaction. Hence less than 1% oligomers were obtained. At higher

temperatures, up to 10% oligomers were formed. UDCaT-1 was proven to be recyclable, with

conversions being marginally lower by 5%from the previous use.

Table 1.9: Activity of various acids for butylation of p-cresol with MTBE

Catalyst Conversion (%)

Selectivity to 2-tert-butyl-p-cresol (%)

UDCaT-1 45 97

Indion 130a 39 92

Fitrol-24b 19 96

Sulfated zirconia 15 91

K-10b 12 96

DTPA/K-10 30 96

aion exchange resin; bclay; c20% dodecatungstophosphoric acid (DTPA) on clay

Page 50: Applications of metal triflates and assisted acids as

30

The isopropylation of m-cresol over three different mesoporous molecular sieves was carried

out by Murugesan and colleagues.92 In their study, the reaction of m-cresol with isopropanol

was performed in the presence of Al-MCM-41 molecular sieves with Si/Al ratios of 59, 103 and

202. Three products were obtained from the reaction, i.e., 2-isopropyl-4-methylphenol 78,

isopropyl-3-methylphenyl ether 79 and isopropyl-(2-isopropyl-5-methylphenyl) ether 81, Scheme

1.25. Of the three catalysts studied, Al-MCM-41 (59) was found to be more stable than Al-MCM-

41 (103) and Al-MCM-41 (202). At the mole ratio of 1:2 of m-cresol to alcohol, good conversion

of m-cresol was obtained with 100% selectivity to 78.

OHOH

OOH

OHOH

O

1.73

1.77

1.78 1.79

1.80

1.81

Scheme 1.25: Alkylation of m-cresol with isopropanol

A number of patents have been granted on the alkylation of cresols with olefins to form alkylated

cresols. This patent overview will focus on the catalysts used for alkylation of cresols with

olefins. Wetzel et al.93 have reported the use of highly acidic aryl sulfonic acids including para-

chlorobenzenesulfonic acid, nitrobenzenesulfonic acid, 4,4-diphenyldisulfonic acid, 2,4,6-

trinitrobenzenesulfonic acid, trifluoromethane sulfonic acid and meta-benzenedisulfonic acid for

this purpose. They performed butylation of o-cresol (647 g) with isobutylene (300 g) in the

presence of m-benzenedisulfonic acid (5 g) at 150 °C for 1 hour and 15 minutes. The product

was found to contain 79% 4-tert-butyl-ortho-cresol.

The butylation of m/p-cresol mixture (67% m-cresol and 33% p-cresol) with isobutylene in the

presence of H2SO4 (0.5 wt%) as catalyst was disclosed by Hess.94 The butylation reaction was

Page 51: Applications of metal triflates and assisted acids as

31

performed at 70 °C, under atmospheric pressure and starting off with 100 moles of the cresol

mixture. The product was neutralised with NaOH. The analysis showed the presence of mono-

alkylated m- and p-cresol to be 47.5% and 13.5% respectively, 17.0% and 19.0% mole of di-

alkylated m- and p-cresols respectively and finally, 2.5% mole of m-cresol and 0.5% mole of p-

cresol remained un-reacted. The reaction was stopped at 68.5% extent of butylation (to reduce

oligomerisation of isobutylene).

In a similar invention disclosed by Stevens et al.,95 the alkylation of m- and p-cresol mixture

(60% m-cresol and 40% p-cresol) with isobutylene in the presence of H2SO4 or AlCl3, BCl3 and

FeCl3 was carried out. Their invention consists of first alkylation of both m- and p-cresol

quantitatively, forming di-alkylated products. Unlike in the invention by Wetzel (where di-

butylation is allowed to proceed to some extent followed by separation), the latter invention

leads to a complete di-butylation of cresols, followed by separation of the two di-butylated

products via distillation. The di-butylated-p-cresol is already a valuable antioxidant, and the less

valuable di-butylated m-cresol is further treated by mixing it with pure m-cresol in the presence

of H2SO4 and heating the mixture at 70-80 °C. Since Friedel Crafts alkylation is reversible, de-

butylation to yield one equivalent of isobutylene molecule from di-butylated meta-cresol occurs.

The isobutylene molecule then alkylates the added m-cresol. The end product is mostly

composed of mono-butylated m-cresol.

In a separate patent by Stevens and Bowman,96 the reaction of isobutylene with m-cresol was

performed with an aim of synthesising 4-tert-butyl-m-cresol in the presence of H2SO4 as

catalyst. However, it is mentioned that other acids such as H3PO4; AlCl3; BF3; FeCl3 and HCl

could be used. Apart from the desired product, a mono-butylated isomer 6-tert-butyl-m-cresol

and the di-butylated product 4,6-di-tert-butyl-m-cresol were obtained. In their invention, a 0.8:1

mole ratio of m-cresol to isobutylene was used with a catalyst concentration being between

0.001-0,1 wt% with respect to m-cresol. A reaction temperature between 0 and 40 °C is

preferred: according to the invention, temperatures higher than 40 °C favour the formation of 6-

tert-butyl-m-cresol. When the reaction is complete, caustic is added to neutralise the acids.

They have found that the desired 4-tert-butyl-m-cresol is soluble in caustic together with un-

reacted m-cresol, which then provides a means of separation. The undesired products (6-tert-

butyl-m-cresol and 4,6-di-tert-butyl-m-cresol) may be subjected to dealkylation to reproduce m-

cresol and isobutylene, which is then recycled to form 4-tert-butyl-m-cresol.

A patent by Gershanov97 disclosed the use of aluminium catalysts for the preparation of di-

alkylated p-cresol via a reaction of p-cresol with olefins. Dodd et al. filed a patent where the

Page 52: Applications of metal triflates and assisted acids as

32

xylenols are alkylated with isobutylene in the presence of Amberlyst 15.98 The separation of a

mixture of m/p-cresol via the reaction of cresols with isobutylene in the presence of acid

catalysts such as aluminium chloride, ferric chloride, boron trifluoride, phosphoric acid, and

sulfuric acid was disclosed by Stevens et al.99 In their invention, the separation of di-butylated

m- and p-cresols via distillation is performed. The di-butylated p-cresol is a valuable antioxidant

(BHT) and is retained, while the dibutylated-m-cresol is of lower value and may be subjected to

de-alkylated which leads to reformation of valuable m-cresol as an essentially pure isomer.

The use of hydrogen fluoride as catalyst for the alkylation of m-cresol with isobutylene in the

presence of carbon dioxide was disclosed in a patent by Hervert.100 His research showed that

the butylation reaction performed in the absence of CO2 leads to lower yields. The invention by

Kurek,101 shows that butylation of p-cresol with isobutylene in the presence of a highly cross-

linked polystyrene divinylbenzene resin with an internal surface area of about 540 m2/g and an

average pore side of 51 Å, at 100 °C and 100 psig afforded 94% 2,6-di-tert-butyl-p-cresol, 9% 2-

tert-butyl-p-cresol and 99% of the reactant was converted to a product. On the contrary,

Amberlyst 15 yielded 99% conversion of p-cresol, but only 27% 2,6-di-tert-butyl-p-cresol and

68% 2-tert-butyl-p-cresol under the same set of conditions. A process for separation of a

mixture consisting of 30% p-cresol and 70% m-cresol via alkylation with isobutylene in the

presence of H2SO4 was also resported by Luten and De Benedictis.102 The reaction was

performed at 90 °C, for 12 minutes and the resulting mixture was neutralised with caustic, the

resulting products, yielding the un-reacted cresols; mono-butylated cresols and di-butylated

cresols which were eventually subjected to fractional distillation.

1.3.4.2 Friedel-Crafts alkylation of anisole

Alkylated anisoles are important industrial compounds which are used as antioxidants, dye

developers and stabilisers for fats, oils, plastic rubbers, etc.103 The preparation of 4-tert-

butylanisole via alkylation of anisole with tert-butyl alcohol in the presence of ZrCl4 and

trifluoroacetic acid was reported by Sartori and co-workers.104 Alkylation of anisole with tert-butyl

acetate as an alkylating agent in the presence of H2SO4 and also with tert-butyl nitrate in the

presence of SnCl4105 as a catalyst, was also reported by Fernholz et al.106 Other catalysts such

as triflic acid,107 Amberlyst 15,108 silica-alumina,109 zeolites110 and activated clays111 have been

reported for the alylation of catechol.

Page 53: Applications of metal triflates and assisted acids as

33

The reaction of anisole with isobutylene generated from MTBE in the presence of 20%

dodecatungstophosphoric acid supported on K10 montmorillonite clay (20% DTPA/K10) was

reported by Yadav and colleagues.112 They found that the reaction of anisole with isobutylene

yield three products, two of which are mono-butylated anisoles (2-tert-butylanisole and 4-tert-

butylanisole), and one di-butylated product (2,4-di-tert-butylanisole) was also present,

(Scheme 1.26:).

1.821.59

1.83

1.84

1.85

O O O O

H+

Scheme 1.26: Butylation of anisole with isobutylene

The catalyst (20% dodecatungstophosphoric acid supported on K10 montmorillonite) was found

to be the best among other catalysts evaluated. The best reaction conditions for the reaction of

anisole were found to be, 170 °C and anisole to MTBE mole ratio of 1:3 and catalyst loading of

1.5 g. The results reported on the alkylation of anisole with n-propanol, using Amberlyst 15 as

catalysts in the presence of supercritical CO2 showed high selectivity (96.4%) to mono-alkylated

products in a monophasic system.113 The alkylation of anisole with primary and secondary

olefins in the presence of niobium phosphate was reported by Lachter and co-wrokers.114

Niobium phosphate gave high selectivity to monoalkylated anisole products and no dialkylated

anisoles were formed under the conditions. Lachter has also reported the use of niobium

phosphate and niobic acid as catalysts for alkylation of anisole with 1-octene-3-ol.115 Their study

showed that niobium phosphate gave higher activity than niobic acid. A quantitative conversion

of the alcohols was obtained with high selectivity (>80%) to mono-alkylated products. The two

mono-alkylation products were observed including linear and branched alkylated anisoles.

The alkylation of anisole with benzyl alcohol in the presence of niobia supported on alumina was

also reported by Lachter et al.116 In this study, the unsupported Nb2O5, AlO3 and alumina-

supported niobia were evaluated. The study showed that alumina exhibited low activity and poor

selectivity to alkylated products. On the other hand, niobia and alumina supported niobia gave

high activity and selectivity to alkylated products. The reaction of anisole with various olefins

Page 54: Applications of metal triflates and assisted acids as

34

was reported by Kretchmer et al.117 in the presence of AlCl3. The reaction resulted in the

formation of mostly ortho-alkylated products. They have found that the extent of ortho-alkylation

is dependent of the solvent and functionality of the alkylating agent.

The Friedel-Crafts alkylation of anisole with n-propanol in the presence of supercritical carbon

dioxide (scCO2) was investigated by Poliakoff and colleagues.118 In their study, the activity of

five different heterogeneous acids, including ion exchange resins (Amberlyst 15 and purolite

CT-175), inorganic supported catalysts (Nafion SAC-13 and Deloxan ASP I/7) a zeolite (Zeolyst

CBV 600) were investigated. The results showed that Amberlyst 15 and Purolite CT-175

provided the best performance between 100-150 °C. At temperatures above 150 °C,

desulfonation of the catalysts was evident, resulting in a decrease of the yield. The inorganic

supported catalysts (Zeolyst CBV 600, Nafion SAC-13 and Deloxan ASP I/7) showed a higher

optimum temperature range, but the reaction became less selective to mono-alkylated products.

The alkylation of anisole with various dienes such as 2-methyl-1,3-butadiene (isoprene), 2,3-

dimethyl-1,3-butadiene and 1,3-cyclohexadiene using K10 montmorillonitres exchanged with

different cations as catalyst was reported.119 Their study showed that para-mono-alkylation is

the preferred reaction and the regio selectivity in the diene is controlled by the attack on the

least hindered position of the most stable carbocation. It is also mentioned that the catalysts

were partially deactivated under the conditions used, and that the yield may be increased by

adding the dienes in small portions.

The alkylation of anisole with tert-butanol in the presence of mesoporous aluminosilicate (Al-

MCM-41) molecular sieves with Si/Al ratios of 25, 50, 75 and 100 was reported by Pandurangan

et al.120 The alkylation reactions were performed at temperatures between 150 and 250 °C

under atmospheric pressure and the results showed that Al- MCM-41 (25) was the most active

catalyst. The major products were found to be 4-tert-butyl anisole, 2-tert-butylanisole and 2,4-di-

tert-butylanisole. The maximum conversion of anisole was observed at 175 °C and thereafter it

decreased with an increase of reaction temperature.

In the alkylation of anisole with isoamylenes in the presence of zinc chloride on alumina at 200

°C and with Kationit KU-1 at 100 °C,121 two products were obtained from the reaction, i.e., p-

pentylanisole and 2,4-di-pentylanisole. Among the catalysts evaluated, zinc chloride showed

better catalytic activity.

The use of tungsten oxide supported on zirconia (WOx/ZrO2) as catalyst for alkylation of anisole

with 1-dodecene between 95-100 °C was reported.122 The reactions were performed using an

Page 55: Applications of metal triflates and assisted acids as

35

olefin to anisole mole ratio of 2.25:1 and 3.2% WOx/ZrO2 with respect to anisole. After the

removal of a used catalysts and purification of the crude product by distillation, 94% yield

colourless oil was obtained consisting of mono-, di-, and tri-alkylated anisole.

The alkylation of anisole with tert-butanol using molybdenum pentachloride (MoCl6) as catalyst

was reported by Qiaoxia et al.123 The mole ratio of anisole to tert-butanol to MoCl6 used in the

study was 1:1.2:1. MoCl6 exhibited high yield of 4-tert-butylanisole and only 6% of 2,4-di-tert-

butylanisole, 2-tert-butylanisole was not present.

1.3.4.3 Friedel-Crafts alkylation of diphenylether (DPE)

The alkylation of diphenyl ethers forms part of the present study where the Lewis acids and

Lewis assisted Brønsted acids were evaluated for butylation of substituted phenols. Hence, a

literature review of the catalysts used to date for the reaction was undertaken. The alkylation of

diphenyl ether with olefins, alcohols and alkyl halides yields commercially important products.

There are numerous applications of alkylated DPE such as heat transfer fluids.124 Alkylated DPE

can be used as a dielectric agent in transformers as claimed in a patent of Westinghouse

Electric Corporation.125 This patent discloses the use of up to 99 wt% of mono ethylated, mono

propylated, mono-butylated DPE and up to 20 wt% dialkylated DPE as dielectric fluid in a

transformer.

Coleman et al.126 disclosed the use of AlCl3 as catalyst for alkylation of DPE with olefins

(propylene and diisobutylene), alkyl halides (ethyl chloride, sec-butyl chloride, tert-butyl chloride

and tert-butyl-amyl-chloride) and lauryl alcohol. The reactions of tert-butyl chloride with DPE

were performed using 1:1 mole ratio of tert-butyl chloride to DPE in the presence of 0.2 mol of

catalyst at 100 °C. When the reaction mixture was distilled, the first fraction consisting of mono-

tert-butyl-DPE (clear and clourless liquid) with a boiling point of 150.5-152.5 °C at 6 mm

pressure was obtained, followed by an isomeric crystalline mono-tert-butyl-DPE boiling at 153.5-

155.5 °C under 6 mm pressure. This pure isomer recrystallised from ethanol and had a melting

point of 54-54.5 °C. The liquid di-tert-butyl-DPE was also isolated via distillation at 191.3 °C

under 6 mm pressure. The quantities of mono-alkylated products were 26% for the liquid isomer

and 41% for the crystalline isomer. The former isomers are believed to be ortho- and para-

isomers, respectively. The remaining 31% is for the di-tert-butyl-DPE. Klingel and Ellison also

used AlCl3 as catalyst for alkylation of DPE with diisobutylene to form di-tert-octyl DPE.127

Page 56: Applications of metal triflates and assisted acids as

36

A process for the alkylation of DPE with 1-hexadecene in the presence of zeolite catalyst (MCM-

22 and the Engelhard catalyst (USY)) was disclosed by Wu and Lapierre.128 The reaction was

performed using 1:1 mole ratio of DPE to 1-hexadecene at 175 °C and the mixture containing

DPE (42.9%) and 1-hexadecene (57.1%) was fed into the reactor at 4g/hour. Wu and Lapierre

observed that the catalyst activity remained constant for 180 hours, and that when the reaction

was terminated the catalyst was still active. In their study they obtained overall conversion of

83% and the conversion of 1-hexadecene was 96%, while that of DPE was 66%. The selectivity

to alkylated products was >95%. Furthermore, their study showed that zeolite MCM-22 exhibit

higher activity than USY catalyst. USY catalyst gave a total conversion was <10%.

Rudnick et al.129 has disclosed a patent where the alkylation of DPE with C14, C16 and C18 olefins

was carried out using zeolites as catalyst. The examples given demonstrated that the zeolite

gives 98-100% monoalkylated compounds as compared to non-zeolite catalysts such as BF3,

which give 56% monoalkylated products and 44% polyalkylated products. Three zeolites have

been found to be active for the alkylation of DPE with 1-hexadecene, namely:130

(1). BB10, a USY-zeolite with 20% exchange sites occupied by sodium cations and the

remaining occupied by protons.

(2). 500PN, a USY-zeolile with 100% ammonium ions in the exchange sites

(3). DD-12, a USY-zeolite with fully protonated exchange sites.

The reactions were carried out using 1:2 mole ratio of 1-hexadecene to DPE, with 1wt% of

catalyst. The results showed that the zeolites are highly selective to mono-alkylation.

Furthermore the controlled additions (metering valve) of the olefin to the reaction mixture gave

complete reaction of olefins at low level of zeolite. On the other hand, bulk addition of the olefin

resulted in quick deactivation of the catalyst prior to completion of the reaction. The alkylation of

DPE with C1-C7 alcohols using zeolites as catalysts was reported131

The alkylation reaction of DPE with isobutylene in the presence of concentrated aqueous H2SO4

and AlCl3 was reported by Cobb and Mitchell.132 In one instance they performed a reaction of

isobutylene with DPE using 9 g of AlCl3 as catalyst and the olefin to ether mole ratio of 4.18:1 of

at 60 °C. The results showed the presence of 2% mono-alkylated DPE, 41% di-alkylated DPE

and 20% tri-alkylated DPE. The product was pale-yellow oil with low viscosity. When they

performed a similar reaction with an olefin to ether mole ratio of 6.4:1, only 2% mono-alkylated,

31% di-alkylated and 27% tri-alkylated products were obtained. The product obtained from the

latter reaction also had low viscosity (918 CPS at 25 °C). Cobb and Mitchell have also evaluated

Page 57: Applications of metal triflates and assisted acids as

37

other Lewis acid type catalysts such as BF3•H3PO4, Amberlyst 15 and H3PO4•P2O5, which also

gave products of low viscosity (50-300 CPS at 25 °C).

The use of H2SO4 (10 ml) as catalyst for the reaction of DPE with isobutylene at 27 °C gave 1%

mono-alkylate, 19% di-alkylate and 63% tri-alkylated products, the viscosity of the products was

5600 CPS at 25 °C. When the olefin to ether mole ratio was increased to 7.15:1, 1% mono-

alkylate, 25% di-alkylate and 57% tri-alkylated products were obtained with viscosity of 2447

CPS at 25 °C. Hence Cobb and Mitchell have showed that the viscosity of alkylated DPE can be

varied by changing the olefin to ether mole ratio. The use of AlCl3 as catalyst for of DPE with

hexadecane at 90 °C for 6 hours was reported by Yunqin et al.133 The mole ratio of DPE, AlCl3

and hexadecane is 16:1.14-2.28:9.4-11.4. The hexadecyl DPE was subsequently sulfonated

with Na2SO4 to yield sodium hexadecyl-DPE-disulfonate which is used as a surfactant of the oil

displacing agent for the tertiary oil recovery. The alkylation of DPE with fatty alcohols (alkylating

agent) in the presence of H2SO4 at 40-100 °C for 1-10 hours was reported.134

The use of chloroaluminate ionic liquids with an anion constituent being aluminium trichloride

dimer (Al2Cl7) and the cation constituent being quaternary ammonium salt (pyridine, chloro-

alkylpyridien or alkylimidazole) with a mole ratio of anion to cation being 1.5-3.1:1 as catalysts

for alkylation of DPE with alkenes was reported.135 In that study, a DPE to olefin mole ratio of

30:1 was used and the reaction was performed at 120-220 °C. The mono-alkylation of DPE

with 1-dodecene in the presence of an ionic liquid was also reported elsewhere.136, 137 The ionic

liquid used in the study was 1-butyl-3-imidazolium aluminium chloride {[BMIM]Cl-AlCl3} and the

results showed that a selectivity of 34% of 5- and 6- dodecyl-DPE could be obtained at 80 °C,

when the mole ratio of ionic liquid to the olefin was 0.5 and mole ration of AlCl3 in ionic liquid

was 0.67.

The use of sulfated zirconia as catalyst for alkylation of DPE with enzylchloride was reported by

Yadav and co-workers.138,139 In their study, it was found that optimum yield is obtained when

using the mole ration of DPE to benylchloride of 7:1 at 110 °C with a catalyst loading of 50

kg/m3 and stirring the reaction at 20-25 rps. It was also found that the particle size of 74-88

microns gave the best activity and that the catalyst is reusable.

The application of heteropolyacids such as tungstosilicic acid on SiO2 as catalyst for alkylation

of DPE with propylene was reported by Hasebe et al.140 Undecane was used as solvent in the

reactions at 160 °C for 2 hours to give 24% isopropyl-DPE and 67.4% di-isoporpyl DPE. Rennie

and colleagues have reported the use of SiO2-Al2O3 as catalyst for alkylation of DPE with

isobutene or propene at 140 °C and 5 bar (propene) to give 71% propylated DPE containing

Page 58: Applications of metal triflates and assisted acids as

38

35% o- propylated DPE.141 The alkylation of DPE with isobutylene using an ion exchange resin

(KU-2) in the H form as catalyst was reported.142

1.4 Etherification of alcohols and olefins Oxygenated compounds such as MTBE, tert-amyl methyl ether (TAME) and tert-amyl ethyl

ether (TAEE) are known to improve the burning efficiency of gasoline and reduce the carbon

dioxide emissions because of their low volatility.143 These oxygenates also help to reduce the

formation of atmospheric ozone resulting from gasoline emissions. In the past MTBE was the

most common oxygenate used as an additive to gasoline for improving the octane number.

However, the water pollution problems created by extensive use of MTBE diverted the focus of

researchers to alternative fuel additives such as tert-amyl methyl ether (TAME) and tert-amyl

ethyl ether (TAEE) as octane enhancing gasoline blending components.144 The ethers such as

TAME and TAEE derived from the reaction of C6 olefins with methanol have other properties

(e.g. higher energy density, low water solubility, hence, higher molecular weight) that are

desirable for reformulated gasoline.145 These tertiary ethers are produced via the reaction of

isoamylene olefins {2-methyl-2-butene (2M2B) or 2-methyl-1-butene (2M1B)} with methanol and

ethanol, respectively, in the presence of homogenous or heterogeneous acid catalyst systems,

Scheme 1.27.

1.88

1.86

1.87

CH3OH O

(2M1B)

(2M2B)

H+

Scheme 1.27: Etherification reaction of methanol and isoamylene

Page 59: Applications of metal triflates and assisted acids as

39

This section of literature review will focus mainly on the catalysts reported for the production of

TAME via etherification of alcohols and isoamylenes. Amberlyst 15 as a catalyst for the

etherification of olefins and alcohols has been widely reported in literature146,147 as a catalyst for

the production of TAME. Wang and James have also reported the use of Amberlyst 15 and

sulphated zirconia for the same purpose.148,149 They have carried out the reactions in a 25 ml

stainless steel batch reactor under helium atmosphere (1.8 MPa) at 80 °C for 2 hours. An

alcohol to olefin mole ratio of 1:1 was used with a catalyst loading of 0.5 g and the solvent being

heptane (4 g). It was found that the olefin undergoes isomerisation to a low extent, i.e. 2M1B

can isomerise to 2M2B, but both olefins react with methanol to form the same ether (TAME). A

small amount of 2,3-dimethyl-2-butanol was detected which is potentially formed via the reaction

of olefins and trace amounts of water in catalysts and reactant solution. Amberlyst 15 afforded

91% conversion of 2M1B and an ether yield of 18.8% with selectivity of 78.7% to 2M2B and only

20.6% to the ether product, while sulphates zirconia supported on SiO2 gave 2M1B conversion

of 51.8%, the ether yield of 29.8% and selectivities of 57.6% and 41.7% to the ether and 2M2B

respectively. Hence the results showed that while Amberlyst 15 gave high conversion of 2M1B,

lower ether, selectivity was afforded by the catalyst. Sulphated zirconia gave low conversion

and higher selectivity to the ether, making sulphated zirconia a catalyst of choice for the

process.

The production of C7 ethers via the reaction of methanol with 2,3-dimethyl-1-butene (DM1B)

and 2,3-dimethyl-2-butene (DM2B) in the presence of a microporous cation exchange resin

(Amberlyst 15) was investigated by Guin et al.150 The reactions were performed in a batch

reactor at 50-70 °C, under helium atmosphere (1.7 MPa). A high methanol to olefin mole ratio

(46:1) and 0.1g of catalyst (0.4-0.7 mm) were used. It was found in their study that 2,3-dimethyl-

1-butene was more reactive than DM2B and that isomerisation was a prominent reaction when

using DM1B as starting material. Under the set of conditions employed for this reaction, the

ether product obtained at equilibrium was about 75% at 50 °C and 61% at 70 °C, suggesting

that the reaction is exothermic.

The synthesis of TAEE and (TAA from a mixture containing isoamylene, ethanol and water via

etherification and hydration reaction of 2M2B in a batch reactive distillation column was reported

by Varisli and Dogu.151 The reactions were performed using Amberlyst 15 as catalyst, it was

showed that increasing the reaction temperature from 90-124 °C resulted in a significant

conversion of 2M2B, reaching 99%. The formation of TAA via reaction of 2M2B and water

occurred. In some cases, selectivity to TAA was higher than to the desired TAEE. Varisli and

Page 60: Applications of metal triflates and assisted acids as

40

Dogu suggest that the high selectivity to TAA is due to the higher adsorption equilibrium

constant of water than ethanol on Amberlyst 15. However, a significant increase in conversion

of 2M2B to TAEE was observed in the absence of water. The comparison of the activities of

various commercial cation exchange resin beads supplied by Rhom & Haas (Amberlyst 16,

Amberlyst 35 and XCE586) and fibre catalyst from Smoptech (SMOPEX-101, a sulfonated

polyethylene fibre with styrene) during production of TAME was reported.152 The etherification

reaction of methanol with isoamylenes (2M2B (93%) and 2M1B (7%)) with methanol were

performed at 60-80 °C. Mole ratios of methanol to isoamlyene of 0.5, 1.0 and 2.0 were used in a

batch reactor set-up and isopentane was used as solvent. The activity of the catalysts was

found to follow the trend Amberlyst 35>Amberlyst 15>SMOPEX-101> XE586. Amberlyst 35 was

found to be most active especially at elevated temperature (≥75 °C). This could be due to the

fact that

The evaluation of a number of solid acid catalysts for etherification of DM1B and DM2B with

methanol was reported by Wang and Guin.153 The catalysts included sulphated zirconia,

tungstated zirconia, two silica-supported sulfated zirconia catalysts (SZ/SiO2-N and SZ/SiO2-S)

together with two commercial ion exchange resins (Nafion NR50 and silica supported Nafion

SAC-13). Additionally, H-ZSM-5 zeolite was also tested. All the catalysts were benchmarked

against the commonly used Amberlyst 15. The reactions were performed in a stainless steel

batch reactor at 80 °C, for 2 hours under helium pressure of 1.8 MPa. A catalyst loading of 0.5 g

was used in all instances and an alcohol to olefin mole ratio of 1:1 was used and the reactions

were carried out in hexane (4 g) as a solvent. The results obtained from their study are shown in

Table 1.10.

Page 61: Applications of metal triflates and assisted acids as

41

Table 1.10: Comparison of various solid acid catalysts during etherification of DM1B with methanol.

Catalyst DM1B Conv. (%)

DM2B Select. (%)

Ether Select. (%)

Ether Yield (%)

Amberlyst 15 91.0 78.7 20.6 18.8

Nafion NR50 29.0 44.0 53.9 15.6

Nafion SAC-13 3.2 21.7 58.9 1.9

HZSM-5 23.2 93.5 1.1 0.3

ZrO2-WO3 0.9 19.4 64.2 0.6

SZ 4.7 23.2 66.2 3.1

SZ/SiO2-S 51.8 41.7 57.6 29.4

SZ/SiO2-N 38.2 34.4 63.8 24.4

SZ/SiO2-NP 21.2 28.7 67.4 14.3

According to Wang and Guin, the isomerisation product 2,3-dimethyl-2-butene and the ether

were major products of the reaction. A small amount of 2,3-dimethyl-2-butanol was detected,

due to the reaction of olefin with trace amount of water from the catalyst and the reactant

mixture. The commercial catalyst (Amberlyst 15) showed high conversion of 2,3-dimethyl-1-

butene, but low selectivity to the desired ether, while silica-supported sulphated zirconia showed

comparable and even higher ether yields than Amberlyst 15.

A comparative study of MTBE production via the etherification reaction of methanol with

isobutylene in the presence of various zeolite catalysts (Scheme 1.28) was under taken by

Hunger et al.154 They performed all the reactions in a fixed bed reactor, using 0.2 g of catalyst,

and the reactions were performed at 60-80 °C, using a methanol to isobutylene ratio of 1:1 and

2:1. The reaction mixture was analysed by GC coupled to an NMR spectrometer.

Page 62: Applications of metal triflates and assisted acids as

42

1.59 1.89

CH3OHzeolites

O

MTBE

Scheme 1.28: Zeolites catalysed etherification of methanol with isobutylene

The results of their study showed that for the reactions performed using a methanol to

isobutylene mole ratio of 2:1, at 60 °C, fluorinated HBeta- and HBeta zeolites gave comparable

activity to those afforded by a commercial ion exchange resin (Amberlyst 15). The equilibrium

yield obtained by the fluorinated HBeta zeolite and Amberlyst 15 was 47-50 mol%. HY and

HZSM zeolites showed relatively less activity, giving low yield of 8-10 mol% under the same set

of conditions. At higher temperature (80°C), the activity of fluorinated HBeta, HBeta and

Amberlyst 15 was significantly low, with an MTBE yield of 20 mol%, while the activity of HY and

HZSM-5 zeolites increased to yield 20 mol% of product. In another study carried out by Chu

and Kühl,155 the synthesis of MTBE over HZSM-5 and HZSM-11 zeolites was performed and the

activity thereof of was compared to that of Amberlyst 15. The reactions were performed in the

liquid phase below 100 °C at 200 psig, with 10 g of the catalyst being used. The by-products

observed in their study were diisobutylene formed from dimerisation of isobutylene and tert-

butanol from the reaction of water with isobutylene. The results showed that both zeolites gave

the higher activity and selectivity than Amberlyst 15. The advantage of zeolites over the

commercial Amberlyst 15 is that zeolites have high thermal stability. Furthermore, zeolites gave

high selectivity to the ether and less oligomerization product.

The synthesis of MTBE over titanium silicate (Ti-silicate) catalysts was also reported156 and the

activity thereof was compared to that of HZSM-5 zeolite. The vapour phase production of MTBE

was performed in a fixed bed reactor, at atmospheric pressure. The reaction was performed at

90 °C, using a 6.5:1 isobutylene to methanol mole ratio. Both catalysts gave steady total

conversion of the substrates to product and good MTBE selectivity without deactivation for 20

hours. ZSM-5 gave a small amount of dimethyl ether, which was formed via methanol

dehydration, whereas Ti-silicate gave high selectivity to MTBE (100%) and no by-products such

as dimethylether or tert-butylether. These reactions demonstrated that Ti-silicate was a more

active and selective catalyst for this chemistry than HZSM-5. Their study also showed that the

reaction of isobutylene with methanol is highly exothermic, hence increasing the reaction

Page 63: Applications of metal triflates and assisted acids as

43

temperature resulted in a considerable decrease in the production of MTBE, but it was also

shown that at low temperature the reaction is sluggish.

Tejero and co-workers have reported the synthesis of isopropyl-tert-butyl ether via the reaction

of isopropanol and isobutylene in the presence of various solid acids such as Amberlyst 15,

Amberlyst 35, Purolite CT275 and HZSM-5 zeolites (Scheme 1.29).157 According to Tejero, the

inherent characteristics of isopropyl-tert-butyl ether such as high octane blending values, lower

oxygen content and low vapour pressure make the compound attractive as a fuel blending

ether.

OH H+

O

0.7-2 ROH/alcohol70-90 oC1.6 MPa

1.77 1.59 1.90

Scheme 1.29: Etherification of isopropanol with isobutylnene

They conducted the reactions using alcohol to olefin mole ratios between 0.7 and 2, at 70-90

°C, and a pressure of 1.6 MPa for the resin-catalysed experiments, 0.4-1 g catalyst loading

(equivalent to 1% with respect to liquid mixture). On the other hand, 10-30 g of zeolites were

used representing 5-15% catalyst loading. Their results showed that various by-products such

as diisobutylene, tert-butyl alcohol and diisopropyl ether are produced in the course of the

reactions catalysed by ion exchange resins. It is also mentioned that at fixed temperature, the

equilibrium conversion increases with an increase of the alcohol to olefin mole ratio. On the

other hand, the equilibrium conversion decreases with an increase of temperature, since the

reaction is exothermic (ΔH0 = -22.9kJmol-1). For the reactions catalysed by zeolites, it was

evident that the rate of reaction was low, with the main by-product being tert-butyl alcohol, which

was formed by water released during the formation of diisopropyl ether. Diisobutylene was also

obtained. Their study has also shown that HZSM-5 was less active and selective to than resins.

The gas phase synthesis of MTBE from methanol and isobutylene in the presence of

dealuminated zeolites clays and Amberlyst 15 was reported by Poncelet and colleagues. 158The

reactions were performed in a fixed bed glass reactor at atmospheric pressure, using a

Page 64: Applications of metal triflates and assisted acids as

44

methanol to isobutylene mole ratio of 1.02 and the space velocity of 3.25 h-1. The reactions were

studied at temperatures between 30 and 120 °C. The results obtained showed that beta zeolites

are the most active and the activity is comparable to that of Amberlyst 15.

