Thesis s4

Download as pdf or txt
Download as pdf or txt
You are on page 1of 191

QUEENSLAND UNIVERSITY OF TECHNOLOGY

SCHOOL OF PHYSICAL AND CHEMICAL SCIENCES

SYNTHESIS AND REACTIONS OF

ORGANOMETALLIC PORPHYRINS

Submitted by Farzad ATEFI (Bachelor of Applied Science Honours) to the School of


Physical and Chemical Sciences, Queensland University of Technology, in partial
fulfilment of the requirements of the degree of Doctor of Philosophy.

August 2007
ABSTRACT

This thesis reports progress in three major aspects of σ-bonded organometallic


porphyrins that are described in the published papers found in chapters 4, 5 and 6.
meso-Iodoporphyrins, which were prepared in a rapid, selective and high yielding
methodology from the respective η1-palladioporphyrins or bromoporphyrins, are
important starting materials for further functionalisations of porphyrins. Their utility
was confirmed in a palladium-catalysed coupling reaction and this novel synthetic
strategy could potentially be applied for iodine/bromine exchange on other organic
substrates..

A η1-palladioporphyrin was also utilised to optimise the reaction conditions leading


to the formation of porphyrinylphosphine oxides. This synthetic strategy simplified
the challenging optimisation of the palladium-catalysed reaction and has great
potential to be applied in other catalytic processes. Subsequently a suite of
porphyrinylphosphine oxides was prepared under the optimised catalytic conditions.
These macrocycles, which represent a new class of porphyrins, were isolated cleanly
in very high yields. Detailed spectroscopic investigations as well as X-ray single
crystal analysis demonstrated their structures unambiguously and established their
potential as ligands for supramolecular chemistry.

The coordinating properties of phosphine oxides in general and


porphyrinylphosphine oxides in particular, towards Mg(II) centred porphyrins were
examined in further experiments. Triphenylphosphine oxide showed a strong affinity
towards Mg(II) porphyrins and the calculated displacement constant of 5.3 × 105 M-1
was two orders of magnitude larger than any other Mg(II) porphyrin-ligand binding
constant reported thus far. Di- and triporphyrin arrays consisting of Mg(II) porphyrin
coordinated to free base and Ni(II) porphyrinyl mono- and bis-phosphine oxides
were also prepared in high yields. Spectroscopic studies indicated that these
porphyrin oligomers exhibit strong inter-porphyrin electronic interaction.

A Mg(II) porphyrinylphosphine oxide dimer was also isolated in a satisfactory yield.


The large self-association constant of 5.5 × 108 M-1 confirmed the strong affinity of

iii
porphyrinylphosphine oxides towards Mg(II) porphyrins and established these
complexes as the first strongly bound synthetic Mg(II) porphyrin analogues of the
“special pair” of the photosynthetic reaction centre.

KEYWORDS
porphyrinoids, iodination, palladium, organometallic, crystal structure, ligand design,
phosphine oxides, oligomers, special pair, photosynthesis

iv
TABLE OF CONTENTS

ABSTRACT iii
KEYWORDS iv
TABLE OF CONTENTS v
LIST OF FIGURES ix
LIST OF SCHEMES xiii
LIST OF TABLES xvi
ABBREVIATIONS xvii
DECLARATION xx
NOTES TO THE READER xxi
ACKNOWLEDGEMENTS xxii

CHAPTER 1 1
1. INTRODUCTION 2
1.1 Research problem investigated 2
1.2 Overall aims of the project 2
1.3 Specific objectives of the project 3
1.4 Relationship of the research papers 3

CHAPTER 2 5
2. General review on porphyrins 6
2.1. Introduction to porphyrins 6
2.2. Electronic absorption properties of porphyrins 9
2.3. Emission properties of porphyrins 11
2.4. Synthetic porphyrins and functionalisation thereof 13
2.5. Introduction to multiporphyrin arrays 16
2.6. Synthetic multiporphyrin arrays 19
2.6.1. Covalently linked multiporphyrin assemblies 19
2.6.2. Multiporphyrin assemblies linked by axial coordination 24
2.7 References 28

v
CHAPTER 3 36
Porphyrins with metal, metalloid or phosphorus atoms directly bonded
to the carbon periphery 36
ABSTRACT 38
INTRODUCTION 39
CLASS F ORGANOMETALLIC PORPHYRINS 40
Mercury 40
Palladium and Platinum 44
Boron 55
Phosphorus 66
Reported intermediates: Te, Mg, Zn, Ni 73
CLASS G ORGANOMETALLIC PORPHYRINS 74
CLASS H ORGANOMETALLIC PORPHYRINS 74
CONCLUSION AND FUTURE OUTLOOK 75
Acknowledgements 76
REFERENCES 76

CHAPTER 4 81
meso-Iodo- and meso-iodovinylporphyrins via organopalladium
porphyrins and the crystal structure of 5-iodo-10,20-diphenylporphyrin 81
ABSTRACT 83
INTRODUCTION 84
EXPERIMENTAL 85
General 85
Synthesis 86
X-ray single crystal structure determination of 5-iodo-10,20-
diphenylporphyrin 91
RESULTS AND DISCUSSION 91
Acknowledgements 98
REFERENCES 98

vi
CHAPTER 5 102
meso-Porphyrinylphosphine Oxides: Mono- and Bidentate Ligands for
Supramolecular Chemistry and the Crystal Structures of Monomeric
{[10,20-Diphenylporphyrinatonickel(II)-5,15-diyl]-bis-[P(O)Ph2]} and
Polymeric Self-Coordinated {[10,20-Diphenylporphyrinatozinc(II)-5,15-
diyl]-bis-[P(O)Ph2]} 102
ABSTRACT 104
INTRODUCTION 106
EXPERIMENTAL SECTION 108
General Procedures and Instrumentation 108
RESULTS AND DISCUSSION 114
Synthesis and Characterization of meso-Porphyrinylphosphine
Oxides 114
Variable-Temperature 1H NMR Studies of 6 121
Electronic Absorption Spectra, Emission Spectra, and Redox
Properties 123
Crystal Structures 127
CONCLUSION 132
Acknowledgements 133
REFERENCES 133

