solvent-free synthesis of bisferrocenylimines

135

Click here to load reader

Upload: ioana-nicolae

Post on 14-Apr-2015

36 views

Category:

Documents


3 download

DESCRIPTION

sfs

TRANSCRIPT

Page 1: Solvent-free Synthesis of Bisferrocenylimines

SOLVENT-FREE SYNTHESIS OF BISFERROCENYLIMINES

AND THEIR COORDINATION TO RHODIUM(I)

PHUMELELE ELDRIDGE KLEYI

Submitted in partial fulfilment of the requirements for the degree of

MAGISTER SCIENTIAE in the Faculty of Science at the Nelson Mandela

Metropolitan University

Supervisor: Prof. Christopher Imrie

Co-Supervisors: Prof. T. I. A. Gerber

Prof. C. W. McCleland

January 2009

Page 2: Solvent-free Synthesis of Bisferrocenylimines

i

ACKNOWLEDGEMENTS

The author would like to express his gratitude to the following, for the contributions

made in the thesis:

1. Dr. Christopher Imrie for guidance, support and encouragement throughout

the project.

2. Prof. C. W. McCleland for help with NMR problems.

3. Prof. T. I. A. Gerber assistance with the project and thesis.

4. Dr. P. Mallon from the University of Stellenbosch for the help with Gel

Permeation Chromatography.

5. Mr. Harold Marchand, Mr H. Schalekamp and Mr. J. Booi for technical

assistance.

6. Mr. Irvin Booysen for assistance with UV-vis, CV and conductometry.

7. Mr. M. Mtyopo and Mr. B. Mpuhlu for assistance with GC.

8. Dr. E. R. T. Elago, Dr. V. O. Nyamori, Dr. Z. Tshentu and Mr. P. Hlangothi for

friendly advice.

9. My colleagues, Mr. D. Onyancha and Mrs. N. Adams for friendship and

support. My brother, Ayanda, for being there when I needed him most.

10. NRF and NMMU for financial support towards my studies.

11. Last but not least, The Almighty God for making my dreams come true.

Page 3: Solvent-free Synthesis of Bisferrocenylimines

ii

ABSTRACT

Solvent-free reactions possess advantages compared to the solvent route, such as

shorter reaction times, less use of energy, better yields, etc. Herein, the synthesis

and characterization of bisferrocenylimines and arylbisamines are described.

Reduction of the above compounds with LAH resulted in the formation of

bisferrocenylamines and arylbisamines, respectively. The coordination chemistry of

all the above compounds to rhodium(I) is also discussed in the prepared complexes

[Rh(COD)(NN)]ClO4, where NN = bisferrocenylimines, and [Rh(COD)(NN)]BF4,

where NN = bisferrocenylamines and arylbisamines. X-ray crystal structures of the

complexes [Rh(COD)(NN)]ClO4 ([3.2] and [3.3]) have been obtained. Complexes of

the type [Rh(COD)(NN)]BF4 were characterized with IR and UV-vis spectroscopy,

cyclic voltammetry and conductometry. The catalytic activity of the complexes was

also investigated: [Rh(COD)(NN)]ClO4 for the polymerization of phenylacetylene and

[Rh(COD)(NN)]BF4 for the hydroformylation of styrene.

Keywords: bisferrocenylimines, coordination chemistry, rhodium(I).

Page 4: Solvent-free Synthesis of Bisferrocenylimines

iii

PRESENTATIONS AND PUBLICATIONS

Publications

• Further solvent-free reactions of ferrocenylaldehydes: Synthesis of 1,1’-

ferrocenyldiimines and ferrocenylacrylonitrile, C. Imrie, P. Kleyi, V. O.

Nyamori, T. I. A. Gerber, D. C. Levendis and J. Look, J. Organomet. Chem.,

692 (2007) 3443-3454.

Conference proceedings

• Solvent-free synthesis and coordination chemistry of

diferrocenyldiazaalkanes, P. Kleyi, C. Imrie and C. W. McCleland, 37th

International Conference on Coordination Chemistry (ICCC37), Cape Town,

South Africa, August 2006.

• Ferrocenylnitrogen-donor ligands for homogeneous catalysis, P. Kleyi, D.

Saku, C. Imrie and C. W. McCleland, 15th International Symposium on

Homogeneous Catalysis (ISHCXV), Sun City, South Africa, August 2006.

• Synthesis and use of bisferrocenylimines as new catalysts for olefin

polymerization, P. Kleyi, C. Imrie and C. W. McCleland, Inorganic Chemistry

Conference (INORG007), Club Mykonos, Western Cape, South Africa, July

2007.

Page 5: Solvent-free Synthesis of Bisferrocenylimines

iv

CONTENTS

Page

ACKNOWLEDGEMENTS ........................................................................................... i

ABSTRACT ................................................................................................................ ii

PRESENTATIONS AND PUBLICATIONS ................................................................ iii

LIST OF FIGURES................................................................................................... viii

LIST OF SCHEMES ................................................................................................... x

LIST OF TABLES ..................................................................................................... xii

ABBREVIATIONS.................................................................................................... xiii

CHAPTER 1 ........................................................................................................... 1

INTRODUCTION ........................................................................................................ 1

1.1 Solvent-free synthesis ...................................................................................... 1

1.1.1 Background................................................................................................... 1

1.2 FERROCENES ................................................................................................... 5

1.2.1 Solvent-free synthesis of ferrocenes ............................................................. 5

1.2.1.1 Synthesis of ferrocenylenones ............................................................... 6

1.2.1.2 Reaction of ferrocenecarboxaldehyde with methylene active compounds ............................................................................................................................ 7

1.2.1.3 Reaction of ferrocenecarboxaldehyde with an ylid ................................. 9

1.2.1.4 Synthesis of ferrocenyl-1,5-diketone derivatives .................................. 10

1.2.1.5 Synthesis of ferrocenoate esters .......................................................... 10

1.2.1.6 Synthesis of ferrocenylimines ............................................................... 11

1.3 SOLVENT-FREE SYNTHESIS OF LIGAND SYSTEMS ................................... 12

1.4 NITROGEN-DONOR LIGAND CHEMISTRY .................................................... 17

1.4.1 Polymerization reactions ............................................................................. 19

1.4.2 Cross-coupling reactions ............................................................................ 22

1.4.2.1 Heck reactions ..................................................................................... 22

1.4.2.2 Suzuki cross-coupling reactions ........................................................... 24

Page 6: Solvent-free Synthesis of Bisferrocenylimines

v

1.4.3 Epoxidation reactions ................................................................................. 26

1.4.4 Asymmetric allylic substitution reactions ..................................................... 29

1.4.5 Ring-opening metathesis polymerization (ROMP) ...................................... 30

1.5 FERROCENYL-NITROGEN DONOR LIGAND CHEMISTRY........................... 32

1.5.1 Ferrocenyl-pyridine ligands ......................................................................... 32

1.5.2 Ferrocenyl-Schiff base ligands .................................................................... 36

1.6 OBJECTIVES OF THE PROJECT .................................................................... 40

1.7 REFERENCES .................................................................................................. 41

CHAPTER 2 ......................................................................................................... 50

RESULTS AND DISCUSSION ................................................................................ 50

2.1 SOLVENT-FREE SYNTHESIS OF BISFERROCENYLIMINES........................ 50

2.1.1 Introduction ................................................................................................. 50

2.1.2 Synthesis and characterization of bisferrocenyimines ................................ 51

2.1.3 Solvent-free synthesis of arylbisimines ....................................................... 57

2.2 REDUCTION REACTION OF BISFERROCENYLIMINES .............................. 60

2.2.1 Reduction of bisferrocenylimines ................................................................ 60

2.2.2 Reduction of arylbisimines .......................................................................... 63

2.3 ELECTRONIC SPECTROSCOPY .................................................................... 63

2.4 CYCLIC VOLTAMMETRY ................................................................................ 66

2.5 EXPERIMENTAL .............................................................................................. 68

2.5.1 Purification procedures ............................................................................... 68

2.5.2 Instrumentation ........................................................................................... 69

2.6 SYNTHESIS OF BISFERROCENYLIMINES AND ARYLBISIMINES .............. 70

2.6.1 General procedure for the synthesis of bisferrocenylimines ....................... 70

2.6.2 Reduction of bisferrocenylimines and arylbisimines ................................... 75

2.7 REFERENCES .................................................................................................. 79

Page 7: Solvent-free Synthesis of Bisferrocenylimines

vi

CHAPTER 3 ......................................................................................................... 80

RESULTS AND DISCUSSION ................................................................................ 80

3.1 SYNTHESIS OF CATIONIC RHODIUM(I) COMPLEXES ................................. 80

3.1.1 Rhodium(I) complexes containing bisferrocenylimines ............................... 80

3.1.2 X-ray Crystallography ................................................................................. 84

3.1.3 Rhodium(I) complexes containing bisferrocenylamines .............................. 89

3.2 ELECTRONIC SPECTROSCOPY .................................................................... 93

3.3 CYCLIC VOLTAMMETRY ................................................................................ 95

3.4 EXPERIMENTAL .............................................................................................. 97

3.4.1 Purification procedures ............................................................................... 97

3.4.2 Instrumentation ........................................................................................... 97

3.5 SYNTHESIS OF RHODIUM(I) COMPLEXES ................................................... 98

3.5.1 Rhodium(I) complexes containing bisferrocenylimines ............................... 98

3.5.1.1 General procedure3 .............................................................................. 98

3.5.2 Rhodium(I) complexes containing bisferrocenylamines ............................ 100

3.6 REFERENCES ................................................................................................ 102

CHAPTER 4 ....................................................................................................... 103

RESULTS AND DISCUSSION .............................................................................. 103

4.1 POLYMERIZATION OF PHENYLACETYLENE ............................................. 103

4.1.1 Introduction ............................................................................................... 103

4.1.2 Polymer characterization .......................................................................... 104

4.2 CATALYTIC POLYMERIZATION STUDIES .................................................. 105

4.2.1 Spectroscopic properties of polymers ....................................................... 106

4.2.2 Thermal analysis ....................................................................................... 108

4.2.3 Mechanistic pathways for polymerization of phenylacetylene ................... 109

Page 8: Solvent-free Synthesis of Bisferrocenylimines

vii

4.3 HYDROFORMYLATION OF STYRENE ......................................................... 111

4.3.1 Introduction ............................................................................................... 111

4.3.2 Catalytic hydroformylation studies ............................................................ 113

4.3.3 Mechanism for hydroformylation of styrene .............................................. 115

4.4 EXPERIMENTAL ............................................................................................ 117

4.4.1 Purification procedures ............................................................................. 117

4.4.2 Instumentation .......................................................................................... 117

4.4.3 Polymerization of phenylacetylene ........................................................... 117

4.4.4 Hydroformylation of styrene ...................................................................... 118

4.5 REFERENCES ................................................................................................ 118

CHAPTER 5 ....................................................................................................... 121

CONCLUSION ....................................................................................................... 121

5.1 Conclusion ...................................................................................................... 121

Page 9: Solvent-free Synthesis of Bisferrocenylimines

viii

LIST OF FIGURES

Page

Figure 1.1: DSC analysis of ferrocenecarboxaldehyde ............................................ 6

Figure 1.2: Some N-donor ligands with sp2-hybridised nitrogen atoms .................. 18

Figure 1.3: Bis(alkylphenylaminopyridinato) titanium complexes ........................... 20

Figure 1.4: Pyridyl-imine complexes of iron (Fe) and palladium (Pd) ..................... 21

Figure 1.5: Examples of Pd-pyridyl complexes used for the Heck reactions .......... 23

Figure 1.6: Examples of Ni(0) complexes used for cross-coupling of arylchlorides.

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

Figure 1.7: Palladium imine and amine complexes for coupling of aryl bromides .. 26

Figure 1.8: Three isomers of Ru(pap)2Cl2 .............................................................. 27

Figure 1.9: Mn(III) Schiff-base complexes for electrocatalytic epoxidation of olefins

................................................................................................................................. 28

Figure 1.10: The cis-β isomer of the binaphthyl-bridged Schiff base titanium

complex .......................................................................................... 28

Figure 1.11: Example of chiral bidentate thiazolyl-pyridine ligands………………30 Figure 1.12: Chiral diimine palladium(II) catalyst for asymmetric alkylation ........ 30

Figure 1.13: The reaction pathway of ROMP ...................................................... 31

Figure 1.14: Re and Pt complexes with ferrocenylpyridine ligands ..................... 33

Figure 1.15: Tungsten complexes with ferrocenylpyridine ligands ...................... 33

Figure 1.16: Examples of ferrocenylpyridyl and pyrimidyl complexes ................. 34

Figure 1.17: Pd and Pt complexes of 1,1’-bis(2-pyridyl)ferrocene ....................... 34

Figure 1.18: Palladium and nickel complexes ..................................................... 35

Figure 1.19: Dimeric cyclopalladated ferrocenylimine complex for catalytic Heck

reaction ........................................................................................... 36

Figure 1.20: Cyclopalladated ferrocenylimine for Mirozoki-Heck reaction ........... 37

Figure 1.21: 1,1’-N-substituted ferrocenediyl Pd(II) complex for Suzuki cross-

coupling reaction ............................................................................. 37

Figure 1.22: Dimeric cyclopalladated ferrocenylketimine complexes for Suzuki

reaction ........................................................................................... 38

Figure 1.23: Chiral ferrocenylphosphine-imine ligand ......................................... 38

Figure 1.24: Cationic Rh(I) and Ir(I) complex for olefin polymerization ................ 39

Page 10: Solvent-free Synthesis of Bisferrocenylimines

ix

Figure 1.25: Cationic Rh(I) complexes for hydroformylation reactions. ................ 40

Figure 2.1: General structure of bisferrocenylimines to be synthesized ............. 51

Figure 2.2: The pictorial stages of the solvent-free synthesis of N,N’-

octylenebis(ferrocenylmethylidine)amine ........................................ 52

Figure 2.3: IR spectrum of [2.5].......................................................................... 54

Figure 2.4: 1H NMR spectrum of [2.5] ................................................................. 55

Figure 2.5: 13C NMR spectrum of [2.8] ................................................................ 60

Figure 2.6: UV-vis spectrum of unsubstitued ferrocene in dichloromethane. ...... 64

Figure 2.7: UV-vis spectra of bisferrocenylimines in dichloromethane. ............... 65

Figure 2.8: UV-vis spectra of bisferrocenylamines in dichloromethane. .............. 65

Figure 2.9: Cyclic voltammogram of ferrocene in acetonitrile. ............................. 67

Figure 2.10: Cyclic voltammograms of [2.2], [2.5], [2.12] and [2.14]. ................... 67

Figure 3.1: 1H NMR spectra of [3.2] (top) and [3.3] (bottom) in CDCl3................ 83

Figure 3.2: ORTEP diagram of [3.2]. ................................................................... 85

Figure 3.3: Crystal packing of [3.2], projection viewed along [100] ..................... 87

Figure 3.4: ORTEP drawing of [3.3] .................................................................... 88

Figure 3.5: Crystal packing of [3.3], projection viewed along [100]. .................... 88

Figure 3.6: Cationic rhodium(I) diamine complexes with the [Rh(COD)Cl2]- anion.

................................................................................................................................. 89

Figure 3.7: Cationic rhodium(I) complexes ........................................................... 90

Figure 3.8: IR spectrum of [3.5] ........................................................................... 91

Figure 3.9: IR spectrum of [3.8] ........................................................................... 93

Figure 3.10: UV-vis spectra of [3.1], [3.2] and [3.3]. .............................................. 94

Figure 3.11: UV-Vis spectra of [3.4], [3.5] and [3.6] .............................................. 94

Figure 3.12: UV-vis spectra of [3.7] and [3.8]. ....................................................... 95

Figure 3.13: Cyclic voltammograms of [3.2], [3.4], [3.5] and [3.6] ......................... 96

Figure 4.1: Stereoisomers of polyphenylacetylene ............................................. 104

Figure 4.2: 1H NMR spectrum of PPA, catalyzed by [3.2]. ................................. 107

Figure 4.3: IR spectrum of PPA prepared using [3.2]......................................... 108

Figure 4.4: TGA and DSC curves of PPA obtained with [3.1]. ............................ 109

Figure 4.5: 1H NMR spectrum of the products of hydroformylation of styrene

catalyzed by [3.6] .............................................................................. 114

Page 11: Solvent-free Synthesis of Bisferrocenylimines

x

LIST OF SCHEMES

Page

Scheme 1.1: Examples of solvent-free reactions .................................................... 3

Scheme 1.2: Solvent-free synthesis of 3-carboxycoumarins ................................... 3

Scheme 1.3: Photoirradited solvent-free dimerization of cholest-4-en-3-one .......... 4

Scheme 1.4: Solvent-free synthesis of ferrocenylenones ........................................ 7

Scheme 1.5: Solvent-free Knoevenagel condensation reaction .............................. 8

Scheme 1.6: Solvent-free Wittig reaction of ferrocenecarboxaldehyde ................... 9

Scheme 1.7: Solvent-free synthesis of ferrocenyl 1,5-diketone derivatives ........... 10

Scheme 1.8: Solvent-free synthesis of ferrocenoate esters .................................. 11

Scheme 1.9: Solvent-free reactions of ferrocenylaldehydes with aromatic amines 11

Scheme 1.10: Synthesis of binaphthol .................................................................... 12

Scheme 1.11: Solvent-free palladium-catalyzed phosphination reaction................. 13

Scheme 1.12: Solvent-free oxidation of thiols to disulfides ...................................... 13

Scheme 1.13: A mechanism for the synthesis of unsymmetrical sulfides from thiols

and alkyl halides using hydrotalcite clays .......................................... 14

Scheme 1.14: Solvent-free metal mediated synthesis of homoallyl alcohols ........... 14

Scheme 1.15: Microwave assisted solvent-free synthesis of β-aminoalcohols........ 15

Scheme 1.16: Sc(OTf)3 catalyzed solvent-free synthesis of β-aminoalcohols ......... 15

Scheme 1.17: Solvent-free synthesis of tetrasubstituted imidazoles on silica gel

support ............................................................................................ 16

Scheme 1.18: Solvent-free synthesis of tetrasubstituted imidazoles on SiO2/NaHSO4

support .............................................................................................. 16

Scheme 1.19: Solvent-free synthesis of 2,4,6-triarylpyridines ................................. 17

Scheme 1.20: Solvent-free synthesis of Schiff bases .............................................. 17

Scheme 1.21: Synthesis of bis(phenoxyketimine) zirconium complexes ................. 20

Scheme 1.22: Synthesis of silica-supported imine palladacycles. ........................... 25

Scheme 1.23: Suzuki cross-coupling of aryl bromides with phenylboronic acid. ..... 26

Scheme 1.24: Allylic alkylation of 1,3-di[henyl-2-enyl acetate with dimethyl malonate

................................................................................................................................. 29

Scheme 1.25: Synthesis of Ru complexes derived from 1st generation Grubbs

catalyst. ........................................................................................... 32

Page 12: Solvent-free Synthesis of Bisferrocenylimines

xi

Scheme 1.26: Synthesis of N-ferrocene salicylaldimine ligand ............................... 39

Scheme 2.1: Solvent-free synthesis of ferrocenylimines ......................................... 50

Scheme 2.2: Solvent-free synthesis of bisferrocenylimines ..................................... 51

Scheme 2.3: Solvent-free synthesis of arylbisimines in the presence of a catalyst. 57

Scheme 2.4: Solvent-free synthesis of arylbisimines............................................... 58

Scheme 2.5: Hydrogenation of ferrocenylbisimines. ............................................... 61

Scheme 3.1: Procedure for the synthesis of cationic rhodium(I) complexes. .......... 81

Scheme 4.1: Polymerization of phenylacetylene with Rh(I) catalysts. ................... 105

Scheme 4.2: Insertion mechanism for polymerization of phenylacetylene ............ 110

Scheme 4.3: Metallacyclic mechanism for the polymerization of phenylacetylene 110

Scheme 4.4: Hydroformylation of olefins. .............................................................. 111

Scheme 4.5: Synthesis of precursor to the indolizidine alkaloid ............................ 113

Scheme 4.6: Hydroformylation of styrene catalyzed by [3.4]-[3.8] ........................ 113

Scheme 4.7: Possible mechanism for hydroformylation of styrene catalyzed by [3.4]-

[3.8] ................................................................................................... 116

Page 13: Solvent-free Synthesis of Bisferrocenylimines

xii

LIST OF TABLES

Page

Table 2.1: Yields of bisferrocenylimines from the solvent-free reaction of diamines

and ferrocenecarboxaldehyde ............................................................... 53

Table 2.2: Chemical shifts for protons on the carbon directly bonded to nitrogen

group. .................................................................................................... 56

Table 2.3: Yields of arylbisimines from a reaction of substituted benzaldehyde and

ethylenediamine .................................................................................... 59

Table 2.4: Yields of bisferrocenylimines .................................................................. 62

Table 2.5: UV-vis data for ferrocene, [2.1]-[2.5] and [2.10]-[2.14]. ......................... 66

Table 2.6: Half-wave potentials of [2.1]-[2.5] and [2.12]-[2.14]. .............................. 68

Table 3.1: The summarized NMR data for [3.1]-[3.3]. ............................................. 82

Table 3.2: Crystal data and structure refinement of [3.2] and [3.3]. ....................... 84

Table 3.3: Selected bond distances, bond angles and torsion angles of [3.2] ......... 86

Table 3.4: Selected bond distances, bond angles and torsion angles of [3.3] ......... 89

Table 3.5: Table of yields and conductivity measurements ...................................... 92

Table 3.6: UV-vis data for complexes [3.1]-[3.8]. .................................................... 95

Table 3.7: Half-wave potentials of [3.1]-[3.6] ........................................................... 96

Table 4.1: Polymerization of phenylacetylene with Rh(I) complexes. .................... 106

Table 4.2: Determination of cis-content of polymers .............................................. 108

Table 4.3: Hydroformylation of styrene catalyzed by rhodium(I) complexes .......... 115

Page 14: Solvent-free Synthesis of Bisferrocenylimines

xiii

ABBREVIATIONS

Å Angstrom

ca. approximately

cat catalyst

COD 1,5-cyclooctadiene

Cp cyclopentadienyl, C5H5

CDCl3 deuterated chloroform

˚ degree

DSC differential scanning calorimetry

ε extinction coefficient

EI electron impact

FAB fast atom bombardment

J coupling constant

Lit. literature

M+ parent molecular ion

MeOH methanol

M.p. melting point

NMR nuclear magnetic resonance

Ph phenyl

PP phenylpropanal

PPA poly(phenylacetylene)

ppm parts per million

RT room temperature

Page 15: Solvent-free Synthesis of Bisferrocenylimines

1

CHAPTER 1

INTRODUCTION

1.1 Solvent-free synthesis

1.1.1 Background

Chemistry has played a leading role in changing people’s lives, due to its impact in

areas such as agrochemicals, the clothing industry, food technology, energy and

transport, the pharmaceutical industry and most recently in the manufacture of

electronic devices. However, discoveries about ecotoxic effects such as endocrine

disruption1 indicated that synthetic chemicals released into the environment have a

negative impact on the world ecosystem. Industrial incidents involving explosions at

a major South African petrochemical company,2,3 the discovery of persistent organic

pollutants and the global warming are examples of chemical disasters. It is because

of this reason that chemists are compelled to shoulder the responsibility for the

consequences and thus develop new synthetic protocols that are environmentally

benign. These new synthetic protocols should comply with green chemistry

principles.4 As an alternative to organic solvents, chemists should employ other

strategies to perform chemical reactions, namely ionic liquids, supercritical fluids,

water as a solvent and solvent-free conditions.

Conventionally, chemical transformations have been carried out in the presence of a

solvent to provide a homogeneous medium for the reagents to interact effectively as

well as for the isolation and purification of the desired product.5,6 It was believed that

in solution the reagents have higher mobility, hence increased molecular collisions

leading to faster chemical reactions. Unfortunately, organic solvents are high on the

list of toxic or hazardous compounds because of large volumes used in industry, and

difficulties in containing volatile organic compounds (VOC’s).7

Page 16: Solvent-free Synthesis of Bisferrocenylimines

2

The development of solvent-free organic synthetic procedures has become an

important and popular research area.8 Chemical reactions under solvent-free

conditions have been practised for many years. It has been reported that the first

written document on a solvent-free reaction came from a book ‘De Lapidibus’ by

Theophrastus (371-286 B.C).9 He reported that when cinnabar (HgS) was ground in

a brass motor with a brass pestle in the presence of vinegar, metallic mercury was

obtained. Afterwards, reports on chemical transformations under solvent-free

medium were rare, until the work of Carey Lea10,11 as well as Ling and Baker at the

end of the 19th century.12 Ling and Baker were successful in preparing quinhydrone

by grinding together two solids. It was again only more than half a century later that

researchers became interested in chemical transformations in the absence of a

solvent medium.

In the 1960’s, Rastogi et al.13-15 carried out detailed investigations about factors

governing the reaction between two solids, such as mobility on the interface, kinetics

and mechanism. In 1984, Patil et al.16 reported on the successful solvent-free

synthesis of unsymmetrically substituted quinhydrones, which was an alternative to

the problematic solution procedure due to self-oxidation reactions. In 1987, Toda et

al.17 introduced the concept of ‘host-guest’ complex formation in the solid state. It

was proposed that the quinine sublimes and its vapour attacks certain hydroquinone

sites. Toda et al.18,19 went on to report on their successful Grignard reaction and

aldol condensation under solvent-free conditions.

Owing to these successes, many researchers began to recognise the advantages of

carrying out chemical reactions in the absence of solvents. In the 1990’s,

researchers became strongly involved in carrying out studies on different types of

reactions without the use of classical solvents. Examples were the work of Kaupp et

al. (1993)20 (Scheme 1.1a), Villemin et al. (1995)21 (Scheme 1.1b) and Ranu et al.

(1997)22 (Scheme 1.1c) as well as other researchers.23-25 Ranu and coworkers were

successful in synthesizing a Michael addition type product under microwave-assisted

solvent-free conditions.

Page 17: Solvent-free Synthesis of Bisferrocenylimines

3

R CHO +

NH

N

O

NH

CH3

Solvent-free

MWNH

N

O

NH

CH3

H

R

R =

R R'

O O

+n

O

Al2O3

MW

n

O

R

R'

O

O

(i): R =CH3; R' = OEt; n = 1, 2(ii): R, R' = CH3; n = 1,2

O

O ;

Cl

;O

NH

N

S

O

R

R'

+ N

R''

solvent-free NN N

R'

O

SR''

HH

(i) R = R' = R'' = H(ii) R = R' = H, R" = CH3(iii) R = R'' = H, R' = Ph

a)

b)

c)

100%

84-91%

78-90%

Scheme 1.1: Examples of solvent-free reactions20-22

In 2000, Scott and Raston26 reported on the synthesis of 3-carboxycoumarins, via

the Knoevenagel condensation reaction, by gently grinding the starting materials in a

mortar with a pestle (Scheme 1.2).

R

OH

CHO

+

O

O

O

OR

OH

O

O

O

O

R

OH

O

OH

O

O

NH4+MeCO2

-

Scheme 1.2: Solvent-free synthesis of 3-carboxycoumarins26

Page 18: Solvent-free Synthesis of Bisferrocenylimines

4

DellaGreca et al27 reported, a year later, that it is possible to dimerize cholest-4-en-

3one under photoirradiation. Powdered cholest-4-en-3-one was simply irradiated

with a UV lamp to yield the desired product (Scheme 1.3).

O

hv

OH

O

+

O OH

Scheme 1.1: Photoirradited solvent-free dimerization of cholest-4-en-3-one27

Solvent-free reactions are of utmost interest from the ecological point of view, and

they offer advantages, such as reduced reaction times, increased product yields,

reduced environmental pollution, simple equipment (lab scale), increased selectivity,

and low cost compared with reactions carried out in solvents.4 The formation of hot

spots (solvents can act as heat sinks) and the prospects of runaway reactions is one

of the few disadvantages of the solvent-free reactions. Another disadvantage is the

difficulty in designing suitable reactors for these reactions to be applied in the

industrial scale.4 However, these problems can be solved by using engineering

reactor technology.7

Page 19: Solvent-free Synthesis of Bisferrocenylimines

5

1.2 Ferrocenes

1.2.1 Solvent-free synthesis of ferrocenes The discovery and characterization of the structure of ferrocene or cyclopentadienyl

iron, Fe(C5H5)2 in the early 1950’s,28 led to an explosion of interest in d-block metal-

carbon bonds and stimulated the development of organometallic chemistry.29-33

Ferrocene derivatives are extremely important since they can be used in a variety of

functions, such as the synthesis of non-linear optical materials, organometallic

complexes, and in catalysis.34-37 However, the preparation of ferrocene derivatives

has usually been performed under homogeneous conditions in the presence of

classical solvents. With the current upsurge of interest in performing chemical

transformations under solvent-free reactions, researchers have successfully

prepared ferrocene derivatives under solvent-free conditions.

In general, the reaction is induced by a method of mechanochemical activation.38-40

This merely means mechanical mixing (grinding or stirring) of the chemical reagents

to bring about the reaction at room temperature. In some cases, this is coupled with

microwave irradiation,41 especially to accelerate reactions that are very slow at room

temperature. A mortar and pestle set is used for simple mechanical mixing and a

commercial microwave oven for irradiation.