The synthesis of octadienyl ethers and butenyl ethers via etherification of 1,3-butadiene with

methanol, followed by hydrogenation of the unsaturated linear ether to the corresponding

saturated ether, in the presence of homogeneous palladium complexes {Pd(di-benzylidene-

acetone)2 and [Pd(allyl)Cl]2} was reported by Patrini et al.159 The reactions were performed in a

300 mL autoclave reactor at 60 °C and in the presence of phosphine ligands. In another study,

the polymer supported palladium (II) complex was used to catalyse etherification of 1,3-

butadiene and methanol.160

A comparative study for the reactivity of branched C6 olefin (2,3-dimethyl-1-butene and 2,3-

dimethyl-2-butene), C8 olefins (2,4,4-trimethyl-1-pentene and 2,4,4- trimethyl-2-pentene) and

linear olefins {1-hexene, 1-pentene and 1-octene } with methanol in the presence of Amberlyst

15 showed that branched olefins have greater reactivity than linear olefins. It was also found

that the reactivity of C6 olefins with methanol was higher than that of C8 olefins. Furthermore, it

was shown that increasing the methanol to olefin mole ratio resulted in increased ether

production. The reaction temperature range of 70-100 °C is suitable for the reaction.161

For the Fisher-Tropsch products 1-hexene and 1-pentene, purification is normally hampered by

the presence of close-boiling branched olefins such as 2-methyl-1-butene; 2-ethyl-1-butene and

2-methyl-1-pentene. A report by de Klerk162 showed that it is possible to selectively etherify the

branched olefins with methanol forming gasoline ethers with high octane number in the

presence of α-olefins, with insignificant loss of α-olefins. In his study, two ion exchange resins

(Amberlyst 15 and 35) were used and it was found that isomerisation of olefins and

etherification took place. Furthermore, it was also found that the temperature range of 65-80 °C,

high space velocity and methanol to tert-olefin ratio of 2 would be beneficial to suppress the 1-

hexene isomerisation.

The use of silica-supported heteropoly acids as catalysts for the gas phase synthesis of MTBE

via the reaction of methanol and isobutylene was reported by Shikata et al.163 In their study, the

Dawson-type tungsophosphoric acid (H6P2W18O62/SiO2), Kiggen-type heteropoly acid

(H3PW12O40/SiO2), unsupported heteropoly acids (H3PW12O40 and H6P2W18O62) and Amberlyst

15 were evaluated. The results showed that silica-supported catalysts exhibited comparable

activity to a commercial Amberlyst 15 and are more active than unsupported heteroply acids.

Page 65: Applications of metal triflates and assisted acids as

45

Other heteropoly acids such as carbon-supported Ag3PW12O40;164 20 wt% H4SiMoO14/SiO2

165 are

also reportedly used during synthesis of ethers.

1.5 Condensation of phenols with dienes

Assisted acid systems have been evaluated as catalysts for the condensation of dienes with

phenolic compounds. The condensation reaction involves both alkylation and etherification

reactions sequentially. 2,2-Dimethylbenzopyran is frequently encountered in natural products

such as vitamin E, its derivatives166 and in flavonoids,167 some of which exhibit important

biological activities.168,169 This system is also present in the recently discovered class of HIV-

inhibitory benzotripyrans.170 Several approaches for the synthesis of such compounds are

reported in the literature. The reactions can be promoted by Brønsted171 or Lewis acids172 and

require high temperature, with moderate yields reported. Ahluwalia et al.173 have reported usage

of H3PO4 as catalysts for isopentylation of phenols with isoprene to form 2,2-dimethylchromans

(3), Scheme 1.30.

1.17 1.91 1.92

H3PO4

OH OH O

X

H2PO4-

H+

1.93

1.94 1.95 1.96

Scheme 1.30: Mechanism for synthesis of 2,2-dimethylchromane

According to Ahluwalia, the reaction is initiated by protonation of isoprene with H3PO4 to form

the mesomer 1.91 and 1.92. The mesomers can alkylate phenol to give 1.94 and 1.95. The allyl

phenol 1.95 is more likely to form, because it is thermodynamically stable, hence 1.95

undergoes cyclisation in the presence of an acid catalyst to give the desired chroman. The

reaction was performed at 30-35 °C, with slow addition of isoprene to a mixture of phenol and

H3PO4 in an inert solvent. The reaction of resorcinol with isoprene resulted in the formation of

Page 66: Applications of metal triflates and assisted acids as

46

three products 1.97, 1.98 and 1.99. In a ratio of 1:2:5, with an overall yield of 80% (Scheme 1.31).

O O

O OH1.97 1.98

1.99

HO O

Scheme 1.31: Products from the reaction of phenol and isoprene.

The use of Amberlyst 15 as catalyst for a one step prenylation of phenols was reported by

Banerji and co-workers,174 where it was found that the catalyst efficiently catalysed the

condensation of substituted phenols 1.100 with isoprene giving improved yields and high

positional selectivity for the synthesis of 2,2-dimethylcromans 1.101. They extended the method

to the synthesis of 2,2-dimethylcromenes 1.103 via a reaction of 9 with 3-hydroxy-3-methylbut-

1-ene 1.102.

OH

R

O

R

O

RHO

+Amberlyst 15

Amberlyst 15

1.102

1.101

1.103

1.100

1.17

Scheme 1.32: Production of chromans and chromenes catalysed by Amberlyt 15

The use of [(acac)2Mo(SbF6)2] complex as a catalyst for synthesis of 2,2-dimethyl-6-methyl-

chroman 1.107via the reaction of p-cresol with prenyl alcohol 1.104 or 1.106 was carried out by

Page 67: Applications of metal triflates and assisted acids as

47

Kočovský and colleagues.175 Under the set of conditions used they have used, 1.107 was

isolated in 28% yield, (Scheme 1.33).

OHOH

OH

or [(acac)2Mo(SbF6)2]OH

O

CH2Cl2, rt

1.1051.60

1.106

1.104

1.107

Scheme 1.33: [(acac)2Mo(SbF6)2]-catalysed preparation of chroman

The reaction of phenol with conjugated dienes in the presence of aluminium phenoxide was

investigated by Dewhirst and Rust.176 Four products could be isolated during the reaction of

phenol with isoprene (Scheme 1.34). The product distribution of 1.96:1.108:1.109:1.110 was

found to be 0.38:0.25:0.24:0.13 at 80 °C, with phenol to isoprene mole ratio of 1.2:2 and in the

absence of a solvent. Dewhirst and Rust also performed a reaction of phenol with 1,3-

butadiene, and the reaction gave 1-methylchroman, ortho- and para-alkylated phenols, with

yields of 5%, 57% and 8%, respectively. The very same reaction was also studied by Bader and

Bean using H3PO4 as catalyst.177

O OHOH

O

1.96 1.108 1.109 1.110

Scheme 1.34: Products for the reaction of isoprene with phenol in the presence of Al(OPh)3.

Page 68: Applications of metal triflates and assisted acids as

48

The sequential addition/cyclisation of phenols with dienes catalysed by Ag(OTf) was recently

reported by Youn and Eom.178 They have investigated the activity of various metal salts

including Ag(OTf), AgSbF6, AgBF4, AgClO4, AgNO3, Ag(OTf), Cu(OTf)2,Sc(OTf)3, etc., as Lewis

acid catalysts for the addition of isoprene onto 4-methoxy phenol 1.111 as a model reaction.

They have found that Ag(OTf) was optimal for the reaction, giving 61% yield of benzopyran

1.112 (Scheme 1.35). Other silver salts were found to be ineffective for the reaction, while Sc,

Ru(III) and Au(I) triflates were moderately active. Cu(OTf)3 gave the lowest activity and triflic

acid gave only 18% yield. Thereafter they evaluated a few dienes and Ag(OTf) was found to be

efficient in promoting the reactions.

1.17 1.112

OH

O

O

OClCH2CH2Cl, rt, 24 h

5 mol% AgOTf

1.111

Scheme 1.35: Ag(OTf) catalysed reaction of isoprene with 4-methoxy phenol

Adrio and Hii179 have recently shown that Cu(OTf)2 is also an active catalyst for the reaction

shown in Scheme 1.35 when the reaction is performed at elevated temperature (50 °C).

However, a significant increase in the product yield was observed after the addition of PPh3

ligand to the Cu(OTf)2 catalysed reaction, especially when the metal to ligand ratio was

increased to 2:1. The highest yield (69%) was obtained when using Cu(OTf)2-bipy catalyst, with

ligand loading of 2.5 mol% in dicholroethane as a solvent. Thereafter various phenolic

substrates and dienes were evaluated where Cu(OTf)2 was used in conjunction with PPh3 and

with bipy ligand.

The synthesis of chromans via cyclocoupling of phenols with allylic alcohols in the presence of

the molybdenum complex [CpMo(CO)3]2 under microwave heating was investigated by

Yamamoto and Itonaga.180 As a model reaction, p-cresol (24) was reacted with prenyl alcohol

(14) to yield a chroman (25). Initially the reaction was carried out at 60 °C in the presence of 5

mol% [CpMo(CO)3]2 and 10 mol% o-chloranil for 3 hours and 69% of the desired chroman was

isolated. A similar yield was obtained at higher temperature (80 °C), and subsequently the

reaction was performed under microwave heating at 150 °C for 1 hour, 84% yield of chroman

was obtained. They found that the use of isoprene, instead of prenyl alcohol resulted in a

decrease in the yield.

Page 69: Applications of metal triflates and assisted acids as

49

1.6 The concept of combined acid systems

The relatively recent (2002) studies by Corey and co-workers have shown that chiral Brønsted

assisted-Lewis acids are exceptionally effective and versatile chiral Lewis acids for catalysing

enantionselective Diels-Alder reactions.181 There are four categories into which the combined

acids may be classified:

(1) Brønsted acid-assisted Lewis acid (BLA), this is the enhancement of Lewis acidity by the

combination with Brønsted acid.

(2) Lewis acid-assisted Lewis acid (LLA), which is the enhancement of Lewis acidity by

combination with Lewis acid.

(3) Lewis acid-assisted Brønsted acid (LBA), the enhancement of Brønsted acidity by

combination with Lewis acidity and finally.

(4) Brønsted acid assisted Brønsted acid (BBA), enhancement of Brønsted acidity by

combination with Brønsted acid.

When strong Lewis acids are combined with a proton donor species such as H2O and other

protic acids such as HCl, the acidity of the medium usually increases. For example, 100% HF

has a Hammett number of -15,182 whereas a medium containing equi-molar concentrations of

HF and SbF6 has an H0 value in excess of -30.183 Thus, a combination of a Brønsted acid with a

Lewis acid sometimes results in the formation of a superacid. The mixtures of HCl/AlCl3 as well

as HBr/AlBr3, (widely used in alkylation reactions) are super acidic ionic liquid systems. The

strength of HCl/AlCl3 Brønsted super acid has a Hammett number (H0)~ -15 which is similar to

that of dry HF, and this combination is very active as a catalyst for alkylation.184 The systems of

HCl/AlCl3 have been used in the industry since the 1940s for alkylation purposes.185

The main objective of this literature chapter is to review the types of acids that have been

explored as catalysts for various chemical reactions. This chapter reveals that among the

combined acid systems used thus far, combinations of metal triflate Lewis acids with Brønsted

acids have not been comprehensively studied, and this presented an opportunity to explore

such combinations as catalysts for a number of reactions reported in the forthcoming chapters

of this thesis.

Page 70: Applications of metal triflates and assisted acids as

50

1.7 Summary

The literature review has shown that a large number of Brønsted and Lewis acids can be used

to catalyse industrially important reactions such as Friedel-Crafts alkylation and etherification of

alcohols and olefins. The literature review has also revealed that solid acids are mostly used at

the commercial scale, for example, Amberlyst 15 is widely reported as catalyst for etherification

reaction of olefins and alcohols.

It has been revealed that metal triflate salts are preferred over metal halide Lewis acids, mostly

due to moisture stability, recyclability and ease of handling. The metal triflate Lewis acids have

been reported to promote numerous reactions such as Diels-Alder reactions, Friedel-Crafts

alkylation reactions of aromatic compounds with alkyl halides and Mukaiyama aldol reactions.

However, it is important to note that the literature has not mentioned the application of metal

triflates as catalysts for the alkylation of cresol, anisole and diphenyl ether with olefins.

Therefore this presented an opportunity to explore the catalytic activity of these Lewis acids for

industrially important reactions, one aspect of study for the present project.

The concept of assisted acid systems has recently been an area of focus in the field of acid

catalysis. However, the main emphasis has been on the use of assisted acids for asymmetric

synthesis. For example, a combination of metal halide or metal triflate Lewis acids with weak

acidic compounds such as bis-naphthol, etc. has been explored. The acid combination of metal

chlorides with Brønsted acids to yield super acids was explored intensively. But the combination

of metal triflate Lewis acids with Brønsted acids has not been widely studied. Hence this

provided another opportunity to investigate these combinations as catalysts for Friedel-Crafts

alkylation of cresols, anisole and diphenylether with isobutylene, a second focus area of the

current work.

The objective of this literature review was to identify the types of acids that have been used thus

far to catalyse alkylation, etherification and condensation reactions, with an aim to use metal

triflates and Bronsted acids as catalysts for such reactions. Furthermore, another important

objective of the study is to investigate whether combined acid systems (Brønsted acid/metal

triflate salt) would effect enhancement of the reaction rate.

The subsequent chapters will discuss the results obtained during evaluation of the activity of

various Lewis acids (mostly metal triflate salts) and Brønsted acids (mineral and organic acids).

The comparison of the activity for Lewis acids, Brønsted acids and assisted acid or combined

acid systems is also discussed.

Page 71: Applications of metal triflates and assisted acids as

51

Three model reactions were chosen to evaluate the activity of the acids. The first reaction

discussed in Chapter 2 involved etherification of alcohols and tert-olefins to form branched

ethers. The second model reaction (Chapter 3) was the Friedel-Crafts alkylation reaction of

phenolic compounds with isobutylene to form various alkylated phenolic compounds. The third

reaction (Chapter 4) was the condensation of phenolic compounds with dienes to form

chromanes.

Page 72: Applications of metal triflates and assisted acids as

52

CHAPTER 2

Hydroalkoxylation of olefins

This chapter focuses on the results obtained with the use of assisted acids made up of mineral

Brønsted acids and metal triflate Lewis acids as catalysts for etherification of various olefins with

alcohols.

2.1 Introduction Ether formation is an important transformation in fine chemicals synthesis as well as in

commodity chemicals. Ether formation may be effected by making use of the traditional

Williamson ether transformation186 or by the hydroalkoxylation of alkenes.187 Typically, the latter

requires rather harsh conditions such as strong Brønsted acids, for example triflic acid or

sulfuric acid, to ensure a successful outcome.188 The synthesis of tertiary ethers is preferably

performed with an alkene and the corresponding alcohol. Branched ethers such as tertiary-amyl

methyl ether (TAME) and tertiary-amyl ethyl ether (TAEE) are useful octane-boosting additives

for petrol and are considered to be viable alternatives to the more harmful additives currently

used. 189 Methyl tertiary-butyl ether (MTBE) is prepared via the reaction of isobutylene and

methanol in the presence of an acid catalyst.190 MTBE is used as a fuel octane rating improver.

However, its solubility in water causes environment pollution problems due to migration of this

material through soils and natural water bodies.191 Consequently, alternative oxygenates are

being researched as gasoline additives. The tertiary-amyl ethers are attractive alternatives for

MTBE to use for gasoline blending.192 They enhance gasoline burning properties which,

amongst others, reduces carbon monoxide emissions. 193,194 Furthermore, tertiary ethers have

high octane numbers, low viscosities and low densities, all of which are properties that are

essential for gasoline blending.195

The etherification of alkenes with alcohols is mostly carried out in the presence of sulfonic acid-

based cation exchange resins such as Amberlyst 15.196,197,198,199 Other catalysts such as

supported sulfated zirconia,200 Si-MCM-41201 and zeolites202 are also used.

Page 73: Applications of metal triflates and assisted acids as

53

Metal triflates are known to promote numerous organic reactions.203 These catalysts have been

shown to be extremely useful in epoxide ring-opening reactions with various nucleophiles,204,205

to be capable of effecting highly efficient acetal formations,206,207 and to readily produce highly

active Pd catalysts for the co-catalysed methoxycarbonylation reaction.208 Importantly, metal

triflates may be readily recycled,209,210 and are capable of catalysing organic reactions in a

hydrous environment as opposed to their corresponding metal halides that are unstable in the

presence of even minute amounts of water.211

This chapter will focus on the results obtained during the evaluation of various Lewis, Brønsted,

and assisted acid systems as catalysts for the etherification of olefins with various alcohols

Scheme 2.1, where 2M2B is 2-methyl-2-butene). The Lewis acids evaluated include metal

triflate salts (such as the triflates of aluminium, zirconium, and scandium) and selected metal

triflate salts of the lanthanide series of elements. The activity of metal chloride Lewis acids of

zirconium, aluminium and lanthanum were also evaluated for comparison purposes. A variety of

Brønsted acids including mineral acids (such as H2SO4, H3PO4 and HNO3) and organic acids

(such as methanesulfonic acid, benzoic acid and para-toluenesulfonic acid) were also evaluated

for the etherification reactions.

The main objective of evaluating these acids is to subsequently prepare various combinations of

Lewis and Brønsted acids, and evaluate their activity for etherification reactions in order to

confirm if there is any improvement on the reaction rate.

ROHmetal triflate RO

R = C1-4 aliphatic group2.1 2.2

Scheme 2.1: Etherification of alcohols with 2M2B

The proposed mechanism, through which metal triflates catalyse the etherification reaction, is

shown in Scheme 2.2. The alcohol is first coordinated to the metal via a lone pair of electrons

from the oxygen atom, followed by a nucleophillic attack of the olefin onto the proton of the

coordinated methanol, which leads to the formation of a tert-carbocation and methoxide anions.

The ionic species eventually bond together forming the desired ether.

Page 74: Applications of metal triflates and assisted acids as

54

CH3OHM(OTf)n

[CH3O-]O

H3C OH

M(OTf)n

M = Al, Zr, Sc or lanthaniden = 3 or 4

Scheme 2.2: Lewis acid catalysed etherification

It is also possible that the metal triflate directly activates the alkene as is the case in some

recently reported Friedel-Crafts reactions of alkynes with aromatic systems, Scheme 2.3. 212

CH3OH

O

M(OTf)n

M(OTf)nO

HH3CM(OTf)n

Scheme 2.3: Another possible mechanism for a Lewis acid catalysed etherification

Page 75: Applications of metal triflates and assisted acids as

55

2.2 Screening of Lewis acids

The time-limited reactions of methanol and 2-methyl-2-butene (2M2B), designed to highlight

rate differences between the various metal triflate catalysts, were performed in a 300 mL

autoclave batch reactor equipped with gas entrainment stirrer and a dip tube for sampling. All

experiments were performed at 100 ˚C, 6 bar N2 pressure over a period of 150 minutes using

methanol (65.72g; 1.960 mol) and 2-methyl-2-butene (11.39 g; 0.163 mol), in a mole ratio of

13:1 and catalyst concentration of 0.1 mol% with respect to the olefin. The results obtained

during the screening of Lewis acids and their correlation to the metal ion radius is shown in

Table 2.1.

Table 2.1: Reaction of methanol with 2M2B and correlation to ionic radiia

Entry M(OTf)n Yield

(Mol%)

Metal Ionic Radius213

(Å)

1 Al(OTf)3 55 0.675

2 Zr(OTf)4 55 0.860

3 Sc(OTf)3 16 0.885

4 Yb(OTf)3 7 1.008

6 Sm(OTf)3 3 1.078

5 La(OTf)3 2 1.098

7 Gd(OTf)3 2 1.172

8 ZrCl4 0 0.860

9 AlCl3 0 0.675

The triflate salts of Zr and Al were found to efficiently catalyse ether formation from the reaction

of 2-methyl-2-butene with methanol. Sc(OTf)3 showed significantly lower activity (Table 1, entry

3) while lanthanide (Yb, La, Sm, Gd) triflates, AlCl3 and ZrCl4 were found to be inactive.

Amongst the triflates, there is a loose correlation between activity and the ionic radius of the

Page 76: Applications of metal triflates and assisted acids as

56

metal triflate concerned, which ties the catalytic activity of the triflates in this instance more

strongly to charge density (z/r).214 The harder metals showed improved catalytic activity for

these reactions over the softer lanthanide triflates.

The observation relating to the ineffectiveness of metal chlorides also mirrors the relative

hardness of given metal systems with different counter ions (in this case triflate vs chloride) and

the influence thereof on the activity of the Lewis acid as a catalyst. Thus, for a given metal with

an apparently given charge density, the counterion plays a role in determining the hardness and

hence also the activity of the catalyst.215 It should also be noted, though, that the metal chlorides

involved here are hydrolytically sensitive and may have undergone methanolysis to some extent

leading to inactive M(OMe)x(Cl)n materials. The reaction is highly selective towards the

formation of the desired ether, although minor levels (<1%) of isomerisation of 2-methyl-2-

butene to 2-methyl-1-butene were observed in the process (GC-FID analysis). These two

isomers provide the same ether product and this side reaction is of no consequence here. Small

amounts of 2-methyl-2-butanol (2%) were formed from trace amounts of water present in the

methanol, Scheme 2.4. However, this reaction does not significantly affect the selectivity since

the amount of water in methanol was found to be small (600 ppm or less by Karl Fischer

titration) and if the alcohol was dried prior to the experiment this alcohol product did not form.

H2OHO

2.1 2.3

Scheme 2.4: Hydration reaction of 2M2B

2.3 Screening of Brønsted acids

The Brønsted acid catalysed reactions follow the Markovnikov mechanistic pathway, the

proposed mechanism for the reaction is shown in Scheme 2.5.

Page 77: Applications of metal triflates and assisted acids as

57

HX HOCH3 O HXO

HX-

Scheme 2.5: Brønsted acid catalysed etherification reaction of methanol and 2M2B

The Brønsted acids screening reactions were performed at 100 ºC, using 0.1 mol% of catalyst

and an alcohol to olefin mole ratio of 13:1. The results thereof are presented in

Table 2.2. All the reactions were terminated at 150 minutes of reaction time. At this stage, the

reactions had not attained equilibrium conversion, and hence are crudely indicative of the rate of

reaction afforded by each acid catalyst.

Table 2.2: Catalytic activity of Brønsted acids during etherification reaction

Acid pKa 216

Yield

(Mol%)

HOTf -14.9 22.0

HCl -8.0 1.5

H2SO4 -3.0 18.0

p-TsOH -2.8 18.1

CH3SO3H -2.0 19.4

HNO3 -1.5 8.0

H3PO4 2.0 1.0

PhCOOH 4.2 <1

Page 78: Applications of metal triflates and assisted acids as

58

It is well known that in many cases catalytic activity of Brønsted acids is dependent on their

dissociation constants (Ka), which is commonly reported in terms of its negative logarithm,

pKa.217

pKa = - logKa

This equation implies that a high value of pKa indicates a very small value of Ka, because Ka =

10-pKa and hence a weak acid. The reaction outcomes showed a correlation between pKa values

of different acids with their activities, with triflic acid giving the highest activity as expected. HCl

was almost inactive as catalyst for the reaction and it is well known that HCl may undergo

addition reaction to olfins, which would lead to the formation of an unreactive chloroalkane

(Scheme 2.6). The lack of activity shown by H3PO4 and PhCOOH may be directly linked to their

pKa values. It is thus evident that several of the metal triflate salts (Table 2.1) showed superior

activity compared to Brønsted acids.

HClCl

2.1 2.4

Scheme 2.6: Addition of HCl to 2M2B

2.4 Optimisation of reaction conditions

The optimisation of the reaction conditions was performed using Al(OTf)3. This catalyst

(Al(OTf)3) together with Zr(OTf)4 exhibited good activity among the Lewis and Brønsted acids

evaluated for hydroalkoxylation of olefins and alcohols. The initial set of conditions under which

catalyst screening was performed was chosen arbitrarily, and therefore it became imperative to

optimise the reaction parameters. Three reaction parameters were optimised i.e. temperature,

catalysts concentration and the alcohol to olefin mole ratio.

2.4.1 The influence of varying the alcohol to olefin mole ratio.

The etherification of olefins with alcohols is an equilibrium controlled reaction.218 This was also

reflected in the results of this investigation, where the reaction failed to proceed to completion.

Therefore it is crucial to identify the optimum alcohol to olefin mole ratio required to achieve

maximum conversion or yield. The experiments were conducted at constant temperature (100

ºC), and a fixed amount of catalyst (1.5 mol%), while varying the methanol to olefin mole ratio

Page 79: Applications of metal triflates and assisted acids as

59

(Figure 2.1). The initial rates (0 to 30 minutes) were comparable in all reactions. However, for a

4:1 reaction, a slight decrease of reaction rate after 30 minutes was evident. A similar

equilibrium yield (55 mol%) for 13:1 and 10:1 reactions was achieved, whereas a slight

decrease of equilibrium yield (50 mol%) was evident for the 4:1 experiment. The latter results

were expected, because when the concentration of one of the reactants is decreased in an

equilibrium controlled reaction, the reverse reaction becomes favourable as explained by Le

Chatelier’s Principle. It was therefore decided to perform the optimisation of other parameters

using the 13:1 mole ratio of methanol to the olefin. The curves representing yield over time

resemble those of a first order reaction rate.

0

10

20

30

40

50

60

70

80

90

100

0 20 40 60 80 100 120 140 160

TAM

E Yi

eld

(Mo

l%)

Time (Min.)

13:1

10:1

4:1

Figure 2.1: Influence of varying the methanol to olefin mole ratio using 1.5 mol% Al(OTf)3.

2.4.2 The influence of changing catalyst concentration

The reactions to investigate the effect of varying catalyst concentration were performed at 100

C, with a fixed alcohol/alkene molar ratio (13:1), Figure 2.2. While a slow reaction rate was

observed in the absence of catalyst, the reaction was sensitive to changing Al(OTf)3

concentration at a fixed alcohol/alkene molar ratio of 13:1 at 100 C. Steadily increasing rates

Page 80: Applications of metal triflates and assisted acids as

60

were noted with increasing catalyst concentrations. This trend held true up to approximately 0.7

mol%, after which gains in the rate were minimal and the equilibrium concentrations of the

product remain essentially unaffected.

Figure 2.2: The influence of changing the catalyst concentration

2.4.3 The influence of changing the reaction temperature

The reaction performed at 100 C proceeds rapidly when using 1.5 mol% catalyst concentration.

Hence, it was decided to perform these reactions using a lower catalyst concentration of 0.1

mol% and methanol to olefin mole ratio of 13:1 (Figure 2.3) in order to be able to identify the

changes to the rates of the reaction. The reaction was somewhat sensitive to temperature with a

110 C reaction temperature providing higher reaction rates than those performed at lower

temperatures. Nonetheless, the reactions are sufficiently fast at the latter temperature to allow

the reaction to proceed to completion.

Page 81: Applications of metal triflates and assisted acids as

61

Figure 2.3: Influence of changing the reaction temperature

2.5 Recycling of Al(OTf)3 and Zr(OTf)4

The literature reveals many instances in which metal triflates are recyclable without loss of

activity.205,208 Hence, several experiments were performed to recycle Zr(OTf)4 and Al(OTf)3 for

etherification reactions in order to determine whether deactivation of the catalyst would occur.

The reactions for both catalysts were performed in a 300 mL stainless steel auto-clave reactor

at 100 ˚C, and starting off with a catalyst concentration of 1.5 mol% relative to the olefin. At the

end of each experiment, the catalyst was recovered by distilling off the reaction contents. In the

subsequent experiments, fresh reagents (alcohol and olefin) were charged into the reactor

(containing the catalyst residue), and the mixture was then heated to the operating temperature.

The recycled catalyst (pre-dissolved in methanol) was then added into the reactor via a pre-

pressurised sample bomb (6 bar N2). The results of the reactions are shown in Figure 2.4 and

Figure 2.5.

Page 82: Applications of metal triflates and assisted acids as

62

Figure 2.4: Recycling of Al(OTf)3 during etherification reactions.

In the reactions catalysed by Al(OTf)3, a slight decrease in reaction rate from the original to the

second run is evident. Thereafter the rate is similar in all the catalyst recycling experiments. The

equilibrium conversion is similar in all instances. It could be deduced that catalyst deactivation

did not prevail in these reactions.

For the reactions catalysed by Zr(OTf)4, the rate of reaction is comparable from the original

experiment up to recycle 5. The equilibrium conversion is also similar in all runs. It can therefore

be concluded that both catalysts could be recycled at least five times (excluding the original run)

without any loss of activity. In addition to that, the equilibrium conversion was not affected in all

instances.

Page 83: Applications of metal triflates and assisted acids as

63

Figure 2.5: Recycling of Zr(OTf)4 during etherification reactions.

2.6 The evaluation of Lewis assisted Brønsted acids Up to this point, our study has shown that lanthanide triflate salts and some mineral Brønsted

acids such as H3PO4 are inactive as catalysts for the present etherification reactions of olefins

with alcohols under the set of conditions specified above. It was then decided to use the two

types of acids in combination, in an attempt to capitalise on the now well-known principle of

assisted acidity.219 Accordingly, reactions were set up to include the two reagents (2M2B and

methanol) along with equimolar amounts of the two inactive catalysts (La(OTf)3 and H3PO4).

The results of this study (Figure 2.6) show that the activity of the combined Lewis/Brønsted

catalyst was enhanced significantly when the acids are used as combinations.

Page 84: Applications of metal triflates and assisted acids as

64

Figure 2.6: Etherification reactions catalysed by La(OTf)3/H3PO4 assisted acid

The mechanistic pathway through which the Brønsted acid binds to the Lewis acids is not well

understood at this stage. However, Barrett et al.220 used Yb(OTf)3 as a catalyst during the

nitration of aromatic compounds instead of a commonly used Brønsted acid (H2SO4) catalyst.221

The product distribution (isomers) was in accordance with electrophilic attack by NO2+ of carrier

type 2.5. On the other hand, bidentate binding of the counterion in lanthanide nitrates 2.6 is

known on the basis of well characterised structures as revealed in the literature.222 Therefore

nitration may proceed via direct attack of the protonated species or the arene may attack the

metal bound nitrate species directly.

2.5

O N

O

O

Ln

H

O

OLn N O

2.6

Barrett and co-workers223 have also reported the acylation of alcohols with acetic acid in the

presence of scandium(III) and lanthanide(III) triflates as catalysts. It was found that Yb(OTf)3

and Sc(OTf)3 were good catalysts for the reaction and they obtained remarkable rate

acceleration in their reactions. They therefore suggested that, in a reaction where Brønsted

acids are used as reagents (e.g. HNO3 in the nitration reaction) and metal triflates as catalysts,

Page 85: Applications of metal triflates and assisted acids as

65

the Brønsted acid binds to the Lewis acids forming a complex (Scheme 2.7) that represents a

stronger Brønsted acid than the parent Brønsted acids (nitric or acetic acids). 224

They proposed two possible ways in which the metal triflates may interact with Brønsted acids

to initiate the reactions. The first proposal was that metal triflate salts may bind to the Brønsted

acid forming the complexes 2.7 and 2.8 shown in Scheme 2.7 and that the complexes are

stronger Brønsted acids than nitric and acetic acids and may be responsible for initiating the

reactions. The second proposition was that the Brønsted acids may protonate the triflate ligands

and liberate a strong acid (triflic acid), which then instigates the reaction.

HNO3 Yb(OTf)3

3+

3OTf -O

N(H2O)YbO

OH

Yb(OTf)3

3+

3OTf -CH3COOHO

C(H2O)YbO

H

2.7

2.8

Scheme 2.7: Proposed reaction of Brønsted and Lewis acid

It was an objective of the present study to evaluate assisted acids (combination of metal triflates

and Brønsted acids) as catalysts for the etherification of olefins and alcohols. As part of the

process, 31P NMR studies were performed in order to verify that a reaction between H3PO4 and

La(OTf)3 occurs. Initially, a 31P NMR spectrum of neat H3PO4 dissolved in deuterated methanol

was obtained and gave a chemical shift of 2.13 ppm. In the subsequent experiment, 0.25 mol

equivalents of La(OTf)3 were added into the NMR tube, and the phosphorus peak was noted to

shift to -5.40 ppm. Increasing the amount of La(OTf)3 to 0.5 mol equivalents resulted in a

chemical shift of the phosphorus peak to -4.55 ppm. A further increase of La(OTf)3 to 1

Page 86: Applications of metal triflates and assisted acids as

66

equivalent shifted the signal to δ= -3.18 ppm while the addition of La(OTf)3 in excess (2 mole

equivalents) gave a signal at δ= -0.34 ppm for the phosphorus peak.

Given the ability of lanthanides to expand their coordination spheres, it is likely that a variety of

complexes were formed during this study, including a 2:1 acid/La, 1:1 acid/La and a complex in

which the acid binds two La ions. Therefore, the NMR spectroscopy result suggests that the first

proposal by Barrett is plausible, namely that the acid binds to the metal to generate a strong

Lewis-assisted Brønsted acid. The excellent reaction rate enhancement afforded by the

La(OTf)3/H3PO4 system, while giving high selectivities, prompted an evaluation of other metal

triflates in combination with a variety of Brønsted acids (Figure 2.7 onwards).

Figure 2.7: Activity of metal triflates/H3PO4 acid combinations

It is evident that all lanthanide triflate-based assisted acids gave high yields of the desired

product, with high selectivity (99.9%) and within the same reaction time as for the very active

triflate salts (of Al and Zr), indicative of highly active combination catalysts. On the other hand,

Zr(OTf)4 and Al(OTf)3 had already exhibited excellent activity as individual catalysts. An attempt

Page 87: Applications of metal triflates and assisted acids as

67

to use a combination of these Lewis acids with H3PO4 did not yield to any rate improvements,

with the activities obtained being similar to those afforded by the individual Lewis acids.

The evaluation of lanthanide triflate salts in combination with H2SO4 and HNO3 also yielded

some enhancements of the etherification rate (Figure 2.8). Each of these Brønsted acids and

Yb(OTf)3 showed insignificant activity for etherification as individual acids (remembering that

these are time-limited incomplete reactions in which relative reactivities may be gleaned from

yield data). The results showed that the individual acids (H2SO4 and HNO3) are moderately

active for the etherification reaction, while La(OTf)3 is inactive. The yields obtained when the

triflate salts of Yb and La were used in conjunction with H2SO4 and HNO3 showed appreciable

enhancements.

0

5

10

15

20

25

30

35

40

Yiel

d (M

ol%

)

Figure 2.8: Comparison of metal triflate/mineral acids activity.

Page 88: Applications of metal triflates and assisted acids as

68

Each of the Brønsted acids used for the Lewis/Brønsted combinations afforded some activity.