CHAPTER 6 138
Multiporphyrin coordination arrays based on complexation of
magnesium(II) porphyrins with porphyrinylphosphine oxides 138
ABSTRACT 140
INTRODUCTION 141
RESULTS AND DISCUSSION 142
Synthesis and NMR characterisation 142
Electronic absorption spectra, emission spectra and redox
properties 150
CONCLUSION 153
EXPERIMENTAL 154
General remarks 154

vii
Acknowledgements 158
REFERENCES 159
ELECTRONIC SUPPLEMENTARY INFORMATION 162

CHAPTER 7 179
7. Discussion, conclusion and future work 179
7.1. Discussion and conclusion 180
7.2 Future Work 183
7.3. References 187

APPENDIX A I
A. Further investigations into the preparation of novel σ-bonded
organometallic porphyrins I
A.1. Transmetallation: Sn, Si, Pt, Au II
A.2. Formation of porphyrin anion radicals and Grignard porphyrin
analogues: Pt, Au, Sn VI
A.3. Formation of porphyrinylgold(I) complexes IX
A.4. References X

viii
LIST OF FIGURES

Chapter 1
Figure 1.1. Flow diagram of the correlation of the project aims, the results and
the chapters in the present thesis. Investigations into the formation of other
organometallic porphyrins have been compiled in Appendix A. 4

Chapter 2
Figure 2.1. Structure of porphine and the IUPAC numbering system. 6
Figure 2.2. Delocalised 18 π-electron conjugation pathway and tautomerism of
porphyrins. 9
Figure 2.3. UV-visible spectra of ZnDPP (left) and H2DPP (right). 9
Figure 2.4. The Gouterman four orbital model.10,11 10
Figure 2.5. Jablonski diagram.14 11
Figure 2.6. Schematic representation of the SP from the purple bacterium
Rhodopseudomonas viridis adapted from Deisenhofer et al.32d 16
Figure 2.7. Schematic representation of the RC from the purple bacterium
Rhodopseudomonas viridis adapted from Deisenhofer et al.32d 17
Figure 2.8. Schematic representation of LH-I from the purple bacterium
Rhodopseudomonas palustris adapted from Roszak et al.33 17
Figure 2.9. Schematic representation of LH-I from the purple bacterium
Rhodopseudomonas acidophila adapted from Cherezov et al.34 18
Figure 2.10. Porhyrin oligomers joined by meso-meso linkages. 19
Figure 2.11. Triply fused porphyrin dimers. 20
Figure 2.12. Formation of directly fused porphyrin tetramer. 21
Figure 2.13. Complexation types of porphyrins which act as ligands. 25

Chaper 3
Figure 1. Classification of organometallic porphyrins. 39
Figure 2. η1-Palladio- and η1-platinioporphyrins [11]. 45

ix
Figure 3. X-ray single crystal structures of η1-palladioporphyrin 55 (left) and
η1-platinioporphyrin 58 (right). These figures were generated using data
originally reported by Arnold and co-workers [11a,b]. 46
Figure 4. CD spectrum of 80 (upper, solid line), 81 (upper, dashed line) and
visible absorption spectrum (lower) in CH2Cl2 [11f]. 49
Figure 5. X-ray single crystal structure of Ni(II) porphyrinylphosphine oxide
222 (left) and Zn(II) porphyrinylphosphine oxide polymer 224 (right).
Reproduced with permission from Inorg. Chem. 2006; 45: 6479-6489.
Copyright 2006 American Chemical Society. 70

Chapter 4
Figure 1. View of the molecular structure of 3 in the crystal. Thermal ellipsoids
are drawn for 50% occupancy and hydrogen atoms have been omitted for
clarity. 98

Chapter 5
Figure 1. Quasi-onium ion. 120
Figure 2. Variable-temperature 1H NMR spectra in CDCl3 (left) and CD2Cl2
(right) of 6. 122
Figure 3. Tautomerism of inner N-protons. 123
Figure 4. UV/vis absorption spectra for porphyrinyl phosphine oxides (in
CH2Cl2 solution; with 5% pyridine for 15 and 16). 124
Figure 5. Fluorescence spectra of H2DPP, 6, and 12 (in CH2Cl2 solution). 126
Figure 6. (a) ORTEP representation of {[10,20-
diphenylporphyrinatonickel(II)-5,15-diyl]-bis-[P(O)Ph2]}·H2O. The complex
has 2-fold symmetry (1 - x, y, -z + 3/2). Ellipsoids are depicted with 30%
probability, and water and H atoms have been omitted. (b) Illustration of the
{[10,20-diphenylporphyrinatonickel(II)-5,15-diyl]-bis-[P(O)Ph2]}·H2O
complex showing the almost perfect square planar coordination contrasting
with the distorted macrocycle framework. Note also the cis-like arrangement of
the diphenylphosphine oxide groups. 127

x
Figure 7. ORTEP representation of the asymmetric unit of {[10,20-
diphenylporphyrinatozinc(II)-5,15-diyl]-bis-[P(O)Ph2]}n·{1.5H2O}n 50%
probability ellipsoids, disordered water omitted). The coordination sphere of
the complex is shown in Figure 8. 129
Figure 8. Section of the one-dimensional coordination polymer in {[10,20-
diphenylporphyrinatozinc(II)-5,15-diyl]-bis-[P(O)Ph2]}n·{1.5H2O}n. 129

Chpater 6
Figure 1. UV-visible titration of MgDPP 2 (1.4 × 10−6 M) in toluene at 25 ºC
upon successive addition of Ph3PO (7.9 × 10−7-1.5 × 10−4 M). Inset: Benesi–
Hildebrand plot (solid line: fitted curve).21 144
Figure 2. (Left) 1H NMR spectra of MgDPP 2 (bottom), complex 8 (middle)
and ligand 4 (top); (right) 1H NMR spectra of MgDPP 2 (bottom), complex 11
(middle) and ligand 7 (top). 148
Figure 3. UV-visible spectra of 8 (1.1 × 10−4 to 5.4 × 10−6 M; 10 mm and 1
mm cells). 151
−6
Figure 4. UV-visible titration of 12 (3.1 × 10 M) in toluene at 25 ºC upon
successive addition of Ph3PO (5.7 × 10−5 to 2.9 × 10−2 M). Inset: Binding
isotherm (solid line: fitted curve for 1 : 2 complex). 151
Figure 5. Fluorescence spectra of 2, 2 + 13 and 9 (left); 8, 10-12 (right) excited 152
at 485 nm in toluene at 25 ºC and normalised to 10−5 M (except for 12:
measured at 10−6 M on a more sensitive detector setting).
Figure S2. 1H NMR of 2. 163
Figure S3. 1H NMR of 3. 164
Figure S4. 1H COSY NMR of 3. 165
Figure S5. 1H NOESY NMR of 3. 166
Figure S6. 1H COSY NMR of 8. 167
1
Figure S7. H NOESY NMR of 8. 168
1
Figure. S8 H NMR of 9. 169
Figure S9. 1H COSY NMR of 9. 170
Figure S10. 1H NOESY NMR of 9. 171
Figure S11. 1H NMR of 10. 172
Figure S12. 1H COSY NMR of 10. 173