It is worth noting that after reviewing the literature, it is apparent that most of the

solvent-free reactions of ferrocenes reported are carbonyl condensation reactions,

particularly the two compounds, ferrocenecarboxaldehyde35-37,41-44 and

acetylferrocene.41,43 These two compounds are important precursors for numerous

ferrocene derivatives, and have been shown to be particularly amenable to reactions

under solvent-free conditions.41-44,47 Differential scanning calorimetry (DSC) analysis

(Figure 1.1)42 indicates that ferrocenecarboxaldehyde exhibits a phase transition,

called a ‘plastic crystal phase’, at 45 °C and runs to its melting point at 120 °C. This

phase transition is considered to a point where a reaction takes place.

Acetylferrocene is expected to be less reactive due to the lack of this phase

transition and also due to steric hindrance by the methyl group.

Page 20: Solvent-free Synthesis of Bisferrocenylimines

6

Figure 1.1: DSC analysis of ferrocenecarboxaldehyde

We shall now consider the solvent-free reactions of ferrocenes and comparisons with

other synthetic methods will be provided where possible.

1.2.1.1 Synthesis of ferrocenylenones

The synthesis of ferrocenylenones can be performed by a variety of methods, such

as Claisen-Schmidt41 and aldol reactions43,44 under homogeneous conditions in

ethanol. Other methods that have been reported include the use 18-crown ether as a

phase-transfer catalyst (PTC) and solvent-free aldol reactions using pulverized

potassium hydroxide.41,43 In 1994, Villemin and co-workers41 described the solvent-

free synthesis of ferrocenylenones under microwave irradiation conditions.

Acetylferrocene was stirred with an aromatic aldehyde and ferrocenecarboxaldehyde

was stirred with a ketone, in the presence of a base or a phase-transfer catalyst

(Aliquat 336), or both, under microwave irradiation, to achieve the corresponding

ferrocenylenone (Scheme 1.4).

Page 21: Solvent-free Synthesis of Bisferrocenylimines

7

Fe

C

+ Ar

O

R

O

KOH, A336

RT or MW Fe

CCH

O

H

HC Ar

R = H, CH3

Scheme 1.4: Solvent-free synthesis of ferrocenylenones41

However, this method suffers from a serious disadvantage,41,43 since it was found

that the ferrocene derivatives combust easily in the microwave oven. Consequently,

Liu et al.43 and Méndez et al.44 described an improved method for the preparation of

ferrocenylenones via aldol condensation under solvent-free conditions. Méndez and

co-workers performed the aldol reaction by adding powdered base to a stirred

mixture of the reagents mixture in the presence of a PTC (Aliquat 336) while Liu et

al. managed to achieve excellent yields of the products, in the absence of the PTC,

by grinding a ketone and base using a mortar and pestle. Thereafter, the appropriate

aldehyde was introduced with further grinding for a few minutes. Interestingly, all

products formed for the different methods have an E-configuration as a major

product.41-44

1.2.1.2 Reaction of ferrocenecarboxaldehyde with methylene active compounds

Knoevenagel condensation of ferrocenecarboxaldehyde with methylene active

compounds is usually performed under classical homogeneous conditions in

ethanol.36,37 Of the two isomers obtained, the E-isomer was found to possess more

enhanced non-linear optical (NLO) properties than the Z-isomer.35 Incidentally, a

whole range of these kinds of ferrocene derivatives have been studied for their NLO

properties.35,36 They have also been used as catalysts for the combustion of

composite explosives.37 The classical homogeneous methods frequently result in low

yields and some reactions require the use of Schlenk techniques.36

Page 22: Solvent-free Synthesis of Bisferrocenylimines

8

Fe

CH

O

Fe

CX-CH2-Y, Al2O3C

X

H

Y

X, Y = electron withdrawing group

solvent-free

58-100%

Scheme 1.5: Solvent-free Knoevenagel condensation reaction36

Cooke and co-workers36 have reported on the solvent-free synthesis of novel

ferrocene derivatives from the Knoevenagel condensation reaction of

ferrocenecarboxaldehyde with methylene active compounds, in the presence of an

inorganic support (Scheme 1.5). The role of the inorganic support, basic alumina in

this case, was to increase the surface area and thus improve the conversion of

reagents into the desired products. The reaction was carried out simply by grinding

or stirring (depending on the nature of the methylene active compound) of the

reaction mixture (at room temperature or heated up to 65ºC) and the consequent

addition of a carefully calculated amount of basic alumina. This was accompanied by

an immediate colour change to purple, indicating formation of the product. Isolation

of the product was achieved simply by extraction with dichloromethane and in some

cases further purification by column chromatography was necessary. Stankovic and

co-workers35 obtained better yields, in some cases, when silica was used as an

inorganic support and they postulated that due to the inherent acidity of the silica, it

acted as a catalyst. It is worth mentioning that the method used by Stankovic and co-

worker differs to the one used by Cooke and coworkers. in that they do not grind the

starting reagents. Instead, a solution of ferrocenecarboxaldehyde in dichloromethane

is added to the inorganic support, followed by evaporation of the solvent. A solution

of methylene active compound in dichloromethane is then added and the solvent

evaporated. At this stage the reaction is left to stand at room temperature. The

method of Stankovic takes longer (1-48 h) than that of Cooke (1-5 h), and yet there

is difference in the yields. Bai et al.37 have reported on three methods for the

Knoevenagel condensation of ferrocenecarboxaldehyde and the methylene active

compound, namely in water, grinding under solvent-free conditions and microwave

induced solvent-free reactions. All the methods are performed in the presence of

Page 23: Solvent-free Synthesis of Bisferrocenylimines

9

potassium hydroxide. Most interestingly all the methods had shorter reaction times

than the previous methods, with reaction times in water (45 min), grinding (12-30

min) and microwave irradiation (6-8 min). None of the three methods was superior in

terms of the product yields, although microwave assisted reactions had shorter

reaction times.

1.2.1.3 Reaction of ferrocenecarboxaldehyde with an ylid

Fe

CH

O

Fe

HCRCH2P+Ph3X-

Solvent-free, NaOH

CHR

R = ArX = Cl, Br, I

71-95%

Scheme 1.6: Solvent-free Wittig reaction of ferrocenecarboxaldehyde47

The Wittig reaction is a very popular method for the regio- and stereo-selective

formation of alkenes from carbonyl compounds.45,46 A resonance-stabilized

phosphonium intermediate, called an ‘ylid’, is generated by abstraction of a proton

from a phosphonium salt by a base. The reaction of the ylid with a carbonyl group

results in the formation of an alkene and triphenylphosphine oxide as a by-product.

The Wittig reaction and many other synthetic methods used for the synthesis of

vinylferrocene derivatives, have usually been performed in classical homogeneous

media. The first ever solvent-free Wittig reaction was reported in 2001, by Liu and

co-workers.47 They were successful in preparation of vinylferrocene by grinding

ferrocenecarboxaldehyde and triphenylphosphinylbenzylphosphonium chloride in the

presence of a sodium hydroxide, using a mortar and pestle (Scheme 1.6). The

reactions were performed at room temperature (5 min) and slow reactions were

heated at 65 ºC (40-45 min). The conversions of starting reagent were moderate to

excellent and the major products were E-alkenes.

Page 24: Solvent-free Synthesis of Bisferrocenylimines

10

1.2.1.4 Synthesis ferrocenyl-1,5-diketone derivatives

1,5-Diketones are very useful synthetic intermediates and are desirable starting

materials for the preparation of heterocyclic and polyfunctional compounds.48-50

These compounds have potential applications in coordination chemistry, molecular

sensing, catalytic reactions, the chemical modification of electrodes and redox active

self-assembly devices. Various procedures have been reported for the synthesis of

1,5-diketones, but require the use of expensive reagents, are carried out under

refluxing conditions, and have longer synthetic routes.

Fe

R

O

PhCOMe, NaOH

solvent-freeFe

O R

Ph

O

R = Ar 68-92%

Scheme 1.7: Solvent-free synthesis of ferrocenyl 1,5-diketone derivatives51

Liu and co-workers51 have reported on the solvent-free synthesis of 1,5-diketone

derivatives containing ferrocene via the Michael addition reaction (Scheme 1.7). The

reaction was performed by grinding a mixture of an α,β-unsaturated ferrocenyl

ketone (or ferrocenylchalcone) and acetophenone with an agate mortar and pestle,

and the reaction mixture heated at 45 ºC. The pure products were obtained after a

short and simple work-up procedure.

1.2.1.5 Synthesis of ferrocenoate esters

Elago et al.52 have recently reported on the synthesis of ferrocenoate esters, amides

and other ferrocenoyl derivatives in ionic liquids and under solvent-free conditions

(Scheme 1.8). Both methods yielded excellent yields irrespective of the nature of the

substituent X. However, the ionic liquid required deaeration and three cycles of

freeze-thaw degassing. Moreover, the reaction requires 16 hours of stirring. In the

solvent-free method ferrocenoyl fluoride and the appropriate substituted phenol were

thoroughly ground in a mortar, and the mixture subjected to microwave irradiation for

1 minute.

Page 25: Solvent-free Synthesis of Bisferrocenylimines

11

Fe

CF

O

DMAP, [bmim][BF4] RT, 16 h

or solvent-free, MW

+ HO

XFe

CO

O

X

X = 4-OCH3 = 4-CH3 = 4-H = 4-Cl = 4-NO2 = 4-Br = 4-CHO

60-100%

Scheme 1.8: Solvent-free synthesis of ferrocenoate esters52

1.2.1.6 Synthesis of ferrocenylimines

Several research groups53-55 have prepared ferrocenylimines by heating a solution of

ferrocencarboxaldehyde with aromatic amines under reflux in anhydrous methanol or

ethanol. Ferrocenylimines have been used for metal coordination and have also

been studied for non-linear optical properties.56 The major setback with the solvent

procedure is the decomposition of the imines during the heating process, leading to

moderate yields.

In highlighting the solvent-free procedure, Nyamori et al.42 have successfully

prepared ferrocenylimines under solvent-free conditions (Scheme 1.9). Grinding of a

mixture of ferrocenecarboxaldehyde and the appropriate substituted aniline with a

glass rod in a glass tube resulted in excellent yields of products. The products were

purified by recrystallization from a minimal amount of cold anhydrous methanol. In

cases where the aniline substituent was an electron-withdrawing group or where the

ferrocenecarboxaldehyde was less reactive, gentle heating (50 ºC) was necessary to

obtain good conversions.

Fe

Y C

O

H+

H2N

X

Grind

Solvent-free Fe

Y C

N

H

X

Y = Ph, Ph-Ph, Ph-O-Ph

87-97%

Scheme 1.9: Solvent-free reactions of ferrocenylaldehydes with aromatic amines42

Page 26: Solvent-free Synthesis of Bisferrocenylimines

12

1.3 Solvent-free synthesis of ligand systems

It is generally accepted that a metal complex is a chemical species which contains a

metal atom or ion bonded to a greater number of ions or molecules than would be

expected from simple valency considerations.57 The ions or molecules that are

bonded or coordinated with the metal are termed ligands. A ligand is regarded as an

ion or a molecule that has a pair of electrons that it can easily donate. The actual

atom through which a ligand is bonded to a metal is called the ligand atom.

Ligands have played an extremely important role in human lives due to their use in

various types of functions, such as in the synthesis of biologically active

organometallic compounds58-62 and in catalysis.63-65 However, the synthesis of the

ligands has always been performed in the presence of a solvent medium. For

example, binaphthols have been prepared from 2-naphthol in the presence of a

catalyst, oxygen or air as an oxidant and chlorinated solvents (Scheme 1.10).66,67

R1

R2

OH

CuCl(OH).TMEDA (1mol%)

O2, CH2Cl2

R1

R2

OH

OH

R2

R1

Scheme 1.10: Synthesis of binaphthol66

Many research groups68-71 have reported on the synthesis of several kinds of

phosphorus ligands in solution. Some methods required the use of very toxic

reagents, such as potassium cyanide,70 working at very low temperature71 and many

reaction steps. Therefore, in order to maintain the concept of sustainable chemistry

the design of new synthetic procedures that are benign to the environment became a

necessity. In this case special attention is given to the solvent-free approach to the

synthesis of ligands.

A good example of the solvent-free approach is that described by Kwong et al.72 on

the phosphination reactions (Scheme 1.11). In that work, arylbromides and triflates

were converted into the corresponding arylphosphines by palladium catalyzed

Page 27: Solvent-free Synthesis of Bisferrocenylimines

13

phosphination with triphenylphosphine under solvent-free conditions. Altough the

yields were an improvement on those reported previously,73 they were still modest.

X

Y2.3 eq. PPh310 mol% Pd(OAc)2

solvent-free, 115°CPPh2

Y

X = Br, OTf

Y = CN, OMe, CHO, COCH3, CO2Me

Scheme 1.11: Solvent-free palladium-catalyzed phosphination reaction72

Another interesting example from the green chemistry perspective is the oxidation of

thiols to disulfides. Thiols and disulfides play an important role in biological

processes.74 Lenardão et al.75 discovered a solid-supported catalyst (Al2O3/KF) for

the solvent-free oxidation of thiols to disulfides (Scheme 1.12). The reactions were

performed either at room temperature, by gently heating or by microwave irradiation.

Without any exception, the microwave assisted reactions proceeded faster and with

higher yields.

R S HAl2O3/KF (40%)

RT, ∆ or MWR S S R

R = Ar = n-C12H25 = HO(CH2)2

48-96%

Scheme 1.12: Solvent-free oxidation of thiols to disulfides75

Vijaikumar and Pitchumani76 have also demonstrated the benefits of the solvent-free

approach in the synthesis of unsymmetrical sulfides from thiols and alkyl halides

using hydrotalcite clays (HT). The proposed mechanism is shown in Scheme 1.13.

Page 28: Solvent-free Synthesis of Bisferrocenylimines

14

HT+ OH-

H2O

HX↑↑↑↑

HT+ X- RS- HT+

RSH

R'SRSR'

Scheme 1.13: A mechanism for the synthesis of unsymmetrical sulfides from thiols

and alkyl halides using hydrotalcite clays76

McCluskey77 developed a method for allylation of carbonyl compounds by

tetraallylstannane in the presence of water as a solvent. This procedure successfully

minimised the environmental impact of tetraallylstannane by allowing the isolation of

inorganic stannane salts and recycling the organic solvents used to extract the

homoallylic alcohol products. A more recent method for allylation of carbonyl

compounds is that reported by Andrews et al.78 This involves the synthesis of

homoallylic alcohols via a metal-mediated reaction of carbonyl compounds with allyl

bromide under solvent-free conditions (Scheme 1.14). In some cases the reactions

are heated or subjected to sonification, and also require quenching of the reaction

with water.

Br +R R'

Oi) M

ii) H2O R'

R OH

M = In, Bi, Zn, CuR, R' = H, Aryl, Alkyl

0-97%

Scheme 1.14: Solvent-free metal mediated synthesis of homoallyl alcohols78

Sabitha et al.79 have also shown that the aminolysis of epoxides with ammonium

acetate can proceed efficiently and regioselectively under solvent-free conditions,

especially when subjected to microwave irradiation (Scheme 1.15). This procedure is

Page 29: Solvent-free Synthesis of Bisferrocenylimines

15

faster and results in higher yields than the corresponding solvent-mediated methods.

The regioselectivity stems from the preferential attack of the nucleophile on the less

hindered carbon atom of the epoxide ring.

R

O NH4OAc, MW

40 - 120 s R

OH

NH2+

R

NH2

OH

major minor

Scheme 1.15: Regioselective microwave assisted solvent-free synthesis of β-

aminoalcohols79

A more recent method for the aminolysis of epoxides is that reported by Placzek et

al.80. This involves the synthesis of β-aminoalcohols via a ring opening of epoxides

with amines in the presence of a scandium triflate catalyst under solvent-free

conditions (Scheme 1.16). The aryl oxiranes underwent cleavage by various amines

in a regiospecific fashion with preferential attack at the benzylic carbon. The catalyst

could be recycled and reused several times.

OOH

NSc(OTf)3 (5 mol%)

solvent-free, RT+ HNR

R'

R

R'

Scheme 1.16: Regioselective Sc(OTf)3 catalyzed solvent-free synthesis of β-

aminoalcohols80

The synthesis of substituted imidazoles is of importance due to their biological

properties. Compounds with the imidazole ring system have many pharmacological

properties and play an important role in biochemical processes. One of several ways

of synthesizing substituted imidazoles is the four-component condensation of

arylglyoxals, primary amines, carboxylic acids and cyanides on Wang resin.81

Balalaie et al.82 have described a novel one-pot, three-component condensation of

benzil, benzonitrile derivatives, and primary amines on the surface of silica gel under

solvent-free conditions (Scheme 1.17). The reaction was accelerated by microwave

irradiation and excellent yields were obtained.

Page 30: Solvent-free Synthesis of Bisferrocenylimines

16

Ph

Ph

O

O

+ + NH2

silica gel

N

N

Ph

Ph

R

PhAr CN

MWR

R = Ar, Alkyl

58-92%

Scheme 1.17: Solvent-free synthesis of tetrasubstituted imidazoles on silica gel

support82

In a recent report, Karimi et al.83 have described another facile method for the

synthesis of tetrasubstituted imidazoles with a silica gel-supported sodium bisulfate

as a catalyst and without any solvent. The reaction involved a four-component

condensation of benzil or benzoin, an aldehyde, amine and ammonium acetate

under microwave irradiation or heating (Scheme 1.18). The method showed

superiority over conventional methods in that it gave excellent yields. Another

advantage of this method is that the catalyst is inexpensive and it avoids problems

associated with catalyst cost, handling, safety and pollution.

Ph

Ph

O

O

+

R

H O

+ R' NH2

NH4OAc/NaHSO4-SiO2

MW or ∆ N

N

Ph

Ph

R' R

Scheme 1.18: Solvent-free synthesis of tetrasubstituted imidazoles on SiO2/NaHSO4

support83

More recently, Adib et al.84 have developed a novel and facile method for the

preparation of 2,4,6-triarylpyridines by reaction of chalcones with ammonium

acetate, with heating under solvent-free conditions (Scheme 1.19). Excellent yields

of triarylpyridine products were obtained and required simple purification techniques.

Page 31: Solvent-free Synthesis of Bisferrocenylimines

17

Ar

O NH4OAc, AcOH (cat.)

100 °C, solvent-free, 4 hAr' N

Ar'

Ar Ar

93-97%

Scheme 1.19: Solvent-free synthesis of 2,4,6-triarylpyridines84

Spurred by the success with the above reactions Adib et al.85 went further to develop

a simple and versatile route for the synthesis of 2,4,6-triarylpyrimidines under

solvent-free conditions and microwave irradiation.

The synthesis of Schiff bases has generally been carried out under reflux in

methanol or ethanol solution. The disadvantage with this method is that more

sensitive Schiff bases tend to undergo some degree of decomposition. Recently,

Naeimi et al.86 developed a mild and convenient route of the reaction of carbonyl

compounds with amines, under solvent-free conditions and in the presence of a

catalyst. They also prepared double Schiff bases and obtained low to excellent

yields, depending on the amine used (Scheme 1.20). The only significant

disadvantage of this method is that the catalyst (P2O5/Al2O3) is moisture sensitive.

R

OH

R'

O+ H2N

Y

NH2

P2O5/Al2O3

Solvent-free

R

OH

R'

N

Y

N

R'

HO

R

R = H, NO2

R' = H, CH3

Y = C6H4-O-C6H4, C6H4-CH2-C6H4, C6H4-SO2-C6H4

Scheme 1.20: Solvent-free synthesis of Schiff bases86

1.4 Nitrogen-donor ligand chemistry

P-donor ligands have received a lot of attention in the fields of organometallic

chemistry and catalysis.87 It was only in the mid 1990s that special attention was

Page 32: Solvent-free Synthesis of Bisferrocenylimines

18

drawn towards N-donor ligands.88 N-donor ligands have a larger range than that of

any other atom and their organic chemistry is varied. The best way to classify N-

donor ligands could be based on the hybridisation of the nitrogen atom, i.e., sp3, sp2

and sp. Ligands with sp2-hybridised nitrogen, such as imines and pyridines, have a

significant coordination chemistry.87 Examples of some ligands containing sp2-

hybridised nitrogen atoms are summarized in Figure 1.2.

N NN N N

N NH

N

R N N R N

N N

N

Figure 1.2: Some N-donor ligands with sp2-hybridised nitrogen atoms

The comparison of the well-studied phosphorus ligands and the nitrogen ligands can

be done by noting the differences in coordination behaviours as well as the

properties of the N-donor ligands.87

(i) The coordination bonds formed by N-donor ligands are fairly stronger than

those formed by P-donor ligands and the strength of the bonds depends

largely on their σ covalency with potentially significant contribution from

the ionic character of the bond itself.87

(ii) The strength of the M-N bonds will be affected much more by steric effects

than that of the corresponding M-P bonds.87

(iii) The N-donors are generally not as effective in forming low-spin complexes

and consequently form thermodynamically less stable species but more

kinetically labile than their low-spin P-donor analogues.87

(iv) N-donor ligands generally show only a limited π-back bonding ability,

making these ligands less suitable for the stabilization of low oxidation

state transition metals. However, sp2-hybridised N-donors such as

pyridine, have been known to show some π-back bonding effects between

the nitrogen heterocycle and the metal centre.89,90

Page 33: Solvent-free Synthesis of Bisferrocenylimines

19

(v) The N-donor ligands exert a very small trans effect in comparison to other

ligands used in organometallic chemistry, resulting in a fast rate of

substitution reactions.87

(vi) The observed reactivity of nitrogen donor ligands is generally high,

particularly with alkyl complexes of transition metals containing these

ligands.87

As it has been mentioned above that N-donor ligands have a larger range than any

other atom, therefore, a few examples of N-donor ligand catalyzed reactions will be

chosen, particularly those reactions catalyzed by transition metal complexes

containing pyridines and Schiff bases (imines). The examples of reactions that will

be focused on include the allylic alkylation, olefin polymerization, cross coupling,

epoxidation and ring opening metathesis polymerization (ROMP).

1.4.1 Polymerization reactions

Polymerization reactions required that for the catalysts to be more efficient, the

following conditions for the catalysts must apply.

The catalyst must:

(i) have high-olefin insertion ability.91

(ii) have two available cis-located sites for polymerization.91

(iii) Be stable enough under the usual polymerization conditions.91

The polymerization was catalyzed efficiently by titanium complexes containing

bis(alkylphenylaminopyridinato) ligands (Figure 1.3(a)-(e)).92 The ligands were

prepared by the reaction of 2-chloropyridine and the appropriate alkylaniline

hydrochlorides. The catalysts were formed at room temperature on reaction of the

ligands with titanium tetrachloride (TiCl4).

Page 34: Solvent-free Synthesis of Bisferrocenylimines

20

N N RTi

N NR

ClCl a) R = 2-Etb) R = 3,5-Mec) R = 4-n-Bud) R = 2-t-Bue) R = Ph

I

Figure 1.3: Bis(alkylphenylaminopyridinato) titanium complexes92

The complex Ie (Figure 1.3) exhibited the highest polymerization activity but

produced the lowest molecular weight of polyethylene compared to complexes Ia-d.

The decrease in polymerization activity for complexes Ia-e was attributed to the

electron donating effects of the alkyl substituents. It was also concluded that as the

distance between the alkyl substituent and the metal centre increased, the more

active was the catalyst.92

A series of new zirconium complexes bearing bis(phenoxyketimine) ligands have

been prepared in low to moderate yields (>70%) (Scheme 1.21).93 The effects of the

substituents on the imine carbon and on the phenyl ring of the aniline moiety on

polymerization activity were investigated. It was found that when R1 is H atom the

molecular weight of polyethylene was low (Mw = 12 000) and the activity 1.6 kg PE

mmol-1 Zr .h-1 at 20 ºC. The best results were obtained when the substituent R1 was

electron donating and substituent R2 was an electron withdrawing group.

NR1

R2

O

ZrCl2

2

But But

NR1

R2

OH

But But

1) n-BuLi

2) 0.5 ZrCl4

Scheme 1.21: Synthesis of bis(phenoxyketimine) zirconium complexes93

Cloete et al.94 have synthesized functionalized pyridinylimine complexes of palladium

as precursor catalysts for ethylene polymerization. The nitrogen atom of the imine

functionality contained various substituents, ranging from alkyl to aromatic and allylic

groups (Figure 1.4, II).

Page 35: Solvent-free Synthesis of Bisferrocenylimines

21

N

II

a: R = allylb: R = styrylc: R = phenold: R = phenyle: R = propyl

N

N NAr Ar

R'

RR

R'

a: R = Me, R' = H, Ar = Mesb: R = R' = Me, Ar = MesC: R = CH2Ph, R' = H, Ar = Mesd: R = R' =CH2Ph, Ar = Mese: R = Me, R' = H, Ar = DIPPf: R = R' = Me, Ar = DIPPg: R = R' = H, Ar = Mes

Mes = mesityl (2,4,6-trimethylphenyl)DIPP = 2,6-diisopropylphenyl

NRPd

Cl Cl

Fe

Cl Cl

N

N NFe

R2

R1

X

R1

R2

X

R1, R2 = H, MeX = Br, I

Cln

III

IV

Figure 1.4: Pyridyl-imine complexes of iron (Fe) and palladium (Pd).94,97,98

The palladium complex IIc exhibited the highest activity at Al:Pd of 2000:1 and whilst

complexes IIb reached the optimal activity at Al:Pd of 1500:1. The reason complex

IIc required a higher amount of methylalumoxane (MAO) was due the fact that the

hydroxyl group on the aromatic ring reacted with MAO to form an Al phenoxide

adduct.94

The iron-based bis(imino)pyridine complexes have been prepared95-98 and have

been used as catalysts for olefin polymerization.97,98 Upon treatment with MAO, all

complexes became active in the polymerization of ethylene. Complexes that

contained the 2,4,6-trimethylphenyl (Figure 1.4, IIIa-d and IIIg) attached to the imine

nitrogen atom were more productive than those that contained the 2,6-

diisopropylphenyl (Figure 1.4, IIe and IIf). This effect was attributed to less

congested active sites for 2,4,6-trimethylphenyl derivatives resulting in higher

propagation rates.97 The replacement of the ethyl group in IIe with the isopropyl

group to form IIf led to increased polymer molecular weights [Mw 26 000 (IIe), Mw

263 000 (IIf)].

Page 36: Solvent-free Synthesis of Bisferrocenylimines

22

The influence of the para-substituent in bis(arylimino)pyridine iron complexes on the

catalytic oligomerization and polymerization of ethylene was investigated.98 These

complexes contained phenyl rings bearing halogen (Br, I) substituents at para-

positions (Figure 1.4, IV). The complexes where R1 was a methyl group, R2 a

hydrogen and bearing a strongly electron withdrawing group at para-position were

found to have the highest activities in oligomerization. On the other hand, those

complexes where R1 and R2 were methyl groups and bearing a halogen at a para-

position were highly active in polymerization. The extremely high polymerization

activities of 4-halo-2-methyl substituted complexes was due to the electronic

influence of the halogen at the para-position.98 The Fe(III) analogues of these

complexes show more enhanced activities due to increased Lewis acid strength and

therefore the coordination of an ethylene molecule is facilitated.98

1.4.2 Cross-coupling reactions

1.4.2.1 Heck reactions

Two pyridine bridged dicarbene palladium(II) complexes (Figure 1.5, V) that were

efficient for the Heck coupling reaction have been reported by Nielsen and co-

workers.99 These complexes were prepared via a dinuclear silver(I) complex, which

was synthesized from the reaction of the tridentate 2.6-bis[(3-methylimidazolium-1-

yl)methyl] pyridine dibromide and silver(I) oxide in dichloromethane or N,N’-

dimethylsulfoxide. The compound V (X = Cl) exhibited a higher activity towards the

coupling reaction of n-butyl acrylate and 4-bromoacetophenone when treated with

hydrazine hydrate, in the absence of the quaternary ammonium salt.99 The main

product of the coupling reaction was the expected n-butyl-(E)-4-acetylcinnamate,

without any detection of the Z isomer by 1H NMR and gas chromatography.99

Page 37: Solvent-free Synthesis of Bisferrocenylimines

23

N

N N

N N

Pd

X

X = Cl, Br

BF4

O

O

O

Si(CH2)3N

NPd

Cl

Cl

N N

Pd

ClCl

R1

R2

N N

Pd

R1

R2

E

E

R1 = H, R2 = i-PrR1 = Me, R2 = i-PrR1 = H, R2 = t-Bu E = CO2Me

VVI

VII VIII

Figure 1.5: Examples of Pd-pyridyl complexes used for the Heck reactions99

Horniakova et al.100 have reported on pyridylimine palladium(II) complex immobilized

on a mesoporous silica (Figure 1.5, VI). This complex is heterogeneous in nature

and could be used as an alternative to the homogeneous catalysts for the Heck

reaction as well as the Suzuki reaction.100 The activity of this complex was compared

with quiniline-imine palladium(II) complex. Both catalysts were found to be highly

active in the Heck reaction with 100% conversions of the arylhalides and high

selectivity towards the E isomer was also observed.