Importantly, the sum of the yields afforded by the two individual types of acid is lower than the

yield obtained by the combination, indicative of a synergistic effect. It has already been shown

that organic Brønsted acids such as triflic acid (HOTf), para-toluenesulfonic acid (p-TsOH) and

methanesulfonic acid (MsOH) gave good activity when used on their own. It was decided to also

test their activity as combinations with lanthanide triflates (Figure 2.9).

Figure 2.9: Comparison of metal triflates/ organic acids for etherification

The yield afforded by the Yb(OTf)3/p-TsOH acid combination (blue bars) is comparable to the

quantity obtained when summing the yields afforded by the Lewis acid and the Brønsted acid in

separate reactions (brown bars), indicating parallel reactions rather than an assisted acidity

reaction. In the case of MsOH, the sum of the yields from the separate reactions is slightly

greater than that of assisted acid counterpart. Therefore it can be concluded that combining the

lanthanide triflates and organic Brønsted acids does not result in rate enhancement, with the

possibility of a measure of suppression of the activity. An attempt to evaluate AlCl3/H3PO4 as

assisted acids was unsuccessful since the reaction did not yield any product.

Page 89: Applications of metal triflates and assisted acids as

69

2.7 The influence of varying La(OTf)3/H3PO4 concentration

Having observed the tremendous rate enhancement for the reactions catalysed by lanthanide

triflates in combination with mineral Brønsted acids, it was decided to investigate the effect of

decreasing the catalyst concentration on the etherification reaction catalysed by La(OTf)3/H3PO4

(Figure 2.10).

Figure 2.10: Effect of increasing La(OTf)3/H3PO4 concentration

The results show a considerable increase in the reaction rate as the catalyst concentration was

increased from 0.1 mol% to 0.8 mol%. The reaction performed with 0.1 mol% did not attain a

state of equilibrium within the reaction time.

2.8 The influence of varying temperature on La(OTf)3/H3PO4 catalysed reactions

An evaluation the influence of temperature on the reactions catalysed by La(OTf)3/H3PO4 was

performed using 0.8 mol% catalyst concentration and varying temperature from 80 °C to 100 °C

(Figure 2.11). The results show an increase of the reaction rate with an increase in the reaction

Page 90: Applications of metal triflates and assisted acids as

70

temperature. At higher temperatures (90 °C and 100 °C), the equilibrium conversion was

attained faster, whereas the reaction took longer than 80 minutes to reach equilibrium at 80 °C.

Figure 2.11: Effect of changing the reaction temperature

The TOF of the reactions where the effect of temperature reduction was evaluated has

demonstrated that the catalyst is more active at high temperature (Table 2.3). This is due to

larger amounts of product that are formed at 100 °C in comparison to that obtained at low

temperatures (80 °C and 90 °C). At 80 °C, the catalyst productivity (TOF) is low.

Table 2.3: Turn over frequencies over a period of 15 minutes.

Temp (°C) TOF (mol prod./mol cat. h)

100 182

90 123

80 59

Page 91: Applications of metal triflates and assisted acids as

71

2.9 Effect of varying the catalyst composition

The influence of changing the relative amounts of H3PO4 and La(OTf)3 used to generate the

active catalyst for the etherification reaction was investigated to determine the influence of this

ratio on the reaction rate. This part of the study stemmed from the 31P NMR work which

indicated that the specific ratio of these two entities generated different types of complexes. It

was thought that this study may shed light as to whether a changing ratio would lead to more or

less active assisted catalysts. The reactions were performed at 100 °C, using an alcohol to

olefin mole ratio of 13:1 (Figure 2.11). The ratios of the acids used also reflect their mol% values

as catalysts in the reactions (but where 1:1, for example, represents 0.1 mol% and 0.1 mol% of

the Brønsted and Lewis acids, respectively).

The results revealed some dependency by the reaction rate on the relative amounts of catalyst

present in the reaction mixture. There appears to be a subtle influence on the catalyst’s activity

by the H3PO4, in that higher ratios of the phosphoric acid (8:1 in Figure 2.12) apparently led to

inhibitory effects, while stoichiometric or sub-stoichiometric ratios give the substantial rate

enhancement previously noted. This observation is in accordance with the studies carried out by

Flowers225 on SmI2, where it was shown that HMPA can bind to SmI2 and displace the iodide to

the outer sphere, producing Sm(HMPA)6I2 in the process. In the present instance, if one accepts

that the metal is activating the acid, and that the lanthanide may bind to more than one

equivalent of acid as shown by the NMR study (and indirectly according to the Flowers study),

then excess phosphoric acid may well bind to the metal, donating electron density to the metal

ion and reducing the activating effect of the metal ion on the Brønsted acid in the process.

Page 92: Applications of metal triflates and assisted acids as

72

0

10

20

30

40

50

60

0 20 40 60 80 100 120 140 160

Yiel

d (M

ol%

)

Time (Min.)

1 H3PO4/2 La(OTf)3

0.5 H3PO4/1 La(OTf)3

1 H3PO4/1 La(OTf)3

8 H3PO4/1 La(OTf)3

2 H3PO4/1 La(OTf)3

Figure 2.12: The effect of varying the catalyst composition

The results also demonstrate that increasing the amount of La(OTf)3 from 1 to 2 mole

equivalents, while maintaining the relative ratio of H3PO4 (1:2 run, Figure 2.12, line 1 on the

legend) does not notably influence the reaction rate. Very interestingly, the 1:2 and 0.5:1 runs in

Figure 2.12 afforded extremely high rates of reaction and, notably, that the 0.5:1 rate is higher

than that of the 1:1 reaction. It is thus recommended that a 2:1 mole ratio of La(OTf)3 to H3PO4

is used in a reaction catalysed by the acids combinations to attain the maximum rates of

reaction

2.10 Recycling of La(OTf)3/H3PO4 assisted acids

A set of experiments aimed at investigating whether La(OTf)3/H3PO4 could be recycled after the

etherification reactions was performed. The initial experiment was performed using 0.2 mol%

catalyst concentration (Figure 2.13). The catalyst recovery was done by distilling the reaction

mixture of a given experiment from the reactor. The remaining contents were dissolved in

methanol and returned to the reactor, which was then charged with all the required reagents

and performing the subsequent reaction.

Page 93: Applications of metal triflates and assisted acids as

73

Figure 2.13: Recycling of La(OTf)3/H3PO4

The reaction rates of the subsequent reactions are comparable to the original run up to

recycle 4. This is a strong indication that the catalyst may be recycled without any loss of

activity. In all experiments the reaction reached equilibrium conversion after approximately the

same lapsed period of time and all rate profiles are essentially identical.

2.11 Solid phosphoric acid (SPA) catalysis

Solid phosphoric acid catalysis is traditionally used for the oligomerisation of propene and

butene in crude oil refineries.226 The catalyst is produced by mixing 85% H3PO4 and silica or

kieselguhr, then extruding and calcining the material so obtained at high temperature.227 It has

been shown in the present project that a combination of various metal triflate salts with mineral

Brønsted acids, especially H3PO4, gives excellent activity as homogeneous catalyst for the

etherification of methanol and 2M2B. It was then decided to broaden the study, whereby a

combination of SPA (as source of H3PO4) and La(OTf)3 was prepared and used as catalyst for

the etherification experiments. Initially, SPA was evaluated on its own as a catalyst for the

Page 94: Applications of metal triflates and assisted acids as

74

reaction and was proven to be inactive. This reaction was far from a trivial step in the study

since the support onto which a catalyst is loaded is known to sometimes dramatically influence

the activity of the catalyst. Subsequently, La(OTf)3/SPA assisted acid (1.03 g, which is

equivalent to 0.9 mol% as free La(OTf)3/H3PO4) was evaluated as a heterogeneous catalyst for

the etherification transformation and the results thereof were compared to those of Amberlyst 15

(a well-known commercial catalyst for etherification processes).228 The results obtained from the

study are shown in Figure 2.14.

Figure 2.14: The activity of La(OTf)3/SPA during etherification

SPA was shown to be completely inactive as a catalyst for the etherification of 2M2B with

methanol and La(OTf)3 has already proven to be inactive (see Figure 2.7). On the other hand,

the La(OTf)3/SPA combination showed significant enhancements in the reaction rate. But when

the solid was filtered and reused in a subsequent experiment, the solid was completely inactive.

It was then decided to distil the products from the filtrate and the residue so obtained was used

as a catalyst, which was shown to be highly active. The residues could be recycled at least

three times without loss of activity. It is outlined in the literature that H3PO4 leaches out of the

support into aqueous media, leading to catalyst deactivation.229 Hence, it is highly possible that

Page 95: Applications of metal triflates and assisted acids as

75

La(OTf)3/H3PO4 leached out of the support into the polar methanolic medium, leaving behind the

inactive solid (SiO2).

2.12 Comparison of activity of Amberlyst 15 and La(OTf)3/H3PO4

A comparison of the catalytic activity of La(OTf)3/H3PO4 to one of the most highly investigated

catalysts cited in literature, namely Amberlyst 15, was made on the basis of experimentally

determined data. The reactions were performed at 100 °C, using the methanol to 2M2B mole

ratio of 13:1. Two stirring speed rates were evaluated for the Amberlyst 15 promoted reaction, to

ensure that mass transfer limitations are taken into consideration if occurring. The equi-molar

concentration of the Amberlyst 15 was used (calculated based on the active species of the

catalyst, i.e. 4.8 mol equivalents of acid per kilogram of catalyst), Figure 2.15.

Figure 2.15: Comparison of La(OTF)3/H3PO4 and Amberlyst 15 for etherification reactions

The reactions catalysed by 1.5 mol% Amberlyst 15 proceeded at slow rate compared to that

catalysed by La(OTf)3/H3PO4 at 0.8 and 0.4 mol% catalyst loading. In order to achieve

comparable activity of the catalyst, the concentration of La(OTf)3/H3PO4 had to be reduced to

0.2 mol%. The reactions catalysed by Amberlyst 15 exhibited similar activity at different stirring

speed (1000 and 500 rpm), hence the mass transfer limitations did not affect the rate of the

Page 96: Applications of metal triflates and assisted acids as

76

reaction. This has thus shown that the combined acid system has high efficacy than the ion

exchange resin (Amberlyst 15). However, the main advantage associated with solid acids is on

the ease of catalyst recovery (via filtration) and thus can be reused.

2.13 The reactivity of other olefins with methanol

The scope of the reaction was broadened using various other alcohol and alkene substrates.

Among the olefins evaluated were 2-methyl-1,3-butadiene (isoprene), methylenecyclopentane

and styrene, Scheme 2.8.

Isoprene Styrene Methylenecyclopentane

Scheme 2.8: Other olefins used for etherification with methanol.

2.13.1 The reactivity of isoprene with methanol

The reaction of isoprene with methanol in the presence acid leads to the formation of three

ether products, (Scheme 2.9).

Page 97: Applications of metal triflates and assisted acids as

77

CH3OHMeO

2-methoxy-2-methyl-3-butene(2M2M3B)

OMe1-methoxy-3-methyl-2-butene

(1M3M2B)

OMe

MeO

2,4-dimethoxy butane(2,4DMB)

Acid

2.9

(2.10)

(2.11)

(2.12)

Scheme 2.9: Products distribution during etherification of isoprene

The very first reaction of isoprene with methanol was performed in the absence of a catalyst

(blank run), and the reaction was proven be essentially non-existent. The second experiment

was catalysed by Zr(OTf)4 (1.0 mol%) at 100 °C, using methanol to olefin mole ratio of 13:1. The

results thereof are shown in Figure 2.16. The products obtained from these experiments were

purified via distillation techniques and characterised using GC-MS and nuclear magnetic

spectroscopy (1H NMR and 13C NMR). The quantification of products was performed using gas

chromatography with flame ionisation detection (FID).

Page 98: Applications of metal triflates and assisted acids as

78

0

5

10

15

20

25

30

35

40

45

50

0 20 40 60 80 100 120 140 160

Mo

l%

Time (Min.)

Conversion

2,4DMB

2M2M3B

1M3M2B

Figure 2.16: Etherification of isoprene with methanol in the presence of Zr(OTf)4.

The equilibrium conversion of approximately 46 mol% was achieved in these experiments. Two

intermediate products (2-methoxy-2-methyl-3-butene and 1-methoxy-3-methyl-2-butene) were

formed during the initial stages of the reaction, and were subsequently converted to 2,4-

dimethoxybutane. The selectivities of the products were 59% to 2-methyl-2-methoxy-3-butene,

24% to 1-methoxy-3-methyl-2-butene and 18% to 2,4-dimethoxybutane at 60 minutes. The next

set of reactions was performed using La(OTf)3 and H3PO4 independently under the same set of

conditions as for Zr(OTf)4. La(OTf)3 gave no products, while H3PO4 gave a small amount of

products (<1%) over a period of 150 minutes, each catalyst being present at a loading of 1.0

mol%. A combination of the two acids was prepared (La(OTf)3/H3PO4), 1.0 mol% of the

combined acid system was prepared from 1 mol% of each of the components) and evaluated for

the etherification of isoprene with methanol under the same set of conditions (100 °C, 6 bar

(N2), 1.0 mol%), (Figure 2.17).

Page 99: Applications of metal triflates and assisted acids as

79

Figure 2.17: Reaction of isoprene with methanol catalysed by La(OTf)4/H3PO4

The results afforded by the combined acid system (La(OTf)3/H3PO4) gave a substantial increase

of the reaction rate, which is comparable to that shown by Zr(OTf)4. The equilibrium conversion

(47 mol%) was achieved within 75 minutes, and the two intermediates products were also

obtained, each of which was eventually converted into 2,4-dimethoxybutane. In both instances,

the formation of 2-methoxy-2-methyl-3-butene was fast. The selectivities to intermediate

products were 67% to 2M2M3B and 19% to 1M3M2B while the selectivity to 2,4DMB was 14%

(based on GC analysis). These results have once again showed that two inactive acids can

jointly catalyse the reaction, giving high activity.

2.13.2 Etherification of styrene with methanol

The etherification of styrene with methanol yielded to only one product, namely 1-

methoxyethylbenzene (Scheme 2.10), not unexpectedly given the stability of benzylic carbon in

comparison with the primary analogue. Styrene is a conjugated molecule, and as a result it is

relatively stable compared to other secondary olefins. Therefore the compound can form a

secondary carbocation that reacts with methanol to form the ether product.

Page 100: Applications of metal triflates and assisted acids as

80

CH3OH

O

Acid

2.13 2.14

Scheme 2.10: Reaction of styrene with methanol

The initial set of reactions was performed at 100 °C, 6 bar (N2) and methanol to olefin ratio of

13:1, using 0.1 mol% of Al(OTf)3. This reaction afforded < 1 mol% of the product. Hence, the

subsequent reactions catalysed by Al(OTf)3 and Zr(OTf)4 were performed using the catalyst

loading of 1.0 mol% (under the same set of conditions as the initial experiment), but only 5

mol% of the product was obtained after a reaction period of 150 minutes. There was no product

obtained from the reactions catalysed by 1.0 mol% of La(OTf)3 and 1.0 mol% H3PO4

independently. Due to the sluggish nature of these reactions, a reaction catalysed by 6.5 mol%

of Al(OTf)3 was performed to establish the equilibrium conversion (Figure 2.18), which was

found to be 57 mol% over a period of 60 hours.

Figure 2.18: Results for etherification of styrene with methanol

Page 101: Applications of metal triflates and assisted acids as

81

The assisted acid [La(OTf)3/H3PO4] was prepared and evaluated in this reaction, 1 mol%

catalyst loading was utilised and the acid combination also gave similar activity to the triflate

salts of Zr and Al, while the individual acids showed no activity. This set of experiments also

served to demonstrate that the mixed acid system is applicable to styrene, which may

potentially polymerise under these conditions as a side reaction. A possible explanation to the

low yields obtained from the styrene reactions could be attributed to the fact that styrene is a

secondary olefin, and that its reactivity would be lower than those of tertiary olefins. However,

styrene is a conjugated compound, which would lead to resonance stabilisation of any

carbocation that forms at the benzylic position. This substrate should, therefore, be more

reactive and should undergo the etherification reaction to a greater extent compared to non-

conjugated secondary olefins where resonance stabilisation of cationic species is not possible.

The 60 hour reaction using Al(OTf)3 as catalyst demonstrated that this substrate is indeed

susceptible to this type of chemistry.

2.13.3 Etherifcation of methylenecyclopentane (MCP) with methanol

Methylenecyclopentane is a tertiary olefin that may form relatively stable tertiary carbocations in

the presence of acid. If this is the case, then such carbocation species may undergo addition of

methanol to form the desired tertiary ether. The reaction of methylenecyclopentane with

methanol yields methyl 1-methylcyclopentyl ether. During this process, the olefin also

undergoes a small measure of an isomerisation reaction, forming 1-methylcyclopentene

(Scheme 2.11).

CH3OH

Oacid

Etherification reaction

Isomerisation reaction

2.15 2.16

2.15 2.17

Scheme 2.11: Etherification and isomerisation of methylenecyclopentane

Page 102: Applications of metal triflates and assisted acids as

82

The blank reaction (no catalyst added) was performed at 100 °C, and using methanol to olefin

mole ratio of 13:1. The reaction was performed under N2 atmosphere (6 bar) to retain the

reagents in the liquid phase. The results show that the reaction occurred to a minor and

insignificant extent. Subsequently, Lewis acidic metal triflates were evaluated as catalysts, at a

catalyst concentration of 0.1 mol% in all experiments. Figure 2.19 shows the results obtained

when using Al(OTf)3 as catalyst. This Lewis acid effected high conversion of methylene

cyclopentane, with equilibrium yield of approximately 60 mol% reached within 30 minutes. The

isomerisation reaction leads to the formation of approximately 27% of 1-methylclopentene by

the time equilibrium had been achieved. The olefin isomer should form the same ether product

via an identical tertiary carbocation, in line with Markovnikov’s rule. The isomerisation proceeds

presumably because the internal alkene formed is more stable than a terminal alkene, which is

in accordance with Zaitsef’s rule which states that “if more than one alkene can be formed

during elimination reaction, the more stable alkene is the major product”. This rule is relevant

because the reaction is believed to proceed via a carbocation intermediate. Apart from being

able to react with methanol to form the desired ether, this intermediate may also undergo an

elimination reaction to form an alkene product again. It is during this elimination step that

Zaitsef’s rule comes into play. Since the internal alkene is thermodynamically more stable than

the exocyclic methylene alkene, the former should form by preference.

Page 103: Applications of metal triflates and assisted acids as

83

0

10

20

30

40

50

60

70

80

90

100

0 10 20 30 40 50 60 70

Mo

l%

Time (Min.)

Methylenecylopentane 1-methylcyclopentene 1-methoxymethylether

Figure 2.19: Etherification of MCP with methanol in presence of Al(OTf)3.

The reaction catalysed by Zr(OTf)4 also showed a high rate of conversion of the substrate,

leading to equilibrium within 15 minutes. All of the results obtained when using various Brønsted

and Lewis acids are shown in Figure 2.20. The isomerisation reaction was noted in all instances

where the etherification reaction was found to be promoted by the various catalysts, as the

Figure shows. This should necessarily be the case since the reaction is believed to proceed via

a carbocation intermediate which may also eliminate under the reaction conditions to again

provide an alkene, as already stated.

Page 104: Applications of metal triflates and assisted acids as

84

0

10

20

30

40

50

60

Zr(OTf)4 Al(OTf)3 H2SO4 La(OTf)3

Mo

l%

Methyl-1-methylcyclopentyl ether 1-methylcyclopentene

Figure 2.20: Products distribution afforded by various acids

The Brønsted acid H2SO4 also showed good activity but which was lower than that obtained

with the Zr and Al triflates. On the other hand, H3PO4 failed to catalyse the reaction, giving the

same minimal yield of the ether as that of the blank reaction, thereby suggesting that H3PO4

was essentially inactive as a catalyst for this reaction under these conditions. La(OTf)3 gave

some activity, indicating that the acid was slightly active for this type of chemistry, as had been

previously established in this project.

The reactions catalysed by combinations of La(OTf)3/H3PO4 showed a significant enhancement

of the reaction rate (Figure 2.21) over the individual catalyst components, as was now expected

to be the case. The reaction was performed at 100 °C, using a 13:1 mole ratio of methanol to

MCP, and a mixed catalyst concentration of 0.1 mol% relative to the alkene. The equilibrium

yield was reached within 60 minutes and the selectivity remained >99.9%. The isomerisation

reaction was noted to take place, with 1-methylcyclopentene accounting for an increasing

amount of the alkene component as a function of time, as would normally be expected where a

given reaction (in this case the isomerisation of the alkene) is allowed to run to its

thermodynamic conclusion.

Page 105: Applications of metal triflates and assisted acids as

85

0

10

20

30

40

50

60

70

80

90

0 20 40 60 80 100 120

Mol

%

Time (Min.)

Methylenecylopentane 1-methylcyclopentene Methyl-1-methylcyclopentyl ether

Figure 2.21: Etherification of MCP with methanol catalysed by La(OTf)3/H3PO4.

The results of the experiment catalysed by La(OTf)3/H3PO4 also showed high conversion of

methylenecyclopentane, forming the ether and the alkene isomerisation product (Figure 2.22).

The amount of product formed in a given time frame is higher than the instances where

individual acids were used. It is also noticeable that the rate of the latter experiment is

comparable to that of a reaction catalysed by La(OTf)3/H3PO4.

Page 106: Applications of metal triflates and assisted acids as

86

0

10

20

30

40

50

60

70

80

0 10 20 30 40 50 60 70

Mol

%

Time (Min.)

MCP Conversion 1-methylcyclopentene Methyl-1-methylcyclopentyl ether

Figure 2.22: Etherification MCP with methanol catalysed by La(OTf)3/H2SO4

The results in Table 2.4 show a comparison of TOF afforded by each catalyst over the period of

10 minutes during etherification of methylenecyclopentane with methanol. These results show

that metal triflates of Zr and Al (entry1 and 2) exhibit high activity, whereas the combined acids

(entry 3 and 4) show lower activity than the latter, with La(OTf)3/H3PO4 giving higher activity that

H2SO4 analogue. It is also evident that individual acids exhibited the least activity (entry 5 and

6). The activity of H3PO4 is comparable to the blank run, hence H3PO4 is virtually inactive.

Page 107: Applications of metal triflates and assisted acids as

87

Table 2.4: Comparison of activity of various catalysts for etherification of MCP and methanol.

Entry Catalyst system TOF (mol prod./mol cat. h)

1 Zr(OTf)3 2170

2 Al(OTf)3 1308

3 La(OTf)3/H3PO4 754

4 La(OTf)3/H2SO4 516

5 H2SO4 286

6 La(OTf)3 262

7 blank run 56

2.14 Etherification of tertiary olefins in the presence of other olefins It was anticipated that the etherification of tertiary olefins may be performed selectively in the

presence of other olefins such as 1-hexene. It was thus decided to investigate this proposal by

performing the etherification reaction using a mixture of olefins. The reaction was performed

using an olefin mixture consisting of 1-hexene and 2-methyl-2-butene with methanol. The mole

ratio composition of 1-hexene to 2-methyl-2-butene to methanol was 1:1:15. The reactions were

performed in the presence of 0.1 mol% Zr(OTf)4. The results showed that only TAME was

formed, while 1-hexene remained un-reactive. The GC trace in Figure 2.23 shows a comparison

of the reaction mixture peaks at the beginning of the experiment (before the addition of the

catalyst) and at the end of the experiment (after 150 minutes). Clearly there is only one product

peak being that for TAME, which indicates the etherification of 2M2B only. No secondary

products were noted, from which can be concluded that etherification of 1-hexene does not

proceed under the reaction conditions used during the investigation.

Page 108: Applications of metal triflates and assisted acids as

88

Figure 2.23: GC trace for selective etherification of 2M2B

A similar observation was also evident when using both (primary and secondary olefins, i.e. only

tertiary olefin (2M2B) was reactive with methanol, Figure 2.24. This study has thus shown that

tertiary olefins can undergo selective etherification with methanol in the presence of primary and

secondary olefins, an observation that is important in instances where selectivity may be

required.

Page 109: Applications of metal triflates and assisted acids as

89

Figure 2.24: Reactivity of methanol with 2M2B in the presence of 1-hexene and cyclohexene

2.15 Etherification of other alcohols with 2-methyl-2-butene

The etherification reaction of ethanol with 2-methyl-2-butene in the presence of an acid is shown

in Scheme 2.12. TAEE (tert-amyl ethyl ether) also finds it use as fuel additive, due to its high

octane number.230

OHAcid

O

TAEE2.1 2.18

2.19

Scheme 2.12: Etherification of ethanol with 2M2B

The reaction of ethanol with 2M2B is highly selective to one product. However, in the presence

of water, hydration of the olefin occurs, resulting in the formation of tert-amyl alcohol, which may

lead to a decrease of selectivity. This side reaction was only noted when the water content of

Page 110: Applications of metal triflates and assisted acids as

90

the ethanol was relatively high, at about 0.5% content. By-product formation could be altogether

eliminated if 100% ethanol was used instead. Furthermore, the isomerisation of the olefin

forming 2-methyl-1-butene (2M1B) also occurred to a small extent, but the latter reaction does

not affect selectivity, since the etherification reaction of this material leads to the desired tertiary

ether. The first reaction in this set of experiments was performed in the presence of Zr(OTf)4

(0.1 mol%) at 100 °C and using an ethanol to olefin mole ratio of 13:1 (in line with the reactions

making use of methanol). It was quickly established that these reactions proceed only slowly,

and the catalyst concentration was therefore increased to 0.5 mol% in the subsequent

reactions, thereby allowing product formation in good yields in acceptable time frames. The

comparison of TAEE yields afforded by various acids is shown in Figure 2.25.

Figure 2.25: The activity of various acids during etherification of ethanol and 2M2B.

The blank reaction showed virtually no product formation in the absence of a catalyst.

Furthermore, the usage of H3PO4 did not improve the conversion, while lanthanides led only to

Page 111: Applications of metal triflates and assisted acids as

91

minor improvements, as would now be expected on the basis of prior results from this study. On

the contrary, Zr(OTf)4 gave the best yield and highest rates of reaction, followed by H2SO4 and

Al(OTf)3. The acid combinations (assisted acids) were also evaluated in this study (Figure 2.26).

Figure 2.26: Reaction catalysed by assisted acids.

2.16 Conclusions

The study has shown that the etherification of olefins with alcohols in the presence of

homogeneous acid catalysts (either Brønsted or Lewis acids) is possible. Hence the following

concluding remarks are made based on the results obtained from the study:

The study has shown that the metal triflate Lewis acids of aluminium and zirconium give the

highest activity for etherification reactions. Scandium triflate also showed some activity, while

the triflate salts of the lanthanide series were virtually inactive. The metal halide Lewis acids of

Al, Zr and La were also inactive as catalysts for the etherification reactions.

Among the Brønsted acids evaluated, it was shown that the organic acids such as

methanesulfonic acids para-toluenesulfonic acid and triflic acid showed some activity for the

Page 112: Applications of metal triflates and assisted acids as

92

reaction. Only sulfuric acid and nitric acid were active among the mineral Brønsted acids

evaluated, with H3PO4 and HCl being completely inactive under the reaction conditions. The

solid phosphoric acid (55% H3PO4 supported in silica) was also proven to be inactive for the

reaction.

This study has shown that the use of Lewis acids such as La(OTf)3 in combination with mineral

Brønsted acids such as H3PO4 as catalysts for etherification reactions leads to a significant

enhancement of the reaction rate, which is indicative of the fact that highly active systems are

formed that effectively catalyse the reaction. However, the combinations of the already active

Lewis acids [Zr(OTf)4 and Al(OTf)3] with mineral Brønsted acids did not yield to any rate

enhancement. Furthermore, the combinations based on organic Brønsted acids also did not

afford any rate enhancement. The NMR-spectroscopy experiments showed that a new complex

is formed when a lanthanide triflate salt is mixed with H3PO4.

It has been shown in this study that metal triflate salts of Al and Zr can be recycled at least five

times without significant loss of activity. Furthermore, the combined acid systems (assisted

acids) can also be recycled four times with no substantial loss of activity. However, in the case

of La(OTf)3/solid phosphoric acid assisted acid, the active species, which is La(OTf)3/H3PO4

leaches out of the support (silica), but the leached material remains active in the reaction.

The reactivity of alcohols varies significantly, with methanol being more reactive than ethanol,

while a bulkier alcohol (n-butanol) did not react with 2M2B; di-n-butyl ether was formed instead.

This study has also shown that primary and secondary olefins are not reactive during

etherification reactions, hence the tertiary olefins can be etherified selectively, even if the

primary and secondary alkenes are present

The literature review has shown that solid acids such as Amberlyst 15 (commercial catalyst),

and zeolites are widely used for the etherification of tertiary olefins with alcohols. But, there has

not been any mention of the metal triflates, assisted acids or combined acids systems as

catalysts for etherification reactions. Furthermore, the current study has showed that the

combined acid system is more effective than Amberlyst 15.

Page 113: Applications of metal triflates and assisted acids as

93

CHAPTER 3

Friedel-Crafts Alkylation This chapter discusses the experimental results on the use of assisted acids composed of metal

triflates and mineral Brønsted acids as catalysts for Friedel-Crafts alkylation of phenolic

compounds with isobutylene.

3.1 Introduction

The Friedel-Crafts alkylation reaction is among the most fundamental reactions in the field of

synthetic organic chemistry, and leads to the synthesis of various important aromatic

compounds via formation of C-C bonds. This reaction is widely utilised in both the laboratory

and commercial industrial scale for the production of fine chemical and valuable synthetic

intermediates. The anhydrous AlCl3 has maintained its wide use as catalyst for the process ever

since it was introduced by Friedel and Crafts in 1877.231 Since then, a number of Lewis and

Brønsted acid catalysts have been introduced for the process. The Brønsted homogeneous

acids identified for the reactions include concentrated H2SO4, HF, p-toluenesulfonic acid232,233

and trifluoromethanesulfonic acid (triflic acid).234 Furthermore, metal halides such as BF3,235

SnCl4,236

SbF5 and ZnCl2 and lately, metal salts of trifluoromethanesulfonate237 also known as

metal triflates, were identified as alternative Lewis acid catalysts for the process. Olah et al.238

reported the use of metal triflate salts of boron-, aluminium- and gallium as catalysts for the

alkylation of toluene, the alkylating agents being alkyl halides. It was found in the study that all

three Lewis acids were effective, convenient and safer during the alkylation reactions.

Furthermore, it was observed that aluminium and gallium triflate complexes have low solubility

in the reaction mixture, and thus these Lewis acids were used as heterogeneous catalyst

systems. However, the addition of nitromethane facilitated the solubility of the Lewis acids and

hence the reaction became homogeneous.239

The heterogeneous catalysts for the Friedel-Crafts alkylation reaction that have been reported in

literature include sulfated metal oxides,240 mesoporous molecular sieves,241 clays and

zeolites.242 The common advantages associated with using heterogeneous catalysts is in the

Page 114: Applications of metal triflates and assisted acids as

94

ease of catalyst recovery and recycling, less potential of contamination of the product with the

catalyst, and also the possibility of carrying out the reactions in a continuous reactor set-up,

rather than in batch mode. Poliakoff and co-workers243 have reported the use of solid acids such

as Amberlyst 15 and Purolite CT-175, inorganic supported solid acids (Nafion SAC-13 and

deloxan ASPP1/7), and also a Zeolite (Zeolyst CBV 600) as catalysts for alkylation of anisole

with n-propanol in supercritical CO2. Amberlyst 15 showed maximum conversion of anisole at

150 ˚C, and thereafter the activity dropped due to de-sulfonation of the catalyst. Similar results

to Amberlyst 15 were also achieved with Purolite CT 175, except that this catalyst was inactive

at lower temperatures, possibly due to a high moisture content in the catalyst. Nafion SAC-13

showed high optimum temperature, since 91% conversion was achieved at 200 ˚C (the catalyst

was de-sulfonated above 250 ˚C). A conversion of 73% was achieved with Zeolyst CBV 600 at

200 ˚C, with low selectivity to mono-alkylated products.

In a study undertaken by Yadav et al.244, numerous heterogeneous acids were evaluated for

alkylation of 4-methoxy phenol with MTBE. The solid acids include Filtrol-24, K-10

Montmorillonite clay, dodecatungstophosphoric acid supported on clay (DTP/K-10), sulfated

zirconia and cation exchange resin (Deloxane ASP). The order of activity was found to be

Filtrol-24>DTP/ K-10>Deloxane ASP resin>K-10 Montmorillonite>sulphated zirconia (S-ZrO2). In

this study, only two products were obtained (2-tert-butyl-4-methoxyphenol and 2,6-di-tert-butyl-

4-methoxyphenol) with high selectivity to the monoalkylated product.

Kawi and co-workers245 have reported the alkylation of 4-methoxyphenol with tert-butanol over

Zn-Al-MCM-41 mesoporous solid acid to form 2-t-butyl hydroxyanisole. The highest conversion

of 92.2% with selectivity of 99.7% to 2-tert-butylhydroxy anisole was obtained at 150 ˚C, using

2:1 mole ration t-BuOH to 4-methoxyphenol and at an autogenous pressure of 1030 kPa. This

reaction is normally performed in the presence of a homogeneous Lewis acid such as AlCl3,

BF3, SbF5, ZnCl2 and also with Brønsted acids such as H2SO4, HF, H3PO4 and HCl246 during the

alkylation of 4-methoxyphenol with tert-butanol247 or isobutylene.248

The disadvantages associated with conventional metal halide catalysts for Friedel-Crafts

alkylation reactions is that they are moisture sensitive, corrosive and generally pose a safety

hazard. Brønsted acids such as HF and H2SO4 are also corrosive, and as a result the reaction

set-up has to be constructed using special materials, such as hastelloy. The heterogeneous

acids are safer, but they normally exhibit low activity and require energy intensive recycling of

substrates. Hence, a significant research effort has been focussed on the exploration of

environmentally friendly and efficient catalysts for the Friedel-Crafts alkylation. Thus, it is the

Page 115: Applications of metal triflates and assisted acids as

95

main objective of this part of the study to develop the chemistry around acids that are moisture

stable and safer to use for Friedel-Crafts alkylation reactions. The concept of assisted acidity

forms the basis of the study where various combinations of Brønsted and Lewis acids will be

prepared and evaluated as catalysts for Friedel-Crafts alkylation. In this study, three phenolics

substrates, i.e. cresols, anisole and diphenyl oxide will be alkylated with isobutene to test the

activity of the acids.