xi
Figure S13. 1H NOESY NMR of 10. 174
Figure S14. 1H COSY NMR of 11. 175
Figure S15. 1H NOESY NMR of 11. 176
Figure S16. 1H NMR of 12. 177
1
Figure S17. H COSY NMR of 12. 178

Chapter 7
Figure 7.1. Concept of the preparation of organometallic porphyrins from a
porphyrin anion radical intermediate (charges are not shown and exact reaction
sequence is not implied). 184
Figure 7.2. Preparation of triporphyrinylphosphine oxide. 185
Figure 7.3. Attachment of porphyrinylphosphonic acids to TiO2 surface. 186

Appendix A
Figure A.1. 1H NMR spectrum of the reaction mixture after stannylation in
CDCl3. III

xii
LIST OF SCHEMES

Chapter 2
Scheme 2.1. Formation of 5,15-diarylporphyrins. 14
Scheme 2.2. Formation of 5,10,15-triarylporphyrins. 15

Chapter 3
Scheme 1. Formation of mercury porphyrins by Smith and co-workers [4]. 40
Scheme 2. Palladium-catalysed C-C coupling with mercurials by Smith and
co-workers [4a,4c,4g]. 42
Scheme 3. Mercuration of meso-tetra(p-tolyl)porphyrin [7]. 42
Scheme 4. Mercuration of 5,15-diaryl- and 5,15-dialkylporphyrins [8]. 43
Scheme 5. Reactions of 5,15-diarylporphyrin mercurials [8]. 44
Scheme 6. Formation of η1-palladioporphyrin [11a]. 44
Scheme 7. Formation of η1-platinioporphyrin with chelating amine ligand
[11e]. 48
Scheme 8. Formation of alkynyl-Pt(II) linked porphyrin dimer [11d]. 49
Scheme 9. Self-assembly of porphyrin pentamer [15]. 50
Scheme 10. Selective monobromination of 5,15-diarylporphyrins [16]. 51
Scheme 11. Selective meso-iodination of 5,15-diarylporphyrins [17]. 52
Scheme 12. Intramolecular Pd(0) catalysed cyclisation [18]. 52
Scheme 13. Formation of doubly linked naphthyl-porphyrins [19]. 53
Scheme 14. Formation of meso-β linked porphyrin dimers [20]. 54
Scheme 15. Pd(0) catalysed annulation of porphyrins [21]. 54
Scheme 16. Formation of porphyrinyl boronates [23-31] (the point of
attachment of the aryl groups to the meso carbon is from the lowermost ring
carbon). 56
Scheme 17. Carbazole linked porphyrin dimer [23a]. 57
Scheme 18. Bis(phenyldipyrromethane) linked porphyrin dimers [25b]. 58
Scheme 19. Cofacial porphyrin dimers [27b]. 58
Scheme 20. Phenyl linked porphyrin dimer [29]. 59
Scheme 21. Aniline-ferrocene linked porphyrin dimer [31a]. 60

xiii
Scheme 22. Azobenzene-ferrocene linked porphyrin dimer [31b]. 61
Scheme 23. Directly linked and phenylene bridged porphyrin dimers
[24a,24c]. 62
Scheme 24. Directly linked porphyrin dimers with ester functionality [28,38]. 63
Scheme 25. Directly linked porphyrin trimer with cyanide functionality [30]. 63
Scheme 26. Formation of β-β linked porphyrin dimer [39]. 64
Scheme 27. Formation of chiral β-β linked porphyrin dimers [40]. 64
Scheme 28. β-Boronated diarylporphyrins [41]. 65
Scheme 29. Boronic acid substituted porphyrins [42]. 66
Scheme 30. meso-Porphyrinylphosphonium salt [43]. 67
Scheme 31. β-Porphyrinylphosphonium salt [44,45]. 67
Scheme 32. β-β Linked porphyrinyl phosphonium salt dimers and trimers
[47]. 68
Scheme 33. Formation of diarylporphyrinylphosphine oxide [17]. 69
Scheme 34. Reactions of η1-palladioporphyrin with various phosphines and
diphenylphosphine oxide [48]. 69
Scheme 35. Palladium-catalysed formation of mono- and bis-
porphyrinylphosphine oxides [48]. 70
Scheme 36. CuI catalysed formation of porphyrinylphosphine oxides,
phosphonates and phosphonic esters [49]. 71
Scheme 37. Coordination arrays of Mg(II) diphenylporphyrin with
porphyrinylphosphine oxides [50]. 72
Scheme 38. Mg(II) porphyrinylphosphine oxide [50]. 72
Scheme 39. Formation of doubly-fused porphyrin dimer via Te-porphyrin
intermediate [51]. 73
Scheme 40. Formation of π-organometallic porphyrins (charges, counterions 74
not shown) [54].
Scheme 41. Formation of β,β’-fused metalloceneporphyrins [56]. 75

Chapter 4
Scheme 1. Iodination of η1-palladioporphyrin. 92
Scheme 2. One-pot meso-iodination. 93
Scheme 3. Palladium catalyzed phosphination. 95

xiv
Scheme 4. Formation of iodoethenylporphyrins; conditions: i) pyridinium
tribromide; ii) 1. Pd(PPh3)4, 2. I2; iii) Ni(acac)2; iv) Zn(OAc)2. 96