Palladium(II) (Figure 1.5, VII) and palladium(0) (Figure 1.5, VIII) complexes based on

pyridyl-imine ligands have been synthesized and used as catalyst precursors for the

Heck reaction.101 As expected all the ligands were active and promoted the complete

conversion of iodobenzene into trans-methyl cinnamate (Z isomer). For the same

ligand, the palladium(II) complexes were more active than their palladium(0)

analogues.101 Although not mentioned in some articles,99,102 it was apparent that

leaching of palladium was a major problem for the homogeneously-catalyzed Heck

reactions.101,103

Page 38: Solvent-free Synthesis of Bisferrocenylimines

24

1.4.2.2 Suzuki cross-coupling reactions

A range of nickel(0) complexes (Figure 1.6) that contain pyridine, 2,2’-bipyridine, 2-

(2-oxazolyl)pyridine and 2,2’-bisoxazole ligands have been prepared and their

catalytic activity towards cross-coupling of aryl chlorides by intramolecularly

stabilized dialkylaluminium reagents has been studied.104

N N

Ni

N N

N N

O

Ni

NN

O

N

O

Ni

O

N

N

O O

N

N N

Ni

N N

IX X

XI XII

Figure 1.6: Examples of Ni(0) complexes used for cross-coupling of arylchlorides104

Complexes IX and X exhibited an increase in activity in the presence of THF as a

solvent, which enhanced the homo-coupling problem.104 It was also observed that IX

and X efficiently catalyzed the cross-coupling of chloroarenes without

hydrodehyalogenation occurring.

Some silica-supported imine palladacycles have been reported by Bedford and co-

workers.105 The objective was to study their catalytic activity as potential

heterogeneous catalysts for the Suzuki cross-coupling reaction, and their

recyclability was also tested. The synthesis of silica-supported imine palladacycles is

illustrated in Scheme 1.22.

Page 39: Solvent-free Synthesis of Bisferrocenylimines

25

OHC

Br

H2N(CH2)3Si(OEt)3

BrNR

R = (CH2)3Si(OEt)3

EtOH, molecular sieves

Pd(dba)2, toluene

60°C

RN

R = (CH2)3Si(OEt)3

Pd

Br

2

PPh3, CH2Cl2

RN

R = (CH2)3Si(OEt)3

Pd

Br

PPh3

N Pd

Br

2

silica, toluenereflux temperature

Si

HOOHOH

Si Si Si

N Pd

Br

PPh3

Si

HOOHOH

Si Si Si

PPh3, CH2Cl2

XIIIXIV

XVXVI

Scheme 1.12: Synthesis of silica-supported imine palladacycles.105

The silica-supported palladium complexes were very poor in terms of their catalytic

activity towards the Suzuki coupling reaction.105 In addition, the complexes were

easily recycled but the activity deteriorated in the subsequent reactions. The

deterioration in activity was due to the degradation of the catalysts since deep red-

brown palladium nanoparticles, which subsequently decomposed to palladium black,

were observed.105

Page 40: Solvent-free Synthesis of Bisferrocenylimines

26

SN R

SNR

PdCl

Cl

SN R

SNR

PdCl

Cl

R = Ph

XVIIXVIII

R = Mes = n-Pr = (S)-MeCHPh

Figure 1.7: Palladium imine and amine complexes for coupling of aryl bromides.106

Wiedermann et al.106 have synthesized palladium imine and amine complexes

(Figure 1.7) and investigated their activity towards the Suzuki cross-coupling of aryl

bromides with phenylboronic acid (Scheme 1.23).

Br

R

+

B(OH)2

catalyst, 2 eq. Cs2CO3

dioxane R

Scheme 1.23: Suzuki cross-coupling of aryl bromides with phenylboronic acid.106

Complex XVIII (R = Mes) showed a higher catalytic activity than all the other

complexes.106 The bulkier mesityl group was responsible for rendering the catalyst

more active.106

1.4.3: Epoxidation reactions

A 2-(phenylazo)pyridine rhuthenium(II) complex has been synthesized by Barf and

co-workers.107 and their catalytic activity in the epoxidation of stilbene was

investigated. Barf and co-workers were able to isolate three isomers, namely trans-

trans-trans (γ), trans-cis-cis (α) and cis-cis-cis (β) (Figure 1.8).

Page 41: Solvent-free Synthesis of Bisferrocenylimines

27

N N

N

a

b

2-(phenylazo)pyridine (pap)

Ru

Nb

Nb

Na

Na

Cl

Cl

Ru

Nb

Nb

Na

Na

Cl

Cl

Ru

Nb

Nb

Na

Na

Cl

Cl

αααα (t c c) ββββ (c c c) γγγγ (t t t)

Figure 1.8: Three isomers of Ru(pap)2Cl2107

The two isomers of Ru(pap)2Cl2, β and γ resulted in 100% conversion of stilbene

with good selectivity.107 The α-isomer which was the most stable gave both lower

conversion and selectivity. This was attributed to the steric hinderance of the α-

isomer as was evidenced by the X-ray crystal structure.107

Another interesting example was the use, by Moutet and Ourari,108 of Mn(III) Schiff-

base complexes for the electrocatalytic epoxidation and oxidation of cyclooctene.

The complexes are shown in Figure 1.9 below. The activity of the catalysts was

dependent on the bridging group Z, i.e., the activity decreased from Salen to Salch to

Sal(Cl)2ophen complexes. The effect was due to the decrease in stability of the

ligand as a result of the increased rigidity.108 The best results for epoxidation of

cyclooctene were obtained when salen complex and 2-methylimidazole were used.

Recently, Bruno et al.109 have reported on dioxomolybdenum(IV) complexes bearing

a bidentate and tetradentate salen-type ligands, that were similar to those prepared

by Barf et al.107

Page 42: Solvent-free Synthesis of Bisferrocenylimines

28

NZ

N

O OR1

R2 R2

R1

Mn

Cl

Salen

Z = (CH2)2

Salchx

Z =

Salophen

Z =

Sal(Cl)2ophen

Z =

Cl Cl

R1 = R2 = HR1 =Cl, R = HR1 = R2 = ClR1 = OCH3, R2 = HR1 = NO2, R2 = H

Figure 1.9: Mn(III) Schiff-base complexes for electrocatalytic epoxidation of

olefins108

An octahedral titanium binaphthyl-bridged Schiff-base complex110 has been prepared

and used as a catalyst for the regio- and stereoselective epoxidation of allylic

alcohols, under microwave-mediated solvent-free conditions. The cis-β isomer was

the preferentially formed product according to the X-ray crystal structure (Figure

1.10).110

Ti

N

N O

Cl

Cl

O

But

cis-ββββ isomer

N

N

=N

N

Figure 1.10: The cis-β isomer of the binaphthyl-bridged Schiff base titanium

complex110

The epoxyalcohols could be obtained with very high regio- and chemoselective way

by the microwave irradiation of the mixture under solvent-free conditions.

Furthermore, high diastereoselectivity could be observed for secondary allylic

alcohols.110

Page 43: Solvent-free Synthesis of Bisferrocenylimines

29

1.4.4 Asymmetric allylic substitution reactions

Palladium complexes based on the chiral pyridine ligands have been reported111,112

and their catalytic activity in asymmetric allylic alkylation of 1,3-diphenylprop-2-enyl

acetate with dimethyl malonate has been studied. The catalysts were prepared in

situ by a reaction of allylpalladium chloride dimer [Pd(η3-C3H5)Cl]2 with the ligands

2,2’-bipyridines, 2,2’,2”-terpyridines and 1,10-phenanthrolines (Scheme 1.24).111

Ph Ph

OCOCH3 CH2(CO2CH3)2

[Pd(η3-C3H5)Cl]2 / ligandPh Ph

CH(CO2CH3)2

*

Scheme 1.24: Allylic alkylation of 1,3-di[henyl-2-enyl acetate with dimethyl

malonate111,112

Phenanthroline catalysts were the most active and their enanioselectivity depended

on the distance of the chiral substituent from the heterocyclic nitrogen. On the other

hand, the enantioselectivity of bipyridines increased when there was a bulky

substituent at the 6-position.111

A series of chiral bidentate ligands containing thiazolyl and pyridyl donors have been

synthesized and used for in situ preparation of copper(I) complexes.113 The

subsequent copper(I) thiazolyl-pyridine complexes were used for the

enantioselective allylic oxidation of cyclohexene with t-butyl perbenzoate. Examples

of the chiral bidentate thiazolyl-pyridyl ligands are shown in Figure 1.11. Complexes

XX gave the highest yields of the product, but with low enantioselectivity, while

complexes XXI exhibited higher enantioselectivity. The reason for the higher

enantioselectivity of the latter complexes was attributed to steric hindrance at the 8-

position of the tetrahydroquinoline ring and the thiazole ring. Ligands with a

substituent at the 8-position of tetrahydrquinoline led to the (R)-configuration and

others led to (S)-configurations.113

Page 44: Solvent-free Synthesis of Bisferrocenylimines

30

N

S

N

R

XIX

R = H, CH3

N

S

N

R

XX

R = H, n-Pr, CH2Ph, CH2SiMe3

N

S

N

R1

XXI

R1 = R2 = HR2

R1 = CH3, R2 = HR1 = i-Pr, R2 = HR1 = H2 = n-PrR1 = R2 = n-BuR1 = R2 = CH2PhR1 = R2 = CH2SiMe3

Figure 1.11: Examples of chiral bidentate thiazolyl-pyridine ligands113

The chiral diimine palladium(II) complexes (Figure 1.12) have been prepared for

asymmetric alkylation of 1,3-diphenylprop-2-enyl acetate with dimethyl malonate by

Albano et al.114

N

N

S

S

S

S

R

R

PdBF4

Figure 1.12: Chiral diimine palladium(II) catalyst for asymmetric alkylation114

These complexes were less active than their diamine analogues due to the steric

hindrance that prevents nucleophilic attack and the electronic issues taking part

during the overall oxidative-reductive catalytic cycles.114

1.4.5 Ring-opening metathesis polymerization (ROMP)

Matos et al115 have prepared complexes of the type [RuCl2(PPh3)2L2], where L =

pyridine, 4-methylpyridine (picoline), 4-aminopyridine and isonicotinamide. The

Page 45: Solvent-free Synthesis of Bisferrocenylimines

31

complexes were investigated for catalytic activity in ring-opening metathesis

polymerization. The reaction pathway of ROMP is illustrated in Figure 1.13.

M CHR

C C

n

C C

M CHR

C C

C CM M

C CM C C CHR

n

Figure 1.13: The reaction pathway of ROMP116

All catalysts were active only in the presence of ethyldiazoacetate (EDA).

Isonicotinamide and aminopyridine gave unimodal polymers with highest molecular

weights Mw, but with slightly lower polydispersity indices Mw/Mn. Pyridine and picoline

gave bimodal polymers.115

Schiff base substituted first and second generation Grubbs catalysts have described

for the polymerization of cyclooctadiene.117 These catalysts showed greater stability

and activity than their parent first generation Schiff base catalysts at high

temperature but much lower activity than their parent second generation Grubbs

catalyst.117

Wright et al.118 have prepared two catalysts that were derivatives of the first

generation Grubbs catalyst. Preparation of the catalysts is illustrated in Scheme

1.25. Complex XXIII was prepared via in situ formation of the ligand (Scheme 1.25b).

Page 46: Solvent-free Synthesis of Bisferrocenylimines

32

Ru

P(i-Pr)3

P(i-Pr)3

Cl

Cl Ph

+NN N

N NAr Ar

toluene, RT

68%N

N

N

N

N

Ru

Cl

Cl Ph

Ar

Ar

(a)

(b)

N NN Ar

NKHMDS

THF N NN Ar

N

+

Ru

P(i-Pr)3

P(i-Pr)3

Cl

Cl Ph

THF

Petrol

N

N

N

N

N

Ru

Cl

Cl

Ar

Ar

Ph

XXII

XXIII

Scheme 1.25: Synthesis of Ru complexes derived from 1st generation Grubbs

catalyst.118

Complex XXII exhibited lower activity than the first and second generation catalyst

due to limtations of C-N-C ligand design, particularly on late 4d and 5d transition

metals.118 The limitations originated from low lability of metal-nitrogen heterocyclic

bonds and the blocking of 3d coordination sites by rigid ligands.118 Catalytic studies

of XXIII were hampered by very low yields, coupled with the difficulties to obtain the

free carbene.

1.5 Ferrocenyl-nitrogen donor ligand chemistry

1.5.1 Ferrocenyl-pyridine ligands

Despite the extensive work on the symmetrical ferrocenylphosphine119 and

unsymmetrical ferrocenyl ligands,120 it was apparent from reviewing the literature that

the chemistry of ferrocenyl-nitogen donor ligands, particularly those containing

pyridine, has not been well developed. One of the first ferrocenylpyridine complexes

was described by Miller et al.121 (Figure1.14a). Electrochemical properties of the

Page 47: Solvent-free Synthesis of Bisferrocenylimines

33

complexes were investigated to determine whether the changes in the oxidation

state of redox ferrocenyl ligand could lead to changes in the reactivity of the rhenium

centre.

Fe

Fe

N

N

Re

Cl

CO

CO

CO

Fe

Fe

N

N

Pt

Cl

Cl

(a) (b)

Figure 1.14: Re and Pt complexes with ferrocenylpyridine ligands121,122

A platinum complex has been prepared, that contain 3-ferrocenylpyridine and the

specific role of platinum as a linker between the ferrocenyl groups has been studied

(Figure 1.14b).122 A single redox wave was observed, corresponding to a two

electron oxidation at the platinum electrode indicating little or no communication

between the ferrocene groups.

Sakanishi et al.123 have reported complexes containing ferrocenylpyridine ligands

attached to a tungsten metal centre (Figure 1.15).

Fe

Fe

FeN W(CO)5

R

R = H, Me, Ph

XN W(CO)4

X = CH, N

N

W(CO)5

OFe N W(CO)5

R

R = H, Me2

Figure 1.15: Tungsten complexes with ferrocenylpyridine ligands123

Page 48: Solvent-free Synthesis of Bisferrocenylimines

34

The spectroscopic and electrochemical studies showed that these complexes have

architectural and electronic properties necessary to exhibit second-order non-linear

optical behaviour. Some complexes that were characterised by X-ray methods

crystallized in a centrosymmetric space group and thus could not show second-order

non-linear optical properties in the bulk state.123

Braga et al.38,39,124 have described the preparation of supramolecular organometallic

materials and coordination networks based on 1,1’-bis(4-pyridyl)ferrocene. Example

of these ligands are shown in Figure 1.16.

Fe

Fe

Fe

N

N

N

N

N

N

N

N

N

Figure1.16: Examples of ferrocenylpyridyl and pyrimidyl complexes38,39,124

The platinum and palladium complexes of 1,1-bis(2-pyridyl)ferrocene have been

prepared and their catalytic reactivity towards carbonyl insertion reactions has been

investigated (Figure 1.17).125

Fe

N

N

Pd

Me

ClFe

N

N

Pt

Cl

Cl

Pt

Cl

Cl

Figure 1.17: Pd and Pt complexes of 1,1’-bis(2-pyridyl)ferrocene125

A rapid insertion of carbon monoxide into the palladium-methyl bond was observed.

However, detailed kinetic studies on the reaction were impossible because of weak

ligand-metal bonding.

Page 49: Solvent-free Synthesis of Bisferrocenylimines

35

Rajput et al.126 have prepared some palladium(II), platinum(II), rhodium(I) and

iridium(I) complexes. Cytotoxicity studies were carried out on selected complexes

and these were also screened for activity against oesophageal and cervical cancer

cell lines. Complexes with a significant activity in an initial screening assay, and

which were soluble in the culture media were further tested to determine their IC50

values. These were compared to IC50 values of cisplatin, determined in the same

way and for the same cell line. Interestingly, some of the complexes displayed

similar growth inhibitory activity to that displayed by cisplatin. In a more recent

publication, Rajput et al 127 have reported on the electrochemical properties of the

rhodium(I) complexes. Results obtained showed some communication between

metal centres or a lack of communication in the complexes.

A series of pyridyl- and quinolyl-N-substituted ferrocenyl and ferrocenediyl ligands

have been synthesized by Gibson and co-workers.128 The activity of the ligands in

olefin polymerization was tested after coordination to palladium and nickel metal

centres. Examples of the palladium and nickel complexes are depicted in Figure

1.18.

Fe

N

N

N

N

X Y

X YFe

N

N N

N

X YX Y

Fe

N N

X YFe

N N

X Y

R

R

R

M = Pd or NiR = H, MeX = Y = Cl, Br

M = Pd or NiX = Y = Cl, Br

M = Pd or NiR = H, MeX = Y = Cl, Br M = Pd or Ni

X = Y = Cl, Br

Figure 1.18: Palladium and nickel complexes128

The palladium complexes were found to be inactive, but with the nickel-based

complexes, although no high molecular weight polyethylene was formed.128 The

nickel complexes were highly selective for the formation of short chain oligomers.

Page 50: Solvent-free Synthesis of Bisferrocenylimines

36

1.5.2 Ferrocenyl-Schiff base ligands

Recently, there has been an upsurge of interest in preparing ferrocenyl-nitrogen

donor ligands based on Schiff bases.124 These are very interesting compounds since

the presence of the ferrocenyl moiety imparts special electrochemical

properties129,130 and modes of coordination.130 Some of the ferrocenyl-Schiff base

ligands have been studied for second-order non-linear properties.131 Several

research groups have reported on the preparation of cyclopalladated complexes

based on ferrocenylimines132-138 and these complexes have been known to catalyse

cross-coupling reactions, namely Heck and Suzuki reactions. An example is the

dimeric complex prepared by Wu et al.139. The palladium centre is stabilized by a

five-membered ring that incorporates the C=N bond (Figure 1.19). Triethylamine and

1,4-dioxane were found to be the suitable base and solvent, respectively. The

catalysts were highly efficient for the olefination of iodobenzene and resulted in

highly regioselective products (trans-isomers).

Fe

X = Cl, Br

N CH3

PdX

2

CH3

Figure 1.19: Dimeric cyclopalladated ferrocenylimine complex for catalytic Heck

reaction139

Zhao et al140 have also reported on the catalytic activity of a cyclopalladated

ferrocenylimine complex in the Mizoroki-Heck reaction for the arylation of

iodobenzene (Figure 1.20). In this complex the ring is formed by nucleophilic attack

of the palladium by the carbon on the second Cp ring and does not incorporate the

C=N bond. The complex was highly active and showed high regioselectivity for trans-

coupling.

Page 51: Solvent-free Synthesis of Bisferrocenylimines

37

FeN

Pd S

Ph3P Cl

Figure 1.20: Monomeric cyclopalladated ferrocenylimine for Mirozoki-Heck

reaction140

Weng et al.141 have described the preparation of a 1,1’-N-substituted ferrocenediyl

palladium(II) complex for the catalysis of Suzuki cross-coupling reaction of aryl

iodides bromides with aryl boronic acids, under non-homogeneous conditions in

aqueous medium (Figure 1.21). The catalyst exhibited a high activity in the coupling

of 4-bromoacetophenone and phenylboronic acid. The catalyst could be easily

recovered, recycled and could be used over a number of runs without loss of activity.

Fe

N

N

PdCl

Cl

Figure 1.21: 1,1’-N-substituted ferrocenediyl Pd(II) complex for Suzuki cross-

coupling reaction141

Zhang et al142 have reported on the synthesis and catalytic activity of the

cyclopalladated ferrocenylketimine complexes for the Suzuki cross-coupling reaction,

under ultrasonic irradiation in aqueous medium (Figure 1.22). Compared to

conventional heating, the reaction occurred faster when accelerated by ultrasonic

irradiation although the yields were comparable. Gong et al.143,144 have reported on

the catalytic activity of mono-and dimeric cyclopalladated ferrocenylimine complexes

for the Suzuki cross-coupling reaction143 and the Buchwald-Hartwig coupling of

amines with aryl halides or sulfonates.144

Page 52: Solvent-free Synthesis of Bisferrocenylimines

38

Fe

N

CH3

R

PdX

2

R = X = ClR = Me, X = I

Figure 1.22: Dimeric cyclopalladated ferrocenylketimine complexes for Suzuki

reaction143,144

Chiral ferrocenylphosphine-imine ligands containing a pyridine unit have been

reported by Hu et al.145 On coordination to a palladium metal centre, highly active

catalysts for the asymmetric allylic alkylations. Of particular interest was the

influence of the position of the pyridine nitrogen atom on the reactivity and

enantioselectivity. The presence of a pyridine unit significantly affected the way a

ligand coordinated to the metal centre, which in turn led to dramatic changes in the

reactivity and enantioselectivity of the catalytic reaction.145 The ligand with a 3-

pyridine nitrogen atom (Figure 1.23) resulted in increased reactivity and

enantioselectivity, while the ligand with a 2-pyridine nitrogen atom gave no alkylation

product.

Fe

PPh2

N N

(S,Sp)

Figure 1.23: Chiral ferrocenylphosphine-imine ligand145

Other examples of allylic alkylations catalysts are those prepared by Platero-Prats et

al.146 These compounds were synthesized by treatment of a ferrocenylimine with an

appropriate chloro-bridged dimeric palladium(II)-allyl complex in dichloromethane

solution at 25 ºC. The complexes were active in the allylic alkylation of (E)-3-phenyl-

2-propenyl (cinnamyl) acetate with sodium diethyl-2-methylmalonate as a

nucleophile.

Page 53: Solvent-free Synthesis of Bisferrocenylimines

39

N-ferrocenyl salicylaldimine ligands have been synthesized by Bott et al.147 (Scheme

1.26). After coordination to magnesium, titanium and zirconium metal centres, their

catalytic activity in polymerization of olefins was investigated. It was found that on

activation with methylalumnoxane (MAO), the titanium complex showed moderate

activity for ethylene polymerization, while the zirconium complex was highly active

for ethylene oligomerization.

Fe

N

HO ButFe

NH2

+OHC

OH

But EtOH

-H2O

Scheme 1.26: Synthesis of N-ferrocene salicylaldimine ligand147

Several groups have reported on the synthesis of bisferrocenylimines from

ferrocenecarboxaldehyde148-151 or 1,1’-diformylferrocene149,151,152 a range of amines

or anilines. Lee et al153 have synthesized cationic rhodium(I) and iridium(I)

complexes based on bisferrocenylimine for polymerization of phenylacetylene

(Figure 1.24).

Fe

CH

(CH2)

Fe

CH

N2

N

MClO4

M = Rh, Ir

Figure 1.24: Cationic rhodium(I) and Ir(I) complex for olefin polymerization153

Compared to its iridium(I) analogue, the rhodium(I) complex gave high molecular

weight polymers in excellent yields while the iridium(I) complex gave better

polydispersity indices.

Page 54: Solvent-free Synthesis of Bisferrocenylimines

40

1.6 Objectives of the project

The elimination of the hazardous materials in synthetic processes is important

especially in the chemical industry and in academia. The versatility of the solvent-

free approach to organic synthesis has been extensively illustrated with examples in

the previous sections.

This project focuses on developing a synthetic procedure that reduces or eliminates

the use of organic solvents using the green chemistry principles. The main objective

of the project is to prepare nitrogen-donor ligands based on ferrocene that can be

used in catalysis. Our group42 has previously reported on successful preparation of

ferrocenylimines under the solvent-free environment. We intend taking this work a

step further and prepare bisferroceylimines ligands for olefin polymerization. The

ligands will then be coordinated to a rhodium(I) metal centre. Lee et al.153 have

demonstrated that rhodium(I) complexes containing bisferrocenylimines are very

effective catalysts for olefin polymerization (Figure 1.24). Based on the work of Lee

et al.153 we will investigate the effect of increasing the alkyl chain length on the

catalytic activity of the complexes.

Rhodium(I) complexes containing diamine ligands have been found to be effective in

hydroformylation reactions. Kim and Alper154 have demonstrated that these

complexes are effective for hydroformylation reactions (Figure 1.25).

Rh

N N

Rh

Cl Cl

Figure 1.25: Cationic rhodium(I) complexes for hydroformylation reactions.154

We intend to prepare similar complexes with bisferrocenylamine ligands and

investigate their catalytic activity towards hydroformylation reactions. The

bisferrocenylamine ligands will be obtained by simple reduction of

Page 55: Solvent-free Synthesis of Bisferrocenylimines

41

bisferrocenylimines, which proceeds much cleaner than the lithium aluminium

hydride route.

1.7 References

1. W. G. Foster, M. S. Neal and E. V. Youglai, International Congress Series,

1266 (2004) 126.

2. http://pubs.acs.org/cen/news/8235/8235sasol.html

3. http://www.southafrica.info/doing_business/businesstoday/businessnews/501

166.htm

4. P. T. Anastas and J. C. Warner, Green Chemistry: Theory and Practice,

Oxford University Press Inc., New York, 1998.

5. C. Reichardt, Solvents and Solvent effects in Organic Chemistry, Wiley-VCH,

Weinheim, 2004.

6. J. F. Coetzee and C. D. Ritchie, Solute-Solvent Interactions, Marcel Dekker,

New York, 1969.

7. G. V. W. Cave, C. L. Raston and J. L. Scott, Chem.. Commun., 2001, 2159-

2169.

8. K. Tanaka, Solvent-free Organic Synthesis, Wiley-VCH, Weinheim, 2003.

9. M. K. Beyer and H.Clausen-Schaumann, Chem. Rev., 105 (2005) 2921.

10. M. Carey Lea, Amer. J. Sci., 46 (1893) 413.

11. M. Carey Lea, Phil. Mag., 34 (1892) 46.

12. R. Ling and J. L. Baker, J. Chem. Soc., 63 (1893) 1314.

13. R. P. Rastogi and B. L. Dubey, J. Am. Chem.. Soc., 89 (1967) 2000.

14. R. P. Rastogi, P. S. Bassi and S. L. Chadha, J. Phys. Chem., 66 (1962) 2707.

15. R. P. Rastogi and P .S. Bassi, J. Phys. Chem., 68 (1964) 2398.

16. A. O Patil, D. Y. Curtin and I. C. Paul, J. Am. Chem. Soc., 106 (1984) 348.

17. F. Toda, K. Tanaka and A, Sekikawa, J. Chem. Soc., Chem. Commun., 1987,

279.

18. F. Toda, H. Takumi and H. Yamaguchi, Chem. Exp., 4 (1989) 507.

19. F. Toda, K. Tanaka and H. Hamai, J. Chem. Soc., Perkin Trans., 1 (1990)

3207.

20. G. Kaupp and J. Schmeyer, Angew. Chem., 32 (1993) 1587.

Page 56: Solvent-free Synthesis of Bisferrocenylimines

42

21. D. Villemin and B. Martins, Synth. Commun., 25 (1995) 3135.

22. B. C. Ranu, M. Saha and S. Bhar, Synth. Commun. 27 (1997) 621.

23. R. S. Varma, R. Dahiya and S Kumar, Tetrahedron Lett., 38 (19970 2039.

24. R. S. Varma, K. P. Naicker and P. J. Liesen, Tetrahedron Lett., 39 (1998)

8437.

25. H. Firouzabadi and M. Abbassi, Synth. Commun., 29 (1999) 1485.

26. J. L. Scott and C. L. Raston, Green Chem., 2 (2000) 245.

27. M. DellaGreca, P. Monaco, L. Previtera, A. Zarrelli, A. Fiorentino, F. Giordano

and C. Mattia, J. Org. Chem., 66 (2001) 2057.

28. G. Wilkinson, J. Organomet. Chem., 100 (1975) 273.

29. S. Onaka, A. Mizuno and S. Takagi, Chem. Lett., 1989, 2037.

30. T. S. Hor and L.-T. Phang, J. Organomet. Chem., 390 (1990) 345.

31. M. Bracci, C. Ercolani, B. Floris, M. Bassetti, A. Chiesi-Villa and C. Guastini,

J. Chem. Soc., Dalton Trans. 1990, 1357.

32. N. Dowling, P. M. Henry, N. A. Lewis and H. Taube, Inorg. Chem., 20 (1981)

2345.

33. M. Herberhold, G.-X. Jin, A. L. Rheingold and G. F. Sheats, Z. Naturfosch.,

476 (1992) 1091.

34. C. Imrie, P. Engelbrecht, C. Loubser and C. W. McCleland, Appl. Organomet.

Chem., 15 (2001) 1.

35. E. Stancovic, S. Toma, R. van Boxel, I. Asselberghs and A. Persoons, J.

Organomet. Chem., 637-639 (2001) 426.

36. G. Cooke, H. M. Palmer and O. Schulz, Synthesis, 1995, 1415.

37. Y. Bai, J. Lu, H. Gan, Z. Wang and Z. Shi, Synthesis and Reactivity in

Inorganic and Metal-Organic Chemistry, 34 (2004) 1487.

38. D. Braga, D. D’Addario and M. Polito, Organometallics, 23 (2004) 2810.

39. D. Braga, S. L. Giaffreda, F. Grepioni, A. Pettersen, L. Maini, M. Curzi and M.

Polito, Dalton Trans., 2006, 1249.

40. M. D. Bala and N. J. Coville, J. Organomet. Chem., 692 (2007) 709.

41. D. Villemin, B. Martin, P. Puciova and S. Toma, J. Organomet. Chem., 484

(1994) 27.