3.2 Alkylation of cresol

The alkylated cresols are commercially important compounds due to their broad applications.

For example, 2,6-di-tert-butyl-p-cresol or butylated hydroxytoluene (BHT) is an industrially

important antioxidant used to inhibit gum formation in motor and aviation gasoline,249 as

insulating oil and as an antioxidant for natural and synthetic rubbers. 250,251 3-Methyl-6-tert-

butylphenol is used as an intermediate in the manufacture of musk ambrette, while 6-isopropyl-

3-methylphenol (thymol) is used in perfumes252 and also as a disinfectant.253

Alkyl phenols are generally manufactured by addition of an alkyl group to the corresponding

phenol. The source of the alkyl group can be an alkyl halide, alcohol or an olefin. Olefins are

commercially preferred due to their ease of handling and atom efficiency during alkylation. The

alkylation of cresols is often carried out in the presence of Brønsted or Lewis acid catalysts. In

the presence of an acid catalyst, alkylation occurs on the ortho- or para-position and the product

distribution is largely dependent on the positions already occupied by substituent groups. In the

case of tert-alkylphenols, rearrangement and dealkylation may occur at higher temperatures.

Alkylation reactions are reversible, and this aspect of alkylation is desirable during separation of

close boiling isomers. For example, Sharma et al.254 have reported a process for the separation

of m/p-cresols; 2,5- and 2,4-xylenols, and p-cresol/2,6-xylenol by first alkylating the phenolic

mixture with α-methylstyrene and diisobutylene in the presence of p-toluenesulfonic acid and

Amberlyst 15, followed by separation and subsequent decomposition or dealkylation. According

to Sharma, di-alkylation of m- and p-cresols occurred. However, if the reaction is left for a

prolonged time at reaction temperature, decomposition of the p-alkylated product occurs giving

back meta-cresol. A patent by Koppers Company Inc., 255 disclosed a process for selective de-

alkylation of 4,6-di-tert-butyl-m-cresol to yield mainly 6-tert-butyl-m-cresol (an intermediate for

the synthesis of thymol, used as an antiseptic and as an antioxidant). This process is performed

by heating 4,6-di-tert-butyl-m-cresol in the presence of catalytic amounts of aryloxides of

zirconium, hafnium, niobium and tantalum until de-alkylation occurs. The alkylation of m-cresol

Page 116: Applications of metal triflates and assisted acids as

96

with isobutylene in the presence of sulfuric acid under mild conditions yields 6-tert-butyl-m-

cresol. Unfortunately, the process only provides low yields to the desired product.256 Another

patent by Leston257also provide some examples on the process of de-alkylation of 4,6-di-tert-

alkyl-meta-cresol in the presence of aluminium aryloxide catalysts.

UOP Inc. filed a patent which discloses a process for the preparation of BHT via a reaction of p-

cresol with isobutylene (mole ratio of isobutylene to p-cresol being 10:1 at 100 ˚C) in presence

of macroreticular cation exchange resins bearing sulfonic acid groups (highly cross linked

polystyrene-divinylbenzene with a surface area of about 540 m2/g and average pore size of

about 51 Å) and Amberlyst 15.258 The emphasis of this patent is that high surface area (525-575

m2/g) and low pore size or pore diameter (40-60 Å) of the catalyst leads to high BHT yields

(>90%); the rest being mono-alkylated p-cresol, which implies a quantitative conversion of p-

cresol and good selectivity to the desired product. Wetzel et al.259 disclosed a process for

alkylation of phenol and ortho-cresol with isobutylene in the presence of highly acidic

homogeneous aryl sulfonic acids (Scheme 3.1).

OH OHOH OH

(84%) (16%)

trifluoromethanesulfonic acid

1 h, 120 oC

OHOH

OH OH

(79%) (21%)

m-benzenedisulfonic acid

5 h, 150 oC

3.1 3.2 3.3 3.4 3.5

3.6 3.73.8

3.93.2

Scheme 3.1: Alkylation of phenol and o-cresol in the presence of sulfonic acids

The aryl sulfonic acids referred to in the literature as catalysts by Wetzel for such types of

reactions include para-chlorobenzene sulfonic acid, 4,4-diphenyldisulfonic acid, meta-benzene

disulfonic acid, nitrobenzenesulfonic acid, 2,4,6-trinitrobenzenesulfonic acid and

trifluoromethanesulfonic acid (Scheme 3.2).

Page 117: Applications of metal triflates and assisted acids as

97

3.10 3.11 3.12

3.13 3.14

3.15

CH3

SO OOH

SHO

O

O

S OH

O

O SO O

OH

NO2

SO O

OH

NO2O2N

NO2

SO O

OH

S

O

OOH

S OHF3C

O

O

para-toluenesulfonic acid 4,4'-diphenyldisulfonic acid 4-nitrobenzenesulfonic acid

2,4,6-trinitrobenzenesulfonic acid meta-benzenedisulfonic acid trifluoromethanesulfonic acid

Scheme 3.2: Aryl sulfonic acid catalysts

Wetzel et al. claim that the effectiveness of these aromatic and alkyl sulfonic acids as selective

para-alkylating agents is directly related to their acid strength. These acids can be neutralised

with any base after completion of the reaction, followed by fractional distillation of the reaction

mixture. The catalysts or their salts would remain in the reboiler, due to their high boiling point

characteristics. The patent also claims that these catalysts are effective for rearrangement of

ortho-/meta-substituted phenols to the desired para-substituted compounds (Scheme 3.3).

OHOH

m-benzenedisulfonic acid (1%)

1.5 h, 150 oC

(84%)

3.163.3

Scheme 3.3: Rearrangement of 2-tert-butylphenol to 4-tert-butylphenol

Page 118: Applications of metal triflates and assisted acids as

98

The acids may also function as de-alkylation catalysts (Scheme 3.4).

OHOH

m-benzenedisulfonic acid (1%)

1.5 h, 150 oC

(81%)3.5

3.3

Scheme 3.4: De-alkylation 2,4-di-tert-butylphenol

In Chapter 2 of this thesis, it was demonstrated that Lewis assisted Brønsted acids (metal

triflate salts and mineral acids combined) could be used as efficient catalysts for etherification of

tertiary olefins with alcohols, yielding various ethers. Furthermore, the literature has shown that

metal chlorides can be used in conjunction with Brønsted acids, resulting in super acidic

systems such as SbF5-HSO3F also known as “magic acid”,260 which can be used for alkylation

reactions. However, there is no mention of the catalytic activity of hydrolytically stable metal

triflate Lewis acids combined with common mineral Brønsted acids, and this has then presented

an opportunity to explore the catalytic activity of metal triflates combined with mineral Brønsted

acids as assisted acid systems for alkylation reactions.

The activity of the acid combinations was evaluated in the Friedel-Crafts alkylation of a meta-

and para-cresol mixture with isobutylene as an alkylating agent. The proposed mechanism for

the reaction of isobutylene and cresol is shown in Scheme 3.5. The hydroxyl group activates the

cresol for substitution in both ortho- and para-positions, accounting for the regioselective

outcome of the alkylation reaction. This alkylation reaction is initiated by protonation of

isobutylene, yielding a tertiary carbocation via the Markovnikov mechanistic pathway, followed

by substitution on the ortho-position (in the case of p-cresol). The 2-tert-butyl cresol

subsequently undergoes further substitution on position 6, yielding 2,6-di-tert-butyl-p-cresol

(DBPC).

Page 119: Applications of metal triflates and assisted acids as

99

3.2H+

OH

OHH

Base

OH

MBPC

H+

OH

DBPC or BHT

3.17

3.18

3.19

3.203.21

Scheme 3.5: The acid catalysed Friedel-Crafts alkylation of p-cresol

In the case of meta-cresol, isobutylene is first added onto the para-position, forming 4-tert-butyl-

m-cresol, followed by addition of another isobutylene molecule at the ortho- position

(Scheme 3.6) to form 2,4-di-tert-butyl-m-cresol (DBMC). According to the literature,261 the

alkylation of cresol is reversible. Thus DBMC and BHT can be de-alkylated to their mono-

alkyated products and even further to their cresylic forms. The isobutylene oligomerisation side

reaction occurs in the presence of acid catalyst. This reaction can form isobutylene dimers,

trimers and sometimes even tetrameric hydrocarbon compounds. This side reaction is

undesirable as it consumes the valuable isobutylene reactant.

Page 120: Applications of metal triflates and assisted acids as

100

MBMC DBMC

OH OH OH

3.22 3.23 3.24

Scheme 3.6: Alkylation of m-cresol with isobutylene

3.3 Evaluation of metal triflate Lewis acids as catalysts for the butylation of cresol

The initial set of experiments involved screening of different metal triflate Lewis acids such as

Zr(OTf)4, Al(OTf)3, Sc(OTf)3 and lanthanide triflates salts [La(OTf)3, Gd(OTf)3, Sm(OTf)3,

Yb(OTf)3] as catalysts for butylation of meta- and para-cresol mixtures, with isobutylene as an

alkylating agent. The activities of two mineral Brønsted acids (H3PO4 and H2SO4) were also

evaluated. It is well-known that the catalytic activity of a Brønsted acid is largely dependent on

its dissociation constant (pKa). For instance, H2SO4 with pKa = -3 is expected to exhibit higher

activity than H3PO4 having pKa = 2. The cresol substrate used in these reactions is a mixture of

m- and p-cresols. The m- and p-cresol balance in the mixture is 57.1% and 41.4% respectively,

the other (1.5%) is made up of unidentified substances. All of the cresol butylation reactions

were performed at 70°C under constant isobutylene pressure (1.1 bar) and the mixtures were

stirred at 1200 rpm in a Parr reactor. The catalysts were weighed directly into the reactor at

ambient conditions. Special exclusion of air or moisture was not required given the literature

assertion that metal triflate Lewis acids can tolerate moisture compared to metal halide

counterparts which are deactivated in the presence of moisture. 262

The results on the imminent tables showed that Al(OTf)3 and Zr(OTf)4 gave good activity and

selectivity to the desired products. Sc(OTf)3 exhibited lower activity compared to the triflate salts

of Zr and Al. During the alkylation reactions, it was observed that an exotherm of about 10 ˚C

prevailed soon after the introduction of isobutylene into the reactor with cresols and catalyst.

Therefore, the internal setup of the reactor was modified with a cooling coil for circulation of

water to control the exotherm and maintain the reaction mixture at the required temperature.

The lanthanide metal triflate salts were essentially inactive for the butylation of cresols, and

these Lewis acids did not lead to the production of any product.

Page 121: Applications of metal triflates and assisted acids as

101

The reaction products were characterised by GC analysis using authentic samples purchased

from Sigma Aldrich as standards. The analysis of the reaction mixture showed that selectivity to

mono-alkylated cresols is higher during the initial stages of the reaction. However, the mono-

alkylated products undergo addition of a second isobutylene, leading to a decrease in the

selectivity towards mono-alkylated cresols while the selectivity to di-alkylated cresols increases.

The GC analysis also showed that the oligomerisation side reaction of isobutylene occurred,

leading to the formation of C8, C12 and even small amounts of C16 hydrocarbon compounds. It

is known that olefins can easily form oligomers in the presence of an acid by a cationic

mechanism,263 Scheme 3.7.

3.2 3.17

3.17 3.2

H+

3.25

Scheme 3.7: Dimerisation of isobutylene under acidic conditions

The product distribution for the reaction catalysed by Zr(OTf)4 is shown in Table 3.1, as

determined by GC analysis. Since the two cresol isomers that make up the substrate possess

different boiling points, the individual reaction profile may be readily tracked. The data are

presented in a single table (Table 3.1) which may be used to compare the reactivity of each

cresol substrate as rates of reaction, selectivities etc. A typical graph representing the products

distribution is also shown in Figure 3.1.

The reaction was performed using 0.05 wt% of Zr(OTf)4 as catalyst at 70 ˚C. An excess of

isobutylene was used at a pressure of 1.1 bar. This experiment gave quantitative conversion of

both m-cresol (57.1 mol%) and p-cresol (41.4 mol%) in a reaction that was allowed to proceed

for 120 minutes. During the initial stages of the reaction, mono-butylated products MBMC (29.7

mol%) and MBPC (20.2 mol%) were formed within 10 minutes, as determined by GC analysis of

samples removed from the reactor. At this stage the results indicate that the reactivity of m-

Page 122: Applications of metal triflates and assisted acids as

102

cresol to form MBMC is higher compared to that of p-cresol to form MBPC (comparing the data

presented in Table 3.1). This could be due to the fact that m-cresol is less sterically hindered

than p-cresol, hence m-cresol can undergo substitution much easier. A more likely explanation

is that the functional groups are cooperative in their positions of activation on the ring in m-

cresol, which is not the case in p-cresol. MBMC and MBPC were subsequently converted into

the corresponding DBMC (7.2 mol%) and BHT (2.1 mol%), respectively, over a period of 10

minutes.

The addition of the first isobutylene molecule to the aromatic ring to form mono-butylated

products is facile, while the addition of a second isobutylene molecule is slow. The latter could

be explained by the fact that mono-butylated cresol is already a bulky molecule, and the steric

hindrance retards the rate at which the second isobutylene molecule is added to the compound

to form di-butylated cresols. At the end of the reaction (120 minutes), small amounts of the

mono-butylated products [MBMC (3.4 mol%) and MBPC (6.6 mol%)] were present, while a

significant quantity of di-butylated products [DBMC (47.4 mol%) and BHT (34.7 mol%)] was

formed. Comparable selectivity (~84%) to both DBMC and BHT was obtained after 120 minutes.

The selectivity of each di-alkylated product was calculated based on the yield afforded by each

cresol separately. For example:

The oligomerisation of isobutylene (side reaction) occurred, yielding a total of 3.4 wt% of

isobutylene oligomers at 120 minutes. It is, however, important to keep the oligomerisation side

reaction minimal because the oligomerisation side reaction consumes the valuable starting

material (isobutylene).

Page 123: Applications of metal triflates and assisted acids as

103

Table 3.1: Results for butylation of p-cresol in the presence of Zr(OTf)4

Time (minutes)

Conversion (mol%)

Butylated cresols yield

(mol%)

Selectivity

(%)

Oligomers (wt%)

MC PC DBMC BHT MBMC MBPC DBMC BHT

0 0 0 0 0 0 0 - - 0

5 18.2 12.9 3.5 0.5 14.7 12 19.2 4.0 0.1

10 36.9 24.8 7.2 2.1 29.7 20.2 19.5 9.4 0.1

15 37.1 30.3 18.1 5.9 19 23.9 48.8 19.8 1.0

20 47.3 34.3 28.0 10.5 19.2 24.2 59.3 30.3 1.5

30 52.0 37.9 35.1 17.9 16.8 20.2 67.6 47.0 1.9

40 54.0 39.4 39.2 23.6 14.7 15.9 72.7 59.7 2.0

50 55.0 40.3 41.6 25.8 13.2 14.5 75.9 64.0 3.3

60 55.1 40.5 43.3 28.3 12 12.3 78.3 69.7 3.1

120 57.1 41.4 47.5 34.7 9.1 6.6 83.9 84.0 3.4

Page 124: Applications of metal triflates and assisted acids as

104

Figure 3.1: Butylation of m/p-cresol in the presence of Zr(OTf)4 (0.05wt%) at 70 °C.

The reactions catalysed by Al(OTf)3 gave conversions of meta- and para-cresols of 56.6 mol%

and 41.0 mol% respectively within 120 minutes, Table 3.2. The comparable amounts of MBMC

(11.2 mol%) and MBPC (10.2 mol%) were present at the end of the reaction. Considerable

amounts of di-butylated products [DBMC (44.3 mol%) and BHT (30.8 mol%)] were obtained

after 120 minutes. In this specific experiment, the catalyst showed slightly higher selectivity to

DBMC (79.8%) than BHT (75.1%) at the end of the run. The amount of oligomers (3.3 wt%)

obtained with Al(OTf)3 are comparable to those obtained with Zr(OTf)4 (3.4 wt%). However, the

experiment catalysed by Zr(OTf)4 exhibited higher reaction rates, and hence higher selectivity to

di-alkylated products as compared to that catalysed by Al(OTf)3.

Page 125: Applications of metal triflates and assisted acids as

105

Table 3.2: Butylation of m/p-cresol in the presence of Al(OTf)3 (0.05 wt%), at 70 ˚C.

Time (minutes)

Conversion (mol%)

Butylated cresols yield

(mol%)

Selectivity

(%)

Oligomers (wt%)

MC PC DBMC BHT MBMC MBPC DBMC BHT

0 0 0 0 0 0 0 - - 0.0

5 12.1 8.7 2.3 0.2 10.4 7.9 18.3 2.3 0.1

10 25.9 18.7 5.8 0.9 23.5 14.4 19.9 5.7 0.2

15 30.3 21.9 13.6 2.3 16.7 19.5 44.9 10.4 0.9

20 36.1 26.2 18.6 4.1 19.2 21.9 49.2 15.7 1.5

30 43.1 31.2 25.9 8.0 16.9 23.3 60.5 25.4 2.4

40 47.5 34.4 31 12.0 16.1 22.5 65.8 34.9 2.9

50 50.5 36.6 34.7 16.6 15.6 19.9 69 45.5 2.2

60 52.4 37.9 36.7 19.0 15.2 19.0 70.7 50.0 3.0

120 56.6 41.0 44.3 30.8 11.2 10.2 79.8 75.1 3.3

While Zr(OTf)4 completely dissolved in the reaction mixture, the majority of the Al(OTf)3

remained insoluble, noted by the catalyst deposits that were observed when the reactor was

opened upon completion of a run. The latter observation is in agreement with the literature264

(the metal triflate salts of boron, aluminium and gallium were insoluble catalysts during

alkylation of toluene with alkyl halides, according to this paper). The latter observation suggests

that metal triflates may function as heterogeneous catalysts during the butylation of cresols. The

heterogeneous catalysts are attractive due to ease of recovery and recycling purposes. It is also

stated in the literature, that improvements in the solubility of metal triflates can be facilitated by

addition of nitromethane (CH3NO2).265

Page 126: Applications of metal triflates and assisted acids as

106

The reaction catalysed by Sc(OTf)3 also showed high conversion of m/p-cresol, i.e. 53.9 mol%

and 39.1 mol% conversions of MC and PC, respectively, within 120 minutes (Table 3.3). When

the reaction was terminated, higher amounts of mono-butylated products (21.2 mol% and 11.8

mol% of MBPC and MBMC) were present, and only 17.9 mol% and 41.1 mol% of BHT and

DBMC were obtained respectively. Thus, it is evident that the rate of the reaction catalysed by

Sc(OTf)3 was relatively slow compared to that of Zr(OTf)4 and Al(OTf)3. It was noticeable that

the catalyst gave higher selectivity towards DBMC (77.7%) than BHT (45.8%). This catalyst also

gave greater amounts of oligomers (5.6 wt%) compared to those obtained from the reactions

catalysed by Al(OTf)3 and Zr(OTf)4.

Table 3.3: Butylation of m/p-cresol in the presence of 0.05 wt% Sc(OTf)3, at 70 ˚C.

Time (minutes)

Conversion (mol%)

Butylated cresol yields

(mol%)

Selectivity

(%)

Oligomers (%)

MC PC DBMC BHT MBMC MBPC DBMC BHT

0 0 0 0 0 0 0 - - 0

5 13.8 8.7 2.8 0.2 11 8.5 20.3 2.3 0.8

10 30.1 15.4 5.9 0.9 24.2 14.5 19.6 5.8 1.5

15 30.2 21.4 12.9 1.9 17.3 19.5 42.7 8.9 2.0

20 36.4 25.5 17.2 3.1 19.2 22.4 47.3 12.2 2.4

30 41.6 30.1 23.7 5.5 17.9 24.6 57.0 18.3 3.1

40 44.9 32.9 28.1 7.6 16.8 25.3 62.6 23.1 3.4

50 47.6 34.8 31.5 9.7 16.1 25.1 66.2 27.9 4.7

60 49.3 36.1 34.1 11.5 15.2 24.6 69.2 31.9 5.1

120 52.9 39.1 41.1 17.9 11.8 21.2 77.7 45.8 5.6

Page 127: Applications of metal triflates and assisted acids as

107

All of the lanthanide triflate salts [La(OTf)3, Sm(OTf)3 and Gd(OTf)3] evaluated in the study did

not yield any product, hence the lanthanide triflates were not active for the reaction under the

conditions used. The results have in essence shown that there is a significant difference in the

activity of the Lewis acids evaluated in this study [(Zr(OTf)4, Al(OTf)3, Sc(OTf)3] and lanthanide

triflates. Zr(OTf)4 and Al(OTf)3 gave the highest activity with conversion >99% for both m- and p-

cresols. Sc(OTf)3 also gave high conversion (~90%) of both-cresols. The comparison of activity

of different catalysts over time is shown in Table 3.4.

Table 3.4: Comparison of activity of the triflate salts of Zr, Al and Sc.

Catalyst Time (minutes)

Conversion (mol%)

Butylated cresol yields

(mol%)

Selectivity

(%)

Oligomers (wt%)

PC MC MBPC MBMC BHT DBMC BHT DBMC

Zr(OTf)4

10 22.3 36.9 20.2 29.7 2.1 7.2 9.4 19.5 0.1

30 38.1 51.9 20.2 16.8 17.9 35.1 47.0 67.6 1.9

120 41.3 56.5 6.6 9.1 34.7 47.4 84.0 83.9 3.4

Al(OTf)3

10 15.3 29.3 14.4 23.5 0.9 5.8 5.9 19.8 0.2

30 31.3 46.4 23.3 16.9 8.0 29.5 25.6 63.6 2.4

120 41.0 55.5 10.2 11.2 30.8 44.3 75.1 79.8 3.3

Sc(OTf)3

10 15.4 30.1 14.5 24.2 0.9 5.9 5.8 19.6 1.8

30 30.1 41.5 24.6 17.9 5.5 23.6 18.3 56.9 3.0

120 39.1 52.9 21.2 11.8 17.9 41.1 45.8 77.7 5.6

Zr(OTf)4 gave similar selectivity of BHT (84%) and DBMC after 120 minutes. The remaining

products are MBPC and MBMC, respectively. Al(OTf)3 gave only 75% selectivity to BHT and

79% to DBMC, which is comparable to the results obtained with Zr(OTf)4. The outcomes of the

Page 128: Applications of metal triflates and assisted acids as

108

reactions using Sc(OTf)3 reflected its lower activity. Furthermore, the results have shown that

butylation of the m-cresol proceeds much faster than that of p-cresol and that the second

butylation of mono-butylated products is a rate-limiting step. It was evident that tripositive metal

triflate salts are less or completely insoluble in the cresol and isobutylene reaction mixture,

because the deposits of these salts were observed at the bottom of the reactor at the end of

each experiment. On the contrary, Zr(OTf)4 was mostly soluble in the reaction mixture under the

conditions, forming a homogeneous mixture with the reaction matrix.

3.3.1 Catalyst recycling studies

The high activity and selectivity and low solubility characteristic of Al(OTf)3 in the reaction

mixture (affording the possibility of recycling) prompted exploration of the influence of various

reaction parameters on the outcome of the reaction. Thus, it was decided to first explore

whether the catalyst could be recycled without any loss of activity and selectivity. All of the

experiments were performed at 70 ˚C under 1.1 bar isobutylene pressure and the stirring speed

was kept constant at 1200 rpm. The catalyst concentration used for the initial experiment was

0.09 wt% (0.053 g, relative to the cresol substrate). The catalyst was recovered by means of

decanting the reaction mixture from the previous experiment, leaving the solid catalyst in the

reactor, followed by charging the reactor with fresh m/p-cresol substrate and carrying out the

subsequent experiment, Table 3.5.

Page 129: Applications of metal triflates and assisted acids as

109

Table 3.5: Recycling of Al(OTf)3 (0.09 wt%) during butylation of m/p-cresol at 70 °C, stirring at 1200 rpm and using 1.1 bar of isobutylene.

Exp. # Time

(minutes)

Cresols conversion

(mol%)

Selectivity

(mol%)

Oligomers* (wt%)

PC MC BHT DBMC

Original

15 24.9 40.6 20.6 47.4

60 41.3 57.1 87.9 82.8

120 41.7 57.7 95.5 87.5 4.7

Recycle 1 15 28.5 39.4 19.6 43.7

60 41.4 56.8 83.5 74.9

120 41.7 56.7 87.2 77.9 5.0

Recycle 2 15 28.8 39.8 24.5 50.5

60 41.4 57.2 86.2 76.7

120 41.7 57.7 95.5 84.0 4.5

Recycle 3 15 26.8 37.0 19.2 51.1

120 41.8 57.8 94.4 77.0 5.2

Recycle 4 60 41.3 56.9 87.9 79.1

120 41.8 57.8 95.4 85.2 4.3

*The oligomers reported were obtained after 120 minutes.

The results reported in Table 3.5 demonstrate that a quantitative conversion of both cresol

substrates was obtained in all the reactions after 120 minutes. The general trend showed that

the catalyst give higher selectivity to DBMC during initial stages of the reaction (15 minutes). At

60 and 120 minutes, the reaction showed higher selectivity to BHT than DBMC. This

observation implies that the first addition of isobutylene molecule to m-cresol is faster than to p-

Page 130: Applications of metal triflates and assisted acids as

110

cresol as has already been noted. However, the second isobutylene addition to MBPC is faster

than to MBMC under the conditions used.

When 0.05 wt% Al(OTf)3 was used, the selectivity to BHT and DBMC was similar (75% and

80%, respectively, Table 3.2, indicative of the fact that higher Al(OTf)3 concentration favours the

formation of BHT rather than DBMC, but this effect is incremental only. The results in Table 3.5

show that Al(OTf)3 could be recycled at least four times without any significant loss of catalytic

activity and it retained its overall selectivity profile.

3.3.2 The influence of catalyst concentration

The influence of catalyst concentration was the second parameter to be studied using Al(OTf)3.

In this part of the study, other reaction parameters such as temperature, stirring speed (1200

rpm) and isobutylene pressure was kept constant. Hence all reactions were performed at 70 °C

under 1.1 bar pressure of isobutylene and the stirring speed was maintained at 1200 rpm. A

constant amount of the m/p-cresol mixture (60.0 g; 0.55 mol) was used in all of the reactions.

The results obtained from this study are shown in Table 3.6.

Page 131: Applications of metal triflates and assisted acids as

111

Table 3.6: The influence of catalyst concentration during butylation of m/p-cresol in the presence of Al(OTf)3 at 70 °C, 1.1 bar isobutylene pressure and stirring at 1200 rpm.

Catalyst

(wt%)

Time

(minutes)

Conversion

(mol%)

Selectivity

(mol%)

Oligomers (wt%)

PC MC BHT DBMC

0.8

15 29.1 40.1 21.2 51.3

60 40.8 56.4 85.8 82.3

120 41.7 57.5 94.7 86.6 3.9

0.35

15 29.8 41.2 19.9 49.2

60 40.7 56.3 79.1 68.0

120 41.7 57.5 93.3 80.8 2.9

0.18

15 27.1 37.5 17.1 48.3

60 39.7 54.8 69.2 71.6

120 41.1 56.7 85.9 75.7 2.9

0.05

15 21.9 30.3 10.1 44.9

60 37.9 52.4 50.1 70.1

120 41.0 56.6 75.1 78.3 3.3

The results obtained after 15 minutes showed that the rate of m/p-cresol conversion decreased

as the catalyst concentration was decreased. The same also holds for the results obtained at 60

and 120 minutes. It is, however, important to note that improvements to the rates of conversions

at catalyst loadings above 0.05 wt% are all marginal, indicative of mass transfer limitations

occurring at 0.05 wt% reaction. This is due to the fact that isobutylene is in the gas phase and is

at relatively low pressure (1.1 bar). Nevertheless, it was imperative to maintain low pressure in

Page 132: Applications of metal triflates and assisted acids as

112

order to form the mono-alkylated cresol products, which would virtually be impossible to form at

high pressures.

These reactions all show high conversion (>95%) of m- and p-cresol after 120 minutes. At

higher catalyst loadings, the rate of the reaction was faster, such that almost all the MBPC was

converted to BHT, accounting for the high selectivity to BHT. The selectivities to BHT for

reactions containing 0.8 wt% and 0.35 mol% catalyst were 94.7 and 93.3 mol%, respectively,

compared to 86.6 and 80.8% selectivity to DBMC after 120 minutes. Apart from a reduction of

the rate of the reaction, a reduction in the catalyst concentration favoured the formation of

DBMC. For instance, at 0.05 wt%, the selectivity to DBMC was 78.3 wt% compared to 75.1 wt%

to BHT. This shift in the selectivity presumable arises due to subtle reactivity differences that are

exaggerated at lower catalyst loading and concomitantly lower reaction rate.

3.3.3 The influence of varying stirring speed

When studying the influence of changing the stirring speed, one gathers data on whether the

reaction rate is influenced by mass transfer limitations related to stirring speed. This is important

in the present context because the catalyst is heterogeneous in nature. It is well known that the

rate of heterogeneous reactions is dependent on the rate at which the substrate comes into

contact with the catalyst, an aspect which is facilitated by stirring. These experiments were

performed at 70 ˚C using 60.0 g (0.55 mol) of m/p-cresol mixture. A high excess of isobutylene

(1.1 bar with constant feeding) was used and Al(OTf)3 was present at a loading of 0.09 wt% with

respect to cresol. The speed of the stirrer was varied from 512 to 1512 rpm (Table 3.7).

Page 133: Applications of metal triflates and assisted acids as

113

Table 3.7: Influence of stirring speed on butylation of m/p-cresol in the presence of 0.09 wt% Al(OTf)3 at 70 °C.

Stirring speed (rpm)

Time (minutes)

Conversion

(mol%)

Selectivity

(mol%)

Oligomers (wt%)

PC MC BHT DBMC

1512 15 27.5 38.0 18.7 56.5

60 40.5 55.9 71.4 77.8

120 41.7 57.6 90.7 85.7 2.7

1052 15 25.1 34.6 15.8 47.1

60 39.1 54.0 60.2 72.8

120 41.0 56.6 81.9 81.7 2.8

512 15 22.9 31.7 9.9 27.6

60 40.3 55.6 72.3 75.6

120 41.7 57.5 91.5 84.8 3.1

A steady increase of the reaction rate with an increase of stirring speed from 512 to 1512 rpm

was evident. The results for the experiment performed at 512 rpm showed high rate of

conversion of m-cresol compared to that of p-cresol. Consequently, the selectivity to DBMC was

higher than that of BHT during the initial stages of the reaction. At 120 minutes, the selectivity to

BHT and DBMC are similar with cresol substrates being converted quantitatively. Thus, it can

be concluded that stirring speed affects the rate of these reactions (mass transfare limitations).

The amount of oligomers formed in all instances was comparable, which indicates that the rate

of a side reaction is not affected by the stirring speed.

Page 134: Applications of metal triflates and assisted acids as

114

3.4 Screening of Brønsted acids

The Brønsted acids evaluated for the alkylation of the m/p-cresol mixture with isobutylene

involved two mineral Brønsted acids, i.e. H2SO4 (98%) and ortho-H3PO4 (99%). The dissociation

constants (pKa) of these two acids vary significantly, with H2SO4 being a stronger acid (pKa = -3)

than H3PO4 (pKa = 2). These Brønsted acids were evaluated under the same set of conditions

as before [70 ˚C, using 60.0 g (0.55 mol) m/p-cresol], with a catalyst loading of 0.19 wt% with

respect to the cresol substrate and high excess of isobutylene (1.1 bar pressure). The stirring

speed was 1200 rpm. The results afforded by H2SO4 are shown in Table 3.8. There was no

conversion of the substrates when using H3PO4 under these conditions and so no data for these

reactions are presented in the Table.

Table 3.8: Butylation of the m/p-cresol in the presence of H2SO4

Time (minutes)

Conversion (mol%)

Alkylated cresols Yield

(mol%)

Selectivity

(%)

Oligomers (wt%)

MC PC MBMC MBPC DBMC BHT DBMC BHT

0 0 0

0 0 0 0 - - 0

5 10.0 5.9

9.8 5.7 0.2 0.2 2.0 3.3 0.2

10 19.6 13.2

18.0 11.6 1.6 1.2 8.2 9.1 0.3

20 34.7 25.8

28.5 20.2 6.2 5.6 17.9 21.7 0.6

30 42.5 30.2

28.1 19.0 14.4 11.2 33.9 37.1 0.7

40 40.2 36.9

26.1 17.9 19.0 14.1 47.3 38.2 1.4

50 48.0 38.4

25.9 16.3 22.1 17.6 46.0 45.8 1.6

60 42.2 33.7

23.6 15.1 24.2 18.6 57.3 55.2 1.8

Page 135: Applications of metal triflates and assisted acids as

115

Sulfuric acid showed some activity; however, retarded reaction rates were obtained with this

acid. Low conversion of both m/p-cresol, i.e. 42.2 and 33.7 mol%, respectively were obtained

within 60 minutes. This acid also led to faster formation of MBMC, as compared to that of

MBPC, as was also found to be the case with the Lewis acid catalysts. While the formation of

BHT and DBMC proceeded slowly, the formation of DBMC was faster than that of BHT. On the

other hand, the selectivity to DBMC (50.7 mol%) was slightly lower than that of BHT (55.2

mol%) after 60 minutes. There was no product obtained when using H3PO4 as catalyst under

the conditions. The significantly divergent activities of the two Brønsted acids can be attributed

to their different pKa values. The results demonstrate that the metal triflates of Zr and Al exhibit

higher activity than Brønsted acids (H3PO4 and H2SO4) under the same set of conditions used.

3.5 Evaluation of assisted acids

An objective of this study was to evaluate a variety of assisted acids (derived from the Lewis

and Brønsted acids that have already been discussed) as catalysts for Friedel-Crafts alkylation

reactions of m/p-cresol with isobutylene. During that part of the study involving the etherification

of olefins with alcohols, it was shown that the activity of Lewis and Brønsted acids, such as

La(OTf)3 and H3PO4, which were completely in-active as individual acids was enhanced by

using those acids in combination (assisted acids). The major discovery was that this

combination showed superior activity for the etherification reactions, delivering catalysts with

activities higher than HOTf, Al(OTf)3 and Zr(OTf)4. Consequently, the assisted acids were

evaluated here for the alkylation of the m/p-cresol mixture with isobutylene. The assisted acids

were prepared by combining the Lewis acids with Brønsted acids either in situ or by mixing the

acids prior to their addition to the reactor. It is proposed that the metal triflate and phosphoric

acid form a complex of the type shown in Scheme 3.8. This will be analogous with what is

claimed in the literature266 where an improved nitration reaction is held to also proceed via a

complex of the type shown in Scheme 3.8, but where the acid is nitric acid. In the present

instance, the protons of the coordinated H3PO4 should be more highly acidic than the parent

H3PO4 by virtue of the electron withdrawing effect of the binding thereof to the metal centre and

these protons should thus be available to catalyse certain types of reactions.