Chapter 5
Scheme 1. Formation of meso-Porphyrinylphosphine Oxide by the Stille
Method. 115
Scheme 2. Synthesis of meso-η1-Palladioporphyrin. 116
Scheme 3. Stoichiometric Formation of meso-Porphyrinylphosphine Oxide 6. 116
Scheme 4. Catalytic Formation of meso-Porphyrinylphosphine Oxides. 118

Chapter 6
Scheme 1. Formation of MgDPP 2 and MgDPP-PPh3PO complex 3. 143
Scheme 2. Formation of di-porphyrin complexes 8, 9 (above) and tri-
porphyrin complexes 10, 11 (below). 146
Scheme 3. Formation of Mg(II) phosphine oxide 12 and possible polymeric
form. 149

Appendix A
Scheme A.1. Proposed formation of porphyrinylstannane and
porphyrinylsilane by transmetallation. II
Scheme A.2. Proposed formation of porphyrinylplatinum(II) and
porphyrinylgold(I) complexes by transmetallation. V
Scheme A.3. Proposed formation of porphyrin anion radical by treatment with
sodium naphthalenide. VI
Scheme A.4. Proposed formation of Grignard porphyrin analogue. VIII
Scheme A.5. Proposed formation of porphyrinylgold(I) complex. IX

xv
LIST OF TABLES

Chapter 2
Table 2.1. Luminescence properties of porphyrins in solution at 300 K.16 12

Chapter 3
Table 1. Electrochemical data for η1-metalloporphyrins and some precursors in
CH2Cl2-Bu4NPF6 vs. Ag/Ag+ using ferrocene as internal standard [E0(ox) =
+0.55 V] at 293K [11c]. 47

Chapter 5
Table 1. Experimental Crystallographic Data for 14 and 16. 113
Table 2. Conditions and Yields (%) for Stoichiometric Reactions to Form 6. 117
Table 3. Conditions and Yields (%) for Catalytic Reactions. 119
Table 4. IR Frequencies (cm-1) of Porphyrinylphosphine Oxides. 120
Table 5. Wavelengths for the UV/vis Absorption Bands for
Porphyrinylphosphine Oxides (in CH2Cl2 Solution; with 5% Pyridine for 15
and 16). 125
Table 6. Selected Bond Lengths (Å) and Angles (deg) for {[10,20-
diphenylporphyrinatonickel(II)-5,15-diyl]-bis-[P(O)Ph2]}·H2O. 128
Table 7. Selected Bond Lengths (Å) and Angles (deg) for {[10,20-
diphenylporphyrinatozinc(II)-5,15-diyl]-bis-[P(O)Ph2]}n·{1.5H2O}n. 131

Chapter 6
Table 1. Oxidation potentials for 1–4, 6, 8 and 10 (10−3 M) measured in
CH2Cl2 containing 0.1 M Bu4NPF6. 153

xvi
ABBREVIATIONS

Note: Porphyrins are often abbreviated, because of their long IUPAC names. The
system used in many articles and books is to state at first, whether the compound is a
free base porphyrin or a metalloporphyrin. H2 indicates a free base porphyrin, while
in case of a metalloporphyrin the chemical symbol of the metal is used. The next set
of letters (usually three or four) abbreviates the parent porphyrin system.
Abbreviations used for parent porphyrins are listed below. The last letter or letters, if
any, indicates substituents that have been introduced to the porphyrin moiety. As an
example H2OEP is the abbreviation for free base octaethylporphyrin and ZnDAPBr
describes zinc diaryl-bromoporphyrin.

AR analytical reagent
Ar aryl
BAHA tris(4-bromophenyl)aminium hexachloroantimonate
BCh bacteriochlorophyll
BPh bacteriopheophytin
Bpy 2,2’-bipyridyl
Bu butyl
Car carotenoid
Cat catalyst
CHIRAPHOS 2,2’-bis(diphenylphosphino)butane
Cod cyclooctadiene
Conc concentration
COSY correlation spectroscopy
Cp* 1,2,3,4,5-pentamethylcyclopentadiene
Cymene p-isopropyltoluene
DAP 5,15-diarylporphyrin
Dba dibenzylideneacetone
DBU 1,5-diazabicyclo[5.4.0]undec-7-ene
DEST Department of Education, Science and Training
Diphos 1,2-bis(methylphenylphosphino)benzene
Dppe 1,2-bis(diphenylphosphino)ethane

xvii
Dppf 1,1’-bis(diphenylphosphino)ferrocene
Dppp 1,3- bis(diphenylphosphino)propane
DDQ 2,3-dichloro-5,6-dicyano-1,4-benzoquininone
DDT threo-1,4-dimercapto-2,3-butanediol
DPP 5,15-diphenylporphyrin
Dqf double-quantum filtered
Dtbpy di-t-butyl-2,2’-bipyridyl
ESI electrospray ionisation
Et ethyl
FT Fourier transformed (transformation)
H hour
HOMO highest occupied molecular orbital
IR infrared
L ligand
LDI laser desorption/ionisation
LH light-harvesting centre
LR laboratory reagent
LSI liquid secondary ion
LUMO lowest unoccupied molecular orbital
M metal
Me methyl
MeOH methanol
Mes mesityl
Min minutes
MS mass spectrum (spectroscopy)
NBS N-bromosuccinimide
NMR nuclear magnetic resonance
NOESY nuclear Overhauser-effect spectroscopy
OAP 2,3,7,8,12,13,17,18-octaalkylporphyrin
OEP 2,3,7,8,12,13,17,18-octaethylporphyrin
ORTEP Oak Ridge thermal ellipsoid plot program
OTf triflate (trifluoromethanesulfonate)
Ox oxidation

xviii
Q quinone
Ph phenyl
Por porphyrin
RC reaction centre
Red reduction
SP special pair
TAP 5,10,15,20-tetraarylporphyrin
TEA triethylamine
Temp temperature
TFA trifluoroacetic acid
THF tetrahydrofuran
TLC thin-layer chromatography
Tmeda N,N,N’,N’-tetramethylethylenediamine
Tol tolyl
Tol-BINAP 2,3-bis(di-p-tolylphosphino)-1,1’-binaphthyl
TPP 5,10,15,20-tetraphenylporphyrin
TriAP 5,10,15-triarylporphyrin
UHP ultra-high purity
UV ultraviolet
Vis visible

xix
DECLARATION

The work contained in this thesis has not been previously submitted to meet
requirements for an award at this or any other higher education institution. To the
best of my knowledge and belief, the thesis contains no material previously
published or written by another person except where due reference is made.