42. C. Imrie, V. O. Nyamori and T. I. A. Gerber, J. Organomet. Chem., (2004)

1617.

Page 57: Solvent-free Synthesis of Bisferrocenylimines

43

43. W.-Y. Liu, Q.-H. Xu, B.-H. Chen and Y.-X. Ma, Synth. Commun., 32 (2002)

171.

44. D. I. Méndez, E. Klimova, T. Klimova, L. Fernando, S. O. Hernandez, and M.

G. Martinez, J. Organomet. Chem., 679 (2003) 10.

45. R. Adams, A. H. Blatt and V. Boekeheide, Organic Reactions, 13, John Wiley

& Sons Inc., 1965.

46. B. S. Furniss, A. J. Hannaford, P. W. G. Smith and A. R. Tatchell, Vogel’s

Textbook of Practical Organic Chemistry, Longman Scientific and Technical,

England, 5th Ed., 1989.

47. W.-Y. Liu, Q.-H. Xu, Y.-X. Ma, Y.-M. Liang, N.-L. Dong and D.-P. Guan, J.

Organomet. Chem., 625 (2001) 128.

48. Z. S. Arigan and H. Suschitiky, J. Chem. Soc., 1961, 2242.

49. S. S. Hirsch and W. J. Bailey, J. Org.Chem., 43 (1978) 4091.

50. F. Kröhnke, Synthesis, 1976, 1.

51. W.-Y. Liu, Q.-H. Xu, Y.-M. Liang, B.-H. Chen, W.-M. Liu and Y.-X. Ma, J.

Organomet. Chem., 637-639 (2001) 719.

52. C. Imrie, E.R.T. Elago, N. Williams, C. W. McCleland and P. Engelbrecht, J.

Organomet. Chem.,690 (2005) 4959.

53. S. K. Pal, K. Alagesan, A. G. Samuelson and J. Pebler, J. Organomet. Chem.,

575 (1999) 108.

54. R. Bosque, C. López, J. Sales, X. Solans and M. Font-Bardia, J. Chem. Soc.,

Dalton Trans., 1994, 735.

55. A. Hutton, N. Jasim, R. M. G. Roberts, J. Silver, D. Cunningham, P. McArdle

and T. Higgins, J. Chem. Soc., Dalton Trans., 1992 , 2235.

56. I. Ratera, D. Ruiz-Molina, C. Sánchez, A. Alcalá, C. Rovira and J. Veciana,

Synth. Met., 121 (2001) 1834.

57. R. P. Houghton, Metal Complexes in Organic Chemistry, Cambridge

University Press, Cambridge, 1979.

58. A. I. Stetsenko, M. A. Presnov and A. L. Konovalova, Russ. Chem. Rev., 50

(1981) 353.

59. I. M. El-Mehasseb, M. Kodaka, T. Okada, T. Tomohiro, K. -I Okamoto and H.

Okuno, J. Inorg. Biochem., 84 (2001) 157.

60. M. S. Seo, Y. S. Sohn, B.-S. Yang, W. Nam and K. M. Kim, Inorg. Chem.

Comm., 7 (2004) 1178.

Page 58: Solvent-free Synthesis of Bisferrocenylimines

44

61. S. Tzanopoulou, I. C. Pirmettis, G. Patsis, C. Raptopoulou, A. Terzis, M.

Papadopoulos and M. Pelecanou, Inorg, Chem., 45 (2006), 902.

62. K. A. Stephenson, S. R. Banerjee, T. Besanger, O. O. Sogbein, M. K.

Levadala, N. McFarlane, J. A Lemon, D. R. Boreham, K. P. Mresca, J. D.

Brennan, J. W. Babic, J. Zubieta and J. F. Valliant, J. Am. Chem. Soc., 126

(2004) 8598.

63. J. Tsuji, Palladium Reagents and Catalysts: Innovation in Organic Chemistry,

John Wiley & Sons, New York, 1995.

64. R. H Grubbs, Handbook of Metathesis, Wiley-VCH, New York, 2003.

65. P. Wipf, Handbook of Reagents for organic Synthesis: Reagents for High-

throughput solid phase and Solution phase Organic Synthesis, John Wiley &

Sons, New York, 1999.

66. M. Noji, M. Nakajima and K .Koga, Tetrahedron Lett. 35 (1994) 7983.

67. M. Nakajima, I. Mayoshi, K. Kanayama, K. hashimoto, S. Noji and M. Koga, J.

Org. Chem., 64 (1999) 2264.

68. L. D. Field and I. P. Thomas, Inorg. Chem., 35 (1996) 2546.

69. H. Tye, D. Smyth, C. Eldred and M. Wills, Chem. Commun., 1997, 1053.

70. P.-H. Leung, S. Selvaratnam, C. R. Cheng, K. F. Mok, N. H. Rees and W.

McFarlane, Chem. Commun., 1997, 751.

71. Y. Chi and X. Zhang, Tetrahedron Lett., 43 (2002) 4849.

72. F.-Y Kwong, C.-W. Lai and K.-S. Chan, Tetrahedron Lett., 43 (2002), 3537.

73. F.-Y. Kwong and K. S. Chan, Chem. Commun., 2002, 1069.

74. R. J. Cremlyn, An Introduction to Organosulfur Chemistry, Wiley & Sons, New

York, 1996.

75. E. J. Lenardão, R. G. Lara, M. S. Silva, R. G. Jacob and G. Perin,

Tetrahedron Lett., 48 (2007) 7668.

76. S. Vijaikumar and K. Pitchumani, J. Mol. Cat. A: Chemical, 217 (2004) 117.

77. A. McCluskey, Green Chemistry. 1 (1999) 167.

78. P. C. Andrews, A. C. Peatt and C. L. Raston, Green Chemistry, 3 (2001) 313.

79. G. Sabitha, B. V. S. Reddy, S. Abraham and J. S. Yadav, Green Chemistry, 1

(19990 251.

80. A. T. Placzek, J. L. Donelson, R. Trivedi, R. A. Gibbs and S. K. De,

Tetrahedron Lett., 46 (2005) 9029.

Page 59: Solvent-free Synthesis of Bisferrocenylimines

45

81. C. Zhang, E. J. Moran, T. F. Woiwode, K .M. Short and A. M. M. Mjalli,

Tetrahedron Lett., 37 (1996) 751.

82. S. Balalaie, M. M. Hashemi and M. Akhbari, Tetrahedron Lett., 44 (2003)

1709.

83. A. R. Karimi, Z. Alimohammadi, J. Azizian, A. A. Mohammadi and M. R.

Mohammadizadeh, Catalysis Commun., 7 (2006) 728.

84. M. Adib, H. tahermansouri, S. A. Koloogani, B. Mohammadi and H. R.

Bijanzadeh, Tetrahedron Lett., 47 (2006) 5957.

85. M. Adib, N. Mahmoodi, M. Mahdavi and H. R. Bijanzadeh, Tetrahedron Lett.,

47 (2006) 9365.

86. H. Naemi, F. Salami and K. Rabiei, J. Mol. Cat. A: Chemical, 260 (2006) 100.

87. A. Togni and L. M. Venanzi, Angew. Chem. Int. Ed. Engl., 33 (1994) 497.

88. C. Pettinari and N. Maschiocchi, J. Organomet. Chem., 690 (2005) 1871.

89. C. Janiak, J. Chem. Soc., Dalton Trans., 2000, 3885.

90. R. Meyer, P. L. Wessels, P. H. van Rooyen and S. Lotz, Inorg. Chim. Acta,

284 (1999) 127.

91. K. C. Gupta and A. K. Sutar, Coord. Chem. Rev., 2007,

doi:10.1016/j.ccr.2007.09.005.

92. M. Talja, M. Polamo and M. Leskelä, J. Mol. Cat. A: Chemical, 280 (2008)

102.

93. S. Chen, X. Zhang, H. Ma, Y. Lu, Z. Zhang, H. Li, Z. Lu, N. Cui and Y. Hu, J.

Organomet. Chem., 690 (2005) 4184.

94. J. Cloete and S. F. Mapolie, J. Mol. Cat. A: Chemical, 243 (2006) 221.

95. V. L. Cruz, J. Martinez-Salazar, J. Ramos, M. L. Reyez, A.Toro-Labbe and S.

Gutierrez-Oliva, Polymer, 48 (2007) 7672.

96. M. E. Bluhm, C. Folli and M. Döring, J. Mol. Cat. A: Chemical, 212 (20040 13.

97. S. McTavish, G. J. P. Britovsek, T. M. Smit, V. C. Gibson, A. J. P. White and

D. J. Williams, J. Mol. Cat. A: Chemical, 261 (2007) 293.

98. C. Görl and H. G. Alt, J. Organomet. Chem., 692 (2007) 4580.

99. D. J. Nielsen, K. J. Cavell, B. W. Skelton and A. H. White, Inorg. Chim. Acta,

327 (20020 116.

100. J. Horniakova, T. Raja, Y. Kubota and Y. Sugi, J. Mol. Cat. A:

Chemical, 217 (2004) 73.

Page 60: Solvent-free Synthesis of Bisferrocenylimines

46

101. P. Pelagatti, M. Marcelli, M. Costa, S. Ianelli, C. Pelizzi and D. Roglino,

J. Mol. Cat. A: Chemical, 226 (2005) 107.

102. W. Chen, C. Xi and Y. Wu, J. Organomet. Chem., 692 (2007) 4381.

103. D. W. Dodd, H. E. Toews, F. d. S. Carneiro, M. C. Jennings and N. D.

Jones, Inorg. Chim. Acta, 359 (2006) 2850.

104. D. Gelman, S. Dechert, H. Schumann and J. Blum, Inorg. Chim. Acta,

334 (2002) 149.

105. R. B. Bedford, C. S. J. Cazin, M. B. Hursthouse, M. E. Light, K. J. Pike

and S. Wimperis, J. Organomet. Chem., 633 (2001) 173.

106. J. Wiedermann, K.Mereiter and K. Kirchner, J. Mol. Cat. A: Chemical,

257 (2006) 67.

107. G. A. Barf and R. A. Sheldon, J. Mol. Cat. A: Chemical, 98 (1995) 143.

108. J.-C. Moutet and A. Ourari, Electrochim. Acta, 42 (1997) 2525.

109. S. M. Bruno, S. S. Balula, A. A. Valente, F. A. Almeida Paz, M.

Pillinger, C. Souza, J. Klinowski, C. Freire, P. Ribeiro-Claro and I.

Gonçalves, J. Mol. Cat. A: Chemical, 270 (2007) 185.

110. A. Soriente, M. De Rosa, M. Lamberti, C. Tedesco, A. Scettri and C.

Pellecchia, J. Mol. Cat. A: Chemical: Chemical, 235 (2005) 253.

111. G. Chellucci, V. Caria and A. Saba, J. Mol. Cat. A: Chemical, 130

(1998) 51.

112. G. Chelucchi, S. Gladiali, M. G. Sanna and H Brunner, Tetrahedron:

Asymmetry, 11 (2000) 3419.

113. P.-F. Teng, C.-S. Tsang, H.-L.Yueng, W.-L. Wong. H.-L. Kwong and I.

D. Williams, J. Organomet. Chem., 691 (2006) 2237.

114. V. G. Albano, M. Bandini, M. Monari, F. Piccinelli, S. Tommasi and A.

Umani-Ronchi, Inorg. Chim. Acta., 360 (2007) 1000.

115. J. M. E. Matos and B. S. Lima-Neto, Catalysis Today, 108-108 (2005)

282.

116. D. Saku, MSc Thesis, Nelson Mandela Metropolitan University, 2007.

117. B. Albert, N. Dieltiens, N. Ledoux, C. Vercaemst, P. van der Voort, C.

V. Stevens, A. Linden and F. Verpoort, J. Mol. Cat. A: Chemical, 260

(2006) 221.

118. J. A. Wright, A. A. Danopoulos, W. B. Motherwell, R. J. Carrol and S.

Ellwood, J. Organomet. Chem., 691 (2006) 5204.

Page 61: Solvent-free Synthesis of Bisferrocenylimines

47

119. A. Togni and T. Hayashi, Ferrocenes: Homogeneous Catalysis,

Organic Synthesis, Materials Science, VCH, Weinheim, 1995.

120. R. C. Atkinson, V. C. Gibson and N. J. Long, Chem. Soc. Rev., 33

(2004) 313.

121. T. M. Miller, k. J. Ahmed and M. S. Wrighton, Inorg. Chem., 28 (1989)

2347.

122. O. Carugo, G. De Santis, L. Fabbrizzi, M. Licchelli, A. Monichino and P.

Pallvicini, Inorg. Chem., 31 (1992) 765.

123. S. Sakanishi, D. A. Bardwell, S. Couchman, J. C. Jeffery, J. A.

McCleverty and M. D. Ward, J. Organomet. Chem., 528 (1997) 35.

124. D. Braga, M. Polito, D. D’Addario and E. Tagliavini, Organometallics,

22 (2003) 4532.

125. J. G. P. Delis, P. W. N. M. van Leeuwen, K. Vrieze, N, Veldman, A. L.

Spek, J. Fraanje and K. Goubitz, J. Organomet. Chem., 514 (1996)

125.

126. J. Raput, J. R. Moss, A. T. Hutton, D. T. Hendricks, C. E. Arendse and

C. Imrie, J. Organomet. Chem., 689 (2004) 1553.

127. J. Rajput, A. T. Hutton, J. R. Moss, H. Su and C. Imrie, J. Organomet.

Chem., 691 (2006) 4573.

128. V. C. Gibson, C. M. Halliwell, N. J. Long, P. J. Oxford, A. M. Smith, A.

J. P. White and D. J. Williams, Dalton Trans., 2003, 918.

129. R. Bosque, C. López and J. Sales, Inorg. Chim. Acta, 244 (1996) 141.

130. S. Pérez, C. López, A. Caubet, X. Solans, M. Font-Bardia, M. Gich and

E. Molins, J. Organomet. Chem., 692 (2007) 2402.

131. C. Imrie, C. Loubser, P. Engelbrecht, C. W. McCleland and Y Zheng, J.

Organomet. Chem., 665 (2003) 48.

132. S.-Q. Huo, Y.-J. Wu, C.-X. Du, Y. Zhu, H.-Z. Yuan and X.-A. Mao, J.

Organomet. Chem., 483 (1994) 139.

133. R. Bosque, C. López and J. Sales, J. Organomet. Chem., 498 (1995)

147.

134. G. Zhao, Q.-G. Wang and T. W. C. Mak, Polyhedron, 17 (1998) 1.

135. C. López, A. Caubet, S. Pérez, X. Solans and M. Font-Bardia, J.

Organomet. Chem., 651 (2002), 105.

Page 62: Solvent-free Synthesis of Bisferrocenylimines

48

136. C. López, R. Bosque, J. Arias, E. Evangelio, X. Solans and M. Font-

Bardia, J. Organomet. Chem., 672 (2003) 34.

137. H.-X. Wang, Y.-J. Li, R.-Q. Gao, H.-F. Wu, F.-Y. Geng and R. Jin,

Inorg. Chem. Commun., 9 (2006) 658.

138. C. Xu, J.-F. Gong, Y.-H. Zhang, Y. Zhu, C.-X. Du and Y.-J. Wu, Inorg.

Chem. Commun., 9 (2006) 456.

139. Y.-J. Wu, J.-J. hou, H.-Y. Yun, X.-L. Cui and R.-J. Yuan, J. Organomet.

Chem., 637-639 (2001) 793.

140. X. M. Zhao, X. Q. Hao, B. Liu, M. L. Zhang, M. P. Song and Y.-J. Wu,

J. Organomet. Chem., 691 (2006) 255.

141. Z.-Q. Weng, L. L. Koh, and T. S. A. Hor, J. Organomet. Chem., 689

(2004)18.

142. J.-L. Zhang, F. Yang, G. Ren, T. C. W. Mak, M.-P. Song and Y.- J. Wu,

Ultrason. Sonochem., 2007, doi:10.1016/j.ultasonch.2007.02.002

143. J.-F. Gong, G.-Y. Liu, C.-X. Du, Y. Zhu and Y.-J. Wu, J. Organomet.

Chem., 690 (2005) 3963.

144. C. Xu, J.-F. Gong and Y.-J. Wu, Tetrahedron Lett., 48 (2007) 1619.

145. X.-P. Hu, H.-C. Dai, C.-M. Bai, H.-L. Chen and Z. Zheng, Tetrahedron:

Asymmetry, 15 (2004) 1065.

146. A. E. Platero-Prats, S. Pérez, C. López, X. Solans, M. Font-Bardia, P.

W. N. M. Leeuwen, G. P. F. Strijdonck and Z. Freixa, J. Organomet.

Chem., 692 (2007) 4215.

147. R. K. J. Bott, M. Schormann, D. L. Hughes, S. J. Lancester and M.

Bochmann, Polyhedron, 25 (2006) 387.

148. C. M. Asselin, G. C. Fraser, H. K. Hall Jr., W. E. Lindsell, A. B. Padias

and P. N. Preston, J. Chem. Soc., Dalton Trans., 1997, 3765.

149. E. Bullita, U. Casellato, F. Ossola, P. Tomasin, P. A. Vigato and U.

Russo, Inorg. Chim. Acta, 287 (1999) 117.

150. V. K. Mippidi, T. Htwe, P. S. Zacharias and S. Pal, Inorg. Chem.

Commun., 7 (2004) 1045.

151. C. Imrie, P. Kleyi, V. O. Nyamori, T. I. A. Gerber, D. C. Levendis and J.

Look, J. Organomet. Chem., 692 (2007) 3443.

152. H.-X. Wang, R.-Q. Gao, X.-L. Yang, L. Wan, H.-F. Wu, F.-Y. Geng and

R. Jin, Polyhedron, 26 (2007) 1037.

Page 63: Solvent-free Synthesis of Bisferrocenylimines

49

153. S.-I. Lee, S.-C. Chul and T.-J. Kim, J. Polym. Sci., Part A: Polymer

Chemistry, 34 (1996) 2377.

154. J.-J. Kim and H. Alper, Chem. Commun., 2005, 3059.

Page 64: Solvent-free Synthesis of Bisferrocenylimines

50

CHAPTER 2

RESULTS AND DISCUSSIONS

2.1 Solvent-free synthesis of bisferrocenylimines

2.1.1 Introduction

There has been an upsurge of interest in performing chemical transformations under

solvent-free conditions. In comparison to methods that employ solvents, the solvent-

free approach proceeds more cleanly and provides higher yields. Examples of the

solvent-free method have been highlighted in the previous chapter. In the literature,

Nyamori et al.1 have reported on the solvent-free synthesis of ferrocenylimines by

grinding ferrocenecarboxaldehyde and a substituted aromatic amine. This procedure

provided excellent yields and the reactions occurred readily at room temperature.

However, the reactions in which the aromatic amine contained an electron-

withdrawing substituent were slow and thus required heating at 50 ºC.

Fe

Y C

O

H+

H2N

X

Grind

Solvent-free Fe

Y C

N

H

X

Y = Ph, Ph-Ph, Ph-O-Ph

Scheme 2.1: Solvent-free synthesis of ferrocenylimines

The objective of this research was to extend the work and synthesize

bisferrocenylimines (Figure 2.1), by reacting ferrocenecarboxaldehyde with

diamines.

Page 65: Solvent-free Synthesis of Bisferrocenylimines

51

Fe

CH

(CH2)

Fe

CH

Nx

N

Figure 2.1: General structure of bisferrocenylimines to be synthesized

The synthesis of bisferrocenylimines by a reaction of diamines and

ferrocenecarboxaldehyde has been reported by several research groups.2,3 The

imines are useful ligands for metal complexation and many ferrocenylimines have

been investigated for non-linear optical properties.4

2.1.2 Synthesis and characterization of bisferrocenyimines

The reaction of ferrocenecarboxaldehyde and the appropriate diamines gave the

corresponding bisferrocenylimines in excellent yields (Scheme 2.2).

Fe

CHOH2N(CH2)xNH2

Grind Fe

CH

(CH2)

Fe

CH

Nx

N

2

x = 2, 3, 4, 6, 8

Scheme 2.2: Solvent-free synthesis of bisferrocenylimines

The solvent-free reaction involved the mixing and grinding of two mole equivalents of

ferrocenecarboxaldehyde and the appropriate diamine. A pictorial description of the

process is shown by a reaction of ferrocenecarboxaldehyde and 1,8-diaminooctane

in Figure 2.2. On grinding the mixture of reagents, a melt was obtained which

eventually solidified at room temperature, to give the bisimines in excellent isolated

yields (Table 2.1). However, for [2.2] the melt obtained only gradually solidified at

room temperature, after removal of water formed in the condensation reaction under

vacuum.

Page 66: Solvent-free Synthesis of Bisferrocenylimines

52

Figure 2.2: The pictorial stages of the solvent-free synthesis of N,N’-

octylenebis(ferrocenylmethylidine)amine

(A) Ferrocenecarboxaldehyde (brown)

1,8-Diaminooctane (white)

(B) Ground mixture of

ferrocenecarboxaldehyde

and 1,8-diaminooctane

MELT

PRODUCT

(C) Solidified N,N’-

octylenebis(ferrocenylmethylidine)-

amine

Page 67: Solvent-free Synthesis of Bisferrocenylimines

53

Table 2.1: Yields of bisferrocenylimines from the solvent-free reaction of diamines

and ferrocenecarboxaldehyde

Compound

number

Ferrocene-

carboxaldehyde

(Fc-Y-CHO)

Diamine

H2N(CH2)xNH2

Yield of

Bisferrocenyliminea

(%)

[2.1]

FcCHO

H2N(CH2)2NH2

Fe

CH

(CH2)

Fe

CH

N2

N

(99%)

[2.2]

FcCHO

H2N(CH2)3NH2

Fe

CH

(CH2)

Fe

CH

N3

N

(94%)

[2.3]

FcCHO

H2N(CH2)4NH2

Fe

CH

(CH2)

Fe

CH

N4

N

(92%)

[2.4]

FcCHO

H2N(CH2)6NH2

Fe

CH

(CH2)

Fe

CH

N6

N

(94%)

[2.5]

FcCHO

H2N(CH2)8NH2

Fe

CH

(CH2)

Fe

CH

N8

N

(97%)

[2.6]

Fc CHO

H2N(CH2)2NH2

Fe Fe

CH N N CH

(97%)

a The isolated yields are based on starting materials.

Page 68: Solvent-free Synthesis of Bisferrocenylimines

54

The solidified melt was initially analyzed by infrared (IR) spectroscopy using a

potassium bromide (KBr) disc, in order to determine that the reaction occurred under

solvent-free conditions. Infrared analysis showed the disappearance of the carbonyl

(C=O) band at approximately 1700 cm-1 and the appearance of the strong imine

(C=N) band at 1646 cm-1. Figure 2.3 shows the IR spectrum of [2.5] immediately

after the melt had solidified. The absence of the carbonyl band in the IR spectrum

was evidence that the reaction was completed under solvent-free conditions.

Figure 2.3: IR spectrum of [2.5]

The solidified melt was ultimately recrystallized from a minimal amount of cold

anhydrous methanol to provide a pure product. 1H NMR spectra of all the

bisferrocenylimines [2.1]-[2.6] showed proton signals in the region δ 8.1-8.2 ppm

which were indicative of the presence of imine (CH=N) protons. Compounds [2.1]2

and [2.6]5 have been synthesized previously; however, they are reported here for the

first time under solvent-free conditions. 1H NMR spectra of [2.1]-[2.5] were expected

to possess some similarities since the only difference in their molecular structures

was the length of the alkyl chain of the diamines. A singlet was observed at δ 3.78

ppm for [2.1], indicatve of the four protons for the two (CH2) groups. Compound [2.2]

has an additional CH2 group in the β-position and exhibited a triplet at δ 3.56 and a

Page 69: Solvent-free Synthesis of Bisferrocenylimines

55

multiplet at δ 2.03 ppm due to the four protons of the two terminal (α) CH2 groups

and two protons of the middle (β) CH2 group, respectively. As expected, [2.3] with

four CH2 groups exhibited a triplet at δ 3.7 ppm due to the protons on the terminal (α)

CH2 groups and a multiplet at δ 1.71 ppm due to the protons on the middle (β) CH2

groups. Three signals were observed for [2.4], a triplet at δ 3.45 ppm for the protons

on the terminal (α) CH2 groups and two multiplets at δ 1.66 and 1.44 ppm for the

protons on the inner (β) CH2 groups and the middle (γ) CH2 groups, respectively. An

interesting scenario was observed with [2.5] where instead of four signals that would

have been expected, only three were observed. A triplet at δ 3.42 ppm due to the

protons on the terminal (α) CH2 groups, a multiplet at δ 1.62 ppm for the inner (β)

CH2 groups and a singlet for the protons on the four middle CH2 groups were

observed (Figure 2.4).

Figure 2.3: 1H NMR spectrum of [2.5]

This suggested that the deshielding effect of the nitrogen groups became less

pronounced as the length of the alkyl chain increased. The substituted

cyclopentadienyl (Cp) ring of the ferrocene moiety for all the compounds was

represented by two triplets and a very intense singlet for the unsubstituted

cyclopentadienyl ring. For all the compounds [2.1]-[2.5] the chemical shifts of the

protons on the carbons directly bonded to the nitrogen groups were expected to be

Page 70: Solvent-free Synthesis of Bisferrocenylimines

56

very similar. However, this was not the case and the results are summarized in Table

2.2. This table shows that the proton signals are shifted to lower frequencies (less

deshielding) as the length of the alkyl chains increased. This effect could be

attributed to the increased free rotation and flexibility of the chain as the length

increased. A significant change was observed between [2.1] and [2.2] as well as

[2.2] and [2.3], while there was less change from [2.3] to [2.4] to [2.5]. Additionally,

for [2.2]-[2.5], these protons appeared as triplets, indicating some vicinal proton-

proton coupling. For [2.1] a singlet was observed showing that there was no vicinal

coupling between the neighbouring protons.

Table 2.2: Chemical shifts for protons on the carbon directly bonded to nitrogen

group

Compound Chemical shift δ

(ppm)

[2.1]

3.78

[2.2]

3.56

[2.3]

3.47

[2.4]

3.45

[2.5]

3.42

13C NMR spectra for all compounds showed signals, for the imine carbons, in the

region δ 160-163 ppm. The observed carbon signals for all compounds [2.1]-[2.5]

were as expected. Compound [2.1] exhibited a single signal at δ 62.80 ppm for two

CH2 groups while for [2.2] two signals were observed at δ 59.25 and 32.90 ppm

Page 71: Solvent-free Synthesis of Bisferrocenylimines

57

representing the carbons directly bonded to nitrogen groups and the middle carbons,

respectively. Compounds [2.3], [2.4] and [2.5] exhibited two, three and four carbon

signals, respectively, in the expected chemical shift regions for the CH2 groups. All

the compounds are very stable in the solid state at room temperature and in open air

while they showed some degree of instability in solution. This was in accordance

with findings of Benito et al.2 Attempts to perform the same reaction using

acetylferrocene were unsuccessful, even after heating the reaction mixture up to 65

ºC. The inability of acetylferrocene to react was attributed to steric hindrance by the

methyl group.

2.1.3 Solvent-free synthesis of arylbisimines

The synthesis of the arylbisimines in a solvent-free environment has been reported

by Naeimi and co-workers.6 The method included the use of a solid-supported

catalyst P2O5/Al2O3 (Scheme 2.3).

R

OH

R'

O+ H2N

Y

NH2

P2O5/Al2O3

Solvent-free

R

OH

R'

N

Y

N

R'

HO

R

R = H, NO2

R' = H, CH3

Y = C6H4-O-C6H4, C6H4-CH2-C6H4, C6H4-SO2-C6H4

Scheme 2.3: Solvent-free synthesis of arylbisimines in the presence of a catalyst.6

Similar types of arylbisimines have been synthesized in this study, where the

aromatic ring contained methyl substituents at 2,3-, 2,5- and 2,4,6-positions. The Y-

spacer was a two-carbon alkyl chain and no catalyst was required. The reactions

involved mixing and stirring of two mole equivalents of an appropriate substituted

benzaldehyde, and ethylenediamine. Stirring the reagents for a few seconds gave a

white solid, an indication that the reaction had taken place since both starting

reagents are liquids (Scheme 2.4).