Page 136: Applications of metal triflates and assisted acids as

116

H3PO4 La(OTf)3 (OTf)3LaO

OP

OH

OH

H

3.26

Scheme 3.8: Formation of a complex via combination of H3PO4 and La(OTf)3

The following set of reactions was performed using a variety of Lewis acid and Brønsted acid

combinations. The La(OTf)3/H3PO4 acid combination was the first assisted acid to be evaluated

in the alkylation reaction at 70 ˚C. An equal molar amount of each acid (4 mmol) of La(OTf)3 and

H3PO4 (99%), were used, hence the total mass of catalyst loaded was 0.015 wt% which is much

less than the instance where individual acids were screened. The stirring speed was kept at

1200 rpm and high excess of isobutylene (1.1 bar) was utilised. The results obtained from this

experiment are showed in Table 3.9.

Page 137: Applications of metal triflates and assisted acids as

117

Table 3.9: Butylation of m/p-cresol in the presence of [La(OTf)3/H3PO4]

Time (minutes)

Conversion (mol%)

Butylated cresols Yield

(mol%)

Selectivity

(%)

Oligomers (wt%)

MC PC MBPC MBMC BHT DBMC BHT MBPC

0 0 0 0 0 0 0 0 0 0

5 15.2 9.6 9.3 11.9 0.3 3.3 3.1 21.7 0.7

10 32.0 12.6 11.3 25.1 1.3 6.9 10.3 21.6 1.5

15 35.7 22.8 19.9 20.9 2.9 14.8 12.7 41.5 3.7

20 38.7 25.9 21.3 19.2 4.6 19.5 17.8 50.4 4.7

30 41.8 29.8 21.9 17.6 7.9 24.2 26.5 57.9 4.4

40 44.5 32.3 21.8 16.7 10.5 27.8 32.5 62.5 4.2

50 47.8 34.2 20.4 16.3 13.8 31.5 40.4 65.9 3.4

60 47.9 38.6 20.0 16.1 18.6 31.8 48.2 66.4 5.5

120 50.2 37.4 15.9 14.6 21.5 35.6 57.5 70.9 6.7

It was evident during the screening of various acids that individual lanthanide triflates and

H3PO4 exhibited no activity for the butylation of cresols. However, when the acids were used as

a combination, a significant enhancement of activity was achieved. This is an indication of a

synergistic effect which prevails upon combining the two acids. However, the assisted acid gave

lower activity towards alkylation reactions than Zr(OTf)4 and Al(OTf)3. Although the assisted

catalyst loading was low in this particular experiment, the reaction rate was high compared to

the instance where the individual Brønsted acids were used. The conversions of m- and p-

cresols were 50.7 and 36.7 mol% respectively after 120 minutes. The assisted acid showed

faster alkylation of both m-cresol and MBMC, resulting in a high selectivity to DBMC (70.9%),

Page 138: Applications of metal triflates and assisted acids as

118

while the selectivity to BHT was only 57.7% after 120 minutes. The assisted acid gave 6.7 wt%

of oligomers over a period of 120 minutes, and this amount of oligomers is much higher than

those obtained when using Zr(OTf)4 and Al(OTf)3. It was remarkable that the assisted acid was

able to promote the reaction effectively as compared to individual acids.

The results achieved with the La(OTf)3/H3PO4 assisted acid prompted further evaluation of other

triflate salts combined with H3PO4. The second assisted acid evaluated was Gd(OTf)3/H3PO4.

The experiment was also performed using the same set of conditions as for La(OTf)3/H3PO4.

Again, this acid system exhibited significant enhancement of activity for alkylation reactions,

compared to the inactive individual acids. A similar amount of m/p-cresol conversion was

accomplished in both reactions, Table 3.10.

Table 3.10: Butylation of m/p-cresol in the presence of [Gd(OTf)3/H3PO4]

Time (minutes)

Conversion (mol%)

Butylated cresols yield

(mol%)

Selectivity

(%)

Oligmers (wt%)

MC PC MBMC MBPC DBMC BHT BHT DBMC

0 0.0 0.0 0.0 0.0 0.0 0.0 - - 0.0

5 16.6 11.9 13.5 11.5 3.1 0.4 3.3 18.7 0.1

10 34.2 20.6 25.8 18.8 8.4 1.8 8.6 24.6 0.6

15 34.2 26.3 18.2 22.7 16.0 3.6 13.7 46.8 1.8

20 38.7 30.1 17.2 23.4 21.5 6.7 22.2 55.6 2.7

30 43.6 32.4 16.5 23.0 27.1 9.4 29.1 62.2 3.0

40 46.2 34.0 16.1 21.6 30.1 12.4 36.5 65.2 3.3

50 47.5 35.1 15.2 20.8 32.3 14.3 40.8 68.0

60 48.7 35.9 14.9 19.7 33.8 16.2 45.1 69.4 3.3

120 51.9 38.1 13.0 14.8 38.9 23.3 61.2 75.0 3.6

Page 139: Applications of metal triflates and assisted acids as

119

Gd(OTf)3/H3PO4 also showed higher activity for alkylation of m/p-cresol compared to individual

acids. This reaction has exhibited high selectivity to DBMC (70.9%) and 63.1% to BHT.

Therefore, the intermediate products (MBMC and MBPC) did not convert quantitatively to

DBMC and BHT. The amount of oligomers (3.6 wt%) formed were comparable to that obtained

when using Zr(OTf)4 and Al(OTf)3 and less than that of the assisted acid formed from

La(OTf)3/H3PO4. It has already been shown that Al(OTf)3 accomplished excellent activity for

alkylation of the cresol mixture. Furthermore, the results showed that a catalyst loading of >0.05

wt% afforded high selectivity towards the formation of BHT. It was then decided to investigate

whether a combination of Al(OTf)3 with H3PO4 will result in further enhancement of reaction rate.

The results of a reaction catalysed by Al(OTf)3/H3PO4 are shown in Table 3.11.

Table 3.11: Butylation of m/p-cresol catalysed by Al(OTf)3/H3PO4

Time (minutes)

Conversion

(mol%)

Butylated cresols Yield

(mol%)

Selectivity

(%)

Oligomers (wt%)

MC PC DBMC BHT MBMC MBPC BHT DBMC

0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

5 15.3 10.4 2.9 0.3 12.4 10.1 2.9 19.0 0.2

10 31.9 18.4 6.5 1.2 25.4 17.2 6.5 20.4 0.6

15 33.7 24.9 15.6 3.3 18.1 21.6 13.3 46.3 2.1

20 39.4 28.5 20.2 5.5 19.2 23 19.3 51.3 2.6

30 43.6 32.1 25.8 9.5 17.8 22.6 29.6 59.2 3

40 46.3 34 29.5 12.5 16.8 21.5 36.8 63.7 3.2

50 48.7 35.6 32.6 15.9 16.1 19.7 44.7 66.9 3

60 49.7 36.3 33.9 18 15.8 18.3 49.6 68.2 3.5

120 52.9 38.8 39.5 25.8 13.4 13 66.5 74.7 3.8

Page 140: Applications of metal triflates and assisted acids as

120

The results obtained when using Al(OTf)3/H3PO4 as a catalyst are comparable to those obtained

when using Al(OTf)3 alone. The assisted acid also gave high selectivity to DBMC (74.7%) and

only 66.5% to BHT. The selectivity trend is similar to that of a reaction catalysed by Al(OTf)3

alone, where selectivities of 79.8% and 75.1%, respectively, were noted. The quantity of the

products afforded by this assisted acid are comparable to those afforded by the assisted acids

of La(OTf)3 and Gd(OTf)3.

3.6 Evaluation of La(OTf)3/SPA assisted acid

The excellent activity afforded by the La(OTf)3/H3PO4 assisted acid also led to the evaluation of

La(OTf)3/SPA assisted acid. SPA was first evaluated on its own for the butylation reaction, and

was proven to be inactive. Thereafter, La(OTf)3/SPA was prepared and its activity evaluated

during m/p-cresol butylation reactions. The reaction was performed in a 300 mL batch autoclave

at 70 ˚C and at 1.1 bar isobutylene. The total amount of the catalyst (including silica support)

used was 0.57g (0.95 wt). This reaction was performed under the same set of conditions as the

one where SPA was used, but stirred at 700 rpm. The low stirring speed was employed to avoid

catalyst attrition, which resulted in the blockage of the sampling line. The results obtained are

shown in Table 3.12.

Page 141: Applications of metal triflates and assisted acids as

121

Table 3.12: Butylation of m/p-cresol in the presence of La(OTf)3/SPA.

Time (minutes)

Conversion (mol%)

Butylated cresol yields

(mol%)

Selectivity (%)

Oligomers (wt%)

MC PC MBMC MBPC DBMC BHT BHT DBMC

0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

5 24.5 18.5 18.4 14.9 6.1 3.6 19.5 24.9 1.0

10 40.2 25.4 29.9 21.4 10.3 4.0 15.7 25.6 1.3

15 41.0 31.0 21.2 24.4 19.8 6.6 21.3 48.3 1.4

20 46.4 34.9 19.2 24.1 27.2 10.8 30.9 58.6 1.8

30 51.1 37.8 16.6 19.9 34.5 17.9 47.4 67.5 2.4

40 53.1 39.1 15.0 16.5 38.1 22.6 57.8 71.8 2.7

50 54.1 39.7 13.9 12.7 40.2 27.0 68.0 74.3 3.1

60 55.0 40.3 12.8 10.8 42.2 29.5 73.2 76.7 3.2

120 56.7 41.2 9.0 4.4 47.7 36.8 89.3 84.1 4.4

The La(OTf)3/SPA catalysed reaction gave a quantitative conversion of m- and p-cresol, i.e.

56.7 mol% and 41.2 mol%, to product. This reaction showed extremely high reaction rates. The

selectivity to BHT (89.3%) is comparable to that of DBMC (84.1%) after 120 minutes. In the

reaction catalysed by La(OTf)3/SPA, the mono-butylated products (MBPC and MBMC) were

almost all converted to their corresponding di-alkylated products (BHT and DBMC). It is also

crucial to note that each individual acid was completely inactive when employed as catalyst in

this type of reaction, but when combined, the activity was significantly enhanced, as is apparent

from the results. The oligomerisation reaction afforded 4.6 wt% hydrocarbon products, resulting

Page 142: Applications of metal triflates and assisted acids as

122

from isobutylene oligomerisation reactions. A comparison of the products afforded by

La(OTf)3/SPA and La(OTf)3/H3PO4 combination catalysts is shown in Table 3.13.

Table 3.13: Comparison of cresols conversions and product selectivities afforded by La(OTf)3/H3PO4 and La(OTf)3/SPA after 120 minutes.

Catalyst Conversion

(mol%)

Alkylated cresol yield

(mol%)

Oligomers

(wt%)

Selectivity

(%)

MC PC MBMC MBPC DBMC BHT Oligomers BHT DBMC

La(OTf)3/SPA 57.0 41.3 9.0 4.4 47.7 36.8 4.4 89.3 84.2

La(OTf)3/H3PO4 50.2 37.4 15.9 14.6 21.6 35.6 6.7 57.5 70.9

High conversions were achieved in reactions catalysed by La(OTf)3/SPA as compared to those

catalysed by La(OTf)3/H3PO4. High selectivites to the di-butylated cresols were also achieved

with La(OTf)3/SPA. La(OTf)3/SPA also gave low amount of oligomers. The high activity and the

heterogeneous nature of La(OTf)3/SPA prompted evaluation of a possibility to recycle the

catalyst. The initial run of the catalyst recycling studies was performed at 70 ˚C, with 1.1 bar

isobutylene and m/p-cresol (61.9 g; 0.57mol), starting off with 0.50 g (0.09 wt%) of catalyst. The

catalyst recycling was performed in a 300 mL stainless steel autoclave reactor. At the end of

each experiment, the catalyst was recovered by decanting all the previous reaction contents

and the solid catalyst remained in the reactor. For the subsequent experiments, fresh m/p-cresol

was charged into the reactor containing the catalyst residue and the reactor was assembled and

heated to the operating temperature. Upon reaching the temperature, isobutylene was allowed

into the reactor, with samples taken over time. The results of the reactions are shown in Table

3.14.

Page 143: Applications of metal triflates and assisted acids as

123

Table 3.14: The results obtained during recycling of La(OTf)3/SPA.

Experiment # Time (minutes)

Conversion

(mol%)

Selectivity

(mol%)

Oligomers (wt%)

PC MC BHT DBMC

Original 15 29.2 40.3 11.9 12.4

60 41.8 57.7 85.9 73.6

120 41.6 57.5 80.2 65.1 4.8

Recycle 1 15 30.3 41.8 21.2 48.3

60 40.1 55.4 73.2 76.8

120 41.3 57.2 89.3 84.2 4.3

Recycle 2 15 22.2 30.7 14.2 44.8

60 34.6 47.8 42.7 68.1

120 38.0 52.5 64.6 76.3 4.8

Recycle 3 15 17.4 24.1 14.6 37.5

60 30.3 41.8 28.0 61.4

120 36.2 51.1 51.1 72.2 4.0

The results showed consistent catalyst activity for the original run and the first recycle run, but a

considerable decrease of activity to the following runs. It is known in the literature that H3PO4

leaches out of the support (SiO2), which leads to catalyst deactivation.267 The current study has

thus showed that La(OTf)3/SPA could be recycled at least four times (including the original run)

without being completely deactivated. The deactivation noted presumably arises due to leaching

Page 144: Applications of metal triflates and assisted acids as

124

away of the H3PO4 from the SiO2 support into the reaction medium, in line with literature

observations. A summary of the overall reaction rates afforded by various assisted catalyst

systems is depicted in Figure 3.2. It is thus evident that La(OTf)3/SPA gave the highest reaction

rate for the conversion of cresols (MC+PC) to mono-butylated products (MBMC+MBPC), which

were effectively converted to di-butylated products (DBMC+BHT). The other catalyst systems

showed comparable reaction rates.

0

10

20

30

40

50

60

70

80

90

100

110

0 20 40 60 80 100 120 140

Mo

l%

Time (Min.)

MC+PC Conv.: La(OTf)3/SPA

MC+PC Conv.: La(OTf)3/H3PO4

MC+PC Conv.:Gd(OTf)3/H3PO4

MC+PC Conv.: Al(OTf)3/H3PO4

MBMC+MBPC: La(OTf)3/SPA

MBMC+MBPC: La(OTf)3/H3PO4

MBMC+MBPC: Gd(OTf)3/H3PO4

MBMC+MBPC: Al(OTf)3/H3PO4

DBMC+BHT: La(OTf)3/SPA

DBMC+BHT: La(OTf)3/H3PO4

DBMC+BHT: Gd(OTf)3/H3PO

DBMC+BHT: Al(OTf)3/H3PO4

Figure 3.2: Summary of reaction rates for various assisted catalysts

3.7 Friedel-Crafts alkylation of anisole

Alkylated anisole is an important industrial compound which is used as an antioxidant, in dye

developers and in stabilisers for fats, oils and plastic rubbers.268 The preparation of 4-tert-

butylanisole via alkylation of anisole with tert-butyl alcohol in the presence of ZrCl4 and

trifluoroacetic acid was reported by Sartori and co-workers.269 Alkylation of anisole with tert-

butylacetate as an alkylating agent in the presence of H2SO4, and also with tert-butyl nitrate in

the presence of SnCl4270 as a catalyst, was also reported by Fernholz et al.271

Page 145: Applications of metal triflates and assisted acids as

125

It was therefore decided to evaluate the activity of assisted acids for alkylation of anisole with

isobutylene. The starting point in these reactions involves screening of individual acids. The

results obtained from individual acids will eventually be compared to those afforded by the

assisted acid systems. There are three major products that were obtained from the reaction of

anisole and isobutylene, i.e. two mono-alkylated anisoles (2-tert-butylanisole and 4-tert-

butylanisole), while a third product is obtained when the mono-alkylated products undergo

further alkylation to form 2,4-di-tert-butylanisole, Scheme 3.9.

O

2,4-di-t-butyl anisole4-t-butyl anisole2-t-butyl anisoleAnisole

H+

O OO

3.27 3.28 3.29 3.303.2

Scheme 3.9: Alkylation of anisole with isobutylene

MTBE (methyl tert-butyl ether) or tert-butanol can be used to generate isobutylene in situ

(Scheme 3.10), with MTBE being the most widely used and recommended reagent. The

drawback associated with tert-butyl alcohol as an alkylating agent is that the overall reaction

forms water, which may affect the activity of the catalyst and may even lead to complete catalyst

deactivation.

O

H+CH3OH

MTBE

OHH2O

H+

t-BuOH3.31

3.32

3.2

3.2

Scheme 3.10: Generation of isobutylene from tert-BuOH and MTBE

Page 146: Applications of metal triflates and assisted acids as

126

Cracking of MTBE is a viable source of isobutylene, resulting in the formation methanol as by-

product. Furthermore, Yadav and co-workers have shown that the rate of alkylation with MTBE

is much faster than with t-BuOH.272

3.7.1 Evaluation of individual Lewis and Brønsted acids

The reaction of anisole with isobutylene is expected to occur with ease because the methoxy

functional group on the benzene ring donates electron density to the ring, thereby making it

more highly nucleophilic. Because of the activating group, anisole undergoes alkylation at its

ortho- and para-positions. The initial set of reactions involved evaluation of metal triflate Lewis

acids and Brønsted acids as catalysts for the alkylation of anisole with isobutylene. The Lewis

acids evaluated include Al(OTf)3, Zr(OTf)4 and two lanthanide triflate salts [La(OTf)3, and

Gd(OTf)3]. The Brønsted acids included H3PO4 and triflic acid (HOTf). The first reaction was

performed at 70 ˚C, using anisole (560 mmol) and high excess of isobutylene (1.1 bar constant

feed) in the presence of Zr(OTf)4. This catalyst showed good activity for the reaction, as may be

seen in Figure 3.3.

Figure 3.3: Alkylation of anisole (560 mmol) with 1.5 bar isobutylene catalysed by 1.46 mmol Zr(OTf)4 at 70 ˚C, and stirred at 1000 rpm.

The overall conversion of anisole under this set of conditions was 55 mol%. Three major

alkylated products were identified, i.e. 4-tert-butylanisole, 2-tert-butylanisole and 2,4-di-tert-butyl

Page 147: Applications of metal triflates and assisted acids as

127

anisole. In the course of the reaction, 3 wt% of oligomers of isobutylene were obtained. The

results showed that 4-tert-butyl anisole and 2-tert-butylanisole were formed faster during the

initial stages of the reaction, with 4-tert-butyl isomers being formed at the highest rate. The

mono-alkylated anisole products subsequently undergo further butylation to form a di-butylated

product (2,4-di-tert-butylanisole). Al(OTf)3 together with lanthanide triflate Lewis acids [La(OTf)3

and Gd(OTf)3] were also evaluated under the same set of conditions and, interestingly, all were

shown to be inactive. Although, it was probable that lanthanide triflate salts may be inactive,

Al(OTf)3 was expected to show some activity as was the case during butylation of cresol. The

Lewis acid was unreactive for the present reaction, despite numerous attempts. The reasons for

this reactivity remain unknown.

The alkylation reactions described here are reversible under some conditions, and a large

excess of isobutylene should be used to avoid de-butylation. The reaction performed at 70 °C

was relatively slow, and hence the reaction temperature was increased to 100 °C in an attempt

to speed up the reaction, Figure 3.4.

Figure 3.4: Alkylation of anisole (560 mmol) with 1.5 bar isobutylene catalysed by Zr(OTf)4 (1.46 mmol ) at 100 ˚C, and stirred at 1000 rpm.

Page 148: Applications of metal triflates and assisted acids as

128

Increasing the reaction temperature resulted in an increase of the reaction rate, leading to a

high conversion of anisole (85 mol%) in the allocated period of two hours. The results of this

experiment also showed that mono-butylated products are formed at a higher rate and that

these products are subsequently converted into a di-alkylated product. The oligomerisation of

isobutene (5 wt%) was also accelerated relative to the reaction performed at lower

temperatures. However, the major products were those of the alkylated anisole.

Brønsted acids were also evaluated under the same set of conditions as the Lewis acids. These

included H3PO4 and triflic acid (HOTf). The strength or the pKa values of these acids vary

significantly, Table 3.15, as has already been indicated.

Table 3.15: Comparison of the acid strengths of H3PO4 and HOTf

Acid Acid strength (pKa) 273 Hammett (H0)

HOTf -14.1 -15.0

H3PO4 (85%) 2.0 -4.3274

It is well known that the activity of a Brønsted acid is dependent on its dissociation constant Ka,

which is normally expressed as a negative logarithm (pKa) i.e.:

pKa = -log Ka

Thus, a high value of pKa represents a very small value of Ka, and hence a very weak acid.275

Furthermore, if the Hammett number of a particular acid is large and negative, it implies that the

acid is stronger. Therefore, HOTf is expected to exhibit higher activity between the two Brønsted

acids evaluated, and undeniably, the better results were obtained from the reaction catalysed by

triflic acid (Figure 3.5).

Page 149: Applications of metal triflates and assisted acids as

129

Figure 3.5: Alkylation of anisole (560 mmol) with 1.5 bar isobutylene catalysed by HOTf (0.2 mol%) at 100 ˚C, and stirred at 1000 rpm.

Triflic acid afforded quantitative conversion of anisole (>99%) to the desired products. The

reaction proceeded at an extremely high rate, and almost all of the anisole was converted into

2- and 4-tert-butyl anisole within 30 minutes. The mono-butylated products were essentially

completely converted into 2,4-di-tert-butylanisole. The rate of conversion of mono-butylated

products was high, such that the reaction was complete within 60 minutes. However, the

reaction was allowed to proceed until 120 minutes (with a post run of 60 minutes).

The second Brønsted acid evaluated was H3PO4. Initially, the reaction was performed using 0.2

mol% catalyst at 100 °C, and the rate was extremely low. Thereafter the catalyst concentration

was increased to 2.0 mol%. Interestingly, this acid showed some activity, although the rate was

low relative to that obtained with triflic acid, Figure 3.6. Despite H3PO4 exhibiting a slow

reaction, high conversion of anisole (90 mol%) was achieved by prolonging the reaction time.

Page 150: Applications of metal triflates and assisted acids as

130

Figure 3.6: The activity of 2.0 mol% H3PO4 during alkylation of anisole.

3.7.2 The evaluation of assisted acids as catalysts for the butylation of anisole

The previous studies showed that assisted acids (prepared by combining the lanthanide triflate

salts with mineral Brønsted acids) exhibit superior catalytic activity compared to individual Lewis

and Brønsted acids. Hence the ultimate objective was to eventually evaluate the catalytic

activity of assisted acids for the butylation of anisole. The first reaction was performed using the

La(OTf)3/H3PO4 assisted acid, Figure 3.7.

Page 151: Applications of metal triflates and assisted acids as

131

Figure 3.7: Alkylation of anisole (560 mmol) with 1.5 bar isobutylene catalysed by La(OTf)3/H3PO4 (0.8 mol%) at 100 ˚C, and stirred at 1000 rpm.

The La(OTf)3/H3PO4 acid exhibited excellent activity and quantitative conversion of anisole was

obtained within two hours. The catalyst was selective to three products, i.e. two mono-alkylated

products and a di-alkylated product, as was the case for the other catalysts. The two mono-

alkylated products were almost entirely converted into 2,4-di-tert-butylanisole by the end of the

reaction. Isobutylene oligomers (3.5 wt%) were also obtained when using this catalyst. The acid

combination has been proven once again to exhibit excellent activity, with total conversion of

intermediate products to 2,4-di-tert-butylanisole being noted. This is compared to inactive

La(OTf)3 and the low activity noted for H3PO4, when used individually as catalysts for this

process. The evaluation of other assisted acids was consequently performed, Figure 3.8.

Page 152: Applications of metal triflates and assisted acids as

132

Figure 3.8: Alkylation of anisole (560 mmol) with 1.5 bar isobutylene catalysed by Al(OTf)3/H3PO4 (0.8 mol%; 0.8 mol%) at 100 ˚C, and stirred at 1000 RPM.

Interestingly, Al(OTf)3 showed absolutely no activity as catalyst for the butylation of anisole on

its own, but when it is combined with H3PO4 it exhibited excellent activity, where almost all the

anisole was converted. Furthermore, both mono-butylated anisoles were almost converted to di-

tert-butylanisole, with 2-t-butylanisole consumed at a faster rate. It is, however, notable that the

rate of the reaction catalysed by La(OTf)3/H3PO4 is much higher than that of the reaction

catalysed by Al(OTf)3/H3PO4. Subsequently, another reaction catalysed by Gd(OTf)3/H3PO4 was

carried out, Figure 3.9.

Page 153: Applications of metal triflates and assisted acids as

133

Figure 3.9: Alkylation of anisole (560.01mmol) with 1.5 bar isobutylene catalysed by Gd(OTf)3/H3PO4 (0.8 wt%; 0.2 mol%) at 100 ˚C, and stirred at 1000 rpm.

This catalyst showed a much higher reaction rate than all other combined acid catalyst systems,

the starting material and intermediates were completely converted to the final product. The rate

of this experiment is comparable to that afforded by triflic acid.

3.8 Alkylation of diphenylether (DPE)

The alkylation of diphenyl ether (DPE) with olefins, alcohols and alkyl halides yields

commercially important products. There are numerous applications of alkylated DPE, including

its use as heat transfer fluids.276 Alkylated DPE can be used as a dielectric agent in

transformers as claimed on a patent of Westinghouse Electric Corporation.277 This patent

discloses the use of up to 99 wt% of mono-ethylated, mono-propylated, mono-butylated DPE

and up to 20 wt% dialkylated DPE as dielectric fluid in a capacitor.

Mobil Oil Corporation filed a patent278 on the use of alkylated DPE as lubricant fluid with

excellent low temperature viscometrics. The later patent pertains to the alkylation of DPE with

olefins such as 1-tetradecene, 1-dodecene and 1-octene in the presence of zeolite at 200 ˚C,

and the catalysts exhibited high selectivity to mono alkylated products. Coleman et al.279

Page 154: Applications of metal triflates and assisted acids as

134

disclosed a process where DPE was alkylated with alkyl halides, olefins and alcohols in the

presence of alkylation catalysts such as AlCl3, AlBr3, FeCl3, and heterogeneous catalysts. The

best results were obtained when using AlCl3. This patent claims that alkylation with alcohols is

favourable when using clay catalysts. The product mixture does not need to be separated, since

the mixture contains properties that are readily suited to be used as, for example, dielectric

agents in transformers, condensers and other electrical equipment or as plasticisers for

synthetic resins. The mixture has relatively low viscosity and freezing point compared to its

individual compounds; therefore mixtures are more valuable than individual products. The

mono-alkylated, di-alkylated and tri-alkylated products were reported in this invention.

In example 4 of this patent, alkylation of DPE with tertiary butyl chloride in the presence of AlCl3

using the reagents in a 1:1 mole ratio at 100 ˚C yielded three products: a liquid mono-tertiary-

butyl-DPE (26%) boiling at 150.5-152.5 ˚C (6 mmHg); a crystalline mono-tertiary-butyl-DPE

isomer (41%) boiling at 153.5-155.5 ˚C (6 mmHg), with a melting point of 54-54.5 ˚C and a liquid

di-tertiary-butyl-DPE (31%) boiling at 191.3 ˚C (6 mmHg). In example 7, diisobutylene was used

as an alkylating agent at 60 ˚C. The reaction produced four products: a crystalline mono-tertiary-

DPE, boiling point 154.5-159.3 ˚C (6 mmHg) and melting at 53.3-54 ˚C; mono-octyl-DPE boiling

at 184.5-185.5 ˚C (6 mmHg); a compound with empirical formula C24H34O, boiling at 218-223 ˚C

(6 mmHg) and a compound with empirical formula C38H42O, boiling at 229.5-239.5 ˚C (6 mmHg).

Phillips Petroleum Company filed a patent280 on the alkylation of DPE with isobutylene in the

presence of sulphuric acid as catalyst. Example 2 of this patent discloses alkylation of DPE with

isobutylene in the presence of AlCl3 at 60 ˚C and mole ratio of 4.2 of olefin to DPE. The product

distribution showed presence of 2% mono-alkylated, 41% di-alkylated and 20% tri-alkylated

products. The product mixture was a pale yellow oil with low viscosity. In another example,

where isobutylene was used (example 6) at 27 ˚C, with DPE to olefin mole ratio of 4.4, less than

1% mono-alkylated; about 19% di-alkylated and about 63% tri-alkylated products were obtained.

It has also been claimed that when the mole ratio of olefin to DPE is increased to 7.15, less than

1% mono-alkylated, 25%-di-alkylated and 57% tri-alkylated products were obtained.

The alkylation of DPE with α-dodecene to produce mono-alkylated DPE, which is a precursor for

synthesis of di-alkyl-DPE sulfonates (an important class of surfactant produced by Dow

Chemical Company under the trade name Dowfax®)281 catalysed by an ionic liquid (IL) has been

reported by Kou et al.282 The catalyst could not be recycled due to leaching of the active species

(Al2Cl7)-.

Page 155: Applications of metal triflates and assisted acids as

135

The alkylation of diphenyl ether with benzyl chloride was studied by Yadav et al.283 In this study,

sulfated zirconia was used as catalyst. The reaction was performed at 90 ˚C. This reaction

produced an isomeric mixture of ortho- and para-benzyl-diphenyl ether having a molecular

weight of 260 g/mol. The HCl generated in situ (by-product) does not take part as catalyst in the

reaction due to its low solubility in the reaction mixture. The other example, where the alkylation

of DPE with 1-decene in the presence of sulfated zirconia was carried out by Yadav et al. 284

(Scheme 3.11). Conversions of about 85% were obtained from the reaction, with high selectivity

to ortho- and para-mono-alkylated products.

When n= 0, m= 7 n= 1, m= 6 n= 2, m= 5 n= 3, m= 4

OH3C(CH2)7CH CH2

O

H3Cn(H2C) (CH2)mCH3

O

(CH2)nCH3

(CH2)mCH3

Sulfated zirconia

150 oC; 2hrs

3.33

3.34

3.35

Scheme 3.11: Alkylation of DPE with 1-decene

In the present study, t-butylation was pursued in order to evaluate the catalytic activity of

assisted acids for butylation of diphenyl ether with isobutylene. The alkylation of DPE was

performed using selected metal triflate salts, H3PO4 and assisted acid systems. The Lewis acids

evaluated included Zr(OTf)4, Al(OTf)3, La(OTf)3, and Gd(OTf)3, with only H3PO4 being evaluated

as a Brønsted acid. The alkylation of DPE with isobutene resulted in the formation of an oily

viscous mixture of products. The viscosity of the reaction mixture depends on the selectivity of

the catalyst used as well as on the extent of alkylation of the substrate. For example, if the

catalyst forms only mono-butylated compounds, the viscosity of the product mixture will be

lower than when the catalyst yields both mono- and di-butylated compounds. Furthermore, the

viscosity of the former product mixture will also be lower than when tri-butylated species are

formed. There were three major products identified from the study, i.e. two mono-butylated

diphenylether products and a di-alkylated-tert-butyl-phenoxy ether, a tri-alkylated phenoxy ether

Page 156: Applications of metal triflates and assisted acids as

136

and tetra-alkylated phenoxyether (which has to be confirmed with NMR spectroscopy). Up to

this stage we have not come across literature where the tetra-alkylated phenoxyether was

reported. The proposed structures of butylated DPE compounds are shown in Table 3.16.

Page 157: Applications of metal triflates and assisted acids as

137

Table 3.16: Properties of the products obtained during alkylation of diphenylether

O 12

3

45

6

7

8

910

11

1213

14

1516

3.36

O123

5

6

7

89

10

11

1213

14

1516

4

3.37

O 1 2

34

5

6

7

8 9 10

11

12

13

1415

16

17

18

19

20

3.38

4-tert-butyl-phenylpheyl ether 2-tert-butyl-phenylphenylether 2,4-di-tert-butyl-phenylphenyl ether

1H NMR (400 MHz, CDCl3): δ=1.32 (s,

9H, H-6 to H-8); δ=6.92-7.34 (m, 9H,

ph).

1H NMR (400 MHz, CDCl3): δ=1.42 (s, 9H,

H-4, H-5 and H-6); δ=6.81-7.41 (m, 9H,

ph).

1H NMR (400 MHz, CDCl3): δ=1.23 (s, 9H, H-

3, H-5, H-6); δ=1.35 (s, 9H, H-10, H-11, H-12);

δ=6.60-7.41 (m, 8H, ph)

13C NMR (400 MHz, CDCl3): δ= 31.8

(C-6 to C-8); δ=34.5 (C-5); δ= 118.5

(C-2); δ= 118.7 (C-12 and C-16);

δ=123.1 (C-14); δ=126.8 (C-9 and C-

3); δ=129.8 (C-13 and C-15); δ=146.4

(C-4); δ=155.1 (C-1); δ=157.8 (C-11).

13C NMR (400 MHz, CDCl3): δ=30.4 (C-4

to C-6); δ=34.9 (C-2); δ= 118.9 (C-16 and

C-12); δ=120.3 (C-10); δ=122.8 (C-14):

δ=123.3 (C-8); δ=127.3 (C-7); δ=129.7 (C-

13 and C-15); δ=141.2 (C-2); δ=156.1 (C-

1); δ=158.0 (C-11)

13C NMR (400 MHz, CDCl3): 13C NMR (400

MHz, CDCl3): δ=30.3 (C-3, C-5, C-6); δ=31.7

(C-10, C-11, C-12); δ=34.6 (C-5); δ=35.0 (C-

9); δ=118.6 (C-16 and C-20); δ=119.7 (C-14);

δ=122.4 (C-13); δ=123.8 (C-7); δ=124.1 (C-

18); δ=129.6 (C-17, C-19); δ=140.0 (C-2);

δ=145.8 (C-8); δ=153.4 (C-1); δ=158.1 (C-15)

m/z (GC-MS): 226 (M+, 28%); 211 (M-

CH3, 100%); 195 (M-C2H6, 2%); 183

(M-C3H9, 6%); 91 (M-PhC3H9, 9%); 77

(M- OPhC3H9, 9%).

m/z (GC-MS): 226 (M+,59%); 211 (M- CH3,

100%); 195 (M- C2H6, 14%); 181 (M- C3H9,

13%); 133(M-C7H11, 15%); 91 (M-C10H12,

26%); 77 (M-OC10H12, 13%); 57 (M-

OC12H13, 9%)

m/z GC-MS: m/z (GC-MS): 282 (M+, 27%);

267 (M-CH3, 100%); 254 (M- C2H6, 2%); 211

(M- (C4H9), 2%); 91(M-OC14H22, 5%); 77 (M-

OC14H22, 4%); 57 (M-OC16H23, 14%)

Bp (760 tor): 307+/-11°C, a : 302-304

°C b

Bp (760 torr): 296+/-19 °C, a : 250-253°C b

Bp (°C@760 torr): 334+/-21°C, a : 339-342°C b

Mp: 54-55 °Ca: 53-55 °Cb

a Literature value, b Experimental value

Page 158: Applications of metal triflates and assisted acids as

138

The mono-alkylated product, (4-tert-butyl-phenylphenylether, 3.36), is a colourless crystalline

product with a boiling point of 302-304 ˚C, and 2-tert-butyl-phenylphenylether (3.37) is a

colourless free flowing liquid at room temperature, with a boiling point of 276-279 ˚C at

atmospheric pressure, while the di-alkylated product2,4-di-tert-butyl-phenylphenul ether, 3.38] is

also a colourless liquid at room temperature boiling at 339-342 ˚C at atmospheric pressure. A

highly vicious material left in the re-boiler was not purified any further due to instrument

limitations.