Farzad Atefi
August, 2007

xx
NOTES TO THE READER
Due to the layout required for a thesis by published papers, the numbering scheme of
compounds might lead to some confusion. Within the individual chapters and
manuscripts the compounds are numbered consecutively. Should a compound be
referred to again in a later chapter, for example in the discussions or the appendices,
it will be numbered as chapter number first followed by compound number.
Compound 5.3 for example refers to compound 3 from chapter 5. Some
inconsistencies in the formatting from chapter to chapter are to be expected. Each
chapter is formatted to meet the requirements of the journal. Some diagrams have
been moved and re-sized within these chapters in order to make it easier for the
reader.

xxi
ACKNOWLEDGEMENTS
First and foremost I want to express my gratitude towards my supervisor Ass. Prof.
Dennis P. Arnold for his tremendous support throughout the course of this project.
His great insight into all aspects of chemistry and his conduct of research are a great
inspiration for me.

I want to offer my gratitude to my associate supervisor Dr. John C. McMurtrie for his
assistance with X-Ray single crystal analysis and whose advice and feedback
provided valuable input for my research.

I am especially grateful to Prof. Mathias O. Senge from the University of Dublin,


Prof. Peter Turner from the University of Sydney and Martin Duriska from the
University of Melbourne for their help with X-Ray single crystal analysis.

I want to thank Dr. John Bartley, Dr. Roger Meder, Dr. Mark Wellard and Dr. Chris
Carvalho for their advise with NMR instrumentation, Dr Llew Rintoul for his
assistance with IR instrumentation and Patrick Stevens for his help with UV/visible
and Fluorescence instrumentations.

I am very grateful for the Post-Graduate scholarship provided by the Department of


Education, Science and Training (DEST) and for the financial support of the
Synthesis and Molecular Recognition Program and the Faculty of Science,
Queensland University of Technology. I could have never finished this thesis without
the before mentioned scholarships.

I want to thank the past and present members of the Arnold research group, as well
as all other Post-Graduate students who shared the same laboratory with me, for the
valuable discussions about the practical aspects of chemistry.

I am most grateful for the love and support I received from my father, my brother
and my late mother. They have always been a pillar of strength throughout my life
and always believed in me and my abilities.

xxii
Last, but definitely no least, I want to thank my partner (and hopefully soon to-be
wife) Dr. Alexandra C. Bloch. She never ceases to support me, even in times of
hardship. Her love is the most precious gift in my life and for that I am eternally
grateful.

xxiii
CHAPTER 1

Introduction

1
Chapter 1

1. INTRODUCTION

1.1 Research problem investigated

Porphyrins and related macrocycles are of fundamental importance to all living


beings on our planet. Life as we know it would cease to exist without biological
processes such as the photosynthesis of plants and bacteria and the oxygen transport
in the cardiovascular systems of humans and animals. Porphyrins are the key
components in these processes and have as such attracted great interest from
researchers of diverse scientific background. Naturally occurring porphyrinoids like
chlorophyll, haem or coenzyme B12 have been the centre of many investigations in
the past century. In order to gain more insight in the complicated role models
supplied by Nature, it is essential to create simpler, synthetic systems that can be
more easily studied. meso-η1-Organometallic porphyrins have emerged in recent
years as very useful tools to create such porphyrin systems, as they are crucial
intermediates in many bond-forming reactions involving porphyrins. They have also
very interesting physical and chemical properties on their own, especially when
incorporated into multiporphyrin entities. The present project involves investigations
of organometallic porphyrins as synthetic tools and as components in supramolecular
arrays.

1.2 Overall aims of the project

The overall project aim of this PhD project, as reported in the present thesis, is to
investigate the reactions of organometallic porphyrins such as meso-η1-palladio(II)
porphyrins and the formation of novel meso-organometallic porphyrins. The overall
objectives include:
• Investigation of the reactions of meso-η1-palladio(II) porphyrins as means of
obtaining synthetically useful porphyrins and novel organometallic
porphyrins
• Formation of novel organometallic porphyrins via different routes
• Formation of multiporphyrin systems consisting of these novel
organometallic porphyrins
• Study of these supramolecular entities

2
Chapter 1

1.3 Specific objectives of the project

• To develop and optimise a fast and high yielding pathway in order to prepare
synthetically important meso-iodoporphyrins from meso-η1-palladio(II)
porphyrins
• To investigate the reactions of meso-iodoporphyrins
• To investigate and optimise the formation of porphyrinylphosphine oxides
from meso-η1-palladio(II) porphyrins
• To investigate and optimise the formation of porphyrinylphosphine oxides by
other means
• To investigate the formation of other novel organometallic porphyrins
• To investigate porphyrinylphosphine oxides as novel ligands for
supramolecular chemistry
• To investigate multiporphyrin entities consisting of porphyrinylphosphine
oxides and magnesium(II) porphyrins as synthetic mimics of the “special
pair” in the photosynthetic reaction centre
• To investigate and characterise the structural, spectroscopic and electronic
properties of all complexes mentioned above by:
1
ƒ NMR spectroscopy
ƒ Multidimensional (COSY, NOESY) NMR spectroscopy
31
ƒ P NMR spectroscopy (where applicable)
ƒ UV-visible spectroscopy
ƒ Fluorescence spectroscopy
ƒ IR spectroscopy
ƒ Mass spectroscopy
ƒ Elemental analysis
ƒ Single crystal X-ray crystallography
ƒ Electrochemistry

1.4 Relationship of the research papers

The following flow diagram illustrates the correlation between aims of this PhD
project and the published manuscripts as well as the results of the research project.