Page 72: Solvent-free Synthesis of Bisferrocenylimines

58

H

O

+H2N

NH2

stirsolvent-free

R1

NN

R1

[2.7]: R1, R2 = CH3; R3, R4, R5 = H

[2.8]: R1, R3 = CH3; R2, R4, R5 =H

[2.9]: R1, R3, R5 = CH3; R2, R4 =H

R1

R2

R3 R5

R5 R3

12

3

45

6

1'2'

3'

4'5'

6'

R2

R4

R4

R2

R3

R4

R5

Scheme 2.4: Solvent-free synthesis of arylbisimines

Although it was apparent that the reaction took place under these solvent-free

conditions, the reaction mixture was analysed by IR spectroscopy using KBr discs to

confirm that the reaction had indeed occurred under these conditions. IR analysis

showed the disappearance of the carbonyl band at approximately 1700 cm-1 and the

appearance of the imine (C=N) band in the region 1637-1639 cm-1. The white crude

solid was eventually recrystallized from cold anhydrous methanol to obtain the pure

product in excellent yield (Table 2.3). 1H NMR revealed that there was no aldehydic

proton signal present in the region δ 9.7-10 ppm for all the compounds. The major

differences between the compounds were expected to be in aromatic region of the

spectra and the methyl substituents. Compound [2.7] was identified by the presence

of two singlets at δ 2.33 and 2.27 ppm representing the methyl groups at 2, 2’ and 3,

3’ positions of the aromatic rings, respectively. For [2.8] two singlets were also

observed at δ 2.39 and 2.29 ppm for the methyl groups at 2, 2’ and 5, 5’ positions,

respectively. Compound [2.9] contained three methyl substituents at 2, 2’, 4, 4’ and

6, 6’ positions. For the aromatic rings, [2.7] exhibited two doublets at δ 7.69 and 7.20

ppm due to protons at the 6, 6’ and 4, 4’ positions, respectively, and a triplet at δ

7.10 ppm due to protons at 5, 5’ positions.. A singlet at δ 7.66 ppm for [2.8]

represents the protons at the 6, 6’ positions.

Page 73: Solvent-free Synthesis of Bisferrocenylimines

59

Table 2.3: Yields of arylbisimines from a reaction of substituted benzaldehyde and

ethylenediamine

Substituted

Benzaldehyde

Ethylenediamine

Yield of

Arylbisiminea

(%)

H

O

H2NNH2

NN

[2.7] (86%)

H

O

H2NNH2

NN

[2.8] (84%)

H

O

H2NNH2

NN

[2.9] (88%) a Yields are based on starting materials

Two doublets were also observed at δ 7.11 and 7.07 ppm for the protons at the 3, 3’

and 4, 4’ positions, respectively. As anticipated, the 1H NMR for [2.9] showed some

chemical shift equivalence of the protons on the aromatic rings as only a singlet at δ

2.37 ppm was observed. 13C NMR spectra of all compounds exhibited signals for the

C=N group in the region δ 161-162 ppm. The 13C NMR spectra of [2.7] and [2.9]

showed six and four signals for the aromatic rings, respectively, while the spectrum

of [2.8] gave rise to five signals. Actually, there were six signals since the signal at δ

Page 74: Solvent-free Synthesis of Bisferrocenylimines

60

131.1 ppm (highlighted) represented two overlapping signals, one lower in intensity.

Figure 2.5 illustrates the aromatic region of the 13C NMR spectrum of [2.8].

Figure 2.4: 13C NMR spectrum of [2.8]

According to predicted chemical shifts as implemented in ChemDraw, the signals

were assigned to C3, C3’ and C4, C4’. The signals for the methyl groups were

observed at δ 19.81 and 14.10 ppm for [2.7]. The methyl signals for [2.8] were

observed at δ 20.38 and 18.46 ppm while for [2.9] signals appeared at δ 21.52 and

21.06 ppm. The signals for the CH2 groups were observed at δ 62.17, 62.15 and

63.89 ppm for [2.7], [2.8] and [2.9], respectively. Since the major difference between

the three compounds was the number and the positions of the methyl groups on the

aromatic rings, it seemed as though the more methyl groups there were present, the

greater the downfield shift of the CH2 groups.

2.2 Reduction reaction of bisferrocenylimines and arylbisamines

2.2.1 Reduction of bisferrocenylimines

The reduction of bisferrocenylimines with lithium aluminium hydride has been

reported elsewhere in the literature.2 Our interest was to carry out this transformation

using a much cleaner synthetic route. We therefore opted for the catalytic

Page 75: Solvent-free Synthesis of Bisferrocenylimines

61

hydrogenation reaction procedure (Scheme 2.5), which takes place cleanly since the

only product is the corresponding bisferrocenylamine.

X

Fe

CH

(CH2)

Fe

CH

Nx

N

H2, Pd/C

MeOH Fe

CH2

(CH2)

Fe

CH2

HN

x

HN

x = 2, 3, 4, 6, 8

Scheme 2.5: Hydrogenation of ferrocenylbisimines

However, attempts to carry out the reactions yielded no results. We were then

forced to use the lithium aluminium hydride route. The reaction was carried out by

heating a solution of the bisferrocenylimine in diethylether for 30 minutes. The

reaction was then quenched with ice/water slurry and the product extracted with

diethyl lether. The products were obtained in excellent yield as pale to bright yellow

solids (Table 2.4). Similar types of compounds that are C2-symmetric have been

reported by Woltersdorf et al.7 The 1H NMR spectra of all the reduction products

showed the disappearance of the imine proton signal and the appearance of two

new signals, the amine (NH) and the CH2 proton signals. For all the compounds

[2.10]-[2.14] the NH proton signal was observed as a broad singlet in the chemical

shift region δ 1.53-2.27 ppm. The CH2 proton signal on the other hand, was

observed as singlet in the chemical shift region δ 3.52-3.58 ppm. It was also

observed that one of the substituted Cp ring proton signals was shifted to lower

chemical shifts for all the compounds. The CH2 proton signal in the reduced form of

the bisferrocenylimines was observed in the expected chemical shift region. The

terminal CH2 proton signals of the alkyl chains appeared to have been shifted to

lower frequencies compared to those in the original bisferrocenylimines. In the

bisferrocenylimines the terminal CH2 proton signal appeared at higher frequencies

largely due to anisotropic effects of the double bond (CH=N). All the terminal CH2

signals were shifted to lower frequencies by ca. 0.8 ppm, which is in agreement with

the literature value.8 13C NMR spectra showed the disappearance of the CH=N

carbon signal and a new CH2 carbon signal, in the expected region, for all the

compounds [2.10]-[2.14].

Page 76: Solvent-free Synthesis of Bisferrocenylimines

62

Table 2.4: Yields of bisferrocenylimines

Compound % Yielda

Fe

CH2

(CH2)

Fe

CH2

HN

2

HN

[2.10]

87%

Fe

CH2

(CH2)

Fe

CH2

HN

3

HN

[2.11]

92%

Fe

CH2

(CH2)

Fe

CH2

HN

4

HN

[2.12]

86%

Fe

CH2

(CH2)

Fe

CH2

HN

6

HN

[2.13]

91%

Fe

CH2

(CH2)

Fe

CH2

HN

8

HN

[2.14]

88%

a Yields are based on starting materials

Infrared spectra of all compounds showed the appearance of several peaks

corresponding to the v(NH) stretching vibration which occurred in the region 3300-

3100 cm-1. Furthermore, the v(CH=N) stretching peak which occurs as a strong

absorption in the region 1640-1645 cm-1 for the bisferrocenylimines was lacking. The

above information was in agreement with the proposed molecular structures of the

compounds.

Page 77: Solvent-free Synthesis of Bisferrocenylimines

63

2.2.2 Reduction of arylbisimines

The reduction of arylbisimines was carried out using the same method as for the

reduction of bisferrocenylimines. These compounds were obtained in moderate to

good yield as a white powder, except for [2.15] which was obtained as a colourless

oil. However, it was also found that [2.15] solidified to a white solid after prolonged

drying process under suction. 1H NMR spectra of all compounds [2.15]-[2.17]

showed the disappearance of the CH=N proton signal and the appearance of CH2

and NH protons signals. The peak in the chemical shift region δ 3.94-3.96 ppm

represented the CH2 proton signal for all the compounds. The NH proton signal was

obtained as a broad singlet in the region δ 3.76-3.81 ppm. The signal for the CH2

groups of the ethylene chain in arylbisimines appeared at higher frequencies than

those of the arylbisamines largely due to anisotropic effects of the imine double

bond. The shift to lower resonance frequencies in the CH2 signals in arylbisamines

confirmed the reduction of the CH=N double bond to a CH2-NH single bond. The

formation of the arylbisamines was also confirmed by 13C NMR spectrometry. The

appearance of a new signal in the chemical shift region δ 48.2-51.77 ppm was due to

the presence of the CH2 group. Infrared spectra showed the appearance of several

peaks for the ν(NH) stretching vibration at 3350-3150 cm-1 for all the compounds.

2.3 Electronic Spectroscopy

The UV-visible spectra of bisferrocenylimines and bisferrocenylamines prepared

were obtained in dichloromethane solution. Spectral comparisons with unsubstituted

ferrocene as a reference were also made. Ferrocene exhibited two bands at

wavelengths of λmax 326 and 442 nm, which have been assigned to 1A2g→1E2g and

1A1g→1E1g ligand field d-d transitions.9 The UV-vis spectrum of unsubstituted

ferrocene is illustrated in Figure 2.6.

Page 78: Solvent-free Synthesis of Bisferrocenylimines

64

Figure 2.6: UV-vis spectrum of unsubstitued ferrocene in dichloromethane

The ferrocenyl bands in [2.1]-[2.5] were observed at longer wavelengths λmax (Figure

2.7, Table 2.5) largely due to conjugation with the CH=N bond. Bathochromic shifts

(shifts to longer wavelengths) are anticipated where conjugation increases in length.

Some absorption bands due to π-π* and n-π* transitions of the imine groups CH=N

in [2.1]-[2.5] were also observed at wavelengths lower than 300 nm. The ferrocenyl

bands in [2.10]-[2.15] were observed at slightly shorter wavelengths λmax (Figure 2.8,

Table 2.5) largely due to the reduction of the CH=N bond resulting to decrease in

conjugation. Shifts to lower wavelengths are termed hypsochromic shifts and are

expected where there is a decrease in conjugation length. The extinction coefficients

of [2.1]-[2.5] were higher for the band at lower wavelengths and lower for the band

at higher wavelengths than that of ferrocene (Table 2.5). On the other hand, for

[2.10]-[2.14], the situation was restored to that of the unsubstituted ferrocene.

Page 79: Solvent-free Synthesis of Bisferrocenylimines

65

Figure 2.7: UV-vis spectra of bisferrocenylimines in dichloromethane

Figure 2.8: UV-vis spectra of bisferrocenylamines in dichloromethane

Page 80: Solvent-free Synthesis of Bisferrocenylimines

66

Table 2.5: UV-vis data for ferrocene, [2.1]-[2.5] and [2.10]-[2.14]

Compound λmax(nm) ε [(L.mol-1.cm-1)]

Ferrocene 326 [202] 442 [335]

[2.1] 329 [2586] 454 [914]

[2.2] 323 [3638] 448 [1087]

[2.3] 321 [2934] 450 [852]

[2.4] 323 [3180] 448 [910]

[2.5] 321 [2836] 445 [803]

[2.10] 322 [140] 438 [177]

[2.11] 323 [204] 469 [219]

[2.12] 324 [137] 437 [208]

[2.13] 323 [160] 438 [230]

[2.14] 324 [182] 439 [245]

ε = molar extinction coefficient

2.4 Cyclic Voltammetry

The redox behaviour of bisferrocenylimines and bisferrocenylamines was studied by

cyclic voltammetry. The observed redox behaviour of the compounds was compared

with that of ferrocene as a standard reference. Ferrocene exhibited a one-electron

reversible wave with an E1/2 at 90.5 mV (Figure 2.9). The cyclic voltammograms

were recorded in acetonitrile with tetrabutylammonium perchlorate (0.1 M) as a

supporting electrolyte, in an inert environment. The three-electrode system was a

platinum disk working electrode, a platinum wire auxiliary electrode and Ag/AgNO3

reference electrode.

Some typical cyclic voltammograms of selected compounds are shown in Figure

2.10, and show one-electron reversible redox waves similar to that of ferrocene. The

bisferrocenylimines exhibited a positive shift in potential indicating that these

compounds became more difficult to oxidise (Figure 2.10, Table 2.6). These positive

shifts can be attributed to the presence of the CH=N bond in close proximity to the

ferrocene group, resulting in the reduced electron density at the metal centre.

Page 81: Solvent-free Synthesis of Bisferrocenylimines

67

Figure 2.9: Cyclic voltammogram of ferrocene in acetonitrile

As expected the bisferrocenylamines exhibited a negative shift since the CH2-NH

bond had no effect on the metal centre.

Figure 2.10: Cyclic voltammograms of [2.2], [2.5], 2.12] and [2.14]

Page 82: Solvent-free Synthesis of Bisferrocenylimines

68

Table 2.6: Half-wave potentials of [2.1]-[2.5] and [2.12]-[2.14]a

Compound Epa (mV) Epc (mV) E1/2 (mV)

Ferrocene 140 41 90.5

[2.1] 243 124 133.5

[2.2] 246 167 206.5

[2.3] 243 171 207

[2.4] 238 168 203

[2.5] 242 163 202.5

[2.12] 101 23 62

[2.13] 111 35 73

[2.14] 112 25 68.5 a Due to solubility problems acceptable voltammograms of compounds [2.6], [2.10] and [2.11] could

not be recorded.

2.5 Experimental

2.5.1 Purification procedures

All reagents and solvents were purified using standard purification and drying

methods.10

Table 2.7: General drying agents for solvents

Solvent Drying Agent

Diethyl ether Na wire

Methanol Mg turnings, I2

Ferrocenecarboxaldehyde, 2,3- and 2,5-dimethylbenzaldehyde and 2,4,6-

trimethylbenzaldehyde were purchased from the Sigma Aldrich Chemical Company.

All other common laboratory chemicals were obtained locally and were used without

further purification.

Page 83: Solvent-free Synthesis of Bisferrocenylimines

69

2.5.2 Instrumentation

Melting points were determined on an Electrothermal IA 900 series digital melting

point apparatus and were uncorrected.

NMR spectra were recorded on a Bruker DPX (300 MHz) spectrometer at ambient

temperatures. 1H NMR spectra were referenced against the deuterated solvent

(CDCl3: δ 7.28) and the values reported relative to tetramethylsilane (TMS: δ 0.00). 13C NMR spectra were similarly referenced internally to the solvent resonance

(CDCl3: δ 77.0) with values reported relative to tetramethylsilane (TMS: δ 0.00).

Infrared spectra were recorded on a DigiLab FTS 3100 Excalibur HE series, running

DigiLab Resolution 4.0 software with solid samples prepared as potassium bromide

(KBr) disks. Microanalyses were obtained on a Carlo Erba EA 1108 elemental

analyser at the University of Cape Town. Fast atomic bombardment (FAB) and high

resolution (EI) mass spectra were recorded on a micromass auto-Tof mass

spectrometer at the Witwatersrand University in South Africa.

Uv-vis spectra were recorded on a Hewlett Packard 8452A diode array spectrometer

in dichloromethane (10-3 M) with a cell width of 1 cm.

Cyclic voltammograms were obtained on a BAS 100B electrochemical analyser with

a three-electrode system using Ag/AgNO3 (0.01 M) as a reference electrode,

platinum wire as the auxiliary electrode and platinum disc as the working electrode.

Samples (10-3 M) were prepared and run under nitrogen at ambient temperatures, in

acetonitrile with tetrabutylammonium perchlorate (0.1 M) as a background

electrolyte. The scan rate used was 100 mV.s-1. Solutions were saturated with

nitrogen by bubbling for 10 minutes prior to each run. The system gave ferrocene

E1/2 = 90.5 mV.

Page 84: Solvent-free Synthesis of Bisferrocenylimines

70

2.6 Synthesis of bisferrocenylimines and arylbisimines

2.6.1 General procedure for the synthesis of bisferrocenylimines

Ferrocenecarboxaldehyde (2 mole equivalents) and the diamine (1 mol equivalent)

were added to a pyrex tube fitted with glass ground joint. The two compounds were

ground together at room temperature (ca. 25ºC) using a glass rod The pyrex tube

was then placed under a high vacuum pump overnight. The products were obtained

as yellow to orange solids after recrystallization from cold anhydrous methanol.

2.6.1.1 N,N’-Ethylenebis(ferrocenylmethylidene)imine [2.1]

The general procedure was followed using

ferrocenecarboxaldehyde (360 mg, 1.68

mmol) and ethylenediamine (50 mg, 0.84

mmol). N,N’-Ethylenebis(ferrocenylmethylidine)amine was obtained as a yellow solid

(380 mg, 99%). M.p. 147-150 °C; 1H NMR (CDCl3): 8.18 (2H, s, N=CH), 4.63 (4H, t,

J = 1.8, C5H4), 4.30 (4H, t, J = 1.8, C5H4), 4.16 (10H, s, C5H5), 3.78 (4H, s, 2 x CH2); 13C NMR (CDCl3): 162.74, 80.85, 70.75, 69.52, 68.86, 62.80; IR (KBr): 3113, 3071,

2917, 2897, 2832, 1782, 1705, 1643, 1462, 1412, 1381, 1327, 1281, 1246, 1215,

1211, 1103, 1049, 1011, 961, 891, 822, 768, 644, 594, 517, 486, 475, 436, 401; m/z

(EI): 453 (29%), 452 ([M+], 80%), 321 (13%), 256 (31%), 241 (22%), 227 (19%) 226

(63%), 213 (92%), 199 (27%), 186 (14%), 160 (14%), 121 (63%), 69 (21%), 56

(16%), 43 (11%), 41(11%), 32 (32%), 30 (11%), 28 (100%); Anal. Calc. for

C24H24N2Fe2: MW, 452.15076. Found: MW, 452.063965.

2.6.1.2 N,N’-Propylenebis(ferrocenylmethylidene)imine [2.2]

The general procedure was followed using

ferrocenecarboxaldehyde (200 mg, 0.93

mmol) and 1,3-diaminopropane (41 mg, 0.55

mmol). N,N’-Propylenebis(ferrocenylmethylidine)amine was obtained as an orange

solid (204 mg, 94%). M.p. 127-129 °C; 1H NMR (CDCl3) 8.18 (2H, s, N=CH), 4.67

(4H, t, J = 1.8, C5H4), 4.39 (4H, t, J = 1.8, C5H4), 4.21 (10H, s, C5H5), 3.56 (4H, t, J =

Fe

CH

(CH2)

Fe

CH

N2

N

Fe

CH

(CH2)

Fe

CH

N3

N

Page 85: Solvent-free Synthesis of Bisferrocenylimines

71

7.0, 2 x CH2), 2.03 (2H, m, CH2); 13C NMR (CDCl3) 160.57, 81.88, 70.28, 69.25,

68.64, 59.25, 32.90; IR (KBr) 3108, 3066, 2929, 2946, 2866, 2823, 1640, 1470,

1449, 1323, 1243, 1105, 1004, 823, 544; m/z (EI) 467 (24%), 466 ([M+], 100%), 401

(30%), 335 (56%), 255 (55%), 254 (38%), 253 (30%), 241 (24%), 240 (60%), 233

(26%), 227 (62%), 226 (24%), 225 (21%), 212 (36%), 199 (38%), 186 (52%), 129

(25%), 120 (90%), 56 (52%), 39 (28%); Anal. Calc. for C25H26N2Fe2: C, 64.41; H,

5.62; N, 6.01; MW, 466.17734. Found: C, 63.15; H, 5.82; N, 5.75; MW, 466.082825.

2.6.1.3 N,N’-Butylenebis(ferrocenylmethylidene)imine [2.3]

The general procedure was followed using

ferrocenecarboxaldehyde (200 mg, 0.93

mmol) and 1,4-diaminobutane (44 mg, 0.50

mmol). N,N’-Butylenebis(ferrocenylmethylidine)amine was isolated as an orange

solid (207 mg, 92%). M.p. 152-154 °C; 1H NMR (CDCl3) 8.19 (2H, s, CH=N), 4.65

(2H, t, J = 1.8, C5H4), 4.36 (2H, t, J = 1.7, C5H4), 4.19 (10H, s, C5H5), 3.47 (4H, t, J =

7.0, 2 x CH2), 1.71 (4H, m, 2 x CH2); 13C NMR (CDCl3) 160.29, 81.08, 70.69, 69.47,

68.78, 62.07, 29.09; IR(KBr) 3071, 2930, 2860, 2814, 2364, 1644, 1470, 1439, 1407,

1381, 1369, 1324, 1244, 1104, 1041, 1020, 1001, 962, 930, 819; m/z (EI) 481 (5%),

480 ([M+], 13%), 284 (15%), 268 (23%), 267 (100%), 213 (15%), 199 (17%), 121

(50%), 56 (18%), 55 (15%), 44 (25%), 43 (34%), 41 (26%), 39 (22%), 30 (67%), 28

(61%), 27 (20%). Anal. Calc.1 for C26H28N2Fe2: C, 65.03; H, 5.88, N, 5.83; MW,

480.20392. Found: C, 64.51; H, 6.05; N, 5.74; MW, 480.095172.

2.6.1.4 N,N’-Hexylenebis(ferrocenylmethylidene)imine [2.4]

The general procedure was followed using

ferrocenecarboxaldehyde (100 mg, 0.47

mmol) and 1,6-diaminohexane (24 mg, 0.23

mmol). N,N’-Hexylenebis(ferrocenylmethylidine)amine was isolated as a light yellow

solid (109 mg, 94 %). M.p. 109-111 °C; 1H NMR (CDCl3) 8.16 (2H, s, CH=N), 4.65

(4H, t, J = 1.8, C5H4), 4.36 (4H, t, J = 1.8, C5H4), 4.19 (10H, s, C5H5), 3.45 (4H, t, J =

1 The tendency for the C and N content to be too low and H too high could be due to the presence of traces of water. This also applies to compounds [2.4] and [2.5].

Fe

CH

(CH2)

Fe

CH

N4

N

Fe

CH

(CH2)

Fe

CH

N6

N

Page 86: Solvent-free Synthesis of Bisferrocenylimines

72

6.2, 2 x CH2), 1.66 (4H, m, 2 x CH2), 1.44 (4H, m, 2 x CH2); 13C NMR (CDCl3)

161.08, 81.96, 70.67, 69.46, 68.78, 62.30, 31.32, 27.61; IR (KBr) 3099, 2935, 2862,

2822, 1646, 1468, 1454, 1409, 1381, 1351, 1326, 1243, 1204, 1164, 1103, 1063,

1051, 1038, 1004, 957, 936, 876, 865, 846, 826, 807; m/z (EI) 509 (15%), 508 ([M+],

41%), 312 (20%), 296 (19%), 295 (71%), 214 (23%), 213 (26%), 199 (27%), 186

(30%), 121 (100%), 56 (34%), 55 (22%), 43 (24%), 41 (39%), 39 (41%), 30 (27%),

28 (49%), 27 (33%). Anal. Calc. for C28H32N2Fe2: C, 66.17; H, 6.35; N, 5.51; MW,

508.25708. Found: C, 64.31, H, 6.46; H, 5.40; MW, 508.126146.

2.6.1.5 N,N’-Octylenebis(ferrocenylmethylidene)imine [2.5]

The general procedure was followed using

ferrocenecarboxaldehyde (200 mg, 0.93

mmol) and 1,8-diaminooctane (239mg, 0.47

mmol). N,N’-Octylenebis(ferrocenylmethylidine)amine was obtained as a yellow solid

(231 mg, 97 %). M.p. 97-100 °C; 1H NMR (CDCl3) 8.15 (2H, s, N=CH), 4.65 (4H, t, J

= 1.8, C5H4), 4.36 (4H, t, J = 1.8, C5H4), 4.19 (10H, s, C5H5) 3.42 (4H, t, J = 6.6, 2 x

CH2), 1.62 (4H, m, 2 x CH2), 1.40 (8H, s, 4 x CH2); 13C NMR (CDCl3) 161.00, 81.13,

70.66, 69.46, 68.78, 62.35, 31.33, 29.84, 27.73; IR (KBr) 3065, 2923, 2848, 2819,

1646, 1612, 1494, 1471, 1371, 1327, 1243, 1106, 1043, 1022, 1001, 950, 824, 768,

725, 545; m/z (EI) 537 (29%), 536 ([M+],100%), 471 (20%), 341 (14%), 340 (79%),

268 (19%), 226 (15%), 213 (19%), 199 (26%), 186 (15%), 121 (49%), 55 (16%), 43

(14%), 30 (43%). Anal.Calc. for C30H36N2Fe2: C, 67.19; H, 6.77; N, 5.22; MW,

536.31024. Found: 65.23; H, 7.19; N, 5.59; MW, 536.157307.

2.6.1.6 N,N’-Ethylenebis(4-phenylferrocenylmethylidene)imine [2.6]

The general procedure was followed using 4-

Ferrocenylbenzaldehyde (75 mg, 0.26 mmol)

and ethylenediamine (10 mg, 0.15 mmol).

N,N’-Ethylenebis(4-phenylferrocenylmethylidene)amine was obtained as an orange

solid (73 mg, 97 %). M.p 208-210 °C; 1H NMR (CDCl3) 8.33 (2H, s, N=CH), 7.63 (4H,

dd, J = 8.3, C6H4), 7.49 (4H, dd, J = 8.3, C6H4), 4.69 (4H, t, J = 1.8, C5H4), 4.38 (4H,

t, J = 1.8, C5H4), 4.06 (10H, s, C5H5), 3.99 (4H, s, 2 x CH2); 13C NMR (CDCl3)

Fe

CH

(CH2)

Fe

CH

N8

N

Fe Fe

CH N N CH

Page 87: Solvent-free Synthesis of Bisferrocenylimines

73

162.99, 128.59, 126.41, 85.01, 70.31, 70.12, 69.81, 67.05; IR (KBr) 3098, 2913,

2847, 2230, 1638, 1605, 1566, 1528, 1454, 1420, 1373, 1362, 1308, 1281, 1227,

1181, 1103, 1084, 1018, 995, 887, 864, 822, 648, 513, 509, 451; m/z (EI) 605 (38%),

604 ([M+, 75%), 302 (17%), 287 (14%), 180 (7%), 152 (9%), 139 (7%), 121 (48%),

69 (10%), 63 (22%), 62 (17%), 56 (38%), 43 (18%), 39 (69%), 38 (28%), 37 (15%),

29 (35%), 28 (100%), 27 (66%), 26 (30%); Anal. Calc. for C36H32N2Fe2: MW,

604.34268. Found: MW, 604.134521.

2.6.1.7 N,N’-Bis-(2,3-dimethylbenzylidene)-ethane-1,2-diimine [2.7]

2,3-Dimethylbenzaldehyde (202 mg, 1.5

mmol) and ethylenediamine (45 mg, 0.91

mmol) were added into a 25 cm3 round-

bottomed flask. The two compounds

were stirred using a magnetic stirrer at room temperature (ca. 25ºC). The round

bottomed flask was placed under high vacuum overnight. N, N’-Bis-(2,3-

dimethylbenzylidine)-ethane-1,2-diamine was obtained as a white powder (188 mg,

86 %). M.p. 119-121 °C; 1H NMR (CDCl3) 8.68 (2H, s, CH=N), 7.69 (2H, d, J = 7.6,

Ar-H), 7.20 (2H, d, J = 7.2, Ar-H), 7.10 (2H, t, J = 7.6, Ar-H), 3.95 (4H, s, 2 x CH2),

2.33 (6H, s, 2 x CH3), 2.27 (6H, s, 2 x CH3); 13C NMR (CDCl3) 161.62, 137.42,

136.41, 134.99, 131.76, 125.88, 125.67, 62.17, 19.81, 14.10; IR (KBr) 3063, 3005,

2964, 2972, 2907, 1637, 1591, 1458, 1374, 1281, 1261, 1197, 1183, 1092, 1010,

990, 980, 956, 903, 798, 784, 760, 716, 488, 426; m/z (EI) 293 (10%), 292 ([M+],

38%), 291 (35%), 162 (30%), 161 (100%), 160 (65%), 159 (19%), 158 (28%), 147

(20%), 146 (86%), 144 (26%), 134 (15%), 133 (69%), 132 (90%), 131 (48%), 130

(37%), 119 (70%), 118 (22%), 117 (29%), 116 (23%), 115 (19%), 105 (15%), 103

(16%), 91 (39%), 77 (21%), 69 (19%), 57 (18%), 55 (16%), 43 (16%), 41 (20%), 28

(31%). Anal. Calc. for C20H24N2: MW, 292.41796. Found: MW, 292.193095.

2.6.1.8 N,N’-Bis-(2,5-dimethylbenzylidene)-ethane-1,2-diimine [2.8]

2,5-Dimethylbenzaldehyde (500 mg, 3.7

mmol) and ethylenediamine (112 mg, 1.90

NN

NN

Page 88: Solvent-free Synthesis of Bisferrocenylimines

74

mmol) were added into a 25 cm3 round bottomed flask. The procedure for [2.7] was

followed and N,N’-bis-(2,5-dimethylbenzylidine)-ethane-1,2-diamine was obtained as

a white powder (489 mg, 88 %). M.p. 109-110 °C; 1H NMR (CDCl3) 8.58 (2H, s,

CH=N), 7.66 (2H, s, Ar-H), 7.11 (2H, d, J = 7.1, Ar-H), 7.07 (2H, d, J = 7.8, Ar-H),

3.94 (4H, s, 2 x CH2), 2.39 (6H, s, 2 x CH3), 2.29 (6H, s, 2 x CH3); 13C NMR (CDCl3)

161.21, 135.49, 134.92, 134.51, 131.09, 128.28, 62.15, 20.38, 18.46; IR (KBr) 3019,

2977, 2909, 2878, 2843, 1639, 1609, 1572, 1496, 1463, 1403, 1387, 1372, 1277,

1246, 1211, 1198, 1164, 1117, 1034, 1017, 969, 959, 942, 897, 822, 790, 724, 645,

505, 566, 467, 407; m/z (EI) 293 (20%), 292 ([M+], 47%), 291 (38%), 162 (42%), 161

(100%), 160 (48%), 159 (22%), 158 (39%), 147 (36%), 146 (98%), 145 (24%), 144

(49%), 134 (27%), 133 (85%), 132 (99%), 131 (59%), 130 (51%), 120 (17%), 119

(72%), 118 (26%), 117 (48%), 116 (30%), 115 (33%), 106 (26%), 105 (32%), 104

(23%), 103 (35%), 91 (52%), 79 (22%), 78 (21%), 77 (44%), 65 (19%), 51 (15%), 41

(18%), 39 (17%), 28 (17%). Anal. Calc. for C20H24N2: MW, 292.41796. Found: MW,

292.193616.