The initial reaction was performed in the presence of 0.14 mol% Zr(OTf)4 at 70 ˚C. The reaction

was stirred at 1200 rpm and at isobutylene pressure of 1.1 bar. It was expected that the reaction

would proceed rapidly, as in the case of the cresols and anisole, but surprisingly, it was slow.

Hence the experiment was allowed to proceed for 24 hrs. Over this period of time only 25%

conversion of DPE was achieved and the products were 16.8% and 3.3% of 3.36 and 3.37,

respectively. Only 2.9% of 3.38 and 1.9% and heavy products were obtained. Having observed

such low rates, it was decided to vary some reaction parameters with the aim of increasing the

rate of the reaction, Table 3.17.

Table 3.17: The variation of reaction parameters during butylation of DPE (0.36mol) in the presence of Zr(OTf)4.

Temp. (˚C)

[Cat.] mol%

Time (hours)

DPE conversion

(%)

Products distribution (%) Oligomers (wt%)

3.36 3.37 3.38 unknown

50 0.14 23 12.3 9.6 1.8 0.3 0.6 8.3

70 0.14 24 24.9 16.8 3.3 2.9 1.9 6.2

100 0.28 3.5 35.1 23.6 5.4 2.3 4.1 10.6

100 0.57 2 40.0 27 6.3 3.9 2.5 8.8

Page 159: Applications of metal triflates and assisted acids as

139

The results showed that an increase of the reaction temperature (50 ˚C to 100 ˚C) resulted in a

proportional increase of the reaction rate, as expected. Furthermore, increasing the catalyst

concentration at constant temperature (100 ˚C) yields an increase of the reaction rate. An

exotherm of approximately 6 ˚C was observed for the reaction performed at 100 ˚C. The overall

rate of DPE butylation was relatively slow, but the formation of mono-alkylated products,

especially (3.36) was fast in all instances. Hence, the evaluation of other metal triflates salts and

H3PO4 was performed at 100 ˚C, using 0.57 mol% catalysts concentration, Table 3.18.

Table 3.18: The results obtained during evaluation of other catalysts for butylation of DPE (0.57 mol%).

Catalyst

Time

(hours)

DPE conversion

(%)

Products distribution

(%)

(3.36) (3.37)

La(OTf)3 17 2.9 2.5 0.4

Al(OTf)3 7 0.6 0.5 0.1

Gd(OTf)3 6 1.4 1.2 0.2

H3PO4 7 0.7 0.6 0.1

The catalytic activities of the individual Lewis and Brønsted acid catalysts were extremely low

(Table 3.18) under the same conditions. Only two mono-alkylated products (3.36 and 3.37) were

obtained when using individual acids. Interestingly, Al(OTf)3 exhibited low activity for this

reaction, as was previously noted. However, the acid combinations of the metal triflates and

H3PO4 at a catalyst concentration of 0.57 mol% (Table 3.19) showed an altogether different

picture. A significant enhancement of reaction rate was observed when using the assisted acid

systems. The results obtained with these acids are comparable to those afforded when using

Zr(OTf)4. This is another example among several others in the present study showing activation

of one acid by the other. It is also noticeable that the amount of oligomers was lower compared

to the reaction catalysed by Zr(OTf)4.

Page 160: Applications of metal triflates and assisted acids as

140

Table 3.19: Results obtained during the evaluation of assisted acids for the butylation of DPE.

Catalyst Time

(hours)

DPE conversion (%)

Products distribution

(%)

Oligomers (wt%)

3.36 3.37 3.38 unknown

La(OTf)3/H3PO4 7 47.9 34.4 5.8 1.9 2.5 7.1

Gd(OTf)3/H3PO4 2 24.4 20.6 2.3 1.1 0.6 3.6

Al(OTf)3/H3PO4 5 34.3 26.1 3.3 0.8 3.6 6.5

3.9 Conclusions

Friedel-Crafts alkylation of m/p-cresol isomers

The study has shown that while Zr(OTf)4 is completely soluble in the reaction mixture of cresol

and isobutylene, its activity is comparable to that of less soluble Al(OTf)3 based on conversion of

cresols and selectivities to the di-alkylated products. Sc(OTf)3 gave high conversion of cresols to

mono-butylated products, some of which were further converted into di-butylated cresols. It was

noticeable that Sc(OTf)3 gave high selectivity to DBMC.

The metal triflate salts of the lanthanide series were proven to be inactive as catalysts for the

reaction under the set of conditions employed. The isobutylene oligomers obtained from the

reactions catalysed by aluminium and zirconium triflate salts are comparable. On the contrary,

Sc(OTf)3 afforded a relatively large amount of oligomers. The study showed aluminium triflate

could be recycled at least five times with no significant loss of catalytic activity, capitalising on

the low solubility of Al(OTf)3 in the reaction medium. Among the Brønsted acids evaluated,

H2SO4 gave some activity (although lower than that of Al- and Zr-triflate salts), while H3PO4 was

inactive.

The rate of the reaction was accelerated as the catalyst concentration was increased from 0.05

to 0.35 mol%, but these rate enhancements were mild. There was little or no further increase of

the reaction rate upon increasing the catalyst concentration to 0.8 mol%. At high catalyst

Page 161: Applications of metal triflates and assisted acids as

141

loadings (0.35 and 0.8 mol%), the selectivity to di-alkylated cresols is high due to high

conversion of the cresols and intermediate products (mono-alkylated cresols) to these products.

A significant enhancement of the reaction rate was achieved by the assisted acid systems

compared to individual acids, although the activity of assisted acids is relatively lower than that

of the triflate salts of aluminium and zirconium. The acid combination of Al(OTf)3 and H3PO4

gave similar results to those afforded by Al(OTf)3 on its own, hence there was assisted acidity

occurring. The combination of SPA and La(OTf)3 yielded a superior assisted solid acid catalyst.

This catalyst could be recycled four times, without being completely deactivated.

Friedel-Crafts alkylation of anisole and diphenyl ether

Among the Lewis acids evaluated, only Zr(OTf)4 gave high activity, while the lanthanide triflate

salts were inactive, as was Al(OTf)3. Three products were obtained, i.e. 2-tert-butyl anisole, 4-

tert-butylanisole and 2,4-di-tert-butyl anisole. The Brønsted acids (H3PO4 and HOTf) evaluated

in the study exhibited diverse activity for butylation of anisole, with triflic acid giving a higher

reaction rate than H3PO4. The assisted acids prepared by combining the inactive Lewis acids

with a mildly active Brønsted acid (H3PO4) showed a significant enhancement of the reaction

rate.

The metal triflate salts [Al(OTf)3, La(OTf)3 and Gd(OTf)3] that were evaluated for the butylation

of DPE with isobutylene were inactive, with the exception of Zr(OTf)4. The mineral Brønsted acid

(H3PO4) was also shown to be inactive for this reaction. However, the combination of the metal

triflate salts with H3PO4 gave good activity, indicating that the acid combination is catalytically

active. The butylation reaction of DPE with isobutylene was slow compared to that of cresol and

anisole. During alkylation reaction of DPE, three products were successfully isolated (2-tert-

butyl DPE, 4-tert-butyl DPE and 2,4-di-tert-butyl DPE).

The study has thus shown that some metal triflate Lewis acids can catalyse the Friedel-Crafts

alkylation reactions effectively as individual catalysts, especially Zr(OTf)4 and Al(OTf)3. The

study has also revealed that some individual Brønsted acids and metal triflate Lewis acids may

be inactive as catalysts for the reaction, but when these acids are used concurrently, superior

activity prevailed indicating a synergistic attribute of the acid combinations.

Page 162: Applications of metal triflates and assisted acids as

142

CHAPTER 4 Synthesis of Chromans

This chapter will discuss the results obtained during the evaluation of assisted acid catalysts for

the synthesis of chroman compounds via the reaction of dienes with phenolic substrates.

4.1 Introduction

Benzopyran is a bicyclic organic compound that results from the fusion of a benzene ring to a

heterocyclic pyran ring. Benzopyran systems exist in many instances as natural products, some

of which exhibit biological activity.285 For example, a class of such compounds called

benzotripyrans has recently been found to inhibit HIV.286 Dihydrobenzopyran (chroman) is also

found in vitamin E, its derivatives,287 and flavanoids.288 The most common and straightforward

way of preparing these compounds is via a cyclocoupling reaction of phenol derivatives with

1,3-dienes in the presence of Lewis or Brønsted acids (Scheme 4.1).

OH

R

acid catalyst

O

R4.1

4.2

4.3

R = CH3, Ph, tert-butyl

Scheme 4.1: Addition of 2-methyl-1,3-diene to a phenol derivative.

The catalysts used in such reactions may be homogeneous or heterogeneous acids. 289, 290 High

temperatures are commonly needed, and low to moderate yields of products are reported.

Considerable research efforts have been focussed on obtaining catalyst systems that are

efficient (high activity and selectivity) under mild reaction conditions.291 Among the acids

reported to promote these reactions is H3PO4.292 Youn et al.293 have reported the use of AgOTf

(5 mol%) as catalyst for sequential addition and cyclisation of phenols with dienes and proposed

mechanism shown in Scheme 4.2.

Page 163: Applications of metal triflates and assisted acids as

143

Ag+

OH

R

Ag

OH

R

H+

H+

OH

Ag

R

Ag+

O

Ag

H

R

O

AgR

[Ag+]

Scheme 4.2: Proposed mechanism for AgOTf catalysed synthesis of chromans.

The superior activity of assisted acids prompted us to further evaluate these acids for the

synthesis of chomans via the reaction of dienes with phenolic compounds. The synthesis of

chromans involves the alkylation reaction of a phenolic compound followed by the etherification

reaction to form a cyclic benzofuran or benzopyran. It is mentioned in the literature that such

reactions normally require high temperatures and large amounts of catalysts to occur. Having

observed the high activity of the assisted acids for hydroalkoxylation of olefins and Friedel-

Crafts alkylation reactions, it was decided to apply these catalysts systems to the synthesis of

the bicyclic materials.

Page 164: Applications of metal triflates and assisted acids as

144

4.2 Reaction of phenol with 2-methyl-1,3-butadiene (isoprene)

The very first reaction involved alkylation of phenolic compounds with isoprene in the presence

of the La(OTf)3/H3PO4 assisted acid. The reaction was performed in a 300 mL autoclave reactor

fitted with a gas entrainment stirrer. In the first reaction, p-cresol, isoprene and catalyst (0.1

mol%) were weighed directly into the reactor vessel according to Scheme 4.3. Prior to heating

the reactor to the operating temperature, a significant exotherm was evident, i.e. the reaction

temperature increased from room temperature up to 220 °C. This strong exotherm was clearly

demonstrative of a vigorous reaction which self-initiated even at ambient temperature, and

indicated that perhaps the reactants should be added sequentially and possibly slowly, the one

to the other.

[La(OTf)3/H3PO4]

OH O

4.4 4.5

4.2

Scheme 4.3: Preparation of 2,2,6-trimethylchroman catalysed by La(OTf)3/H3PO4

Having observed such a large exotherm, the subsequent reactions were performed in a glass

reactor setup (round bottom flask fitted with a reflux condenser and equipped with magnetic

stirrer) for the ease of controlling the reaction temperature. In this experiment, the catalyst and

the p-cresol were weighed into a round bottom flask. The olefin (isoprene) was added drop-

wise, while stirring the reaction mixture, and using an ice bath to control the reaction

temperature. The reaction was allowed to proceed for the next 6 hours, and thereafter a sample

was taken and treated with sodium bicarbonate before it was injected into a gas chromatograph

coupled to an FID detector.

The GC results showed the presence of a product peak (79%), which was substantiated with

GC-MS analysis. The product was purified using a Kugelrohr distillation setup, and the sample

was analysed using NMR spectroscopy, and was found to correspond to the desired 2,2,6-

trimethylchoman. The 1HNMR spectrum showed a singlet peak at δ = 1.21 ppm, intergrating for

6 protons of the two methyls on the pyran ring. The triplet peak at δ =1.64 -1.67 ppm,

intergrating for two protons correspond to -CH2- of the pyran ring furtherst from the benzene

Page 165: Applications of metal triflates and assisted acids as

145

ring, The singlet at δ = 2.13 ppm, intergrating for 3 protons is due to the methyl on the para-

position of the benzene ring. The second triplet peak at δ = 2.50-2.65 ppm integrating for two

protons is corresponds to the -CH2- of the pyran ring closest to the benzene ring, while the

multiplet peak at δ =6.57-6.78 ppm is ascribed to the protons of the benzene ring. The 13CNMR

spectrum showed 11 peaks instead of 12 and this is because, the two methyl peaks bonded to a

quaternary carbon at δ = 73.88 ppm are chemically equivalent with δ = 26.86 ppm. The 13C

signal of the methyl on the benzene ring appeared at δ =20.48 ppm, while those of the aromatic

ring appeared at δ =120.57-151.77 ppm.

Having obtained good catalytic activity from La(OTf)3/H3PO4 during the reaction of p-cresol with

isoprene, it was decided to evaluate various phenolic compounds in reactions with isoprene

(Table 4.1). In the current study, the reaction of isoprene with hydroquinone yielded two

isomeric products (4.8 and 4.9), with an overall yield of 84%.. The isoprene molecules were

added on both hydroxyl groups of hydroquinone (Scheme 4.4), thereafter cyclising to form

tricyclic products. In the literature, it has been reported that only 68% of 4.7 forms290 upon using

Amberlyst 15 as catalyst. The initial interpretation of these results is that the Amberlyst 15

system provides a catalyst that is less active and more selective for the monosubstituted

product, while the present catalyst is more active leading to the two disubstituted products. In

another report, 294 the condensation of hydroquinone with isoprene in the presence of ortho-

phosphoric acid yielded three products (4.7, 4.8 and 4.9). Consequently, it can be concluded

that the distribution of the products obtained from this reaction is dependent to some extent on

the activity of the catalyst employed.

In the current study, purification of the isomers was performed using column chromatography

techniques with hexane/ethyl acetate (solvent system 100:1) and using alumina as the

stationary phase. The GC-MS analysis of these compounds gave a molecular ion or parent ion

with overall m/z = 246 (which corresponds to the molar mass of each of the two isomers). The

NMR characterisation technique was useful during characterisation of 4.8 and 4.9. The isomer

4.8 showed, beside other signals,a singlet peak for the aromatic protons at δ =6.58 ppm and

integrating for two protons (H-14 and H-15), while the 4.9 isomer exhibited a singlet peak at δ =

6.49 ppm and also integrating for two protons (H-7 and H-15). However, a clear distinction

between the two isomers could not be made based on the 1HNMR spectra as noted by the

splitting pattern of the aromatic ring protons H-7, H-14 and H-15. The 13CNMR spectra of the

two isomers are similar, hence there was no clear differences in the chemical shifts of the two

Page 166: Applications of metal triflates and assisted acids as

146

isomers. Nevertheless, the number of peaks on the spectrum corresponds to the quantity of

carbon atoms in the compounds.

O

O

O

O

OH

OH

O

OH

H+

4.6 4.74.8 4.9

78

15

14

15

4.2

Scheme 4.4: Reaction of isoprene with hydroquinone in the presence of La(OTf)3/H3PO4

During the reaction of phenol with isoprene, the major product formed is a chroman (4.10). Four

minor product peaks were also present on the GC trace, although the minor peaks were not fully

characterised in the current study. In a publication by Dewhirst and Rust,295 it is reported that

the reaction of phenol with isoprene in the presence of aluminium phenoxide yields the products

(4.10-4.14) presented in Scheme 4.5. Furthermore, Dewhirst and Rust also reported that

bisphenol is formed at higher temperatures (>100 °C) in the presence of aluminium phenoxide.

These by-products correspond to the C-alkylation reactions of phenol with isoprene forming

ortho-, para-, and ortho-/para- alkylated phenol as well as the p-alkylated chroman products.

The average combined yield reported by Dewhirst and Rust was 70%, with the product

districution ratio of 4.10:4.12:4.13:4.14 = 0.38:0.25:0.24:0.13

Page 167: Applications of metal triflates and assisted acids as

147

4.10 4.12

4.13 4.14

O

O

OH

OH

Scheme 4.5: Products obtained from the reaction of phenol and isoprene.

The Brønsted acid (H3PO4) catalysed reaction of isoprene with phenol was also reported by

Bader and Bean.296 In their study, six products were obtained including the alcohols which were

formed via partial hydration of ortho- and para-alkylated phenols. In the current study, it is

proposed that the carbocation that is formed via protonation of the diene by the acid undergoes

nucleophillic attack by the OH group of a phenol, followed by [3,3] sigmatropic rearrangement to

form ortho-alkylated phenol intermediate, which is consequently followed by cyclisation via the

tertiary carbocation to form the ether product as shown by the porposed mechanism in Scheme

4.6.

Page 168: Applications of metal triflates and assisted acids as

148

OH

R

H+ O

OH

H+

OH

-H+

R

RR

O

R

-H+

Scheme 4.6: Proposed mechanism for preparation of chroman using Brønsted acid.

The various chromans obtained in the present study during the reaction of different phenols with

isoprene are presented in Table 4.1. During the reactions of 2-tert-butyl-4-methylphenol and 2-

tert-butylphenol and naphthol with isoprene, high yields of the corresponding chroman products

(4.17, 4.19 and 4.21) were isolated.

Page 169: Applications of metal triflates and assisted acids as

149

Table 4.1: Reaction of various phenols (2.1 mmol) with isoprene (7.0 mmol) in the presence of assisted acid La(OTf)3/H3PO4.

Entry ROH Product Yield (%)

1

OH

4.15

O

4.10

53

2

OH

4.4

O

4.5

74

3

OH

4.16

O

4.17

81

4

OH

4.18

O

4.19

83

5

OH

4.20

O

4.21

92

Page 170: Applications of metal triflates and assisted acids as

150

6

OH

OH4.6

O

O

O

O

4.8 4.9

84

The reaction of phenol with isoprene gave a relatively low yield (53%) of the corresponding

chroman, due to the formation of several by-products, while p-cresols yielded 74% of the

chroman derivative (4.5). All the other substrates gave large quantities of the chroman products

(>80%), the highest yield being 92% afforded from the reaction of 2-naphthol with isoprene.

4.3 Reaction of phenols with 2,3-dimethyl-1,3-butadiene

The reaction of substituted phenols with 2,3-dimethyl-1,3-butadiene was also performed using

La(OTf)3/H3PO4 as catalyst. In these reactions a 1:1 mole ratio of phenolic substrate to the olefin

was used, and the olefin was added drop-wise, into the reaction mixture via an addition funnel.

Here, too, all reactions were also exothermic. The separation of isomers was conducted using

column chromatography. The solvent system used was hexane/ethyl acetate (100:1 v/v)

mixture. The reaction of each phenol with the diene yielded two isomers (Table 4.2), the first

isomer consisting of 5-membered chiral cyclic ether, while the second isomer consists of six-

membered cyclic ether.

The results obtained from the reaction of phenolic substrates with 2,3-dimethyl-1,3-butadiene

instead of isoprene are completely distinct. This is due to the fact that isoprene can form a tert-

carbocation on only one carbon atom, whereas 2,3-dimethyl-1,3-butadiene can form tert-

carbocations on two carbon centres, and hence the reactivity of 2,3-dimethyl-1,3-butadiene is

relatively high. For all the reactions, the GC-MS analysis confirmed the presence of two isomer

peaks. For example, the compounds 4.21 and 4.22 gave an m/z=190, corresponding to the

molecular ion. The 1HNMR spectrum of 4.21 showed only one doublet peak at δ = 1.00 ppm,

integrating for three protons which corresponds to a single methyl group attached to the pyran

ring, while two singlet peaks at δ = 1.14 and 1.36 ppm corresponding to the two methyl groups

of the pyran ring were also present. Furthermore, a sextet peak at δ = 1.85-1.94 ppm for a

Page 171: Applications of metal triflates and assisted acids as

151

single proton next to the -CH2- on the pyran ring. On the other hand, 4.22 showed two doublet

peaks at δ = 0.93 ppm and at δ = 0.99 ppm, corresponding to the methyl groups of the isopropyl

substituent of the furan ring. A singlet reasonating at δ = 1.33 ppm, which correspond to the

methyl functional group of the furan ring, was also noted and a septet at δ = 1.96-2.02 ppm was

indicative of a proton on the isopropyl substituent. These unique characteristics of the isomers

in entry1 are common for the other isomers in entries 2 and 3, except that the chemical shifts

are different.

A possible mechanism for the formation of 4.21 and 4.24 is shown in Scheme 4.7. The initial

step in this mechanism involves the protonation of the olefin, yielding a tertiary carbocation,

followed by the nucleophillic addition of the carbocation to the ortho-position of p-cresol, forming

an olefinic substrate (which is still likely to react further), Thus, the tert-olefinic cresol substrate

get protonated and form a carbocationic fuctional group. Both positions of the olefinic fuctional

group forms tert-carbocations after being protonated. As a result, the reaction of the hydroxyl

functional groups of cresol with carbocations, lead to the formation of two products, i.e. a five-

membered ring (4.22) and a six-membered ring (4.21) chroman isomers. The isomer consisting

with a five membered ring is chiral. The chromans 4.23 and 4.24, 4.25 and 4.26 in Table 4.2 are

formed via the same mechanistic pathway as 4.21 and 4.24.

Page 172: Applications of metal triflates and assisted acids as

152

H+

OH

OH

H+

OH

O O

(4.21) (4.22)

OH

-H+-H+

Scheme 4.7: Brønsted acid catalysed preparation of chroman.

The reaction of phenolic substrates with 2,3-dimethyl-1,3-butadiene also gave high quantities of

the corresponding chromans in all cases. It was noticeable that 2-tert-butyl phenol gave a

relatively low yield (73%) as compared to other substrates, but this remains a respectable yield.

The latter could be attributed to the fact that this compound is activated for further alkylation on

the para-position, which may lead to the formation of p-alkylated 2-tert-butylphenol by-product

or possibly by dealkylation of the t-butyl group leading to by-products..

Page 173: Applications of metal triflates and assisted acids as

153

Table 4.2: Reaction of substituted phenols with 2,3-dimethyl-1,3-butadiene.

Entry ROH Products Overall yield (%)

1

OH

4.4

O

4.21

O

4.22

87 (4:1)

2

OH

4.16

O

4.23

O

4.24

93 (1:1.7)

3

OH

4.18

O

4.25

O

4.26

73 (1:1.5)

The six membered ring isomer (4.21) in entry 1 is formed favourably compared to that with furan

ring (4.22), while in entry 2 and 3, the isomers with furan rings are formed in a slight excess.

This could be due to the fact that tert-butyl groups on the benzene ring increases the steric

hindrance, thereby favouring more distally positioned bulk. These results hint at the possibility of

rapid migration of the carbocation between the two tertiary centres caused by facile1,2-H shift.

This is possible and would probably arise because of the anticipated energy

compatibility/equivalence between the two carbocations. Here, product development control

Page 174: Applications of metal triflates and assisted acids as

154

may be determinative in the outcome of the reaction, arising from a late transition state

according to Hammond postulate.297

4.4 Conclusions

The study has shown that the assisted acids system such as La(OTf)3/H3PO4 can catalyse the

condensation of phenols and butadienes efficiently in one-pot transformations. Furthermore, the

catalyst was able to promote the reactions at room temperature.

The reaction of phenol with isoprene gave relatively low yields, due to the formation of by-

products, while the condensation of hydroquinone with isoprene yielded two product isomers.

The tert-butylated phenols reacted effectively with isoprene, giving high yields of the chroman

products.

The reaction of 1,3-dimethyl-2,3-butadiene with various phenolic substrates gave two chroman

isomers (five-membered dihydrobenzofurans and six-membered chromans), the products being

chiral but racemic. In all instances high yields were obtained. It may be an interesting future

project to determine if the latter set of products may be prepared in an enantioselective fashion

in the presence of chiral Lewis/Brønsted acid pairs. It will further be a useful part of the study to

see if selectivity for the chroman or dihydrofuran may be enhanced.

Page 175: Applications of metal triflates and assisted acids as

155

CHAPTER 5 Experimental Protocol

This chapter describes the experimental protocols, analytical techniques and the reactor set-ups

employed when performing varous experiments

5.1 General protocol for preparation of ethers

The etherification reactions were carried out in a 300 mL autoclave stainless steel reactor

equipped with gas entrainment stirrer. The reagents (methanol and an olefin) were weighed

directly into the reactor. The reactor was then assembled and heated up to the operating

temperature, while stirring at 1000 rpm. Upon reaching the operating temperature, the catalyst

dissolved in methanol was introduced into the reactor using a sample bomb or pressurised

stainless steel vessel (6 bar N2). The reaction was monitored over time from the addition of the

catalyst, with samples taken and analysed in a gas chromatograph (GC). The reactor setup is

shown in Figure 5.1.

Catalyst preparation and addition: The required amount of catalyst was weighed in a glass

beaker, and dissolved in 5 g of methanol. The catalyst solution was then transferred into a 50 ml

stainless steel sample bomb using a syringe. Finally, the vessel was pressurised with N2 (6 bar).

While the reactor was being heated to the operating temperature, one end of the catalyst vessel

was connected onto the reactor inlet, upon reaching the operating temperature, the catalyst was

introduced into the reactor by first opening the outlet valve of the samples bomb followed by

opening the inlet valve of the reactor to allow all the catalyst contents to empty into the reactor.

Example: Methanol (56.03 g; 1900 mmol), and 2M2B (11.02 g; 150 mmol) were weighed into

the reactor. The reactor was assembled and heated up to the operating temperature (100 °C).

At 100 °C, the addition of Al(OTf)3 (0.08 g; 0.17 mmol) dissolved in methanol (5 g; 156 mmol)

was performed.

Page 176: Applications of metal triflates and assisted acids as

156

Figure 5.1: Reactor set-up used for etherification of alcohols and olefins

5.1.1 Reagents

The reagents (Table 5.1) used during the etherification reactions were obtained from different

sources. The isoprene used in these reactions was 95% purity, and was further purified by

flashing it through a column packed with activated alumina. After this step, the purity was found

to be 99% (as analysed in GC). The impurities were identified by GC-MS to as C10 cyclic

compounds. All the other reagents were used as obtained from the supplier. The

methylenecyclopentene used was obtained from a mixture of primary and secondary olefins. All

the metal triflate salts were obtained from Sigma-Aldrich except zirconium and aluminium

triflates, which were obtained from Sasol (Germany). The Brønsted acids were also obtained

from Aldrich, with high purity i.e. >98%.

NV: Non return valve

PI: Pressure indicator

PT: Temperature indicator

SIC: Stirrer control

S: Sampling line

M: stirrer mottor

~: Bursting disc

Page 177: Applications of metal triflates and assisted acids as

157

Table 5.1: Reagents used during etherification reaction of olefins and alcohols

Reagent Supplier Purity (%)

Methanol Aldrich 99

2M2B Saarchem 99

Isoprene Aldrich 99

Styrene Saarchem 99

1-hexene Sasol 99

methylenecyclopentene Sasol 22

Tert-amylalcohol Saarchem 99

Ethanol Saarchem 99

n-Butanol Aldrich 99

Lanthanide triflates Aldrich 99+

Zr(OTf)4 and Al(OTf)3 Sasol 99+

5.1.2 Purification of ethers

The purification of ether products was performed in an Aldershaw distillation set-up (Figure 5.2).

The reaction mixture was transferred from the stainless steel reactor into a 500 mL round

bottom flask and the glass beads were added to avoid spluttering and create smooth boiling of

the reaction mixture. A 15-plate Aldershaw column was fitted onto the round bottom flask. A

condenser was also fitted on top of the column and the distillation set up was assembled to

completion. The reflux timer was used to control the rate of sample reflux and sample

withdrawal. A heating mantle was used to supply heat to the reaction mixture and the

thermocouples were used to monitor the temperature changes and tap-water was used as a

coolant which was circulated in the condenser.

Page 178: Applications of metal triflates and assisted acids as

158

Figure 5.2: Experimental set-up for the purification of ethers.

Page 179: Applications of metal triflates and assisted acids as

159

5.1.3 Gas chromatography analysis

The samples were analysed in using a 5890 Agilent gas chromatography connected to a PC

equipped with Class-VP software for recording the integration of chromatograms. The free fatty

acid (FFAP) column with length of 50 m, internal diameter of 0.2 mm was utilised. The

temperature limits of this column are between 60 and 250 °C. The GC programme was set to

run for 15 minutes, and the flame ionisation detector was used for detection of the reaction

components. The GC programme is shown in (Table 5.2).

Table 5.2: GC operating conditions during analysis of the etherification reaction mixture.

Parameter Condition

Carrier gas Hydrogen

Initial Temperature 60 °C

Initial time 5 min.

Temperature ramping rate 6 °C/min.

Final temperature 240 °C

Final time 5 min.

Injector temperature 250 °C

Detector temperature 250 °C

5.1.4 NMR analysis

The NMR analysis was used to confirm the nature of the ethers or products obtained from the

reactions. The analysis was carried out using a Varian 400 MHz NMR, the spectra were

recorded in deuterated chloroform (CDCl3) at room temperature with TMS as an internal

standard. The chemical shifts are reported in parts per million (ppm) and the coupling constants

(J) in Hertz. Tert-amyl alcohol was characterised using an authentic sample obtained from

Saarchem.

Page 180: Applications of metal triflates and assisted acids as

160

Characterisation of tert-amyl-ether

O21

4

5

3

1H NMR (400 MHz, CDCl3): σ = 0.79-0.83 (t, 3H, J=7.5 Hz, H-3); 1.07 (s, 6H, H-4); 1.41-1.47 (q,

2H, J=7.3, H-2); 3.11(s, 3H, J=0, H-5)

13C NMR (400 MHz, CDCl3): σ = 8.06 (C-3); 24.5 (C-4); 32.1 (C-2); 48.9 (C-5); 74.5 (C-1).

Characterisation of 1,1-dimethyl-2-propenyl methyl ether

O2

1

3

4

5

1H NMR (400 MHz, CDCl3): σ = 1.26 (s, 6H, H-4); 3.20 (s, 3H, H-5); 5.10-5.16 (m, 2H,H-1);

5.76-5.84 (m, 1H, H-2)

13C NMR (400 MHz, CDCl3): σ = 24.5 (C-4); 49.5 (C-5); 74.28 (C-3); 113.0 (C-1); 142.5 (C-2)

m/z (GC-MS): (M+,2%); (M-CH3, 100%); 73 (M-C2H6, 30%); 55 (M-OC2H6), 51%); 41 (M-OC3H9,

37%); 27 (M-OC4H9), 9%).

Characterisation of 2-methyl-3-butenyl methyl ether

O12

345

6

1H NMR (400 MHz, CDCl3): σ=1.59 (s, 3H, H-5); 1.66 (s, 3H, H-6); 3.21 (s, 3H, H-1); 3.80-3.81

(d, 2H, J=7.1, H-2); 5.23-5.72 (t, 1H, J=6.8, H-3)

13C NMR (400 MHz, CDCl3): σ= 17.6 (C-6); 25.66 (C-5); 57.4 (C-1); 68.6 (C-2); 121.5 (C-3);

136.6 (C-4)

Page 181: Applications of metal triflates and assisted acids as

161

m/z (GC-MS): 100 (M+,14%); 85 (M-CH3, 100%); 69 (M-OCH3, 14%); 55 (M-OC2H6), 29%); 41

(M-OC3H9, 32%); 29 (M-OC4H9, 9%).

Characterisation of 1,3-dimethoxy-3-methylbutane

O O1 2

3

4

5

67

1H NMR (400 MHz, CDCl3): σ= 1.17 (s, 6H, H-3 and H-4); 1.77-1.80 (t, 2H, J=7.3, H-5); 3.19 (s,

3H, H-1); 3.32 (s, 3H, H-7); 3.44-3.47 (t, 2H, J=7.3, H-6)

13C NMR (400 MHz, CDCl3): σ= 25.3 (C-3 and C-4); 39.4 (C-5); 49.4 (C-1); 58.5 (C-7); 67.0 (C-

5); 74.1 (C-2)

Characterisation of 1-methoxy-1-methylbenzene

8

9

O 12

3

45

67

1H NMR (400 MHz, CDCl3): σ=1.59-1.57 (d, 3H, J=6.6 Hz, H-2); 3.32 (s, 3H, H-7); 4.34-4.39 (q,

1H, J=6.35, H-1); 7.30-7.44 (m, 5H, H-5 to H-8)

13C NMR (400 MHz, CDCl3): σ= 23.9 (C-2); 56.7 C-3); 80.0 (C-1); 126.3 (C-6 and C-8); 127.6

(C-7); 128.6 (C-5 and C-9); 143.7 (C-4)

m/z (GC-MS): 136 (M+, 3%); 121 (M-CH3, 100%); 105 (M-OCH3, 23%); 91 (M-OC2H6, 13%); 77

(M-OC3H7, 24%); 28 (M-OC5H10), 26%).

Page 182: Applications of metal triflates and assisted acids as

162

Characterisation of 1-methoxy-1-methylcyclopentane

O1 23

6

7

5

4

1H NMR (400 MHz, CDCl3): σ= 1.18 (s, 3H, H-3); 1.32-1.38 (m, 2H, H-4); 1.46-1.56 (m, 2H, H-

7); 1.60-1.68 (m, 2H, H-5); 1.71-1.79 (m, 2H, H-6); 3.11 (s, 3H, H-1)

13C NMR (400 MHz, CDCl3): σ= 22.7 (C-3); 23.9 (C-5 and C-6); 37.8 (C-4 and C-7); 50.0 (C-1);

84.7 (C-2)

m/z (GC-MS): 114 (M+, 13%); 99 (M-CH3), 19%); 85 (M-OCH3, 100%); 72 (M-OC2H6, 47%); 55

(M-OC3H6, 26%); 43 (M-OC4H8), 15%).