3
Chapter 1

Chapter 5, 6 Chapter 5, 6
Incorporation of Synthesis, characterisation and
porphyrinylphosphine oxides in investigation into the properties
multiporphyrin arrays of multiporphyrin arrays

Chapter 5 Chapter 5
Development of the catalytic Method optimisation for the
formation of synthesis of
porphyrinylphosphine oxides porphyrinylphosphine oxides

Chapter 5 Chapter 5
Development of Method development, synthesis
porphyrinylphosphine oxides and characterisation of
from η1-palladio(II) porphyrins porphyrinylphosphine oxides

Chapter 4 Chapter 4
Development of iodoporphyrins Method development, synthesis,
from η1-palladio(II) porphyrins characterisation and reactivity of
iodoporphyrins

Chapter 2 – General Literature Chapter 3 – Review on


review organometallic porphyrins
™ General porphyrin ™ Types of organometallic
introduction porphyrins
™ Introduction to the ™ Organometallic porphyrins
photosynthetic process as synthetic tools
™ Introduction to ™ Multiporphyrin arrays
multiporphyrin arrays containing organometallic
porphyrins

Figure 1.1. Flow diagram of the correlation of the project aims, the results and the
chapters in the present thesis. Investigations into the formation of other
organometallic porphyrins have been compiled in Appendix A.

4
CHAPTER 2

General review on porphyrins

5
Chapter 2

2. General review on porphyrins

2.1. Introduction to porphyrins

The word porphyrin stems from the ancient Greek word porphura, which was used
to describe the colour purple. Indeed all naturally occurring and synthetic porphyrins
are deeply coloured compounds. The basic structure of the parent porphyrin
macrocycles consists of four pyrrole rings joined together by four methine bridges.
Figure 2.1. shows the basic porphyrin, also referred to as porphine, together with the
IUPAC numbering of the ring system.1
18 19 20 2
1
β 17
3

16 NH N 4
24 21
15 5 meso
23 22
α 14 N HN 6

13 7
11 9
12 10 8

Figure 2.1. Structure of porphine and the IUPAC numbering system.

In order to simplify the IUPAC numbering system, carbons 1, 4, 6, 9, 11, 14, 16 and
19 are referred to as the α-positions, carbons 2, 3, 7, 8, 12, 13, 17 and 18 as the β-
positions and carbons 5, 10, 15 and 20 as the meso-positions.2 The planar porphyrin
macrocycle is an 18 π-electron aromatic system. The two inner NH groups (22, 24)
lose protons under basic conditions in order to form a dianion species. Such a
porphyrin dianion is able to coordinate almost every metal within its cavity to form a
metalloporphyrin. This incorporation may result in the distortion of the planar
macrocycle in order to maximise the binding strength towards the metal fragment.3

Naturally occurring metalloporphyrins are key components in some of the most


important biological processes; without them, life as we know it would be
impossible. Iron (II) protoporphyrin IX (1) derivatives in haemoglobin, cytochrome c
and myoglobin are some of the most widely studied biological molecules.

6
Chapter 2

N N
Fe
N N

HO2C 1 CO2H

These naturally occurring chromophores are anchored in a protein matrix and


reversibly bind oxygen. Thus haemoglobin transports oxygen from the lungs and the
skin to muscle tissue, where it is transferred via cytochrome c to myoglobin for
storage.4

One of the most vital and fascinating biochemical processes is the photosynthesis
that is achieved in plants and bacteria. During this process CO2 and H2O are
chemically transformed to glucose and O2 utilising the most abundant natural energy
resource – sunlight. The key chemical entity in this process is chlorophyll, e.g.
chlorophyll a (2).

N N
Mg
Phytyl =
N N
H
Me
H
H 2
O
PhytylO2C CO2Me

In this green pigment Mg(II) is coordinated within the core of the porphyrin. Its role
is the absorption of light in both the ultraviolet (~400 nm) and red (~675 nm) regions
of the electromagnetic spectrum. The porphyrinoid is embedded in a protein matrix
with the help of its substituents which also help to regulate the light absorption
properties of the macrocycle. The central Mg(II) metal also helps to fine-tune the
light absorption properties of the chromophore and also binds water, which is the
source of electrons.5

7
Chapter 2

The amount of knowledge that has been accumulated in the past 100 years is
evidence of the crucial role porphyrins play in our daily life. Hans Fischer, by many
regarded as the godfather of modern porphyrin chemistry, published the first
monographs concerned with these macrocycles.6 For three decades these were the
standard porphyrin references until they were replaced by the first multiauthor
publication on porphyrins in 1975.7 Due to the rapid growth of the multidisciplinary
investigations of both naturally occurring and synthetic porphyrins, in 1979 a seven
volume series edited by David Dolphin was published.8 For the first time the editor
compiled all aspects of porphyrin science, from physical to chemical and biological,
in one portfolio. Just recently The Porphyrin Handbook, a twenty-volume series with
over 7000 pages divided into more than 120 chapters, was released.9 New aspects of
porphyrin research showing possible applications as drugs and catalysts, as well as
findings from modern instrumentation, such as supercomputers and synchrotron X-
ray crystallography, combined with the perpetual quest to mimic the photosynthetic
process, are the forces driving many researchers from various scientific backgrounds.

Although research is on the brink of fully appreciating the principles of


photosynthesis, it is not yet possible for mankind to copy this highly complex
system. In order to simulate such biological processes, simpler synthetic porphyrin
analogues have to be designed and isolated and new synthetic tools have to be
developed. These more tractable macrocycles might offer the insights needed to
understand systems which are presently too complex for our scientific tools.9

8
Chapter 2

2.2. Electronic absorption properties of porphyrins

The delocalised aromatic character of porphyrins, which results from extensive


conjugation (Figure 2.2.), accounts for the one of the most striking features of these
chromophores, namely their intense colour.

NH N N HN

N HN NH N

Figure 2.2. Delocalised 18 π-electron conjugation pathway and tautomerism of


porphyrins.

What the human eye senses as intense colour can be measured by electronic
absorption spectroscopy and explained by molecular orbital theory.10 Modifications
to the basic macrocycle structure as well as the substituent in the core of the
porphyrin result in changes in the optical spectra.11 Figure 2.3. shows the UV-visible
spectra of Zn(II) (ZnDPP, left) and free base 5,15-diphenylporphyrin (H2DPP, right).
The figure illustrates that a typical porphyrin absorption spectrum consists of two
regions, one at around 400 nm, which is referred to as the Soret or B-band region,
and one at approximately 550 nm, which is called the Q-band region. The first
corresponds to a strongly allowed transition from the ground state to the second
exited state (S0 → S2). The latter corresponds to a transition from the ground to the
first excited state (S0 → S1).10d

Figure 2.3. UV-visible spectra of ZnDPP (left) and H2DPP (right).