2.6.1.9 N,N’-Bis-(2,4,6-trimethylbenzylidene)-ethane-1,2-diimine [2.9]

2,4,6-Trimethylbenzaldehyde (500 mg, 3.4

mmol) and ethylenediamine (101 mg, 1.70

mmol) were added into a 25 cm3 round-

bottomed flask. The same procedure for

[2.7] was used and N,N’-bis-(2,4,6-trimethylbenzylidine)-ethane-1,2-diamine was

obtained as a white crystalline needles (457 mg, 84 %). M.p. 126-127 °C; 1H NMR

(CDCl3) 8.64 (2H, s, CH=N), 6.93 (4H, s, Ar-H), 4.02 (4H, s, 2 x CH2), 2.37 (12H, s, 4

x CH3), 2.28 (6H, s, 2 x CH3); 13C NMR (CDCl3) 162.64, 139.03, 137.89, 131.41,

129.69, 63.89, 21.52, 21.06; IR (KBr) 2957, 2941, 2917, 2881, 2843, 2734, 1637,

1610, 1568, 1481, 1463, 1430, 1392, 1376, 1285, 1223, 1154, 1048, 1032, 1008,

975, 932, 897, 861, 841, 789, 727, 597, 565, 542, 512, 440; m/z (EI) 321 (8%), 320

([M+], 21%), 176 (34%), 174 (94%), 173 (32%), 172 (32%), 161 (27%), 160 (83%),

159 (15%), 158 (32%), 148 (18%), 147 (67%), 146 (100%), 145 (49%), 144 (37%),

133 (39%), 132 (26%), 131 (31%), 130 (32%), 120 (18%), 119 (12%), 117 (18%),

116 (14%), 115 (21%), 105 (27%), 91 (26%), 77 (14%), 41 (15%). Anal. Calc. for

C22H28N2: MW, 320.47112. Found: MW, 320.1650.

NN

Page 89: Solvent-free Synthesis of Bisferrocenylimines

75

2.6.2 Reduction of bisferrocenylimines and arybisimines

2.6.2.1 N,N’-Ethylenebis(ferrocenylmethyl)amine [2.10]

To a solution of lithium aluminium hydride (17 mg,

0.44 mmol) in diethyl ether (40 cm3) was added

N,N’-ethylenebis(ferrocenylmethylidene)imine

(100 mg, 0.22 mmol). The resultant mixture was heated under reflux for 1 h, and the

reaction was quenched with ethyl acetate/ice-water slurry. The solution was

extracted with diethyl ether (2 x 30 cm3) and the combined ethereal extracts were

dried over anhydrous sodium sulfate. The solution was filtered and the solvent

removed in vacuo. N,N’-Ethylenebis(ferrocenylmethyl)amine was obtained as a light

yellow powder (88 mg, 87 %). M.p. 82-84 °C; 1H NMR (CDCl3) 4.23 (4H, t, J = 1.8,

C5H4), 4.18 (4H, t, J = 1.9, C5H4), 4.15 (10H, s, C5H5), 3.58 (4H, s, 2 x CH2), 2.84

(4H, s, 2 x CH2), 2.27 (2H, br-s, 2 x NH); 13C NMR (CDCl3) 86.79, 69.18, 68.97,

68.62, 48.72, 30.06; IR (KBr) 3098, 2957, 2925, 2854, 1665, 1626, 1590, 1558,

1472, 1430, 1406, 1364, 1314, 1286, 1263, 1245, 1122, 1105, 1084, 1037, 1025,

1001, 819, 771, 668, 649, 637, 497, 482, 452; m/z (EI) 457 ([M+ +1], 8%), 456 ([M+],

6%), 455 (2%), 289 (9%), 257 (4%), 199 (100%), 154 (48%), 136 (49%); Anal. Calc.

for C24H28N2Fe2: MW, 456.18252. Found: MW, 456.19.

2.6.2.2 N,N’-Propylenebis(ferrocenylmethyl)amine [2.11]

To a solution of lithium aluminium hydride (21 mg,

0.56 mmol) in diethyl ether (40 cm3) was added

N,N’-propylenebis(ferrocenylmethylidene)imine (131

mg, 0.28 mmol). The procedure for [2.10] was

followed and N,N’-propylenebis(ferrocenylmethyl)amine was obtained as a yellow

powder (121 mg, 92 %). M.p. 86-88 °C; 1H NMR (CDCl3) 4.18 (4H, t, J = 1.6, C5H4),

4.16 (10H, s, C5H5), 4.13 (4H, t, J = 1.6, C5H4), 3.53 (4H, s, 2 x CH2), 2.72 (4H, t, J =

6.8, 2 x CH2), 2.02 (2H, br-s, 2 x NH), 1.72 (2H, m, CH2); 13C NMR (CDCl3) 86.78,

68.90, 68.83, 68.25, 49.36, 48.45, 30.08; IR (KBr) 3102, 2928, 2831, 1556, 1468,

Fe

CH2

(CH2)

Fe

CH2

HN

2

HN

Fe

CH2

(CH2)

Fe

CH2

HN

3

HN

Page 90: Solvent-free Synthesis of Bisferrocenylimines

76

1435, 1411, 1402, 1390, 1262, 1103, 1062, 1035, 1018, 1001, 929, 825, 807, 776,

668, 646, 499, 490, 470; m/z 471 ([M+ +1], 16%), 470 ([M+], 8%), 307 (8%), 289

(8%), 199 (100%). Anal. Calc. for C25H30N2Fe2: MW, 470.2091. Found: MW,

469.9878.

2.6.2.3 N,N’-Butylenebis(ferrocenylmethyl)amine [2.12]

To a solution of lithium aluminium hydride (32 mg,

0.83 mmol) in diethyl ether (40 cm3) was added N,N’-

butylenebis(ferrocenylmethylidene)imine (200 mg,

0.42 mmol). The procedure for [2.10] was followed and N,N’-

butylenebis(ferrocenylmethyl)amine was obtained as a yellow powder (174 mg, 86

%). M.p.98-99 °C; 1H NMR (CDCl3) 4.19 (4H, t, J = 1.8, C5H4), 4.13 (10H, s, C5H5),

4.12 (4H, t, J = 1.9, C5H4), 3.54 (4H, s, 2 x CH2), 2.65 (4H, t, J = 6.4, 2 x CH2), 2.07

(2H, br-s, 2 x NH), 1.25 (4H, m, 2 x CH2); 13C NMR (CDCl3) 86.59, 68.97, 68.84,

68.31, 49.56, 49.22, 28.26; IR (KBr) 3098, 3056, 2896, 2866, 2801, 1471, 1453,

1442, 1410, 1396, 1383, 1318, 1259, 1227, 1151, 1114, 1105, 1045, 1037, 1028,

998, 968, 923, 900, 878, 857, 848, 834, 804, 774, 737, 724, 623, 517, 498, 487, 463,

415; m/z 485 ([M+ +1], 21%), 484 ([M+], 9%), 483 (4%), 307 (5%), 285 (6%), 199

(100%). Anal. Calc. for C26H32N2Fe2: MW, 484.23568. Found: MW, 484.09989.

2.6.2.4 N,N’-Hexylenebis(ferrocenylmethyl)amine [2.13]

To a solution of lithium aluminium hydride (30 mg,

0.78 mmol) in diethyl ether (40 cm3) was added N,N’-

hexylenebis(ferrocenylmethylidene)imine (200 mg,

0.39 mmol). The procedure for [2.10] was followed

and N,N’-hexylenebis(ferrocenylmethyl)amine was obtained as a yellow powder (183

mg, 91 %). M.p. 102-103 °C; 1H NMR (CDCl3) 4.20 (4H, t, J = 1.8, C5H4), 4.13 (10H,

s, C5H5), 4.12 (4H, t, J = 1.8, C5H4), 3.52 (4H, s, 2 x CH2), 2.63 (4H, t, J = 7.1, 2 x

CH2), 1.53 (2H, br-s, 2 x NH), 1.35 (4H, m, 2 x CH2), 1.77 (4H, m, 2 x CH2); 13C

(CDCl3) 87.16, 68.90, 68.83, 68.25, 49.90, 49.44, 30.34, 27.69; IR (KBr) 3098, 2924,

2849, 2818, ,1478, 1452, 1434, 1309, 1247, 1227, 1210, 1153, 1121, 1105, 1043,

1021, 1000, 960, 923, 873, 822, 768, 731, 692, 665, 615, 520, 489, 481; m/z 513

Fe

CH2

(CH2)

Fe

CH2

HN

4

HN

Fe

CH2

(CH2)

Fe

CH2

HN

6

HN

Page 91: Solvent-free Synthesis of Bisferrocenylimines

77

([M+ +1], 9%), 512 (10%), 511 (4%), 313 (7%), 307 (3%), 289 (3%), 199 (100%), 154

(22%). Anal. Calc. for C28H36N2Fe2: MW, 512.28884. Found: MW, 511.7.

2.6.2.5 N,N’-Octylenebis(ferrocenylmethyl)amine [2.14]

To a solution of lithium aluminium hydride (19. mg,

0.52 mmol) in diethyl ether (40 cm3) was added

N,N’-octylenebis(ferrocenylmethylidene)imine (137

mg, 0.26 mmol). The procedure for [2.10] was

followed and N,N’-octylenebis(ferrocenylmethyl)amine was obtained as a yellow

powder (121 mg, 88 %). M.p. 65-66 °C; 1H NMR (CDCl3) 4.20 (4H, t, J = 1.7, C5H4),

4.13 (10H, s, C5H5), 4.13 (4H, C5H4), 3.52 (4H, s, 2 x CH2), 2.62 (4H, t, J = 7.0, 2 x

CH2), 1.67 (2H, br-s, 2 x NH), 1.49 (4H, m, 2 x CH2), 1.31 (8H, s, 4 x CH2); 13C NMR

(CDCl3) 87.28, 68.88, 68.79, 68.15, 50.03, 49.46, 30.39, 29.88, 27.71; IR (KBr)

3095, 2923, 2852, 1473, 1432, 1409, 1326, 1240, 1150, 1120, 1105, 1044, 1018,

999, 961, 923, 884, 864, 848, 819, 772, 723, 668, 614, 523, 501, 484, 456; m/z 541

([M+ +1], 15%), 540 ([M+], 23%), 342 (5%), 341 (15%), 215 (4%), 199 (100%). Anal.

Calc. for C30H40N2Fe2: MW, 540.342. Found: MW, 539.9.

2.6.2.6 N,N’-Bis(2,3-dimethylbenzyl) ethane-1,2-diamine[2.15]

To a solution of lithium aluminium hydride (17 mg,

0.44 mmol) in diethyl ether (40 cm3) was added

N,N’-bis(2,3-dimethylbenzylidene)ethane-1,2-

diimine (65 mg, 0.22 mmol). The procedure for

[2.10] was followed and N,N’-bis(2,3-dimethylbenzyl)ethane-1,2-diamine was

obtained as a colourless oil (37 mg, 56%). M.p.; 1H NMR (CDCl3) 7.18-7.15 (2H, m,

Ar-H), 7.11 (4H, br-s, Ar-H), 3.81 (4H, s, 2 x CH2), 2.86 (4H, s, 2 x CH2), 2.32 (6H, s,

2 x CH3), 2.28 (6H, s, 2 x CH3), 1.78 (2H, s, 2 x NH); 13C NMR (CDCl3) 138.4, 137.7,

137.5, 135.4, 129.2, 127.1, 125.8, 52.7, 49.5, 21.0, 15.2; IR (NaCl) 3308, 3066,

3030, 3013, 2937, 2915, 2821, 2732, 2691, 1588, 1463, 1383, 1353, 1330, 1295,

1248, 1183, 1163, 1115, 1090, 1018, 990, 972, 902, 880, 820, 774, 725, 710, 675,

666; m/z; 297 ([M+ +1], 90%), 296 ([M+], 8%), 295 (25%), 171 (5%), 148 (28%), 119

(100%). Anal. Calc. for C20H28N2: MW, 296.44972. Found: MW, 296.1094.

Fe

CH2

(CH2)

Fe

CH2

HN

8

HN

NH

HN

Page 92: Solvent-free Synthesis of Bisferrocenylimines

78

2.6.2.7 N,N’-Bis(2,5-dimethylbenzyl) ethane-1,2-diamine [2.16]

N,N’-Bis(2,5-dimethylbenzylidene)ethane-1,2-

diimine (131 mg, 0.44 mmol) was added to a

solution of lithium aluminium hydride (34 mg,

0.88 mmol) in diethyl ether (40 cm3). The

procedure for [2.10] was followed and N,N’-

bis(2,5-dimethylbenzyl)ethane-1,2-diamine was obtained as a white powder (87 mg,

68%). M.p. 43-46 °C; 1H NMR (CDCl3) 7.13 (2H, s, Ar-H), 7.06 (2H, d, J = 7.6, Ar-H),

6.99 (2H, d, J = 7.6, Ar-H), 3.76 (4H, s, 2 x CH2), 2.86 (4H, s, 2 x CH2), 2.32 (6H, s, 2

x CH3), 2.31 (6H, s, 2 x CH3), 2.00 (2H, s, 2 x NH); 13C NMR (CDCl3) 138.12, 135.77,

133.42, 130.60, 129.64, 128.07, 51.77, 49.24, 21.38, 18.94; IR (KBr) 3265, 3048,

3017, 2971, 2946, 2918, 2894, 2859, 2823, 2758, 2732, 2703, 1610, 1499, 1474,

1457, 1376, 1368, 1303, 1279, 1240, 1229, 1204, 1188, 1156, 1131, 1098, 1051,

1038, 995, 980, 928, 882, 833, 814, 800, 718, 704, 668, 544, 492, 440, 419 ; m/z

297 ([M+ +1], 51%), 296 ([M+], 6%), 295 (18%), 167 (9%), 148 (22%), 119 (100%).

Anal. Calc. for C20H28N2: MW, 296.44972. Found: MW, 296.2178

2.6.2.8 N,N’-Bis(2,4,6-trimethylbenzyl) ethane-1,2-diamine[2.17]

N,N’-Bis(2,4,6-trimethylbenzylidene) ethane-

1,2-diimine (100 mg, 0.31 mmol) was added

to a solution of lithium aluminium hydride

(25 mg, 0.66 mmol) in diethyl ether (40

cm3). The procedure for [2.10] was followed and N,N’-bis(2,4,6-trimethylbenzyl)

ethane-1,2-diamine was obtained as a white powder (79 mg, 79 %). M.p.: 69-72 °C; 1H NMR (CDCl3) 6.85 (4H, s, Ar-H), 3.81 (4H, s, 2 x CH2)), 2.92 (4H, s, 2 x CH2),

2.33 (12H, s, 4 x CH3), 2.26 (6H, s, 2 x CH3), 1.99 (2H, s, 2 x NH); 13C NMR (CDCl3)

137.6, 137.4, 129.6, 129.5, 48.2, 47.2, 21.3, 20.1; IR (KBr) 3202, 3005, 2957, 2926,

2845, 2817, 2789, 1613, 1483, 1461, 1443, 1378, 1355, 1345, 1332, 1262, 1222,

1209, 1129, 1110, 1091, 1029, 1014, 914, 867, 851, 825, 807, 774, 759, 709, 665,

619, 592, 549; m/z 325 ([M+ +1], 37%), 324 ([M+], 4%), 323 (7%), 307 (8%), 297

NH

HN

NH

HN

Page 93: Solvent-free Synthesis of Bisferrocenylimines

79

((14%), 154 (19%), 133 (100%), 120 (44%). Anal. Calc. for C22H32N2: MW,

324.50288. Found: MW, 324.2869.

2.7 References

1. C. Imrie, V. O. Nyamori and T. I. A. Gerber, J. Organomet. Chem., 689 (2004)

1617-1622.

2. A. Benito, J. Cano, R. Martínez-Máňez, J. Soto, J. Payá, F. Lioret, M. Julve, J.

Faus and M. Dolores Marcos, Inorg. Chem., 32 (1993) 1197-1203.

3. V. K. Muppidi, T. Htwe, P. S. Zcharias and S. Pal, Inorg. Chem. Commun., 7

(2004) 1045-1048.

4. I. Ratera, D. Ruiz-Molina, C. Sánchez, R. Alcalá, C. Rovira and J. Veciana,

Synth. Met., 121 (2001) 1834.

5. J. Rajput, PhD Thesis: Platinum group metal coordination complexes of

ferrocenyl-N-donor ligands and their potential application in catalysis and

medicinal chemistry, University of Cape Town, 2003, 34.

6. H. Naeimi, F. Salimi and K. Rabiei, J. Mol. Cat. A, 260 (2006) 100-104.

7. M. Woltersdorf, R. Kranich and H.-G. Schmalz, Tetrahedron, 53 (1997) 7219.

8. R. M. Silverstein, F. X. Webster and D. J. Kiemle, Spectroscopic Identification

of Organic Compounds, 7th Ed., John Wiley & Sons, Inc. 2005.

9. R. C. J. Atkinson, V. C. Gibson and N. J. Long, Chem. Soc. Rev., 33 (2004)

313.

10. B. S. Furniss, A. J. Hannaford, P. W. G. Smith and A. R. Tatchell, Vogel’s

Textbook of Practical Organic Chemistry, Longman Scientific and Technical,

England (5th Ed), 1989.

Page 94: Solvent-free Synthesis of Bisferrocenylimines

80

CHAPTER 3

RESULTS AND DISCUSSION

3.1 Synthesis of cationic rhodium(I) complexes

3.1.1 Rhodium(I) complexes containing bisferrocenylimines

The bisferrocenylimines described in the previous chapter were prepared in order for

them to be reacted with a rhodium(I) metal centre, to form cationic rhodium(I)

complexes. Rhodium(I) is understood to form square planar coordination complexes

with π-acceptor ligands and some five-coordinate complexes are known as well.1

Cationic rhodium complexes are known for their application in the field of catalysis.

An appropriate metal precursor for the synthesis of rhodium(I) complexes is the

chloro-bridged rhodium cyclooctadienyl dimer, chloro(1,5-cyclooctadiene)rhodium(I).

The dimer can readily be obtained from the reaction of 1,5-cyclooctadiene with

rhodium trichloride trihydrate under reflux.1 The type of complex formed from the

dimer is dependent on the nature of the ligand and the ratio of the metal to ligand.2

The cationic rhodium(I) complexes were synthesized using a literature procedure as

illustrated in Scheme 3.1.3 A solution of silver perchlorate in acetone was added to a

solution of the rhodium dimer in acetone. On precipitation of silver chloride, a

solvated complex of general formula [Rh(COD)(acetone)2]ClO4 was formed.3,4 The

addition of a bisferrocenylimine ligand resulted in a cationic complex by

displacement of the coordinated solvent from the rhodium coordination sphere. The

complexes [3.1] (x = 2), [3.2] (x = 3) and [3.3] (x = 4) were obtained in low to

excellent yields by redissolving the residue, after concentration of solvent, with

dichloromethane and precipitated by addition of diethyl ether. The complexes were

further purified by recrystallization. Complex [3.2] had the highest yield while [3.1]

was obtained in lowest yield. This effect was attributed to the stability of the six-

membered ring formed by the bidentate ligand [2.2] with the metal centre compared

to five- and seven-membered rings formed by [2.1] and [2.3], respectively. Attempts

to prepare complexes containing [2.4], [2.5] and [2.7]-[2.9] were not successful.

Page 95: Solvent-free Synthesis of Bisferrocenylimines

81

+ 2 AgClO4

acetone[Rh(COD)(acetone)2]ClO4 + AgCl(s)Rh

Cl

ClRh

Fe

CH

(CH2)

Fe

CHN

xN

Rh

2L

ClO4

[3.0]

x = 2: [3.1] = 3: [3.2] = 4: [3.3]

Scheme 3.1: Procedure for the synthesis of cationic rhodium(I) complexes.3

1H NMR data for [3.1]-[3.3] is summarized in Table 3.1. The imine (CH=N) chemical

shifts for [3.2] and [3.3] were observed in the expected region while for [3.1] the

peak was shifted remarkably to lower frequencies. This effect has also been

observed by Lee et al.3 and could not be explained (Figure 3.1). The CH=N signal in

[3.1] moved from δ 8.17 ppm in the free ligand to δ 7.42 ppm in the complex. The

CH=N signal moved to higher frequencies for [3.2] and [3.3]. This signal was

observed to have shifted from δ 8.17 ppm in the free ligand [2.2] to δ 8.31 ppm in the

complex. On the other hand, the signal for [3.3] moved from δ 8.18 ppm in the free

ligand [2.3] to δ 8.22 ppm in the complex. In the ferrocene region, a sharp singlet

was observed at δ 4.37 and 4.07 ppm for [3.1] and [3.2] respectively, and it was

assigned to the unsubstituted Cp ring. Two singlets at δ 4.73 and 4.80 ppm for [3.1]

were assigned to the substituted Cp ring. The substituted Cp ring signals for [3.2]

were observed at δ 4.23 and 4.72 ppm. The appearance of additional signals in the

ferrocene region of [3.3] complicated the assignments.

Page 96: Solvent-free Synthesis of Bisferrocenylimines

82

Table 3.1: The summarized NMR data for [3.1]-[3.3]

Complex Yield

(%)a

1H (ppm) 13C (ppm)

CH=N COD Fc CH=N COD Fc

[3.1]

32

(47.0)b

7.42 4.16,

2.64,

2.09,

4.80,

4.73,

4.37

168.48 84.49,

30.55

74.70,

73.22,

70.68,

84.65

[3.2]

84

(37.1)b

8.31 5.48,

5.39,

2.58,

1.97

4.72,

4.23,

4.07

170.44 75.61,

30.89

71.73,

70.51,

70.28,

84.89

[3.3]

70

(39.0)b

8.22 6.11,

5.65,

2.62,

1.79

c

170.39 75.63,

31.26

c

a Isolated yields are based on starting reagents.

b Numbers in parentheses are conductivity values in Ohm-1.cm2.mol-1

c The appearance of additional signals in the ferrocene region complicated the assignments.

Two sharp and equally intense singlets were observed at δ 4.26 and 4.14 ppm and

are assigned to the unsubstituted Cp rings. This suggests that the Cp rings could be

chemically inequivalent due to the spatial orientation of the ferrocene, something that

was not observed in [3.1] and [3.2]. Other signals could not be assigned due to the

complexity of the signals in the region. Initially, it was thought that the extra signals

were due to the presence of impurities. However, the signals were persistent even

after recrystallization several times. The COD ligand exhibited the expected patterns

for [3.1] giving rise to a singlet due to CH=CH protons at δ 4.16 ppm, a multiplet δ

2.64 ppm and a doublet δ 2.09 ppm due to CH2 protons. For [3.2] and [3.3] the

CH=CH proton signal was split into two singlets at δ 5.48 and 5.39 ppm and 6.11

Page 97: Solvent-free Synthesis of Bisferrocenylimines

83

and 5.65 ppm, respectively. The CH2 signals were

Figure 3.1: 1H NMR spectra of [3.2] (top) and [3.3] (bottom) in CDCl3

observed in the expected region for both [3.2] and [3.3]. 13C NMR showed the

exhibited C=N signals in the expected region, δ 168.48, 170.44 and 170.39 ppm for

[3.1], [3.2] and [3.3], respectively. The ferrocene signals for [3.1] and [3.2] were

observed in the expected region while the same problem as with the 1H NMR was

experienced for [3.3]. The COD signals were also observed in the expected region

Page 98: Solvent-free Synthesis of Bisferrocenylimines

84

for all complexes. IR spectra of all the compounds showed that the v(C=N) stretching

signals have moved to lower frequencies. The infrared together with the NMR

information explain the coordination of the ligands to the rhodium metal centre.

3.1.2 X-ray Crystallography

Crystals of [3.2] that were suitable for X-ray crystallographic analysis were obtained

by slow diffusion of diethyl ether into a solution of the complex in dichloromethane.

The complex was observed to crystallize in a triclinic space group P1bar with Z = 4

and contains two molecules (structural isomers) in an asymmetric unit. The structure

was refined successfully with the final R factor of 0.0456.

Table 3.2: Crystal data and structure refinement of [3.2] and [3.3]

Complex [3.2] Complex [3.3]

Empirical formula C33H38ClFe2N2O4Rh C34H40ClFe2N2O4Rh

Formula weight 776.71 790.74

Temperature 113(2) K 173(2) K

Wavelength 0.71073 Å 0.71073

Crystal system Triclinic Monoclinic

Space group P1bar P21/c

Unit cell dimensions

a 12.0890(2) Å 12.983(3) Å

b 16.5931(2) Å 13.536(2) Å

c 17.5512(3) Å 17.729(4) Å

α 67.8240(10)˚ 90˚

β 72.9460(10)˚ 92.580(7)˚

γ 72.3860(10)˚ 90˚

Volume 3042.77(8) A3 3112.6(11) A3

Z 4 4

Calculated density 1.696 Mg/m3 1.687 Mg/m3

Reflections collected 70435 / 11534 48758 / 5716

Unique [R (int) = 0.1029 [R (int) = 0.1694

Goodness-of-fit on F2 1.023 1.034

Final R indices R1 = 0.0456, wR2 = 0.0875 R1 = 0.0611, wR2 =0.0934

R indices (all data) R1 = 0.0758, wR2 = 0.0987 R1 = 0.1324, wR2 = 0.1119

Page 99: Solvent-free Synthesis of Bisferrocenylimines

85

All hydrogen atoms were placed geometrically with fixed bond length and refined

with isotropic displacement parameters depending on their carbon atoms. The

parameters for crystal data collection and structure refinements are in Table 3.2.

The ORTEP drawing shown in Figure 3.2 confirms the molecular structure of [3.2].

The bond lengths, angles, torsion angles and other parameters are in Table 3.3. The

rhodium atom is oriented in an essentially square planar geometry defined by two

nitrogen atoms of the bidentate ligand and the two C=C double bonds of the COD

ligand.

Figure 3.2: ORTEP diagram of [3.2]

The six-membered ring formed by the rhodium atom, the two nitrogen atoms of the

ligand and the three carbon atoms of the alkyl chain separating the two nitrogen

atoms is in a chair conformation. The chair conformation is the lowest energy state

that a six-membered ring can be found in, which explains the reason for [3.2] being

obtained in excellent yields. The Rh(1A)-N(1A) and Rh(1A)-N(2A) bond distances for

molecule A are 2.077(4) and 2.095(4) Å, respectively. For molecule B, the Rh(1B)-

Page 100: Solvent-free Synthesis of Bisferrocenylimines

86

N(2B) and Rh(1B)-N(1B) the bond distances are 2.076(4) and 2.088(4) Å,

respectively. The bond distances Rh(1A)-C(2A), Rh(1A)-C(6A), Rh(1A)-C(1A) and

Rh(1A)-C(5A) for molecule A, are 2.144(4), 2.150(5), 2.152(4) and 2.158(4) Å,

respectively. On the other hand, for molecule B the bond distances Rh(1B)-C(2B),

Rh(1B)-C(5B), Rh(1B)-C(6B) and Rh(1B)-C(1B) are 2.130(4), 2.143(4), 2.144(4) and

2.171(4) Å, respectively.