Page 183: Applications of metal triflates and assisted acids as

163

5.2 General protocol for alkylation of phenolic compounds

The butylation reactions were performed in a 300 mL autoclave reactor equipped with a heating

mantle, gas entrainment stirrer, bursting disc, dip-tube and several valves (Figure 5.3).

Figure 5.3: The experimental set-up for the butylation of phenolic compounds

The isobutylene from the cylinder was allowed to pass through a regulator (locked at 2 bar), via

a heat exchanger (70 °C) to ensure that isobutylene passes through the mass flow meter

(coriolis) was in the gas phase. The isobutylene was heated again to 70 °C (to avoid cooling the

reaction mixture when the olefin wa introduced) before it entered the reactor via a dip-tube (also

used for taking samples). A non return valve located after HV2 was installed to avoid any back

pressure. The cooling coil to which the tap water was circulated for maintaining the reaction

temperature was also installed. This was because butylation reactions of some phenolic

compounds such as cresols are exothermic. The reactor was equipped with bursting disc and

depressurising valves to vent any unnecessary pressure build up.

Page 184: Applications of metal triflates and assisted acids as

164

Butylation reaction example: A mixture of meta- and para-cresol (60.0g; 554.9 mmol) together

with a catalyst Zr(OTf)4 (0.61g; 0.89 mmol) were weighed directly onto the reactor. The reactor

was assembled and heated up to the operating temperature (70 °C) while stirring at 1200 rpm.

When the temperature of 70 °C is attained, the isobutylene is allowed to flow into the reactor

and the reaction was allowed to proceed over a period of 60 minutes and taking samples over

fixed time intervals. The isobutylene cylinder was then closed, but, the reaction was allowed to

proceed for another 60 minutes (post run) with residual isobutylene already in the reactor. After

120 minutes of the total reaction time, the reactor was cooled down, followed by venting out any

residual isobutylene that may be present and finally pouring out the reaction mixture and

cleaning the reactor.

Reagents

The reagents used during alkylation reactions were obtained from various suppliers, Table

5.3.The Lewis acids and Brønsted acids used as catalysts for these reactions were obtained

from the suppliers specified in Table 5.1

Table 5.2.

Table 5.3: Reagents used during alkylation or butylation of phenolic compounds

Reagent Supplier Purity (%)

m/p-Cresol Sasol ≥99

Anisole Sigma-Aldrich 99.5

Diphenylether Sigma-Aldrich 99.8

isobutylene Air Liquide 99.9

Purification and analysis of butylated products

The purification of the products obtained during alkylation of anisole and DPE was performed in

the same set-up used to purify ether products in Figure 5.2.

Page 185: Applications of metal triflates and assisted acids as

165

Analytical techniques for characterisation of the reaction products.

The authentic samples obtained from Sigma Aldrich were used as standards during

characterisation of the products obtained from the reaction of cresols with isobutylene. The

products obtained from the reactions where anisole and DPE were used as substrates and were

characterised using GC, GC-MS and NMR techniques after purification.

GC analysis: The analysis of the reaction mixtures was performed in a 6890 Shimadzu gas

chromatography. A DB-1 capillary column with length of 20.0 m, internal diameter of 100 μm

and film thickness of 0.10μm was used at constant pressure. The program was set to run for 20

minutes. The flame ionisation detector (FID) was used for detection of reaction products. The

GC analysis conditions are shown in Table 5.4.

Table 5.4: Analysis conditions of the butylated compounds

Parameter Condition

Carrier gas Hydrogen

Initial Temperature 100 °C

Initial time 0 min.

Temperature ramping rate 10 °C/min.

Final temperature 290 °C

Final time 0 min.

Injector temperature 250 °C

Detector temperature 300 °C

Page 186: Applications of metal triflates and assisted acids as

166

NMR and GC-MS analysis

The reaction of anisole with isobutylene afforded two mono-butylated products (2-tert-

butylanisole and 4-tert-butylanisol) and one di-butylated anisole (2,6-di-tert-butyl anisole)

Characterisation of 2-tert-butylanisole

O1

23

4

56

7

89

10

11

1H NMR (400 MHz, CDCl3): δ=1.38 (s, 9H, H-5,H-6 and H-7); δ=3.80 (s,3H, H-1); δ=6.82-6.94

(m, 2H, J=7.6 H-9 and H-10); δ=7.13-7.19 (d,1H, J=9.8, H-8); δ=7.23-7.26 (d, 1H; J=9.5, H-11)

13CNMR (400 MHz, CDCl3): δ=30.0 (C-5, C-6 and C-7); δ=34.9 (C-4); δ=55.1 (C-1); δ=111.9 (C-

11); δ=120.8 (C-9); δ=127.0 (C-10); 138.4 (C-3); δ=158.8 (C-2)

m/z (GC-MS): 164 (M+, 27%); 149 (M-CH3, 100%); 134 (M-C2H6, 5%); 121 (M-C3H9, 51%); 105

(M-C4H9); 91 (M-C5H12, 35%); 77 (M-OC5H12, 10%).

Characterisation of 4-tert-butylanisole

1

23

456

7

8

9

10

11

O

1H NMR (400 MHz, CDCl3): δ=1.28 (s, 9H, H-7,H-8 and H-9); δ=3.71 (s,3H, H-1); δ=6.80-6.83

(d, 2H, J=8.8, H-4 and H-10); δ=7.26-7.29 (d, 2H, J=8.8, H-3 and H-11)

13C NMR (400 MHz, CDCl3): δ=31.9 (C7-9); δ=34.3 (C-6); δ=55.7 (C-1); δ=113 (C-11 and C-13);

δ=126.6 (C-4 and c-10); δ=143.3 (C-5); δ=157.8 (C-2)

Page 187: Applications of metal triflates and assisted acids as

167

m/z (GC-MS): 164 (M+, 15%); 149 (M-CH3, 100%); 133 (M-C2H6, 3%); 121 (M-C3H9, 27%); 109

(M-C4H9, 13%); 91 (M-C5H12, 10%); 77 (M-OC5H12), 8%).

Characterisation of 2,4-di-tert-butylanisole

O1

23

4

56

7

8910

11

1213

14

15

1H NMR (400 MHz, CDCl3): δ=1.30 (s, 9H, H-11,H-12 and H-13); δ=1.39 (s, 9H, H-5, H-6 and

H-7); δ=3.78 (s,3H, H-1); δ=6.18-6.83 (d, 1H, J=8.6 H-14); δ=7.15-7.19 (m,1H, J=6.1, H-8);

δ=7.31-7.32 (d,1H, J=2.4, H-15)

13C NMR (400 MHz, CDCl3): δ=29.8 (C-5, C-6 and C-7); δ=31.8 (C-11, C-12 and C-13); δ=34.5

(C-4); δ=35.3 (C-10); δ=55.3 (C-1); δ=111.2 (C-15); δ=123.5 (C-8); δ=124.3 (C-14); δ=137.6 (C-

3); 142.5 (C-9); δ=156.4 (C-2).

m/z (GC-MS): 220 (M+, 20%); 205 (M-CH3, 100%); 189 (M-C2H6, 2%); 175 (M-C3H9, 4%);

147(M-C5H12, 3%); 105 (M-C8H18, 2%); 91 (M-C9H21, 2%); 57 (M-OC11H23, 9%)

The Friedel-Crafts reaction of DPE with isobutylene yielded two mono-alkylated products, one

di-alkylated product as well as one tri-alkylated product. The heavy products with high viscosity

were also obtained. The NMR data of the alkylated DPE products is shown below.

Page 188: Applications of metal triflates and assisted acids as

168

Characterisation of 1-tert-butyl-2-diphenylether (2-tert-butyl-DPE)

O123

5

6

7

89

10

11

1213

14

1516

4

1H NMR (400 MHz, CDCl3): δ=1.42 (s, 9H, H-4, H-5 and H-6); δ=6.81-7.41 (m, 9H, ph).

13C NMR (400 MHz, CDCl3): δ=30.4 (C-4 to C-6); δ=34.9 (C-2); δ= 118.9 (C-16 and C-12);

δ=120.3 (C-10); δ=122.8 (C-14): δ=123.3 (C-8); δ=127.3 (C-7); δ=129.7 (C-13 and C-15);

δ=141.2 (C-2); δ=156.1 (C-1); δ=158.0 (C-11)

m/z (GC-MS): 226 (M+,59%); 211 (M-CH3, 100%); 195 (M-C2H6, 14%); 181 (M-C3H9, 13%);

133(M-C7H11, 15%); 91 (M-C10H12, 26%); 77 (M-OC10H12, 13%); 57 (M-OC12H13, 9%)

Synthesis of 1-tert-butyl-2-diphenylether (2-tert-butyl-DPE)

O 12

3

45

6

7

8

910

11

1213

14

1516

1H NMR (400 MHz, CDCl3): δ=1.32 (s, 9H, H-6 to H-8); δ=6.92-7.34 (m, 9H, ph).

13C NMR (400 MHz, CDCl3): δ= 31.8 (C-6 to C-8); δ=34.5 (C-5); δ= 118.5 (C-2); δ= 118.7 (C-12

and C-16); δ=123.1 (C-14); δ=126.8 (C-9 and C-3); δ=129.8 (C-13 and C-15); δ=146.4 (C-4);

δ=155.1 (C-1); δ=157.8 (C-11).

m/z (GC-MS): 226 (M+, 28%); 211 (M- CH3, 100%); 195 (M-C2H6, 2%); 183 (M-C3H9, 6%); 91

(M-PhC3H9, 9%); 77 (M- OPhC3H9, 9%).

Page 189: Applications of metal triflates and assisted acids as

169

Characterisation of 2,4-di-tert-butyl-phenylpheyl ether

O 1 2

34

5

6

7

8 9 10

11

12

13

1415

16

17

18

19

20

1H NMR (400 MHz, CDCl3): δ=1.23 (s, 9H, H-3, H-5, H-6); δ=1.35 (s, 9H, H-10, H-11, H-12);

δ=6.60-7.41 (m, 8H, ph)

13C NMR (400 MHz, CDCl3): δ=30.3 (C-3, C-4, C-6); δ=31.7 (C-10, C-11, C-12); δ=34.6 (C-5);

δ=35.0 (C-9); δ=118.6 (C-16 and C-20); δ=119.7 (C-14); δ=122.4 (C-13); δ=123.8 (C-7);

δ=124.1 (C-18); δ=129.6 (C-17, C-19); δ=140.0 (C-2); δ=145.8 (C-8); δ=153.4 (C-1); δ=158.1

(C-15)

m/z GC-MS: m/z (GC-MS): 282 (M+, 27%); 267 (M-CH3, 100%); 254 (M-C2H6, 2%); 211 (M-

(C4H9), 2%); 91(M-OC14H22, 5%); 77 (M-OC14H22, 4%); 57 (M-OC16H23, 14%)

5.3 General protocol for synthesis of chromans

The synthesis of chromanes was performed in a 250 mL three-neck round bottom flask fitted

with a reflux condenser and equipped with a magnetic stirrer. The reactions were performed at

room temperature, under atmospheric pressure. During the synthesis of chromanes, all

reagents were used as obtained from the suppliers.

Example: p-cresol (30 g; 280 mmol) and the assisted acid [La(OTf)3/H3PO4] (0.1602 g; 0.27

mmol/ 0.0482 g; 0.490 mmol) were weighed directly into the flask, followed by assembling the

reaction set-up. Subsequently, an addition funnel containing isoprene (30.53 g; 448.97 mmol)

was inserted and the olefin (isoprene) was added drop-wise into the flask, while stirring the

reaction mixture. When the reaction temperature started to increase, a cooling bath containing

ice cold water was periodically inserted to maintain the reaction mixture at room temperature.

Page 190: Applications of metal triflates and assisted acids as

170

The purification of some of the compounds was performed using column chromatographic

techniques, loaded with activated, neutral alumina and using hexane: ethylacetate solvent

system in a ratio of 9:1(v/v). Some of the compounds were further purified in a Kugelrhor

distillation set up shown in Figure 5.4. The chromanes were characterised using GC-MS and

NMR spectroscopy

Figure 5.4: Kugelrhor distillation setup used for purification of chromans.

Page 191: Applications of metal triflates and assisted acids as

171

Characterisation of 8-tert-butyl-2,2-dimethylchroman

O

12

3

4

56

78

9

101112

1314

15

1H NMR (400 MHz, CDCl3): δ=1.48 (s, 6H, H-1 and H-3); δ=1.59 (s, 9H, H-12 to H-14); δ=1.88-

1.92 (t, 2H, H-4); δ=2.90-2.93 (t, 2H, H-5); δ=6.85-7.23 (m, 3H, Ph)

13C NMR (400 MHz, CDCl3): δ=23.2 (C-5); δ=27.0 (C-12 to C-14); δ=30.1 (C-4); δ=34.9 (C-11);

δ=74.0 (C-2); δ=118.9 (C-8); δ=121.1 (C-6); δ=124.5 (C-9); δ=127.6 (C-7); δ=137.9 (C-15);

δ=152.6 (C-10)

m/z (GC-MS): 218 (M+, 36%); 203 (M-CH3, 58%); 163 (M-C4H9, 34%); 147 (M-C5H12, 100%);

119(M-OC6H15, 11%); 91 (M-OC8H17, 13%); 77 (M-OC9H19, 7%)

Characterisation of 8-tert-butyl-2,2,6-trimethylchroman

O

12

3

4

56

78

9

101112

1314

15

16

1H NMR (400 MHz, CDCl3): δ=1.22 (s, 6H, H-1 and H-3); δ=1.27 (s, 9H, H-12 to H-14); δ=1.61-

1.64 (t, 2H, H-4); δ=2.13 (s, 3H, H-16); δ=2.61-2.64 (t, 2H, H-5); δ=6.16 (s, 1H, H-9); 6.80 (s,

1H, H-7)

13C NMR (400 MHz, CDCl3): δ=20.8 (C-16); δ=23.1 (C-5); δ=26.9 (C-1 and C-3); δ=29.2 (C-12

to C-14); δ=32.8 (C-4); δ=34.7 (C-11); δ=73.6 (C-2); δ=120.6 (C-6); δ=125.3 (C-9); δ=127.6 (C-

Page 192: Applications of metal triflates and assisted acids as

172

7); δ=127.8 (C-8); δ=137.6 (C-10); δ=150.3 (C-15)

m/z (GC-MS): 232 (M+, 49%); 217 (M-CH3, 41%); 190 (M-C3H9, 1%); 177 (M-C4H9, 35%); 161

(M-C3H12, 100%); 133 (M-C5H18, 9%); 105 (M-OC6H18, 7%); 91 (M-OC7H20, 9%);77 (M-OC8H22,

5%);

Characterisation of 2,2,6-trimethylchroman

O

12

3

4

56

78

9

10

11

12

1H NMR (400 MHz, CDCl3): δ=1.21 (s, 6H, H-1 and H-3); δ=1.64-1.68 (t, 2H, J=6.8, H-4);

δ=2.14 (s, 3H, H-12); δ=2.60-2.63 (t, 2H, J=6.8, H-5); δ=6.57-6.78 (m, 3H, H-7,H-9 and H-10)

13C NMR (400 MHz, CDCl3): δ=20.5 (C-12); δ=22.5 (C-5); δ=26.9 (C-1 and C-3); δ=33.0 (C-4);

δ=73.9 (C-2); δ=117.0 (C-10); δ=120.6 (C-6); δ=127.9 (C-7); δ=128.7 (C-9); δ=129.8 (C-8);

δ=151.8 (C-11)

m/z (GC-MS): 176 (M+, 55%); 161 (M-CH3, 33%); 147 (M-C2H6, 8%); 133 (M-C3H9, 8%); 121(M-

OC3H9, 100%); 105 (M-OC4H9, 8%); 91 (M-OC5H11, 26%); 77 (M+-OC6H13, 10%)

Page 193: Applications of metal triflates and assisted acids as

173

Characterisation of 2,2-dimethylchroman

11O

12

3

4

56

78

9

10

1H NMR (400 MHz, CDCl3): δ=1.32 (s, 6H,H-1 and H-3); δ=1.77-1.80 (t, 2H, J=6.8, H-4);

δ=2.75-2.78 (t, 2H,J=6.8, H-5); δ=6.76-7.09 (m, 4H, H-7 to H-10)

13C NMR (400 MHz, CDCl3): δ=22.7 (C-5); δ=27.2 (C-1 and C-3); δ=33.1 (C-4); δ=74.3 (C-2);

δ=117.5 (C-10); δ=119.8 (C-8); δ=121.1 (C-6); 127.6 (C-7); δ=129.6 (C-9); δ=154.3 (C-11)

m/z (GC-MS): 162 (M+,55%); 147 (M-CH3, 54%); 133 (M-C2H6, 11%); 119 (M-OC2H6, 11%); 107

(M-OC3H6, 100%); 91 (M-OC3H8, 16%); 77 (M-OC4H10, 18%)

Characterisation of 2,2-dimethyl-3.4-dihydro-2H-benzochomene

O

12

3

4

56

78

910

11

1213

14

1H-NMR (400 MHz, CDCl3): δ=1.42 (s, 6H,H-1 and H-3); δ=1.88-1.91 (t, 2H, J=7.0, H-4);

δ=2.86-2.89 (t, 2H, J=6.6, H-5); δ=7.13-7.74 (m, 6H, H-7 to H-13)

Page 194: Applications of metal triflates and assisted acids as

174

Characterisation of 3,3,8,8-tetramethyl-1,2,3,8,9,10-hexahydopyrano[3,2-f]chromene

O

O

12

3

4

567

8

9

101112

1314

1516

1H NMR (400 MHz, CDCl3): δ=1.30 (s, 12H,H-1,H-3,H-11 and H-9); δ=1.78-1.82 (t, 4H, J=6.8,

H-4 and H-12); δ=2.54-2.58 (t, 4H,J=6.8, H-5 and H-13); δ=6.58 (s, 2H, H-7 and H-15)

13C NMR (400 MHz, CDCl3): δ=20.4 (C-4 and C-12); δ=26.8 (C-1,C-3,C-9 and C-11); δ=33.1 (C-

5 and C-13); δ=72.7 (C-2 and C-10); δ=116.2 (C-7 and C-15); δ=119.2 (C-6 and C-14); δ=147.2

(C-8 and C-16)

m/z (GC-MS): 246 (M+, 36%); 231 (M-CH3, 5%); 191 (M-C4H12, 77%); 175 (M-OC4H12, 19%);

161(M-O2C4H12, 13%); 147 (M-O2C5H12, 10%); 133 (M-O2C6H12, 4%); 121 (M-O2C7H14, 5%); 91

(M-O2C9H18, 10%); 77 (M-O2C10H20, 9%)

Page 195: Applications of metal triflates and assisted acids as

175

Characterisation of 2,2,7,7-tetramethyl-2,3,4,7,8,9-hexahydropyrano[2,3-g]chromene

13

11

128

9

14

1516

12

3

4

56

7

10

O

O

1H NMR (400 MHz, CDCl3): δ=1.30 (s, 12H; H-1, H-3, H-10 and H-12); δ=1.73-1.77 (t, 4H,

J=6.8, H-4 and H-9); δ=2.67-2.71 (t, 4H, J=6.8, H-5 and H-8); δ=6.49 (s, 2H, H-14 and H-15)

13C NMR (400 MHz, CDCl3): δ=22.4 (C-5 and C-8); δ=27.1 (C-1; C-3;C-10 and C-12 ); δ=32.9

(C-4 and C-9); δ=73.6 (C-2 and C-11); δ=116.6 (C-14 and C-15); δ=119.9 (C-6 and C-7);

δ=147.2 (C-13 and C-16)

m/z (GC-MS): 246 (M+, 100%); 231 (M-CH3, 3%); 191 (M-C4H12, 74%); 175 (M-OC4H12, 21%);

161(M-O2C4H12, 14%); 147 (M-O2C5H12, 13%); 133 (M-O2C6H12, 4%); 121 (M-O2C7H14, 5%); 91

(M-O2C9H18, 20%); 77 (M-O2C10H20, 8%)

Page 196: Applications of metal triflates and assisted acids as

176

Characterisation of 2,2,3,6-tertramethylchroman

O

1 2 3

4 5

67

89

10

11

1213

1H NMR (400 MHz, CDCl3): δ= 0.91-0.93 (d, 3H,J=6.8, H-5); δ=1.07 (s, 3H, H-3); δ=1.28 (s, 3H, H-

1); δ=1.70-1.87 (m, 1H, H-4); δ=2.16 (s, 3H, H-10); δ=2.31-2.66 (m, 2H, H-6); δ=6.59-6.81 (m, 3H,

H-Ph).

13C NMR (400 MHz, CDCl3): δ=15.8 (C-5); δ=19.1 (C-10); δ=19.5 (C-4); δ=26.4 (C-3); δ=30.1 (C-

1); δ=34.6 (C-6); δ=90.9 (C-2); δ=115.7 (C-12); δ=120.2 (C-7); δ=126.8 (C-8); δ=127.7 (C-9);

δ=128.6 (C-11); δ=150.4 (C-13)

m/z (GC-MS): 190 (M+, 30%); 176 (M-CH3, 55%); 161 (M-C2H6, 9%); 147 (M-C3H9, 66 %);133 (M-

C4H12, 7%); 121 (M-OC4H12,100%); 105 (M-OC5H12, 6%); 91 (M-OC7H13), 25%); 77 (M-OC8H15,

11%)

Page 197: Applications of metal triflates and assisted acids as

177

Characterisation of 2-isoporpyl-2,5-dimethyl-2,3-dihydro-1-benzofuran

O

1

2

34

5

6

78

9

10

11

1213

1H NMR (400 MHz, CDCl3): δ= 0.93-0.94 (d,3H,J=6.8,H-3); δ=0.98-1.00 (d, 3H, J=6.8, H-1);

δ=1.33 (s, 3H, H-5); δ=1.96-2.02 (q, 1H, J=6.8, H-2); δ=2.26 (s, 3H, H-10); δ=2.73-3.12 (m, 2H,

H-6) and δ=6.61 (m, 3H, H-Ph).

13C NMR (400 MHz, CDCl3): δ=17.4 (C-1 and C-3); δ=20.8 (C-10); δ=23.3 (C-5); δ=37.3 (C-2);

δ=38.8 (C-6); δ=91.7 (C-4): δ=108.9 (C-12); δ=125.7 (C-8); δ=127.0 (C-7); δ=128.4 (C-11);

δ=129.2 (C-9); δ=157.2 (C-13).

m/z (GC-MS): 190 (M+, 37%); 175 (M-CH3, 10%); 161 (M-C2H6, 5%); 147 (M-C3H9, 100%); 133

(M-C4H12, 7%); 121 (M-C5H15), 46%); 105 (M-OC5H15, 7%); 91 (M-OC6H15, 15%); 77(M-OC7H18,

6%).

Page 198: Applications of metal triflates and assisted acids as

178

Characterisation of 8-tert-butyl-2,2,3,6-tetramethylchroman

O4

12

3

5

67

89

10

11

12

1314

1516

17

1H NMR (400 MHz, CDCl3): δ=0.98-0.99 (d,3H,J=6.8,H-4); δ= 1.15 (s,3H,H-1); δ=1.37 (s, 9H,H-

14 to H-16); δ=1.40 (s,3H,H-3); δ=1.77-1.86 (m, 2H,H-6); δ=2,24 (s,3H,H-10); δ=6.63 (s, 1H, H-

8) and δ=6.71 (s, 1H,H-11)

13C NMR (400 MHz, CDCl3): δ=16.6 (C4); δ=20.0 (C-10); δ= 20.9 (C-3); δ=27.6 (C-1); δ=29.9

(C-14 to C-16); δ= 31.7 (C-13); δ=34.7 (C-6); δ=35.4 (C-5); δ=90.6 (C-2); δ=121.5 (C-7);

δ=125.1 (C-11); δ=127.6 (C-8); δ=1338.0 (C-9 and C-12); δ=150.0 (C-17)

m/z (GC-MS): 246 (M+, 50%); 231 (M-CH3, 23%); 217 (M-C2H6, 1%); 203 (M-C3H9, 24%); 189

(M-C4H9, 3%); 177 (M-C3H12, 71%); 161 (M-C4H15, 100%); 147 (M-C5H18, 23%); 133 (M-C6H21,

14%); 121 (M-OC6H21, 10%) ; (105 (M-OC7H21, 9%); 91 (M-OC8H23, 15%); 77 (M-OC9H25, 15%)

Page 199: Applications of metal triflates and assisted acids as

179

Characterisation of 7-tert-butyl-2-isopropyl-2,5-dimethyl-2,3-dihydro-1-benzofuran

O

1

23

4

5

67

8

910

11

1214

1516

1713

1H NMR (400 MHz, CDCl3): δ= 0.94-0.95 (d,3H,J= 6.84,H-1); δ= 0.99-1.01 (d, 3H,J=6.6, H-3);

δ=1.30 (s,3H, H-5); δ=1.34 (s,9H,H-14 to H-16); δ=1.95-2.02 (q, 1H,J=6.84, H-2); δ=2.27 (s, 3H,

H-10); δ=2.96-3.08 (m,2H, H-6); δ=6.80 (s,1H, H-11); δ=6.85 (s, 1H,H-8)

13C NMR (400 MHz, CDCl3): δ= 17.5 (C-1 and C-3); δ=21.1 (C-10); δ=23.4 (C-4); δ=29.0 (C-14,

C-15 and C-16); δ=34.0 (C-2); δ=37.0 (C-6); δ=38.9 (C-13); 90.6 (C-5); δ=123.2 (C-11);

δ=125.0 (C-8); δ=127.2 (C-9); δ= 128.5 (C-7); δ= 132.1 (C-12); δ=154.9 (C-17)

m/z (GC-MS): 246 (M+, 60%); 231(M-CH3, 16%); 203 (M-C3H9, 50%); 177 (M-C3H12, 21%); 161

(M-C4H15, 100%); 147 (M-C5H18, 51%); 133 (M-C6H21, 14%); 91(M-OC8H23, 14%); 77(M-OC9H25,

7%)

Synthesis of 7-tert-butyl-2-isopropyl-2,methyl-2,3-dihydro-1-benzofuran

1

2

56

7

8

910

11

1213

1415

16O 3

4

1H NMR (400 MHz, CDCl3): δ= 0.95-0.96 (d,3H,J= 6.8,H-1); δ= 1.00-1.02 (d, 3H,J=6.84, H-3);

δ=1.32 (s,3H, H-4); δ=1.97-2.04 (q, 1H,J=6.8, H-2); δ=2.38-3.12 (m,2H, H-6); δ=6.72-7.05

(m,3H, H8-H10).

Page 200: Applications of metal triflates and assisted acids as

180

13C NMR (400 MHz, CDCl3): δ= 17.5 (C-4); δ=21.6 (C-2); δ=23.2 (C-6); δ=29.4 (C-13 to C-15);

δ=34.0 (C-12); δ=37.2 (C-1); δ=38.9 (C-3); δ=90.6 (C-5); δ=119.3 (C-9); δ=122.9 (C-8); δ=124.3

(C-10); δ= 127.4 (C-7); δ= 132.5 (C-11); δ=157.2 (C-16).

m/z (GC-MS): 232 (M+, 49%); 217 (M-CH3, 16%); 189 (M-C3H9, 74%); 173 (M-C4H9, 7%); 163

(M-C5H12, 15%); 147 (M-C6H15, 100%); 133 (M-C7H15, 29%); 119 (M-C8H18, 14%); 105 (M-

OC8H18, 9%); 91 (M-OC9H18, 19%); 77 (M-OC10H20, 9%).

Page 201: Applications of metal triflates and assisted acids as

181

References

1. Arrhenius, S. A. Z.; Phys. Chem. 1887, 1, 631.

2. Bronsted, J. N.; Recl. Tav. Chim. Pays Bas, 1923, 42, 718.

3. (a) Lowry, T. M.; Chim. Ind. (London), 1923, 42, 43 (b) Lowry, T.M.; Trans. Faraday Soc.

1924, 20,13

4. Lewis, G. N.; Valency and Structure of atoms and Molecules; Wiley: New York, 1923. 5. Kazansky, V. B.; Topics Catal., 2000, 11/12, 55

6. Hammett, L. P.; Physical Organic Chemistry; McGraw-Hill: New York, 1940.

7. Hall, N. F., Conant, J. B., J. Am. Chem. Soc. 1927, 49, 3047.

8. Gillespie, R.J.; Acc. Chem. Res. 1968,1, 202.

9. (a). Olah, G. A.; Prakash, G. K.; Sommer, J., Superacids, Wiley: New York, 1985. (b). Olah,

G. A.; J. Org. Chem. 2005, 70, 2413. (c). Olah, G. A.; J. Org. Chem., 2001, 66, 5943.

10. (a). Reed, C.A.; Kim, K.C.; Bolskar, R.D.; Mueller, L., Science, 2000, 289, 101. (b). Reed,

C.A., Chemm. Commun., 2005, 1669.

11. (a) Olah, G.A.; Prakash, G.K.; Sommer, J. Superacids; Wiley: New York, 1985. (b) Olah,

G.A., J.Org. Chem. 2005, 70,2413. (c) Olah, G.A. J.Org. Chem. 2001, 66, 5943.

12. (A). Pearson, R.G.; J. Am. Chem. Soc., 1963, 85, 3533. (b). Pearson, R.G., Acc. Chem,

Res. 1993, 26, 250.

13. Esteve, P.M.; Ramirez-Solis, A.; Mota, C.J.; J. Am. Chem. Soc. 2002, 124, 2672.

14. Feller, A.; Lercher, J. A.; Adv. Catal. 2004, 48, 229.

15. Frank, H. G.; Stadelhofer, J.W.; Industrial Aromatic Chemistry; Springer-Verlag: Berlin,

Germany,1998.

16. Alson, A.C.; Ind. Eng. Chem. 1960, 52, 833.

17. Olah, G.A.; Nijima, M.; Kerekes, I. Synthesis 1973, 779, 780.

18. Olah G.A.; Mathew, T.; Geopprt, A.; Torok, BBusci, I; Li, X.Y.; Wang, Q.; Marinez, E.R.;

Batamack, P; Aniszfeld, R.; Surya Praksh, G.K. J. Am. Chem. Soc. 2005,127, 5964.

19. Nowak, F.M.; Himes, J.F.; Mehlberg, R.L.; Advances in hydrofluoric acid (HF) catalysed

alkylation , http//www. Uop.com/objects/NPRASpr2003HFAlkyl.pdf.

20. Evering, B. L. Adv. Catal. Relat. Subj. 1954, 6,197.

21. Zhang, C.; Adv. Catal. 2006, 40, 153.

22. R.W. Marshman Aldrichmica Acta. 1995, 28(3), 77.

Page 202: Applications of metal triflates and assisted acids as

182

23. (a) Kobayashi, S. Chem Lett. 1991, 2187. (b) Forsburg, J.H.; Spaziano, V.T.;

Balasubramanian, T.M.; Liu, G.K.; Kinsley, S.A.; Duckworth, C.A.; Poteruca, J.J.; Brown P.S.;

Miller J.L. J.Org. Chem. 1987, 52, 1017.

24. Kishida, T.; Yamauchi, T.; Kubota, Y.; Sugi, Y. Green Chem., 2001, 5, 57

25. Tsuruta, H.; Yamaguchi, K.; Imamoto, T. Chem. Commun. 1999, 1703.

26. Cotton, F.A.; Wilkinson, G. Advanced inorganic Chemistry, 5th ed.; Wileyy: New York, 1988;

973.

27. (a) Thom, K..F.; US Patent 3615169 1971; Chem Abstr.1972, 76, 5436a (b) Collins, S.;

Hong, Y.; Tetrahedron Lett.1987, 28, 4391.

28. Collins, S.; Hong, Y. Tetrahedron Lett. 1987, 28, 4391.

29. Waller, F.J.; Barrett, A.G.M.; Braddock, C.D.; Mckinnell, R.M.; Ramprasad, D. Perkin

Trans.1, 1999, 867.

30. Baes, C.F., Jr.; Mesmer, R.E. The Hydrolysis of Cations; John Wiley & Sons:New York,

1976, p129.

31. Mukaiyama, T.; Narasaka, K.; Banno, T. Chem Lett.1973, 1101.

32. Kobayashi, S. Chem. Lett. 1991, 2087.

33. Kobayashi, S.; Hashiya. I. Tetrahedron Lett. 1992, 33, 1625.

34. Kobayashi, S.; Hichiya, I.; Ishitani, H.; Araki, M. Synlett. 1993, 472.

35. Kobayashi, S.; Hachiya, I.; Takahori. T.; Araki, M.; Ishitani, H. Tetrahedron Lett., 1992,

33(45), 6815.

36. Fringuelli, F.; Taticchi, A., The Diels-Alder Reaction: Selected Practical Methods, Wiley &

Sons, 2002, 652.

37. Kobayashi, S,; Hachiya, I.; Araki, M.; Ishitani, H. Tetrahedron Lett. 1993, 34, 3755.

38. Oakes, R.S.; Heppenstall, T.J.; Shezad, N.; Clifford, A.A.; Rayner, C. Chem. Commun.,

1999, 1459.

39. Oakes, R.S.; Clifford, A.A.; Bartle, K.D.; Thornton,-Pett, M.; Rayner, C.M. Chem. Commun.,

1999, 247.

40. Kobayashi, S.; Tsuchiya, T.; Komoto, I.; Matsuo, J. J. Organomet. Chem. 2001, 624, 392.

41. Kobayashi, S.,Synlett., 1994, 689.

42. Rideout, D.C.; Breslow, R. J. Am. Chem. Soc.,1980, 102 , 7816.

43. Nagayama, S.; Kobayashi, S.; Angew. Che. Int. Ed. 2000, 39, 567

44. Vidis, A.; Ohlin, G.; Laurenczy, E.; Kusters, E.; Sedelmeier, G.; Dyson, P.J.; Adv. Syn.

Catal. 2005,347, 266.

Page 203: Applications of metal triflates and assisted acids as

183

45. Aggarwal, A.; Lancaster, N.L.; Sethi, A.R.; Welton, T. Green Chemistry, 2002, 4, 517

46. Kumar, A.; Sarma, D. Appl. Catal. A: General, 2008, 335, 1

47. Kobayashi, S.; Hachiya, I.; Takahori, T.; Araki, M.; Ishitani, H., Tetrahedron Lett., 1992, 33

(45), 6815.