9
Chapter 2

The spectra of metalloporphyrins consist often of two bands in the visible region.
While the lower energy band Q(0,0) is a result of the excitation to the first excited
state, the higher energy band Q(1,0) is often referred to as a vibronic overtone, as it is
associated with vibrational transitions interacting with the electronic transitions of
the macrocycles.10d Metals with closed electron shells have only a slight effect on the
absorption spectra of the chromophores, due to a small perturbation of the π electrons
of the porphyrin ring. Metal orbitals of open shell metals on the other hand mix much
more strongly with the ring orbitals of the porphyrins, thus having a more
pronounced effect on the absorption spectra.10d If the covalent binding is limited to
the s and p orbitals of the metal ion (e.g. magnesium, zinc) a red-shift results, with a
more intense Q-band compared with the free base analogue.12

In free base porphyrins the symmetry of the macrocycles is lowered from D4h to D2h
due to the presence of the pyrrole protons. Thus the degenerate Q(0,0) is replaced by
transitions polarised along the x and y axes, i.e. Qx(0,0) and Qy(0,0). Both these
bands have also the vibronic overtones Qx(1,0) and Qy(1,0) resulting in a total of four
bands in the visible region for free base porphyrins.10d

S2

eg x,y

B
S1
egy (LUMO) egx (LUMO)

a1u
a2u
S0

a1u (HOMO) a2u (HOMO)

Figure 2.4. The Gouterman four orbital model.10,11

The absorption bands can be theoretically explained by the so-called “Gouterman


four orbital model” developed by Gouterman et al. in the 1960’s10a-c and discussed
by Anderson.13 Figure 2.4. shows the four HOMOs and LUMOs of a typical

10
Chapter 2

metalloporphyrin. Because of the similar energy of a1u and a2u and the degeneracy of
the pair of LUMOs there is a strong interaction between the a1u → eg and the a2u →
eg transitions. Constructive interference leads to the intense short-wavelength B-band
whereas destructive interference leads to the weak long-wavelength Q-band.

2.3. Emission properties of porphyrins

Jablonski diagrams are representaions of molecular electronic state energy levels.14


Figure 2.5. shows such a diagram which is applicable to explaining the emission
properties of most aromatic compounds with singlet ground states, including
porphyrins.

This figure is not available online.


Please consult the hardcopy thesis
available from the QUT Library

Figure 2.5. Jablonski diagram.14

According to Kasha’s rule, after excitation from the ground state S0 to any singlet
excited state Sx, all higher excited states convert radiationlessly to the lowest excited
state S1 before they can emit.14 From S1 the molecule can either radiationlessly decay
to the ground state S0, convert internally to the lowest triplet state T1 or emit
fluorescence radiation. From T1 the molecule can either decay radiationlessly to the
ground state S0, reexcite to the lowest excited state S1 or emit phosphorescence
radiation.14

11
Chapter 2

The systematic study of the influences of the central metal ion on the luminescence
properties of porphyrins by Becker and Allison remains to date the basic reference in
this field.15 The studies show that four different categories of metalloporphyrins can
be discerned, namely non-luminescent, fluorescent, phosphorescent and fluorescent
and phosphorescent (Table 2.1.).16

Table 2.1. Luminescence properties of porphyrins in solution at 300 K.16

This figure is not available online.


Please consult the hardcopy thesis
available from the QUT Library

Fluorescence and phosphorescence of metalloporphyrins occurs ordinarily at shorter


wavelengths than for free base analogues as the singlet- and triplet-state energies are
higher for the former.16 A decrease in fluorescence and increase in phosphorescence
for metalloporphyrins can be attributed to the “heavy atom” effect.17 The low-lying
d-d excited states in Ni(II) porphyrins result in a lack of emission in these complexes,
while all other d8 transition metal complexes exhibit strong luminescence.16 Mg(II)
porphyrins exhibit very high fluorescence quantum yields (~0.25) with very long
excited state lifetimes (~130 ms), while Zn(II) porphyrins show moderate
fluorescence quantum yields (~ 0.05) with long excited state lifetimes (~0.75 ms).10d
The fluorescence spectra of metalloporphyrins consists of two peaks, namely Q(0,0)
and Q(0,1) which are mirror images of the absorption peaks Q(0,0) and Q(1,0).10d

Chelating different metal fragments in the core of the chromophores remains one of
the most powerful tools for synthetic chemists to adapt the chemical, physical and
structural properties of porphyrins.18

12
Chapter 2

2.4. Synthetic porphyrins and functionalisation thereof

Naturally occurring porphyrinoids are often challenging to synthesise. Multistep


reactions with low yields and problematic purification processes lead to the necessity
of simpler synthetic analogues.19 Unsubstituted porphine is almost insoluble in
organic solvents and is therefore little used as a synthetic starting point.20 In order to
increase the solubility, the porphyrin macrocycle has to be functionalised with alkyl
or aryl groups. In the past five decades two main synthetic porphyrin systems were
utilised by synthetic chemists. 5,10,15,20-Tetraarylporphyrins (TAPs) such as
5,10,15,20-tetraphenylporphyrin (TPP, 3) were prepared in good yields of up to 50%
from the acid catalysed condensation of pyrrole with the respective aldehyde.41
While a variety of aldehydes were successfully utilised in this synthetic strategy, the
big disadvantage of TAPs is the absence of any vacant meso position. The a2u
HOMO orbital of porphyrins has an area of high electron density in the meso
positions (see Figure 2.4.) and therefore communication to and from the macrocycle
to substituents occurs more readily in these positions.10,11,22 Also substitution
reactions occur generally more readily in the meso position if both β and meso
positions are available.21b From a synthetic viewpoint functionalisation of the TAPs
has also been challenging, as substitution reactions occur almost randomly in any of
the eight vacant β positions.23

N N N N
M M
N N N N

3 4

In order to overcome the problems associated with TAPs, 2,3,7,8,12,13,17,18-


octaalkylporphyrins (OAPs) were developed. These porphyrins have four vacant
meso positions for functionalisation and all eight β positions are blocked by alkyl
groups. They were isolated from the acid catalysed condensation of the