Table 3.3: Selected bond distances, bond angles and torsion angles of [3.2]

Molecule A Molecule B

Rh(1A)-N(1A) 2.077(4) Rh(1B)-N(2B) 2.076(4)

Rh(1A)-N(2A) 2.095(4) Rh(1B)-N(1B) 2.088(4)

Rh(1A)-C(2A) 2.144(4) Rh(1B)-C(2B) 2.130(4)

Rh(1A)-C(6A) 2.150(5) Rh(1B)-C(5B) 2.143(4)

Rh(1A)-C(5A) 2.152(4) Rh(1B)-C(6B) 2.144(4)

Rh(1A)-C(2B) 2.158(4) Rh(1B)-C(1B) 2.171(4)

N(1A)-C(9A) 1.294(6) N(1B)-C(9B) 1.289(6)

N(2A)-C(13A) 1.277(5) N(2B)-C(13B) 1.286(6)

N(1A)-(Rh(1A)-N(2A) 85.92(14) N(2B)-(Rh(1B)-N(1B) 85.06(14)

N(1A)-(Rh(1A)-C(2A) 92.34(16) N(2B)-(Rh(1B)-C(2B) 162.39(16)

N(2A)-(Rh(1A)-C(2A) 171.25(16) N(1B)-(Rh(1B)-C(2B) 95.29(16)

N(1A)-(Rh(1A)-C(6A) 154.73(17) N(2B)-(Rh(1B)-C(5A) 95.04(16)

N(2A)-(Rh(1A)-N(6A) 89.01(16) N(1B)-(Rh(1B)-C(5A) 174.44(16)

C(2A)-(Rh(1A)-N(CA) 96.11(18) C(2B)-(Rh(1B)-C(5A) 82.94(18)

N(2A)-Rh(1A)-N(2A)-C(9A) 112.4(4) N(2B)-Rh(1B)-N(1B)-C(9B) 112.8(4)

C(2A)-Rh(1A)-N(1A)-C(9A) -59.0(4) C(2B)-Rh(1B)-N(1B)-C(9B) -49.6(4)

C(6A)-Rh(1A)-N(1A)-C(9A) -168.7(4) C(5B)-Rh(1B)-N(1B)-C(9B) 21.5(19)

C(1A)-Rh(1A)-N(1A)-C(9A) -96.7(4) C(6B)-Rh(1B)-N(1B)-C(9B) -163.7(4)

C(5A)-Rh(1A)-N(1A)-C(9A) 3.1(10) C(1B)-Rh(1B)-N(1B)-C(9B) -86.9(4)

N(2A)-Rh(1A)-N(1A)-C(10A) -65.1(3) N(2B)-Rh(1B)-N(1B)-C(10B) -65.5(3)

The bond distances that have just been mentioned were found to be comparable to

with literature-cited bond distances for similar complexes.5 The bite angles N(1A)-

Rh(1A)-N(2A) and N(2B)-Rh(1B)-N(1B) are 85.92(14) and 85.06(14) for molecule A

and molecule B, respectively. The deviation from the ideal 90° bond angles (square

planar) around the Rh-atom was due to the steric bulk of the COD ligand.5 The bond

distances N(1A)-C(9A) and N(2A)-C(13A) are 1.294(6) and 1.277(5) and are typical

Page 101: Solvent-free Synthesis of Bisferrocenylimines

87

bond distances for the C=N bond. The crystal packing in the unit cell of [3.2] is

shown in Figure 3.3 and no intermolecular interactions were exhibited.

Figure 3.3: Crystal packing of [3.2], projection viewed along [100]

Crystals of [3.3] that were suitable for X-ray crystallographic analysis were obtained

by slow diffusion of diethyl ether into a solution of the complex in dichloromethane.

The complex was observed to crystallize in a monoclinic space group P21/c with Z =

4. The structure was refined successfully with the final R factor of 0.0611. All

hydrogen atoms were fixed in geometrically calculated positions with Uiso set at 1.2

or 1.5 those of the parent atoms. The parameters for crystal data collection and

structure refinements are in Table 3.2.

The ORTEP drawing shown in Figure 3.4 confirms the molecular structure of [3.3].

The bond lengths, angles, torsion angles and other parameters are in Table 3.4. The

rhodium atom is oriented in an essentially square planar geometry defined by two

nitrogen atoms of the bidentate ligand and the two C=C double bonds of the COD

ligand.

Page 102: Solvent-free Synthesis of Bisferrocenylimines

88

Figure 3.4: ORTEP drawing of [3.3]

Figure 3.5: Crystal packing of [3.3], projection viewed along [100]

The Rh(1)-N(1) and Rh(1)-N(2) bond distances are 2.089(5) and 2.105(5) Å and are

slightly longer than those for [3.2]. The Rh(1)-C(6), Rh(1)-C(2), Rh(1)-C(5) and

Rh(1)-C(1) are 2.135(6), 2.146(6), 2.150(60) and 2.165(6) Å, respectively. The bite

angle N(1)-Rh(1)-N(2) is 89.04(18)° and is much closer to the ideal 90° for square

planar complex. Figure 3.5 shows the crystal packing in the unit cell of [3.3] and no

Page 103: Solvent-free Synthesis of Bisferrocenylimines

89

intermolecular interactions are present. The crystal packing of [3.2] completely

differs from that of [3.3].

Table 3.4: Selected bond distances, bond angles and torsion angles of [3.3]

Bond distances Bond angles Torsion angles

Rh(1)-N(1) 2.089(5) N(1)-Rh(1)-N(2) 89.04(18) N(2)-Rh(1)-N(1)-C(9) 107.9(5)

Rh(1)-N(1) 2.105(5) N(1)-Rh(1)-C(6) 158.7(2) C(6)-Rh(1)-N(1)-C(9) 12.3(9)

Rh(1)-C(6) 2.135(6) N(2)-Rh(1)-C(6) 92.9(2) C(2)-Rh(1)-N(1)-C(9) -98.8(5)

Rh(1)-C(2) 2.146(6) N(1)-Rh(1)-C(2) 89.7(2) C(1)-Rh(1)-N(1)-C(9) -61.4(5)

Rh(1)-C(5) 2.150(6) N(2)-Rh(1)-C(2) 153.2(2) N(2)-Rh(1)-N(1)-C(10) -76.6(4)

Rh(1)-C(1) 2.165(6) C(6)-Rh(1)-C(2) 97.8(2) C(2)-Rh(1)-N(1)-C(10) .76.7(4)

N(1)-C(9) 1.283(7) N(1)-Rh(1)-C(5) 163.6(2) C(1)-Rh(1)-N(2)-C(14) 7.4(14)

N(1)-C(10) 1.478(7) N(2)-Rh(1)-C(5) 91.8(2)

N(2)-C(14) 1.289(7) C(6)-Rh(1)-C(5) 37.6(2)

N(2)-C(13) 1.473(7) C(2)-Rh(1)-C(5) 82.2(2)

C(9)-C(21) 1.450(8) C(10)-(11)-C(12) 116.1(5)

C(10)-C(11) 1.512(8) C(13)-C(12)-C(11) 114.7(5)

C(11)-C(12) 1.522(8) C(25)-C(21)-C(22) 106.6(6)

C(12)-C(13) 1.522(8) C(25)-C(21)-C(9) 130.7(6)

C(14)-C(41) 1.470(8) C(22)-C(21)-C(9) 122.4(6)

3.1.3 Rhodium(I) complexes containing bisferrocenylamines

Initially, the objective was to synthesize rhodium complexes similar to those reported

by Kim and Alper (Figure 3.6)6 using bisferrocenylamines as ligands. These

complexes are said to be highly effective in hydroformylation reactions.

Rh

N N

Rh

Cl Cl+ -

Figure 3.6: Cationic rhodium(I) diamine complexes with the [Rh(COD)Cl2]- anion.6

Page 104: Solvent-free Synthesis of Bisferrocenylimines

90

The complexes were prepared by stirring equimolar amounts (1 mole) of

[Rh(COD)Cl]2 [3.0] with bisferrocenylamines [2.10]-[2.12], using the Schlenk

technique, at room temperature in an argon environment, for 12 h.6 Yellow

precipitates were formed, almost immediately, on addition of the ligand to a solution

of [Rh(COD)Cl]2 in toluene. However, these complexes were highly insoluble in most

organic solvents and therefore could not be characterized. The insolubility of the

complexes was thought to be due to very high lattice energies in the molecules. It

was then decided that a smaller counterion should be used instead of the bulky

anionic complex. Tetrafluoroborate ion, BF4- was chosen as the anion to be used and

all the complexes were soluble in most organic solvents, making characterization

possible.

The complexes were prepared by a slightly modified literature method.4 Silver

tetrafluoroborate (AgBF4) (2 mol) in acetone was added to [Rh(COD)Cl]2 (1 mol) in

acetone using the Schlenk technique in an argon atmosphere. After filtration of the

precipitated AgCl, the yellow filtrate was treated with the appropriate ligand (2 mol) in

acetone. The mixture was stirred at room temperature for 24 h. Addition of n-pentane

caused the precipitation of a yellow solid after the volume of acetone was reduced to

approximately 5 cm3. The complexes (Figure 3.7) were further purified by

recrystallization and were isolated in moderate to excellent yields (Table 3.5).

Fe

CH2

(CH2)

Fe

CH2

HN

x

HN

RhBF4

X = 2 [3.4] = 3 [3.5] = 4 [3.6]

Figure 3.7: Cationic rhodium(I) complexes

Page 105: Solvent-free Synthesis of Bisferrocenylimines

91

Surprisingly, 1H NMR data was not helpful in terms of the characterization of all

complexes since their spectra showed only broad signals. 13 C NMR gave no signals

for complexes even after running the experiment for 24 h. IR spectra of complexes

[3.4], [3.5] and [3.6] showed absorptions in the region 3180-3280 cm-1 representing

the v(NH) stretching frequencies. The frequencies were similar to those reported by

Garrald et al2 and Beller et al.7 for ν(NH) of similar complexes. For ferrocene, three

bands were observed in all complexes in the regions 3090-3100 cm-1, 1405-1415

cm-1 and 1104-1115 cm-1, which were assigned to v(CH) stretching, v(C-C)

stretching and ring breathing, respectively.1

Figure 3.8: IR spectrum of [3.5]

Complexes [3.4], [3.5] and [3.6] each exhibited a band at 1642, 1634 and 1638,

respectively, which was attributed to the v(C=C) stretching frequency of the COD

ligand. Another sharp band was observed in the region 480-490 cm-1 and it was

assigned to the Rh-N stretching frequency. A signal at around 1000 cm-1 was

observed in all complexes which, according to Beller et al.,8 could be assigned to the

BF4- ion, indicative of a cationic species. The infrared spectrum of [3.5] is shown in

Figure 3.8.

Page 106: Solvent-free Synthesis of Bisferrocenylimines

92

Table 3.5: Table of yields and conductivity measurements

Number Complex Yield (%) Λm

(Ohm-1.cm2.mol-1)

[3.4] Fe

CH2

(CH2)

Fe

CH2

HN

2

HN

RhBF4

66

37.6

[3.5] Fe

CH2

(CH2)

Fe

CH2

HN

3

HN

RhBF4

82

40.4

[3.6] Fe

CH2

(CH2)

Fe

CH2

HN

4

HN

RhBF4

72

33.6

[3.7] CH2

(CH2)

CH2

HN

2

HN

RhBF4

60

32.0

[3.8] CH2

(CH2)

CH2

HN

2

HN

RhBF4

68

42.8

The aromatic v(C=C) stretching frequencies were observed at 1506 cm-1 for [3.7]

and 1509 cm-1 for [3.8]. A Rh-N stretching band was also observed for [3.7] and

[3.8] in exactly the same region as in the previous complexes.

A sharp band that was assigned to the BF4- ion was observed at 1032 and 1036 cm-1

for both complexes. This was an indication of the cationic nature of the complexes.

The conductivity values of the complexes (Table 3.5) were comparable to the

conductivity values obtained by Denise and Pannetier9 for similar types of

complexes.

Page 107: Solvent-free Synthesis of Bisferrocenylimines

93

Figure 3.9: IR spectrum of [3.8]

3.2 Electronic spectroscopy Uv-vis spectra of all complexes were obtained in a dichloromethane solution (10-4

M). In comparison with the free ligands [2.1], [2.2] and [2.3], spectra of complexes

[3.1], [3.2] and [3.3] exhibited an extra band at λmax 350, 382 and 383 nm,

respectively (Figure 3.12, Table 3.6). Moreover, the bands that were observed in the

ligands [2.1]-[2.3] appeared to have shifted to higher wavelengths λmax. This was

clearly indicative of some coordination to the rhodium(I) ion. Both bands at lower

wavelengths, for the complexes, appeared as shoulders. UV-vis spectra of

complexes [3.4], [3.5] and [3.6] exhibited a shoulder at λmax 462, 466 and 468 nm,

respectively, (Figure 3.11, Table 3.6). Bands at λmax 385, 388 and 387 nm in [3.4],

[3.5] and [3.6], respectively, were thought to be due to coordination to the rhodium

metal.

This band appeared in complexes [3.1]-[3.3] as well as in complexes [3.7] and [3.8]

as will been seen later. One band that was observed in the corresponding ligands

[2.10], [2.11] and [2.12] was not observed in the spectra of the complexes. It was

postulated that these bands could have been masked by the bands at λmax 388, 387

and 382 nm.

Page 108: Solvent-free Synthesis of Bisferrocenylimines

94

Figure 3.10: UV-vis spectra of [3.1], [3.2] and [3.3]

Figure 3.11: UV-Vis spectra of [3.4], [3.5] and [3.6]

UV-vis spectra of complexes [3.7] and [3.8] exhibited only one band at λmax 382 and

381 nm respectively (Figure 3.10, Table 3.5). These spectra clearly showed ligand

coordination to the metal since [2.16] and [2.17] were inactive in the UV-vis region

(800-200 nm).

Page 109: Solvent-free Synthesis of Bisferrocenylimines

95

Figure 3.12: UV-vis spectra of [3.7] and [3.8]

Table 3.6: UV-vis data for complexes [3.1]-[3.8]

Complex λmax (nm) ε (L.mol-1.cm-1)

[3.1] 303 (1198) 350 (670) 469 (219)

[3.2] 347 (593) 382 (392) 469 (469)

[3.3] 348 (711) 383 (456) 467 (282)

[3.4] 385 (174) 462 (34)

[3.5] 388 (345) 466 (50)

[3.6] 387 (219) 468 (32)

[3.7] 382 (228)

[3.8] 381 (141)

ε = molar exctinction coefficient.

3.3 Cyclic Voltammetry

Some typical voltammograms of selected rhodium(I) complexes are illustrated in

Figure 3.13, and depict one-electron reversible redox waves.

Page 110: Solvent-free Synthesis of Bisferrocenylimines

96

Figure 3.13: Cyclic voltammograms of [3.2], [3.4], [3.5] and [3.6]

Rhodium(I) complexes containing bisferrocenylimines exhibited positive shifts in

potential meaning that the complexes became more resistant to oxidation than their

corresponding free ligands (Figure 2.11, Table 3.6). Similarly, rhodium(I) complexes

containing bisferrocenylamines exhibited positive shifts in potentials compared to

their corresponding free ligands and thus became more difficult to oxidise (Figure

3.13, Table 3.6). The positive shift in potentials exhibited by the compounds was

evidence that coordination to the rhodium centre had occurred. Rhodium(I)

complexes containing arylbisamines [3.7] and [3.8] were electrochemically inactive

as in the case of their corresponding free ligands.

Table 3.7: Half-wave potentials of [3.1]-[3.6]

Compound Epa (mV) Epc (mV) E1/2 (mV)

[3.1] 413 342 327.5

[3.2] 388 300 344

[3.3] 372 261 316.5

[3.4] 246 105 175.5

[3.5] 222 104 163

[3.6] 157 72 114.5

Page 111: Solvent-free Synthesis of Bisferrocenylimines

97

3.4 Experimental

3.4.1 Purification procedures

All reagents and solvents were purified using standard purification and drying

methods.8 Silver tetrafluoroborate and the chloro-(1,5-cyclooctadiene)rhodium(I)

dimer were obtained from Sigma Aldrich Chemical Company, Milwaukee, USA.

Table 3.7: General drying agents for solvents.

Solvent Drying Agent

Dichloromethane CaH2

Hexane Na wire

Toluene Na wire

3.4.2 Instrumentation

Unless otherwise mentioned, all reactions were carried out using standard Schlenk

techniques in an argon gas environment. All the other instruments employed for

characterizations were the same as stated in Chapter 2.

X-ray crystal intensity data were collected on a Nonius Kappa-CCD diffractometer

using graphite monchromated MoKα radiation at the University of Cape Town.

Temperature was controlled by an Oxford Cryostream cooling system (Oxford

Cryostat). The strategy for the data collections was evaluated using the Bruker

Nonius “Collect” program.9 Data were scaled and reduced using DENZO-SMN

software (Ontwinowski & Minor, 1977). An empirical absorption correction utlilized

the program SADABS (Sheldrick, 1996). The structure was solved by direct methods

and refined by full-matrix least-squares with the program SHELXL-97 (Sheldrick,

1997), refining on F2.10,11 Packing diagrams were produced using the program

PovRay and graphic interface X-seed (Barbour, 2001).12 All the non-H atoms were

refined anisotropically.

Page 112: Solvent-free Synthesis of Bisferrocenylimines

98

3.5 Synthesis of rhodium(I) complexes

3.5.1 Rhodium(I) complexes containing bisferrocenylimines

3.5.1.1 General procedure3

Silver perchlorate [3.0] (0.23 mmol) (Scheme 3.1) in acetone (2 cm3) was added to a

solution of the chloro-(1,5-cyclooctadiene)rhodium dimer (0.11 mmol) in acetone (30

cm3). After removal of the precipitated AgCl, the reaction mixture was then heated

under reflux for 30 minutes. The reaction mixture was treated with bisferrocenylimine

(0.23 mmol) in toluene (20 cm3) and the resultant dark red solution was left to stir at

room temperature (ca. 25 °C) for 3h. Solvents were removed in vacuo and the dark

red solid was recrystallized from a dichloromethane/hexane mixture. The products

were obtained as dark red to orange solids.

3.5.1.2 Complex [3.1]3

The general procedure in 3.5.1.1 was followed using

N,N’–ethylenebis(ferrocenylmethylidene)imine (104

mg, 0.23 mmol) in toluene (20 cm3). Complex [3.1]

was obtained as a dark red solid (27 mg, 32 %). M.p.

215 °C (decomp) (lit. 192 °C, decomp.); 1H NMR (CDCl3) 7.43 (2H, s, N=CH), 4.80

(4H, t, J = 1.7, C5H4), 4.73 (4H, t, J = 1.8, C5H4), 4.37 (10H, s, C5H5), 4.16 (4H, s,

COD-CH), 3.85 (4H, s, CH2), 2.64 (4H, br-s, COD-CH2), 2.09 (4H, d, J = 7.5, COD-

CH2); 13C NMR (CDCl3) 168.48, 84.65, 84.49, 74.71, 73.23, 70.68, 57.01, 30.55; IR

(KBr); 3249, 2951, 2886, 2836, 1612, 1447, 1412, 1377, 1259, 1219, 1098, 1102,

954, 825, 733, 622, 479; m/z (FAB) 665 ([M++2], 3%), 663 ([M+], 29%), 553 (3%),

489 (2%), 453 (9%), 333 (3%), 308 (18%), 289 (16%), 233 (4%), 154 (100).

Anal.Calc. for C32H36N2Fe2Rh: MW, 663.06324. Found: MW, 662.7.

Fe

CH

(CH2)

Fe

CH

N2

N

ClO4

Rh

Page 113: Solvent-free Synthesis of Bisferrocenylimines

99

3.5.1.3 Complex [3.2]

The general procedure in 3.5.1.1 was followed using

N,N’-propylenebis(ferrocenylmethylidene)imine (107

mg, 0.23 mmol) in toluene (20 cm3). Complex [3.2]

was obtained as a reddish orange solid (72 mg, 84

%). M.p. 220 °C (decomp.); 1H NMR 8.31 (2H, s, N=CH), 5.49 (2H, s, COD-CH),

5.39 (2H, s, COD-CH), 4.72 (4H, s, C5H4), 4.68 (4H, s, 2 x CH2), 4.23 (4H, s, C5H4),

4.17 (2H, m, CH2), 4.07 (10H, s, C5H5), 2.59 (4H, br-s, COD-CH2), 1.97 (4H, d, J =

7.7, COD-CH2) 13C NMR (CDCl3) 170.44, 77.62, 75.61, 73.41, 70.51, 70.28, 65.04,

30.89, 30.55; IR (KBr) 3106, 2933, 2852, 1624, 1457, 1411, 1373, 1331, 1253, 1093,

1052, 998, 897, 829, 622; m/z (FAB) 677 ([M+], 11%), 675 (2%), 567 (2%), 467 (5%),

424 (3%), 347 (2%), 307 (18%), 289 (17%), 242 (2%), 154 (100%). Anal. Calc.2 for

C33H38N2Fe2Rh; C, 51.03; H, 4.93; N, 3.61; MW, 677.26372. Found: C, 50.27; H,

4.59; N, 3.48; MW, 676.7.

3.5.1.4 Complex [3.3]

The general procedure in 3.5.1.1 was followed

using N,N’-butylenebis(ferrocenlmethylidene)imine

(111 mg, 0.23 mmol). Complex [3.3] was obtained

as an orange solid (68.3 mg, 70%). M.p. 228 °C

(decomp.); 1H NMR (CDCl3) 8.22 (2H, s, CH=N), 6.11 (2H, s, COD-CH), 5.65 (2H, s,

COD-CH), 4.14 (5H, s, C5H5), 4.12 (5H, s, C5H5), 2.62 (4H, m, COD-CH2), 1.79 (4H,

s, COD-CH2), other signals could not be assigned properly; 13C NMR (CDCl3)

170.41, 84.58, 75.63, 73.74, 73.90, 73.33, 70.04, 69.96, 69.78, 66.46, 58.02, 29.29,

28.73; IR (KBr) 3101, 3013, 2926, 2882, 2838, 1620, 1454, 1436, 1414, 1375, 1331,

1256, 1146, 1103, 1046, 1002, 967, 830, 624, 506, 483; m/z (FAB) 693 ([M+ +2],

9%) 691 ([M+], 65%), 581 (7%), 495 (4%), 481 (6%), 396 (5%), 345 (3%), 307 (21%),

289 (20%), 233 (7%), 154 (100%). Anal. Calc. for C34H40N2Fe2Rh: C, 51.64; H, 5.10;

N, 3.54; MW, 691.09454. Found: C, 51.46, H, 5.43; N, 3.31; MW, 690.7.

2 The tendency for the C and N content to be too low and H too high could be due to the presence of traces of water. This also applies to compounds [3.3]-[3.8].

Fe

CH

(CH2)

Fe

CH

N3

N

ClO4

Rh

Fe

CH

(CH2)

Fe

CH

N4

N

RhClO4

Page 114: Solvent-free Synthesis of Bisferrocenylimines

100

3.5.2 Rhodium(I) complexes containing bisferrocenylamines3

3.5.2.1 Complex [3.4]

Silver tetrafluoroborate [3.0] (33.3 mg, 0.171

mmol) in acetone 2 cm3 was added to a solution

of [Rh(COD)Cl]2 (42.1 mg, 0.086 mmol) in

acetone (25 cm3) and the reaction mixture stirred

vigorously at room temperature for a few minutes. The precipitated AgCl was filtered

off and the yellow filtrate was stirred with N,N-ethylenebis(ferrocenylmethyl)amine

(78.1 mg, 0.171 mmol) in acetone (20 cm3). The reaction mixture was left to stir at

room temperature for 24 h. Reduction of the solvent volume and addition of hexane

caused the precipitation of complex [3.4] as a yellow solid. After filtration, the solid

was washed with diethyl ether and recrystallized from a dichloromethane/hexane

mixture (85.4 mg, 66.2%). M.p. 182 °C (decomp.); IR (KBr) 3268, 3101, 2934, 2881,

2829, 1638, 1454, 1410, 1383, 1335, 1304, 1234, 1107, 1085, 1028, 1002, 918, 822,

484; Anal. Calc. for C32H40N2Fe2Rh: C, 50.97, H; 5.35; N, 3.71. Found: C, 49.60, H,

6.01; N, 3.24.

3.5.2.2 Complex [3.5]

The procedure as for [3.4] was followed using

silver tetrafluoroborate (23.2 mg, 0.122 mmol),

[Rh(COD)Cl]2 (30.1 mg, 0.061 mmol) and N,N-

propylenebis(ferrocenylmethyl)amine (55.8 mg,

0.122 mmol). Complex [3.5] was obtained as a yellow solid (94.3 mg, 81.8%). M.p.

134 °C (decomp.); IR (KBr) 3271, 3092, 2937, 2888, 2831, 1633, 1454, 1381, 1332,

1283, 1238, 1124, 1108, 1084, 1039, 1002, 974, 913, 819, 481; Anal. Calc. for

C33H42N2Fe2Rh: 51.60, H, 5.51, N, 3.65. Found: C, 50.40; H, 5.66; N, 3.30.

3 No MS data was obtained for compounds [3.4]-[3.8] owing to a spectrometer breakdown.

Fe

CH2

(CH2)

Fe

CH2

HN

2

HN

RhBF4

Fe

CH2

(CH2)

Fe

CH2

HN

3

HN

RhBF4

Page 115: Solvent-free Synthesis of Bisferrocenylimines

101

3.5.2.3 Complex [3.6]

The procedure as for [3.4] was followed using

silver tetrafluoroborate (36.1 mg, 0.185 mmol),

[Rh(COD)Cl]2 [3.0] (45.6 mg, 0.093 mmol) and

N,N-butylenebis(ferrocenylmethyl)amine (89.6

mg, 0.185 mmol). Complex [3.6] was obtained as a yellow solid (84.8 mg, 72.3 %).

M.p. 204 °C (decomp.); IR (KBr) 3285, 3259, 3180, 3092, 2996, 2926, 2882, 2829,

1642, 1476, 1432, 1410, 1379, 1331, 1230, 1125, 1107, 1085, 1037, 1028, 997, 953,

923, 817, 488; Anal. Calc. for C34H44N2Fe2Rh: C, 52.21; H, 5.67; N, 3.58. Found: C,

51.38; H, 6.33; N, 2.57.

3.5.2.4 Complex [3.7]

The procedure as for [3.4] was followed using

silver tetrafluoroborate (29.2 mg, 0.150 mmol),

[Rh(COD)Cl]2 [3.0] (37.0 mg, 0.075 mmol) and

N,N-bis(2,5-dimethylbenzyl)ethane-1,2-

diamine (44.5 mg, 0.150 mmol). Complex [3.7]

was obtained as a light yellow solid (53.8 mg, 60.4 %). M.p. 110 °C (decomp.); IR

(KBr) 3250, 2964, 2926, 2882, 2838, 1629, 1506, 1458, 1388, 1339, 1304. 1164,

1125, 1085, 1059, 1037, 958, 817, 765, 528, 483; m/z (FAB); Anal. Calc. for

C28H40N2Rh: C, 56.58; H, 6.78; N, 4.71. Found: C, 55.52; H, 7.50; N, 3.87.

3.5.2.5 Complex [3.8]

The procedure as for [3.4] was followed using

silver tetrafluoroborate (27.2 mg, 0.139 mmol),

[Rh(COD)Cl]2 (34.3 mg, 0.070 mmol) and N,N-

bis(2,4,6-trimethylbenzyl)ethane-1,2-diamine

(45.1 mg, 0.139 mmol). Complex [3.8] was

obtained as a light yellow solid (63.6 mg, 68.1%). M.p. 170 °C (decomp.); IR (KBr)

3259, 2961, 2917, 2873, 2829, 1638, 1616, 1581, 1458, 1432, 1379, 1340, 1304,

Fe

CH2

(CH2)

Fe

CH2

HN

4

HN

RhBF4

CH2

(CH2)

CH2

HN

2

HN

Rh BF4

CH2

(CH2)

CH2

HN

2

HN

RhBF4

Page 116: Solvent-free Synthesis of Bisferrocenylimines

102

1120, 1085, 1063, 1032, 896, 852, 769, 716, 633, 611, 519, 484; m/z (FAB); Anal.

Calc. for C30H44N2Rh: C, 57.89; H, 7.13; N, 4.50. Found: C, 53.71; H, 6.89; N, 3.56.

3.6 References

1. W. P. Griffith, The Chemistry of the Rarer Platinum Metals, John Wiley &

Sons, London 1967.

2. M. A. Garralda and L. Ibarlucea, J. Organomet. Chem., 311 (1986) 225.

3. S. –I. Lee, S. –C. Shim and T. –J. Kim, J. Polym. Sci., Part A: Polymer Chem.,

34 (1996) 2377.

4. P. Pertici, F. D’ Arata and C. Rosini, J. Organomet. Chem., 515 (1996) 163.

5. G. R. Julius and S. Cronje, Helvetica Chim. Acta, 85 (2002) 3737.

6. J. J. Kim and H. Alper, Chem. Commun., (2005) 3059.

7. M. Beller, H. Trauthwein, M. Eichberger, C. Breindl, T. E. Müller and A. Zapf,

J. Organomet. Chem., 566 (1998) 277.

8. B. S. Furniss, A. J. Hannaford, P. W. G. Smith and A. R. Tatchell, Vogel’s

Textbook of Practical Organic Chemistry, Longmann Scientific and Technical,

England (5th Ed.) 1989.

9. Z. Otwinowski and W. Minor in C. W. Carter, J. Sweet and R. M. Sweet (Eds),

Macromolecular Crystallography Part A, Academic Press, New York, 276

(1997) 307.

10. G. M. Sheldrick, SHELX97, Programme for Solving Crystal Structures,

University of Göttingen, Germany, 1997.

11. G. M. Sheldrick, SHELX97, Programme for the Refinement of Crystal

Structures, University of Göttingen, Germany, 1997.

12. L. J. Barbour, X-Seed, University of Missouri-Columbia, USA, 1999.

Page 117: Solvent-free Synthesis of Bisferrocenylimines

103

CHAPTER 4

RESULTS AND DISCUSSION

4.1 Polymerization of phenylacetylene

4.1.1 Introduction

Polyphenylacetylene (PPA) has been found to be a very interesting polymer because

of its photoconductivity,1 photoluminescence,2 non-linear optical3 and membrane

properties.4 Cametti et al.5 have also investigated iodine-doped polyphenylacetylene

for potential application in technology. Polymerization of phenylacetylene has been

carried out in various conditions, including cationic, radical and coordination

mechanisms.6 Transition metal complexes have been used since the early

pioneering work during 1930 to the 1950’s. Masuda et al.7 first polymerized

phenylacetylene in 1974 using WCl6 and MoCl5 as catalysts, to give high molecular

weight polymers. Since then, other transition metal complexes (palladium, rhodium,

iridium, etc) have been investigated for catalytic activity towards polymerization of

polyphenylacetylene.