48. Ohba, M.; Izuta, R.Heterocycles, 2001, 55, 823.

49. Narasaka, K.; Iwasawa, N.; Inoue, M.; Yamada, T.; ;Nakashima, M.; Sugiromi, J. J.Am

Chem Soc., 1989,111, 5340.

50. Maruoka, K.; Itoh, T.; Shirasaka, T.; Yamamoto, H. J. Am.Chem.Soc., 1988, 110, 310.

51. Kaufmann, D.; Boese, R.; Angew. Chem. Int. Ed., 1990, 111, 5439.

52. Kobayashi, S.; Hachiya, I.; Ishitani, H.; Araki, M. Tetrahedron Lett., 1993, 34(28), 4535.

53. Yamamoto, H.; Ishihara, K.: Acid Catalysis in Modern Organic Synthesis, Wiley-VCH,

Weinheim, 2007, 2, 652.

54. Kobayashi, S.; Ishitani, H.; Nagayama, S. Chem. Lett. 1995, 423.

55. Kobayashi, S.; Ishitani, H.; Nagayama, S. Synthesis, 1995, 1195.

56. Crousse, B.; Begue, J.-P.; Bonnet-Delpon, D. J. Org. Chem., 2000, 65, 5009.

57. Gadhwal, S.; Sandhu, J.S. J. Chem. Soc., Perkin Trans 2000, 2827.

58. Ishitani, H.; Kobayashi, S. Tetrahedron Lett. 1996, 37, 7357.

59. Friedel, C.; Crafts, J.M. Comptes Rendus, 1877, 84, 1392.

60. Kotsuki, H.; Ohishi,T.; Inoue, M. Synlett., 1998, 255.

61. Kotsuki, H.; Ohishi,T.; Inoue, M.; Kojima, T. Synthesis, 1999, 603.

62. Tsuchimoto, T.; Tobita, K.; Hiyama, T.; Fukuzawa, S. J.Org. Chem.,1997, 62, 6997.

63. Mine, N.; Fujiwara, Y.; Taniguchi, H.; Chem. Lett., 1986, 375.

64. Olah, G.A. US. Pat. 4721559,1988.

65. Olah, G.A.; Flood, S.H.; Kuhn, S.J.; Moffatt, M.E.; Overchuck, N.A. J. Am. Chem. Soc.,1964,

86, 1046.

66. Sinha, S.; Mandal, B.; Chandrasekaran, S. Tetrahedron Lett., 2000, 41, 9109.

67. Song, C.E.; Shim, W.H.; Roh, E.J.; Choi, J.H. Chem. Commun., 2000,1595.

68. Chen, D.; Yu, L.; Wang, P.G. Tetrahedron Lett. 1996, 37, 4467.

69. Xie, W, Bloomfield, K.M.; Jin, Y.; Dolney, N.Y.; Wang, P.G. Syn. Lett., 1999, 498.

70. Zhang, W.; Wang, P.G.; J. Org. Chem., 2000,65, 4732.

71. Huang, T.; Li, C.-J.; Tetrahedron Lett. 2000,41, 6715.

72. Hachiya, I.;Moriwaki, M.; Kobayashi, S. Bull. Chem. Soc. Jpn.,1995, 68, 2053.

73. Leffingwell, J.C.; Shackelford, R.E. Cosmetics Perfumery, 1974, 89(6), 69.

Page 204: Applications of metal triflates and assisted acids as

184

74. Hopp, R. Recent Adv. Tobacco Sci. 1993, 19, 46.

75. Garvey, B.S.,Jr., Sarback, D.V. Ind. Eng. Chem., 1942, 34, 1314.

76. Milos, M.; Mateli, I.; Jerkovic, I. Food Chem., 2000, 71, 79.

77. Yanishlieva, N.V.; Marinova, E.M.; Gordon,M.H.; Raneva, V.G. Food Chem., 1999, 64, 59.

78. Weinrich, W. J. Ind. Eng. Chem., 1943, 35(3), 264

79. Koenig, Ber., 1891, 24(179), 3889

80. Niederl, Natelson, J. Am. Chem. Soc.,1931, 53, 272.

81. Chaudhuri, B.; Patwardhan, A.A.; Sharma, M.M., Ind. Eng. Chem. Res.,1990, 29, 1025.

82. Santacesaria, E.; Silvani, R.; Wilkinson, P.; Carrà, S. Ind. Eng. Chem. Res., 1988, 27, 541.

83. Santacesaria, E.; Di Serio, D.; Tesser, R.; Cammarota, F., Ind. Eng. Chem. Res., 2005, 44,

9473.

84. Hino, A.; Arata, K. J. Chem. Soc., Chem. Commun., 1979, 245, 1148.

85. Yadav, G.D.; Thorat, S.T., Ind Eng.Chem. Res., 1996, 35, 721.

86. Yadav, G.D.; Ramesh, P. Can. J. Chem. Eng. 2000, 78(5), 917.

87. Devassy, B.; Shanbhab, G.V.; Lefebvre, F.; Halligudi, S.B., J. Mol. Catal. A: Chemical, 2004,

210, 125.

88. Sarish, S.; Devassy, B.M.; Halligudi, S.B., J. Mol. Catal. A: Chemical, 2005, 235, 44.

89. Shinde, A.B.; Shrigadi, N.B.; Samat, S.D. Appl. Catal. A: General, 2004, 276, 5.

90. Yadav, G,D,; Pujari, A.A.; Joshi, A.V. Green Chemistry, 1999, 269.

91. Yadav, G.D.; Murkute, A.D. J. Catal., 2004, 224, 218.

92. Umamaheswari, V.; Palanichamy, M.; Murugesan, V. J. Catal., 2002, 210, 367.

93. Wetzel, W.H.; Nelson, H.G.; Shelton, F.J. US Pat.3932537. 1976.

94. Hess, M. US Pat.3534111,1970.

95. (a) Stevens, D.R.; McKinley, J.B. US Pat.2206924, 1940.

96. Stevens, D.R.; Bowman, R.S. US Pat.2560666, 1951.

97. Gershanov, F.B.; Liakumovich, A.G.; Michurov, J.I.; Pantukh, B.I.; Rutman, G.I.; Sobolev,

V.M.; Grinberg, A.A.; Gurvich, Y.A.; Zakharova, N.V.; Nafikova, A.M. US Pat.3049086, 1976.

98. Dodd, J.R., US Pat.4236031, 1980.

99. Stevens, D.R.; Livingstone, J.C. US Pat.2297588,1942.

100. Hervert, G.L. US Pat.3929912,1975.

101. Kurek, P.R. US Pat.4380677, 1983.

102. Luten, D. B., Jr.; De Benedictis, A. US Pat. 2435087, 1948.

Page 205: Applications of metal triflates and assisted acids as

185

103. Kroupa, J.; Podstata, J.; Matous, V.; Nechybova, V.R. Czech Pat.265262, 1990, (Chem.

Abstr., 1992, 117, 48093).

104. Sartori, G.; Bigi, F.; Casiraghi, G.; Casnari, G.; Cheisi, L.; Arduini, A. Chem. Ind., 1985, 22,

726.

105. Noburu, O.; Tetsuya, Y.; Kamimura, A.; Aritsune, K. J, Chem. Soc., Chem. Commun.,

1986, 16, 1285.

106. Fernholz, H.; Schimdt, H.J. Angew. Chem., 1969, 8, 521.

107. Rajadaksha, R.A.; Chaudari, D.D. Ind Eng. Chem. Res., 1987, 26, 1276.

108. Campbell, D.B.; Onopchenko, A.; Young, D.C. Ind. Eng. Chem. Res., 1990, 24, 642.

109. Saito, T.; Ito, T.; Nishizawa, H. Jap. Pat. 0782197, 1995.

110. Yoo, J.W.; Lee C.W.; Park, S.–E.; Ko, J. Appl. Catal. A, 1999, 187, 225

111. Akoi, T.; Sawa, K.Jap. Pat. 04273838,1992.

112. Yadav, G.D.; Goel, P.K.; Joshi, A.V.E., Green Chem., 2001, 3, 92.

113. Licence, P.; Carter, D.N.; Gomez, R.A.; Ross, S.K.; Polliakoff, M. Contenoeus Reactions in

Supercritical CO2.

114. Pereira, C.C.M.; de la Cruz, M.H.C.; Lachter, E.R. J. Braz. Chem. Soc., 2010, 21(2), 367.

115. Pereira, C.C.M.; Lachter, E.R., Appl. Catal. A, 2004, 266, 67.

116. De la Cruz, M.C.H.; Abdel-Rehim, M.A.; Rocha, A.S.; da Silva, J.F.C.; da Costa Faro Jr,

A.; Lachter, A.R., Catal. Commun., 2007, 8(11), 1650.

117. Kretchmer, R.A.; McCloskey, M.B. J. Org. Chem. , 1989, 37(12), 1972.

118. Amandi, R.; Licence, P.; Ross, S.K.; Aaltonen, O.; Poliakoff, M., Org. Proc. Res. & Dev.,

2005, 9, 451.

119. Cativiela, C.; Garcia, J. I.; Garcia-Matres, M.; Mayoral, J.A.; Figueras, F.; Fraile, J.M.;

Cseri, T; Chiche, B., Appl. Catal., 1995, 123 (2), 273.

120. Kamala, P.; Pandurangan, A., Catalysis Lett., 2006, 110 (1-2), 39.

121. Viktorova, E.A.; Shuikin, N.I.; Korosteleva, G.S.; Baranova, N,G., Russian Chem. Bull.,

1981, 10(8), 1415.

122. Xiong, Y.; Wu, M.M.; Young, L.B.; Chu, A.S., US Pat. 5750480, 1988.

123. Qiaoxia, G,; Jisheng, Y.; Liwei, C. Huaxue Yajiu Yu Yingyong, (Chemical Abstr.), 2008,

20(1),30

124. Sawant, D.P.; Devassy, B.M.; Halligudi, S.B.,J. Mol. Catal. A: Chemical, 2004, 217, 211.

125. Mandelcorn, L.; Miller, R.L.; Mercier, G.E.; Pickett, J.H.; Dakin, T.W., US 4054937, 1977.

126. Coleman, G.H.; Perkins, R.P. US Pat. 2170809, 1939.

Page 206: Applications of metal triflates and assisted acids as

186

127. Klingel, H.; Ellison, F.E. US Pat. 0020828, 1979.

128. Wu, M.M.; Lapierre, R.B. US Pat. 6407292 B1, 2002.

129. Rudnick, L.R.; Kremer, R.A.; Law, D. A. US Pat. 5552071, 1996.

130. McGee, W.D.; Sall, E.D.; Smith, R.G., Jr. WO 2000018707 A1, 2000.

131. Grigierev, V.V.; Prokofev, K.V.; Bazylyuk, T.A., Khimiya i Tekhnologiya Topoliv i Masel,

1991, 6, 21.

132. Cobb, R.L.; Mitchell, M.D. US Pat. 4668296, 1987.

133. Yu, K; Zhao, X.; Yang, X.; LI, J.; Hu, G.; Yang, M.; Shuai, H.; Zhang, D.; Wang, J.; Dou, Y.

CN 101503379 A 20090812, 2009 (Chemical Abstr).

134. Li, D.; Wei, W.; Li, R.; Chang, H.; Bai, Y. CN 1915969 A 20070221, 2007, (Chem. Abstr.).

135. Peng, P.; Xi, L.; Duan, Q. CN 101081806 A 20071205, 2007, (Chemical Abstr.)

136. Huang, Y.; Feng, Y.; Zhao, S.; An, J.; Yu, J. Shiyou Huagong, 2005, 34(7), 622, (Chem.

Abstr.).

137. Piao, L.Y.; Fu, X.; Yang, Y. L.; Tao, G.H.; Kou, Y. Catal. Today, 2004, 93-95, 301.

138. Yadav, G.D.; Sonali, S. Org. Proc. Res. & Dev., 2002, 6, 256.

139. Yadav, G.D.; Kundu, B. Can. J. Chem. Eng., 2001, 79(5), 805.

140. Hasebe, R.; Okuda, N.; Saito, N., JP 06016573 A, 1994.

141. Rennie, W.M.O.; Kenedy, M.H., DE 2132568, 1972.

142. Isagulyants, V.I.; Panidi, E.V., Zhurnal Priklandnoi Khimi, 1961, 34, 1578.

143. Rihko, L.K.; Krause, A.O.I., Ind. Eng. Chem. Res., 1995, 34, 1172.

144. Rihko, L.K.; Krause, A.O.I., Appl. Catal. A:Gen., 1993, 101, 283.

145. Ignatius, J.A.; Jarvelin, H.; Lindqvist, P., Hydrocarbon Process., 1995, 74, 51.

146. Boz, N.; Dogu, T.; Murtezaoglu, K.; Dogu, G., Catal. Today, 2005, (100), 419.

147. Girolamo, M. D.; Lami, M.; Marchionna, M.; Pescarollo, E.; Tagliabue, L; Ancilloti, F., Ind.

Eng. Chem. Res., 1997, 36, 4452.

148. Wang, S.; Guin, J. A.; Fuel Pocess. Technol., 2003, 84, 135.

149. Wang, S.; Guin, J., Energy & Fuel, 2001, 15, 666.

150. Liu, J.; Wang, S.; Guin, W.J.; Fuel Procces. Technol., 2001, 69, 205.

151. Vaisli, D.; Dogu, T., Ind. Eng. Chem. Res., 2005, 44, 527.

152. Paakkonen, P.K.; Krause, A.O., Rective & Functional Polym., 2003, 55, 139.

153. Wang, S.; Guin, J., Chem. Commun., 2000, 2499.

154. Horvath, T.; Seiler, M.; Hunger, M., Appl. Catal. A: General, 2000, 193, 227.

155. Chu, P.; Kuhl, G., Ind. Eng. Chem. Res.,1987, 26, 365.

Page 207: Applications of metal triflates and assisted acids as

187

156. Chang, K.; Kim, G.; Ahn, W., Ind. Eng. Chem. Res.,1992, 31, 125.

157. Tejero, J.; Creus, E.; Iborra, M.; Cunill, F.; Izquierdo, J.F.; Fite, C., Catal. Today, 2001, 65,

381.

158. Colligon, F.; Poncelet, G., J. Catal., 2001, 202, 68.

159. Patrini, R.; Lami, M.; Marchionna, M.; Benvenuti, F.; Galletti, A.M.R.; Sbrana, G., J. Mol.

Catal A: Chem., 1998, 129, 179.

160. Benvenuti, F.; Carlini, C.; Marchionna, M.; Patrini, R.; Galletti, A.M.R.; Sbrana, G., J. Mol.

Catal. A: Chem., 1999, 139, 177.

161. Snelling, J.; Curtis, C.W.; Park, Y., Fuel Process. Technol., 2003, 83, 219.

162. de Klerk, A., Ind. Eng. Chem. Res., 2004, 43, 6349.

163. Shikata, S.; Nakata, S.; Okuhara, T.; Misono, M., J.Catal., 1997, 166, 263.

164. Ono, Y.; Baba, T. 8th Intern. Congr. Catal., Berlin, 1984, 5, 405.

165. Igarashi, A.; Matsuda, T.; Ogino, Y. Sekiyu Gakkaishi, 1994, 37, 632.

166. Schneider, C., Mol. Nitur, Food Res., 2005, 49, 7.

167. Middelton, E., Jr.; Kandaswami, C.; Theoharides, T.C.; Pharmacol. Rev., 2000, 52, 673.

168. Waring, M.A. J. Chem. Soc. Perkin Trans. I 1990, 2979,

169. Bergman, R.; Harting, J.; Lues, L. Schittenhelm, C.; Gericke, R. J. Med. Chem., 1990, 34,

3074.

170. Kashman, Y.; Gustafson, K.R.; Fuller, R.W.; Cardellina II, J.H.; McMahon, J.B.; Currens,

M.J.; Buckheit, Jr., Hughes, S.H.; Cragg, G.M.; Boyd, M.R. J. Med. Chem. 1992, 35, 2735.

171. Ahluwalia, V.K.; Arora, K. K.,Tetrahedron, 1981, 37, 1437.

172. Matsui, M.; Yamamoto, H., Bull. Chem. Soc. Jpn., 1995, 68, 2663.

173. Ahluwalia, V.K.; Krishnan, K.A.; Ravinder, S. J., J. Chem. Soc. Perkin Trans. 1, 1982, 335

174. Kalena, G.P.; Jain, A.; Banerji, A., Molecules, 1997,2,100.

175. Malkov, A.V.; Spoor, P.; Vinader, V.; Kočovský, P., J. Org. Chem., 1999, 64, 5308.

176. Dewhirst, K.C.; Rust, F.F., J.Org.Chem.,1962, 28, 798.

177. Bader, A.R.; Bean, W.C. J. Am. Chem. Soc.,1958, 80, 3037.

178. Youn, S.W.; Eom, J.I., J.Org. Chem., 2006, 71, 6705.

179. Adrio, L. A.; Hii, K.K., Chem. Commun., 2008, 2325,

180. Yamamoto, Y.; Itonaga, K., Org. Lett., 2009, 11(3), 717.

181. Corey, E.J.; Shibata, T.; Lee, T.W., J. Am. Chem. Soc., 2002, 124, 3808.

182. Liang, J.J.-N.. PhD Thesis; McMaster University, 1976. sited in O’Donnell , T.A.. Chem.

Soc. Rev. 1987, 16(1),6.

Page 208: Applications of metal triflates and assisted acids as

188

183. Olah, G.A.; Surya Prakash, Sommer, J. Superacids; Wiley New York, 1985.

184. Smith, G.P.; Dworkin, A.S.; Pagni, R.M.; Zingg, S.P.; J. Chem. Soc. 1989,111,525.

185. Stefanidakis, G.; Gwin, J.E.; In Chemical Proccessing handbook; McKetta J.J., Ed.;

Dekker: New York, 1993; p 80.

186. Fox, M-A; Whitesell, J.K.; Organic Chemistry, Jones and Bartlett Publishers, Boston,

London, 1994, p 271.

187. Lemechko, P.; Fanny, G.; Antiniotti, S.; Dunah, E.;Tetrahedron Lett., 2007, 48(33), 5731.

188. Rosenfeld, D.C.; Shekhar, S.; Takemiya, A.; Utsunomiya, M.; Hartwig, J.F.; Org. Lett.,

2006, 8(19), 4179.

189. Cowely, M.; Energy & Fuels, 2006, 20, 1771.

190. Wang, S.; Guin, J.A.; Energy & Fuels, 2001, 15, 666.

191. Rhodes, A.; J. Oil Gas, 1999, 97, 39.

192. Wang, S.; Guin, J.A.; Fuel Process. Technol., 2003, 84, 135.

193. Collignon, F.; Poncelet, G.; J. Catal, 2001, 202, 68.

194. Boz, N.; Dogu, T.; Murtezaoglu, K.; Dogu. G.; Catal. Today 2005, 100, 419.

195. Snelling, J.; Curtis, C.W.; Park, Y.; Fuel Process. Technol. 2003, 83, 219.

196. Verevkin, S.P.; Krasnykh, E.L.; Vasitsova, T.V.; Heintz, A.; J. Chem. Eng. Data, 2003, 48,

591.

197. Klepáčová, K.; Mravec, D.; Kazonyi, A.; Bajus, M.; Applied Catal. A: General, 2007, 328, 1.

198. Syed, F.H.; Egleston, C.; Datta, R.; J. Chem. Eng. Data, 2000 45. 319.

199. Katariya, A.M.; Moudgalya, K.M.; Mahajani, S.M.; Ind. Eng. Chem. Res. 2006, 45, 4233.

200. Wang, S.; Guin, J.A.; Chem. Commun., 2000, 2499.

201. Wang, S.; Guin, J.A.; Energy & Fuels, 2001, 15, 666.

202. Poncelet, G.; Collignon, F.; J. Catal, 2001, 202, 68.

203. Terblans, Y.; Huyser, M.; Huyser, J.J.; Green, M.J.; Young, D.A.; Sibiya, M.S.; Can. J.

Chem., 2005, 83 (6-7), 854.

204. Ishitani, H.; Kobayashi, S.; Tetrahedron Lett., 1996, 37, 7357.

205. Williams, D.B.G.; Lawton, M.; Org. Biomol. Chem. 2005, 3, 3269.

206. Mitamura, K.S.; Kobayashi, S.; Chem. Commun., 1996, 183.

207. Williams, D.B.G., Lawton, M.; Green Chem.,2008, 10, 914.

208. Williams, D.B.G.; Shaw, M.L.; Green, M.J.; Holzapfel, C.W.; Angew. Chem. Int. Ed. 2008,

47, 560.

209. Weike, S.; Jianjun, L.; J. Chem. Res. 2008, 2004 (8); 552.

Page 209: Applications of metal triflates and assisted acids as

189

210. Williams, D.B.G.; Lawton, M.; Tetrahedron Lett., 2006, 47,6657.

211. Kobayashi, S.; Manabe, K.; Pure Appl. Chem. 2000, 72(7), 1373.

212. Coulombel, L.; Rajzmann, M.; Pons, J. -M.; Olivero, S.; Duñach, E.; Chem. Eur. J. 2006,

12, 6356.

213. Tables of the laboratory, Merck.

214. Waller, F.J.; Barrett, G.M.; Braddock, D.C.; McKinnell, R.M.; Ramprasad, D.; J. Chem.Soc.

Perkin Trans.1, 1999,867.

215. Frost, C.G.; Hartley, J.P.; Griffin, D.; Tetrahedron Lett., 2002, 43(27), 4789.

216. CRC Handbook of Chemistry and Physics, 56th Ed., 1975-1976.

217. Atkins, P.W.; Physical Chemistry, 6th Ed. Oxford University Press, Melbourne, 1998.

218. Rihko, L.K.; Kivirananta-Pääkkönen, P.K.; Krause, O.I.; Ind. Eng. Chem. Res., 1997, 36,

614.

219. Yamamoto, H.; Futatsugi, K.: Angew Chem Int. Ed., 1924, 44, 42.

220. Waller, F.J.; Barrett, G.M.; Braddock, D.C.; Ramprasad, D.; Chem. Commun., 1997, 613.

221. Bruice, P.Y.; Organic Chemistry, Second Ed., Prentice Hall International, New Jersey,

1998, p610.

222. Cotton, F.A.; Wilkonson, G.; Advanced Inorganic Chemistry, Wiley, New York, 4th

Ed.1980, p173, 810,1004.

223. Barrett, A.G.M.; Braddock, D.C. Chem Commun., 1997, 351.

224 Barrett, A.G.M.; Braddock, C.D.; Henschke, J.P.; Walker, E.R.; J. Chem. Soc., Perkin

Trans.1, 1999, 873.

225. Sadasivam, D.V.; Antharjanam, P.K.A.; Prasad, E.; Flowers, R.A. J.Am. Chem. Soc., 130,

2008, 7228.

226. Prinsloo, N.M.; Fuel Proc. Technol., 2006, 87,437.

227. Villadsen, J.; Livbjerg, H.; Catal. Rev., Sci. Eng.17 (2), 1978, 203.

228. De Klerk, A.; Ind. Eng. Chem. Res., 2004, 43, 6349.

229. De Klerk, A.; Leckel, D.O.; Prinsloo, N.M. Ind. Eng. Chem. Res., 2006, 45, 6127.

230. (a) Everson, R.C.; Jansen, W. J.; Chem. Eng. Data, 2001,46(2), 247. (b) Di Girolamo,

M.P.; Lami, M.; Marchionna, M.; Pescarollo, E.; Tagliabue, L.; Ancillotti, F.; Ind. Eng. Chem.

Res., 1997, 36, 4452.

231. Friedel, C.; Crafts, J.M. Comptes Rendus, 1877, 84, 1392.

232. Mahindaratne, M.; Wimalasena, K. J. Org. Chem. 1998, 63, 2858.

233. Khalaf, A.A.; Albar, H.A.; Bahhafi, S.O. Revue Rouamine de Chimie, 2006, 51(9), 895.

Page 210: Applications of metal triflates and assisted acids as

190

234. Olah, G.; Torok, B.; Prakash, S.; Appl. Catal.,A 1996, 146, 107

235. Singh, R.P.; Singh, V.K.; Tetrahedron 2000, 57, 241

236. Bachmann, W.E.; Johnson, J.R.; Fieser, L.F.; Snyder, H.R., Organic Reactions, (Volume

III), John Wiley & Sons, INC., London, 1946, p2.

237. Kobayashi, S.; Suguira, M.; Kitawaga, H.; Lam, W.W.L. Chem Rev. 2002, 102, 2227.

238. Olah, G.A.; Farooq, O.; Farnia, S.M.F.; Olah, J.A.; J. Am. Chem. Soc.,1988, 110, 2560.

239. Olah, G.A. US Pat. 4271559,1988. 240. Song, R.; Gaffney, A.M.; Yeh, C.Y.; Angevine, P.J., US Pat. 20090036721.

241. Chen, C-W; WU, W-J; Zeng, X-S; Jiang, Z-H; LI SHI, Ind. Eng. Chem. Res., 2009, 48(23),

10359.

242. Preethi, M.E.L.; Sivakumar, T.; Palanichami, M. Catal. Commun., 2010, 11(10), 876.

243. Amandi, R.; Licence, P.; Ross, S.K.; Aaltonen, O.; Poliakoff, M. Org. Proc. Develop., 2005,

9, 451.

244. Yadav, D.G.; Rahuman, M.S.M. Appl. Catal. A: General, 2003, 253, 113.

245. Selvaraj, M.; Kawi, S. J. Mol. Catal. A: Chemical, 2007, 265, 250

246 Olah, G.A., Friedel-Crafts Chemistry, Wiley Interscience, New York, 1973.

247. Lipovich, I.M., Med. Prom, S.S.R.1959, 13(1), 33; Chem. Abstr. 1959, 53,1604H.

248. Rosenwald, R.H.; US Pat. 2470902,1949.

249. Domingos, A.K.; Saad, E.B.; Vechiatto,W.W.D.; Wilhem, H.M.; Ramos, L.P., J. Braz.

Chem. Soc., 2007, 18(2), 416.

250. Shinde, A. B; Shrigasim, N. B.; Samat, S.D. Appl. Catal. A: General, 2004, 276, 5.

251. Nishiyama, T.; Sakita, K.; Fuchigami, T.; Fukui, T., Polymer Degradation and Stability,

1998, 62, 529.

252. Leffingwell, J.C.; Shackelford, R.E., Cosmet Perfumery, 1974, 89(6), 69.

253. Hopp, R.; Adv. Tobacco Sci, 1993, 19, 46.

254. Chaudhuri, B.; Patwardhan, A.A.; Sharma, M.M. Ind. Eng. Chem. Res.,1990, 29, 1025.

255. Leston, G., US Pat. 346649, 1967.

256. Stevens, D.R.; Bowmen, R.S., US Pat. 2560666, 1951.

257. Leston, G., US Pat. 3091646, 1963.

258. Kurk, P.R., US Pat. 4380677, 1983.

259. Wetzel, W. H.; Nelson, H.G.; Sheldon, F.J., US Patent 3932537, 1976.

260. Olah, G.A. J.Org. Chem.,2005, 70(7), 2413.

261. Yadav, G.D.; Thorat, T.S. Ind. Eng. Chem. Res. 1996, 35, 721.

Page 211: Applications of metal triflates and assisted acids as

191

262. Kobayashi, S.; Manabe, K. Pure Appl. Chem., 2000, 72(7), 1373.

263. Torres, M.; Gutierez, M.; Mugica, V.; Romero, M.; Lopez, L., Catal. Today, 2011, 166(1),

205.

264. Olah, G.A.; Farooq, O.Farnia, S. M. F.; Olah, J. A., J. Am. Chem. Soc., 1988, 110, 2560.

265. Kawada, A.; Mitamura, S.; Kobayashi, S.,Synlett., 1994, 545.

266. Barrett, A.G.M.; Braddock,. D.C.; Henschke, J.P.; Walker, E.R., J. Chem. Soc., Perkin

Trans.1,1999, 873.

267. Prinsloo, N. M., Fuel Processing Technology, 2006, 87, 437.

268. Kroupa, J.; Podstata, J.; Matous, V.; Nechybova, V.R. Czech Pat.265262, 1990, (Chem.

Abstr., 1992, 117, 48093).

269. Sartori, G.; Bigi, F.; Casiraghi, G.; Casnari, G.; Cheisi, L.; Arduini, A. Chem. Ind., 1985, 22,

726.

270. Noburu, O.; Tetsuya, Y.; Kamimura, A.; Aritsune, K. J, Chem. Soc., Chem. Commun.,

1986, 16, 1285.

271. Fernholz, H.; Schimdt, H.J. Angew. Chem., 1969, 8, 521.

272. Yadav., G.D.; Doshi, N.S. Catal Today, 2000, 60, 263.

273. CRC, Handbook of Chemistry and Physics 56th Ed.,1975-1976.

274. Fǎrcaşiu,D.; Ghenciu, A., J. Am. Chem. Soc., 1993, 115, 10901.

275. Atkins, P.W., Physical Chemistry, 6th Ed. Oxford University Press, Melbourne, 1998, 229.

276. Sawant, D.P.; Devassy, B.M.; Halligudi, S.B.,J. Mol. Catal. A: Chemical, 2004, 217, 211.

277. Mandelcorn, L.; Miller, R.L.; Mercier, G.E.; Pickett, J.H.; Dakin, T.W., US 4054937, 1977.

278. Rudnick, L.R.; Kremmer, R.A.; Law, D.A.,US Patent 552071, 1996.

279. Coleman, G.H.; Perkins, R. P.,US Patent 2170809, 1939.

280. Cobb, R.L.; Mitchell, M.D.,US Patent 4668269, 1987.

281. Bob, D.; Sotwald, T.,J. Can. Pet. Technol. 30, 1991, 133.

282. Piao, L.Y.; Fu, X.; Yang, Y.L.; Tao, H.G.; Kou, Y.,Catal. Today, 2004, 93-94, 301.

283. Yadav, G.D.; Sengupta, S.,Org. Process Research & Development, 2002, 6, 256.

284. Yadav, G.D.; Kundu, B.,Can. J. Chem. Eng. 2001, 79, 805.

285. Brown, P.E.; Lewis, R.; Waring, M.A. J. Chem. Soc. Perkin Trans. I ,1990, 2979.

286. Kashman, Y.; Gustafson, K.R.; Fuller, R.W.; Cardellina II, J.H.; McMahon, J.B.; Currens,

M.J.; Buckheit, Jr., R.W.; Highes, S.H.; Cragg, G.M.; Boyd M.R.; J. Med. Chem.1992, 35, 2735.

287. Schneider, C. Mol. Nutr. Food Res. 2005, 49, 7.

288. Ren, W.; Qiao, Z.; Wang, H.; Zhu, L.; Zhang, L. Med. Res. Rev. 2003, 23, 519.

Page 212: Applications of metal triflates and assisted acids as

192

289. Adrio, L.A.; Hii, K.K. Chem.Commun. 2008, 2325.

290. Kalena, G.P.; Jain, A; Banerji, A. Molecules,1997, 2, 100.

291. a) Bienaymé, H.; Ancel, J.-E.; Meilland, P.; Simonato, J.-P. Tetraheron Lett., 2000, 41,

3339. b) Youn, S.W.; Eom, J. I. J. Org. Chem., 2006, 71, 6705.

292. a) Bader, A.R.; Bean, W.C. J. Am.Chem.Soc., 1958, 80, 3037. b) Ahluwalia, V.K.; Arora,

K.K. J. Chem. Soc. Perkin Trans. 1 1982, 335. c) Ahluwalia, V.K.; Arora, K.K. Tetrahedron,

1981, 37, 1437.

293. Youn, S.W.; Eom, J. I. J.Org.Chem. 2006, 71, 6705.

294. Ahluwalia, V.K.; Arora, K.K.; Jolly, R.S., J. Chem. Soc. Perkin 1, 1982, 335.

295. Dewhirst, K.C.; Rust, F.F. J. Org. Chem. 1963, 28,798.

296. Bader, A.R.; Bean, W.C. J. Am. Chem. Soc., 1958, 80, 3037.

297. Hammond, G.S. J.Am. Chem. Soc., 1955, 77, 334.

Page 213: Applications of metal triflates and assisted acids as

193

Page 214: Applications of metal triflates and assisted acids as

194

Page 215: Applications of metal triflates and assisted acids as

195

Page 216: Applications of metal triflates and assisted acids as

196

Page 217: Applications of metal triflates and assisted acids as

197

Page 218: Applications of metal triflates and assisted acids as

198

Page 219: Applications of metal triflates and assisted acids as

199

Page 220: Applications of metal triflates and assisted acids as

200

Page 221: Applications of metal triflates and assisted acids as

201

Page 222: Applications of metal triflates and assisted acids as

202

Page 223: Applications of metal triflates and assisted acids as

203

Page 224: Applications of metal triflates and assisted acids as

204

Page 225: Applications of metal triflates and assisted acids as

205

Page 226: Applications of metal triflates and assisted acids as

206

Page 227: Applications of metal triflates and assisted acids as

207

Page 228: Applications of metal triflates and assisted acids as

208

Page 229: Applications of metal triflates and assisted acids as

209

Page 230: Applications of metal triflates and assisted acids as

210

Page 231: Applications of metal triflates and assisted acids as

211

Page 232: Applications of metal triflates and assisted acids as

212

Page 233: Applications of metal triflates and assisted acids as

213

Page 234: Applications of metal triflates and assisted acids as

214

Page 235: Applications of metal triflates and assisted acids as

215

Page 236: Applications of metal triflates and assisted acids as

216

Page 237: Applications of metal triflates and assisted acids as

217

Page 238: Applications of metal triflates and assisted acids as

218

Page 239: Applications of metal triflates and assisted acids as

219

Page 240: Applications of metal triflates and assisted acids as

220

Page 241: Applications of metal triflates and assisted acids as

221

Page 242: Applications of metal triflates and assisted acids as

222

Page 243: Applications of metal triflates and assisted acids as

223

Page 244: Applications of metal triflates and assisted acids as

224

Page 245: Applications of metal triflates and assisted acids as

225

Page 246: Applications of metal triflates and assisted acids as

226

Page 247: Applications of metal triflates and assisted acids as

227

Page 248: Applications of metal triflates and assisted acids as

228

Page 249: Applications of metal triflates and assisted acids as

229

Page 250: Applications of metal triflates and assisted acids as

230

Page 251: Applications of metal triflates and assisted acids as

231

Page 252: Applications of metal triflates and assisted acids as

232