The zwitterion complex Rh+(COD)BPh4- has been found to produce stereoregular

cis-PPA with molecular weights up to 35 000, under hydrosilation conditions.8 The

[Rh(norbonadiene)Cl]2 complex has, so far, been reported to yield the highest

molecular weight of approximately 4.3 x 106.9 The catalytic activity of Rh(I), Ir(I) and

Ru(IV) complexes containing 1,3-dimesityl-3,4,5,6-tetrahydropyrimidin-2-ylidene and

1,3-di(2-propyl)-3,4,5,6-tetrahydropyrimidin-2-ylidene ligands has been reported by

Zhang et al.10 Molecular weights ranging between 55 000 and 200 000 have been

obtained in ionic liquids using Rh(I) complexes as catalysts.11

Page 118: Solvent-free Synthesis of Bisferrocenylimines

104

4.1.2 Polymer characterization

Four possible stereoisomers can be formed in the catalytic polymerization of

phenylacetylene (Figure 4.1).10 The stereochemistry of PPA can be generated from

the configuration of the C=C bond and the conformation of C-C single bond of the

polymer main chain. The stereoisomers can be easily distinguished by their

physicochemical and spectroscopic properties.6

Ph H

Ph

HPh

H

cis-cisoidal

PhH H Ph

Ph H Ph H

trans-cisoidal

Ph PhH H

Ph H Ph H

cis-transoidal

Ph Ph Ph Ph

trans-transoidal

Figure 4.1: Stereoisomers of polyphenylacetylene

The solubility of the isomers can also be utilized as an aid to distinguish between

them. For example, cis-transoidal and trans-cisoidal isomers are both highly soluble

in benzene, while the cis-cisoidal isomer is insoluble. NMR and IR spectroscopy

show that the cis-cisoidal and cis-transoidal isomers have similar spectroscopic

behaviour, which differentiates them from the trans-cisoidal isomer.

The most significant difference in the IR spectra of the three isomers is the presence

of a band at 740 cm-1 in the cis-cisoidal and cis-transoidal isomers. This band

represents the out-of-plane stretching of the C-H bonds and can also be associated

with the cis content in the polymer. The trans-cisoidal isomer exhibits a band at 1265

cm-1 due to the out-of-plane deformation vibrations of the trans-C-H bonds. The band

at 740 cm-1, present in the cis-cisoidal and cis-transoidal isomers, is lacking in the

trans-cisoidal isomer. On the other hand, the band at 1265 cm-1, present in the trans-

Page 119: Solvent-free Synthesis of Bisferrocenylimines

105

cisoidal isomer, is lacking in the cis-cisoidal and cis-transoidal isomers. Hence, the

band at 740 cm-1 can be used as a function of the cis content in the polymer and the

band at 1265 cm-1 as a function of the trans content.

The ratio between the bands at 1500 and 1450 cm-1 provides information about the

stereochemistry of the polymer. A polymer with a cis content can be identified by a

ratio of 1 or smaller, together with a strong band at 740 cm-1. A ratio greater than 1

and a band at 1265 cm-1 can be related to a polymer with a trans content. Some

polymers with cis conformation can isomerise to trans conformation, resulting to

ratios greater than 1 and weak bands at 740 cm-1.

1H NMR spectra of the polymer have been used to differentiate between the

respective isomers by looking at differences in the chemical shifts of the aromatic

protons. Apart from these differences, a signal at approximately 5.82 ppm represents

the olefinic proton in the cis-cisoidal and cis-transoidal isomers. The trans-cisoidal

isomer exhibits a smaller or no signal at 5.82 ppm and also displays a broad weak

signal in the region 3-4 ppm due to the aliphatic protons. The area of the signal at

5.82 ppm can be correlared with the intensity of the band at 740 cm-1 to determine

the cis content in the polymer.

4.2 Catalytic polymerization studies

The catalytic polymerization of phenylacetylene was studied with the complexes

[3.1], [3.2] and [3.3] (Table 4.1).

Ph HCatalyst, MeOH

RT, 24 h Ph H

n

Scheme 4.1: Polymerization of phenylacetylene with Rh(I) catalysts

Page 120: Solvent-free Synthesis of Bisferrocenylimines

106

Table 4.1: Polymerization of phenylacetylene with Rh(I) complexes.a

Catalyst Mn Mw Mw/Mn cis-Content (%)b

[3.1] 8129 21203 2.6 98.7

[3.2] 8223 21105 2.6 100

[3.3] 8749 22330 2.5 99.4 a Reaction conditions: 0.3 mol % catalyst in MeOH (15 cm3); at RT for 24 h b Calculated according to reference 12 and 13

The aim was to determine the effect, if any, of increasing the length of the alkylene

chain of the bidentate nitrogen-donor ligands on the catalytic activity of the

complexes.

The number average molecular weight Mn, weighted average molecular weight Mw,

polydispersity index (Mw/Mn) and the cis-content of the polymer samples were

determined (Table 4.1). The molecular weights Mw of the polyphenylacetylene

obtained for all the catalyst were similar, but they were approximately 4 times smaller

than the literature values for similar types of complexes (Table 4.1). For example,

[3.1] has been investigated for its catalytic activity in the polymerization of

phenylacetylene and was found to have the highest catalytic activity with a molecular

weight of approximately 86 000.6 However, according to the results in Table 4.1,

[3.3] produced the highest molecular weight polymer. Moreover, since there was not

much of a difference between the molecular weights and the fact that the molecular

weight decreased from [3.1] to [3.2], it was very difficult to conclude on the influence

of increasing the length of the alkylene chain. The lower polydispersity values (~2.5)

that were obtained imply a more uniform distribution of the polymers.6 All the

catalysts that were investigated resulted in polymers with high a cis content (Table

4.1). Complex [3.2] produced a polymer with the highest cis content.

4.2.1 Spectroscopic properties of polymers

As already mentioned, the stereochemistry of the isomers can be deduced from the

physicochemical and spectroscopic properties. All the catalysts that were

investigated produced polymers with similar 1H NMR and IR spectra. All spectra

Page 121: Solvent-free Synthesis of Bisferrocenylimines

107

exhibited a signal at δ 5.86 ppm, due to olefinic protons, which was indicative of cis-

isomer polymers (Figure 4.2). Two further signals at δ 6.96 and 6.65 ppm were due

to the aromatic protons. According to the literature, the pattern of the signals

indicates a stereoregular polyphenylacetylene with a predominantly cis-transoidal

structure.14 The insolubility of the polymer in methanol made its separation very

easy.

Figure 4.2: 1H NMR spectrum of PPA, catalyzed by [3.2]

Infrared spectra of the polymers that were prepared exhibited a strong absorption

band at λmax 737 cm-1 and lacked the absorption band at λmax 1265 cm-1 (Figure 4.3).

This information confirmed a cis-isomer polymer and that the polymers produced

were linear.14 The ratio of the infrared absorption band at λmax 1500 and 1450 cm-1

can provide information about the stereochemistry of the polymer. A ratio of 1 or

smaller, together with a band at λmax 740 cm-1 indicates the cis content, while a ratio

greater than 1 and a band at λmax 1265 cm-1 indicates the existence of trans content

of the polymer.

Page 122: Solvent-free Synthesis of Bisferrocenylimines

108

Figure 4.3: IR spectrum of PPA prepared using [3.2]

As shown in Figure 4.3, all polymers produced absorption bands at λmax 1488 and

1443 cm-1 (Figure 4.3). Ratios of approximately 1.00 were calculated for these bands

and all were consistent with the cis content of the polymer (Table 4.2).

Table 4.2: Determination of cis-content of polymers

Catalyst Band Ratio cis-Content (%) Polymer

(1488 vs 1443) Colour

[3.1] 1.05 98.7 Yellow

[3.2] 1.03 100 Yellow

[3.3] 1.02 99.4 Yellow

4.2.2 Thermal analysis

The TGA curve showed that the polymer was stable up to 260°C in a nitrogen

atmosphere (Figure 4.4). Decomposition of the polymer continued slowly as the

temperature increased until a residue of 7% remained, at 475°C.

wavenumber (cm-1

)

Page 123: Solvent-free Synthesis of Bisferrocenylimines

109

Figure 4.4: TGA and DSC curves of PPA obtained with [3.1]

The DSC thermogram displayed peaks associated with the cis content of the

polymer and no glass transition was observed. Two exothermic peaks at 172 and

250°C were observed. The two peaks corresponded to the cis-trans isomerization

and the crystallization phenomena, respectively.15 An endothermic peak occurred at

300°C which corresponded to thermal decomposition of the polymer.

4.2.3 Mechanistic pathways for polymerization of phenylacetylene

The transition metal catalyzed polymerization of phenylacetylene is widely known to

occur via two main mechanisms: the four-centre acetylene insertion mechanism

(insertion mechanism)16 and metallacycle (or metathesis) mechanism.17 The first

step of the insertion mechanism is the displacement of the cyclooctadiene in the

catalyst precursor by the solvent, leading to the formation of the species

[Rh(NN)(solvent)2]+ in solution. The formation of a hydridoacetylenic species follows

via the oxidative addition of phenylacetylene to the rhodium(I) metal centre. The

coordination is followed by a migratory insertion, resulting in the formation of a vinylic

rhodium intermediate species, consequently resulting in the formation of the polymer

(Scheme 4.1).

Page 124: Solvent-free Synthesis of Bisferrocenylimines

110

C C

Ph H

M P

HC CH** C C

Ph H

M P

C

C

H

H*

*

CC

H

M

P

H

CC

H

H**

CC

C

H

M

P

H

CC

H

H

**

CH

H

P = polymer chain

M = Rh

Scheme 4.2: Insertion mechanism for polymerization of phenylacetylene16,18

The metallacycle occurs via the formation of a metal-carbene complex with a vinylic

metallacycle intermediate. A key step is the rearrangement of a monomer unit

formed by the stepwise addition of a second metal-carbene complex. The polymer is

formed by the repetition of the above steps (Scheme 4.2).

C

M

HP

P = polymer chain

M = Rh

HC CH**

C C

M CH

H

*

*P

H

C

M

H

P

C

C

H**

H

HC CH

C C

C C

C

P

H HH

H H

M*

*

Scheme 4.3: Metallacyclic (metathesis) mechanism for the polymerization of

phenylacetylene19

Page 125: Solvent-free Synthesis of Bisferrocenylimines

111

An apparent distinction between the two mechanisms is that the insertion

mechanism predicts that two carbons of a monomer unit become doubly bonded to

each other, while the metallacycle mechanism predicts that two carbons of a

monomer unit end up singly bonded to one another in the resulting polymer.20

However, it is generally believed that the polymerization of acetylene catalyzed by

rhodium complexes proceeds via the insertion mechanism, resulting in stereoregular

polymers.

4.3 Hydroformylation of styrene

4.3.1 Introduction

The hydroformylation or “oxo” reaction was discovered by Otto Roelen in 1938,

through modification of the Fischer-Tropsch synthesis to produce aldehydes and

ketones rather than hydrocarbons as the main products.21-23 He observed that

ethylene, H2 and CO were converted into propanal and diethyl ketone (high

pressures) in the presence of Co2(CO)8 as a catalyst. However, aldehydes are the

primary products of the hydroformylation of olefins or alkenes (Scheme 4.3).

RH2, CO

CatalystR

CHO+ R

CHO

linear branched

*

Scheme 4.4: Hydroformylation of olefins

Hydroformylation reactions are very important industrial processes in that most

aldehydes (linear) produced can be reduced to alcohols or oxidized to carboxylic

acids.23 Alcohols are used for the synthesis of phthalate plasticizers by esterification

reaction with phthalic anhydride. In turn, the phthalate plasticizers are used primarily

for polyvinyl chloride plastics. The aldehydes are also used for the production of

detergents, surfactants, solvents, lubricants, cosmetics and other widespread

chemicals.23,24 Branched aldehydes are also very useful for stereoselective and

asymmetric synthesis.24 Thus, hydroformylation reactions have attracted more

attention from both industrial and academic research groups.

Page 126: Solvent-free Synthesis of Bisferrocenylimines

112

Until the early 1970s, cobalt carbonyl complexes were the most employed

homogeneous catalysts in the hydroformylation of alkenes.21,23 The dimeric

dicobaltoctacarbonyl complex Co2(CO)8 was believed to be rapidly converted into a

cobalt hydrocarbonyl species HCo(CO)4, by H2/CO under pressure and was the

active species in the hydroformylation of alkenes. However, the regioselectivity of the

HCo(CO)4 towards linear aldehydes depends on the reaction conditions and the

alkene substrate used. Modification of the HCo(CO)4 catalyst by replacing one

carbonyl ligand with a trialkylphosphine PR3 to produce HCo(CO)3(PR3), led to an

improvement in the rate of reaction and regioselectivity.25 The replacement of the CO

ligand with trialkylphosphine causes stronger Co-CO bonding and consequently

decreases the CO partial pressure, thereby stabilizing the catalyst. It also prevents

the formation of Co metal.

The success of rhodium catalysts led to a decrease in the use of cobalt carbonyl

catalysts. The reason was mainly because rhodium catalysts were more catalytically

active than cobalt catalysts. The use of rhodium carbonyl complexes favoured

formation of a higher proportion of linear aldehydes at comparable temperatures.

Hydroformylation of alkenes26 and alkynes27 using

tris(triphenylphosphine)chlororhodium as a catalyst have been achieved by Osborn

et al.26,27 Union Carbide developed a ligand BIPHEPHOS, which in conjunction with

Rh(CO)2(acac), resulted in regioselective formation of linear aldehydes from various

functionalized terminal alkenes under mild conditions (Scheme 4.4). The aldehyde

was then converted to the indolizidine alkaloid. Many rhodium complexes have been

employed as catalysts for hydroformylation reactions and examples can be obtained

in the review article of Clarke.24

Page 127: Solvent-free Synthesis of Bisferrocenylimines

113

N

BOC

+ CO + H2

Rh(CO)2(acac)

BIPHEPHOS

60 °C, 5 atm, 83%

N

BOC

CHO

BIPHEPHOS =

MeO OMe

tButBu

O

PO O

P

O

OO

Scheme 4.5: Synthesis of precursor to the indolizidine alkaloid

4.3.2 Catalytic hydroformylation studies The catalytic activity of the rhodium(I) complexes was investigated for the

hydroformylation of styrene. Scheme 4.5 illustrates the possible aldehydes expected

from the hydroformylation of styrene, that is, 2-phenylpropanal (branched) and 3-

phenylpropanal (linear).

+ CO + H2

catalyst

800 psi, 20 h

CHO

CHO

+

2-phenylpropanal 3-phenylpropanal

Scheme 4.6: Hydroformylation of styrene catalyzed by [3.4]-[3.8]

The two products and unreacted styrene were identified and distinguished by 1H

NMR spectroscopy. 2-Phenylpropanal was identified by a high intensity doublet at δ

9.72 ppm due to the CH=O group, a a doublet of quartets at δ 3.67 ppm due to the

α-proton and a high intensity doublet at δ 1.48 ppm due to the methyl group (Figure

4.5). 3-Phenylpropanal exhibited very low intensity signals, a triplet at δ 9.84 ppm

due to the CH=O group and two triplets at δ 2.99 and 2.81 ppm due to the β- and α-

CH2 groups, respectively. Unreacted styrene was identified by a doublet of doublets

at δ 6.76 ppm due to the vinylic CH group and two doublets at δ 5.80 and 5.31 ppm

due to the cis- and trans-protons of the vinylic CH2 group.

Page 128: Solvent-free Synthesis of Bisferrocenylimines

114

Figure 4.5: 1H NMR spectrum of the products of hydroformylation of styrene

catalyzed by [3.6]

All complexes [3.4]-[3.8] exhibited excellent catalytic activity and selectivity towards

the hydroformylation of styrene (Table 4.3). Conversions of styrene were

comparable; however, [3.8] resulted in the highest conversion of styrene. All

complexes selectively favoured the formation of the branched aldehyde, 2-

phenylpropanal. It has been shown that the bite angle of chelating ligands have a

direct effect on the regioselectivity of hydroformylation reactions.28 Chelating ligands

with bite angles greater than 90° favour linear over branched aldehydes. Complexes

[3.4]-[3.8] are expected to be in a square planar configuration, and thus would have

bite angles of approximately 90°, which explains why all compounds favoured

branched over linear aldehydes. Complexes [3.7] and [3.8] provided excellent yields

with [3.8] recording the highest yield of all the complexes. Complexes [3.4]-[3.6]

provided excellent conversion of styrene and selectivity towards branched

aldehydes, however, the yields were moderate. The reason was thought to probably

be that the aldehyde formed was further hydrogenated to an alcohol or that the

styrene was hydrogenated to ethylbenzene. In addition, the GC chromatograms of

[3.5] and [3.6] showed an extra peak before 13 minutes and after 17 minutes,

Page 129: Solvent-free Synthesis of Bisferrocenylimines

115

respectively. These peaks were not observed in any of the other chromatograms.

However, the side products were not isolated from the reaction mixture.

Table 4.3: Hydroformylation of styrene catalyzed by rhodium(I) complexes

Entry Catalyst Conversionb

(%)

Selectivity

B/L Ratioc

Yield (%)d

2-PP 3-PP

1 [3.4] 90 27 74 2.4

2 [3.5] 92 25 70 2.7

3 [3.6] 89 26 69 1.9

4 [3.7] 95 27 91 1.8

5 [3.8] 99 25 95 1.8 aReaction conditions: cat. (10 mg), toluene (150 cm3), CO/H2 (400/400 psi), alkene/cat. ratio (1000),

RT, 20h. bDetermined by GC. cDetermined by 1H NMR and GC. dDetermined by GC.

Kim and Alper have reported that sterically hindered rhodium catalysts resulted in

low conversions of styrene.29 This effect was due to the bulkiness of the chelating

nitrogen donor ligands in the rhodium catalysts. The presence of the bulky ferrocenyl

substituents on the chelating ligands in [3.4]-[3.8] explains the reason for the lower

conversions, compared to the smaller benzene substituents on the ligands in [3.7]

and [3.8]. In conclusion, [3.7] and [3.8] gave the best results since both recorded

high conversions of styrene and excellent aldehyde yields.

4.3.3 Mechanism for hydroformylation of styrene

The mechanism for the hydroformylation of styrene is proposed here although it was

not investigated in this project. It is well documented that the COD ligand is readily

replaced by the carbonyl ligand even at room temperature.29,30 Therefore, it has

been suggested that in the presence of CO and H2, the COD in the rhodium catalyst

is replaced by CO, resulting in the pentacoordinated rhodium species

RhH(NN)(CO)2. The RhH(NN)(CO)2 is believed to be active in the hydroformylation

reaction. Scheme 4.6 illustrates the possible reaction pathways for the

hydroformylation of styrene.

Page 130: Solvent-free Synthesis of Bisferrocenylimines

116

N

N

Rh CO

H2

N

N

Rh

CO

CO

H

Ph

N

N

Rh

CO

CO

H PhN

N

Rh

CO

CO

Ph

N

N

Rh

CO

CO

Ph

less favoured

highly favoured

N

N

Rh

CO

Ph

C

O

H2

N

N

Rh

CO

C

O

HH

Ph

+

Ph

CHO

N

N

Rh

CO

H

CO

Scheme 4.7: Possible mechanism for hydroformylation of styrene catalyzed by [3.4]-

[3.8]

The mechanism is also thought to occur in an analogous manner as the one

proposed by Wilkinson.30 The initial step involves the addition of styrene to the

RhH(NN)(CO)2 species, followed by the insertion of styrene resulting in the rhodium

alkyl complex that undergoes migratory insertion of CO to form the rhodium acyl

complex. This step is followed by the oxidative addition of H2 to give the dihydrido

acyl rhodium complex. This step is rate determining and is the only one that results

in the change in oxidation state of the rhodium. The final step involves the reductive

elimination of the product and the reformation of the active rhodium hydride species.

Page 131: Solvent-free Synthesis of Bisferrocenylimines

117

4.4 Experimental

4.4.1 Purification procedures

All solvents used were purified and dried as in the previous Chapters.

Phenylacetylene and styrene were obtained from the Sigma Aldrich Chemical

Company, Milwaukee, USA. The synthetic gases H2 and CO were obtained locally

from Afrox.

4.4.2 Instumentation

Gel Permeation Chromatography was conducted at the Stellenbosch University with

a Waters 717 autosampler, a Waters 619 flow unit, a Waters 410 difractometer, a

Waters 600E system controller, a Waters 515 HPLC pump, a Valveco 8-port switch

for high and low temperature HPLC applications and a Wyatt technology laser

photometer. The programme was controlled using a MilleniumTM software.

Gas Chromatography was conducted on a Focus GC Thermo Finnigan instrument

(Model no: AI 3000) equipped with a flame ionization detector (FID) and a DB 1701

column (film thickness 0.25 µm, internal diameter 0.25 mm, length 30 m). The Delta

Chromatography software was used for recording and integration of chromatograms.

4.4.3 Polymerization of phenylacetylene6

Polymerization of phenylacetylene was carried out in a 2-necked round-bottomed

flask (25 cm3) under an argon atmosphere using Schlenk techniques. The catalyst

(0.02 mmol) was added to a solution of phenylacetylene (0.55 g, 5.44 mmol) in

methanol (15 cm3). The reaction mixture was stirred at RT for 24 h. The yellow

precipitates which formed were filtered off, washed with methanol and then dried

under the vacuum. The polymers were analysed by NMR and IR spectroscopy,

thermal analysis and gel permeation chromatography.

Page 132: Solvent-free Synthesis of Bisferrocenylimines

118

4.4.4 Hydroformylation of styrene29

All hydroformylation reactions were conducted in a 2 L Parr stirred reactor. The

reactor was charged with an appropriate amount of styrene (1000 equiv.), catalyst

(10 mg) and toluene (150 ml). The reactor was flushed three times with CO and

pressurized to 400 psi. After attaching and purging of the H2 line, the reactor was

pressurized to 800 psi. After stirring the reaction for 16 h, the excess CO and H2

were released. The reaction mixture was analysed with 1H NMR and GC. Similar

conditions were used for hydroformylation reactions 4.4.4.1-4.4.4.5.

4.4.4.1 Hydroformylation of styrene (1381 mg, 13.26 mmol) with [3.4] (10 mg,

0.01326 mmol). Conversion: 90%.

4.4.4.2 Hydroformylation of styrene (1356 mg, 13.02 mmol) with [3.5] (10 mg,

0.01302 mmol). Conversion: 92%.

4.4.4.3 Hydroformylation of styrene (1332 mg, 12.79 mmol), with [3.6] (10 mg,

0.01279 mmol). Conversion: 89%.

4.4.4.4 Hydroformylation of styrene (1753 mg, 16.83 mmol) with [3.7] (10 mg,

0.01683 mmol). Conversion: 92%.

4.4.4.5 Hydroformylation of styrene (1674 mg, 16.07 mmol) with [3.8] (10 mg,

0.01607 mmol). Conversion: 95%.

4.5 References

1. E. T. Kang, K. G. Neoh, T. M. Masuda, T. Higashimura and M. Yamamoto,

Polymer, 30 (1989) 1328.

2. C. W. Lee, K. S. Wong, W. Y. Lam and B. Z. Tang, Chem. Phys. Lett., 307

(1999) 67.

3. D. Neher, A. Wolf, C. Bubeck and G. Wegner, Chem. Phys. Lett., 163 (1989)

6705.

4. T. Kaneko, T. Yamamoto, H. Tatsumi, T. Aoki and E. Oikawa, Polymer, 41

(2000) 4437.

5. C. Cametti, A. Furlani, M. V. Russo and G. Lucci, Synth. Met., 83 (1996) 77.

Page 133: Solvent-free Synthesis of Bisferrocenylimines

119

6. S. –I. Lee, S. –C. Shim and T. –J. Kim, J. Polym. Sci. A: Polymer Chemistry,

34 (1996) 2377.

7. T. Masuda, K. Hasegawa and T. Higashimura, Macromolecules, 7 (1974) 728.

8. Y. Goldberg and H. Alper, Chem. Commun., (1994) 1209.

9. M. Tabata, W. Yang and K. Yokoda, Polym. J., 20 (1990) 1105.

10. Y. Zhang, D. Wang, K. Wurst and M. K. Buchmeiser, J. Organomet. Chem.,

690 (2005) 5728.

11. P. Mastrorilli, C. F. Nobile, V. Gallo, G. Suranna and G. Farinola, J. Mol.

Catal. A: Chemical, 184 (2002) 73.

12. C. I. Simionescu, V. Percec and S. Dimitrescu, J. Polym. Sci. Polym. Chem.

Ed., 15 (1977) 2497.

13. A. Furlani, S. Licoccia, M. V. Russo, A. Camus and N. Marsich, J. Polym. Sci.

A: Polym. Chem., 24 (1986) 991.

14. I. M. Barkalov, A. A. Berlin, V. I, Goldanskii and G. Mingao, Vysokomol.

Soedin., 5 (1963) 368.

15. F. D. Kleist and N. R. Byrd, J. Polym. Sci. A-1, 7 (1969) 3419.

16. S. Shirakawa and S. Ikeda, J. Polym. Sci. Chem. Ed., 12 (1974) 929.

17. T. Masuda and T. Higashimura, Adv. Polym. Sci., 81 (1986) 121.

18. G. Wegner, Angew. Chem. Intl. Ed. Engl., 20 (1981) 361.

19. C. –C. Han and T. J. Katz, Organometallics, 4 (1985) 2186.

20. C. S. Yannoni and R. D. Kendrick, J. Chem. Phys., 74 (1981) 747.

21. M. M. T. Khan and A. E. Martell, Homogeneous Catalysis by Metal

Complexes, Academic Press, Inc., New York, 1974.

22. J. Tsuji, Transition Metal Reagents and Catalysis: Innovations in Organic

Synthesis, John Wiley & Sons, LTD, New York, 2000.

23. http:/ chemistry.Isu.edu/Stanley/webpub/4571-Notes/Chap16-

hydroformylation.doc.

24. M. L. Clarke, Current Organic Chemistry, 9 (2005) 701.

25. L. H. Slaugh and D. Mullineaux, J. Organomet. Chem., 13 (1968) 469.

26. J. A. Osborn, F. H. Jardine, J. F. Young and G. Wilkinson, J. Chem. Soc. A.

(1965) 1711.

27. F. H. Jardine, J. A Osborn, G. Wilkinson and J. F. Young, Chem. Ind.

(London), (1965) 560.

Page 134: Solvent-free Synthesis of Bisferrocenylimines

120

28. C. P. Casey, G. T. Whiteker, M. G. Melville, L. M. Petrovich, J. A. Gavney, Jr.

and D. R. Powell, J. Am. Chem. Soc., 114 (1992) 5535.

29. J. J. Kim and H. Alper, Chem. Commun., (2005) 3059.

30. C. K. Brown and G. Wilkinson, J. Chem. Soc. A, (1970) 2753.

31. W. Kläui, D Schramm and G. Schramm, Inorg. Chim. Acta, 357 (2004) 1642.

Page 135: Solvent-free Synthesis of Bisferrocenylimines

121

CHAPTER 5

CONCLUSION

5.1 Conclusion

Bisferrocenylimines and arylbisimines were successfully prepared under solvent-free

conditions and were fully characterized. The solvent-free reactions carried more

advantages than the solvent reactions, and these included shorter reaction times,

better yields and required no heat. The reduction of the above compounds was

achieved by lithium aluminium hydride (LAH).

Cationic rhodium(I) complexes containing bisferrocenylimines were successfully

prepared and fully characterized. Suitable single crystals grown allowed the

determination of the structures of [3.2] and [3.3] using X-ray crystallography. The

successful synthesis of cationic rhodium(I) complexes containing

bisferrocenylamines and arylbisamines was also achieved. Characterization was

only possible with IR and UV-vis spectroscopy, cyclic voltammetry and

conductometry. For example, UV-vis spectra of [3.1], [3.2] and [3.3] showed an

extra band at λmax 350, 382 and 383 nm, respectively compared to the corresponding

free ligands [2.1], 2.2], and [2.3]. The bands that also appeared in the free ligands

were shifted to higher wavelengths. Similarly, [3.4], [3.5], [3.6], [3.7] and [3.8]

exhibited bands at λmax 385, 388, 387, 382 and 381 nm. All these bands were

attributed to coordination of the ligands to the rhodium(I) ion.

Complexes [3.1], [3.2] and [3.3] were catalytically active in the polymerization of

phenylacetylene. The molecular weights of the polymers were very low, with Mw =

22330 being the highest molecular weight recorded for [3.3]. All polymers produced

with the catalysts were in the cis-transoidal configuration. Complexes [3.4]-[3.8]

were catalytically active in the hydroformylation of styrene. All complexes favoured

the formation of the branched (iso) aldehyde 2-phenylpropanal. Complexes [3.7] and

[3.8] were most active catalysts with the highest conversions of styrene.