13
Chapter 2

corresponding pyrrole with formaldehyde in high yields of up to 75%. To ensure


solubility, ethyl groups were often incorporated as substituents in the β positions to
give 2,3,7,8,12,13,17,18-octaethylporphyrins (OEP, 4). However steric hindrance
due to the alkyl chains was often encountered in the subsequent substitution of the
vacant meso positions.24

5,15-Diarylporphyrins (DAPs) overcame the synthetic and structural problems


associated with both TAPs and OAPs. They were prepared from the acid catalysed
condensation of dipyrromethane with the corresponding aryl aldehyde in a two step
synthetic strategy shown in Scheme 2.1. High yields of up to 60% and two vacant
meso positions for subsequent functionalisation, as well as scaleable synthetic
strategies to isolate gram quantities of the desired porphyrins, make DAPs attractive
building blocks for synthetic chemists.21b,25
Ar

N NH N
H NH HN
+
+
O H N HN
H
Ar
O H

Ar = aryl (e.g. phenyl, tolyl, mesityl, etc.) Ar

Scheme 2.1. Formation of 5,15-diarylporphyrins.

As expected for all aromatic compounds, electrophilic substitution reactions account


for the majority of the classical porphyrin functionalisation pathways. The meso-
positions of porphyrins were nitrated using nitric acid in either acetic- or sulphuric
acid, nitronium salts (e.g. NO2BF4), PhSeNO2 in THF or with nitrate salts [e.g.
Cu(NO3)2]. The resulting nitroporphyrins were reduced to the corresponding amines
by reduction with sodium borohydride and Pd/C in methanol. The formylation of
porphyrins under Vilsmeier conditions with POCl3 and dimethylformamide was for
many years the standard entry point for porphyrin functionalisation. Only
metalloporphyrins were successfully utilised for this reaction, as free base porphyrins
were protonated under the acidic reaction conditions. Formylporphyrins were
converted into a variety of functionalised porphyrins, including the cyanoporphyrins,
vinylporphyrins under Wittig conditions, and carboxyporphyins.23

14
Chapter 2

The porphyrin periphery was also successfully functionalised by nucleophilic


substitution reactions. The first strategy involved the generation of a
metalloporphyrin cation radical, either chemically or electrochemically, by the one-
electron oxidation of the neutral macrocycle. These radical species were isolated in
crystalline form and are stable in solution for up to fourteen days, as the unpaired
electron is delocalised over the π electron cloud of the chromophore.26 The radical
cation porphyrins were subsequently nitrated, chlorinated, pyridinated, cyanated and
phosphinated.27 They were also used to prepare directly linked porphyrin oligomers,
which will be discussed in section 2.5.

In 1998 Senge and co-workers developed a nucleophilic substitution reaction on


porphyrins by utilising organolithium reagents.28 Various OAPs were functionalised
with butyllithium and aryllithium reagents.29 The reactions occurred always
selectively on one of the four vacant meso positions. Under similar reaction
conditions DAPs were substituted on the meso position as no attack on the β carbon
took place (Scheme 2.2.).30 This reaction pathway has been of immense importance
to obtain 5,10,15-triarylporphyrins (TriAPs) with only one vacant meso position, as
subsequent substitutions can be selectively accomplished without obtaining reaction
mixtures containing mono- and di-substituted porphyrins.
Ar Ar

N N N N
M + Ar'Li M Ar'
N N N N

Ar M = 2H, Ni, Zn, Cu, Co Ar

Scheme 2.2. Formation of 5,10,15-triarylporphyrins.

In the past two decades the palladium-catalysed substitution has emerged as the most
versatile functionalisation pathway for porphyrins. These reactions will be discussed
in Chapters 3, 4 and 5.

15
Chapter 2

2.5. Introduction to multiporphyrin arrays

In 1984 Deisenhofer and co-worker reported the X-ray structure of the


photosynthetic reaction centre (RC) of Rhodopseudomanis viridis, a purple
bacterium.31 The centre of the RC contains two bacteriochlorophyll molecules
(BCh), which are referred to as the special pair (SP). The special pair is the primary
electron donor in the RC and has an almost perfect two-fold symmetry with face to
face overlap of the macrocycles (Figure 2.6.).31,32

Figure 2.6. Schematic representation of the SP from the purple bacterium


Rhodopseudomonas viridis adapted from Deisenhofer et al.32d

The electron transfer from the SP is facilitated by a neighbouring BCh to the primary
electron acceptor, a bacteriopheophytin (BPh). The electron is relayed in the final
stages via menaquinone and a non haem iron protein to the final electron acceptor, a
ubiquinone. The RC consists of two almost symmetrical branches (M and L),
however small conformational changes favour the electron transfer pathway via the
L-branch. The RC is embedded in a cytoplasmic membrane and the large distance
between the redox sites (~70 Å) generates extremely long-lived charge separated
states (t1/2 > 1 s) with a remarkable quantum efficiency ≈ 1. Unwanted backtracking
charge recombination is negligible, as each small step in the process releases
adequate amounts of energy (Figure 2.7.).

16
Chapter 2

SP
M side L side

BChl
BChm

BPhl

BPhm

Fe
ubiquinone menaquinone
Figure 2.7. Schematic representation of the RC from the purple bacterium
Rhodopseudomonas viridis adapted from Deisenhofer et al.32d

Although the SP is suited to absorb light, it is not capable of saturating the maximum
turnover rate for the photosynthetic process. The structure of light-harvesting centre I
(LH-I) of the purple bacterium Rhodopseudomonas palustris reveals that the RC is
the centre of so-called light-harvesting antennae. In this case 32 BChs capture the
sunlight and funnel the energy towards the RC (Figure 2.8.).33

RC

Figure 2.8. Schematic representation of LH-I from the purple bacterium


Rhodopseudomonas palustris adapted from Roszak et al.33

17
Chapter 2

LH-I is usually surrounded by another set of slightly smaller antennae, which are
referred to as the LH-II. In the purple bacterium Rhodopseudomonas acidophila LH-
II consists of 27 BChs which are divided into two orthogonal rings. The inner ring
constitutes 18 BChs while the outer slightly larger ring encompasses nine BChs. LH-
II and LH-I are co-planar and in close proximity which leads to the optimal
organisation for energy transfer (Figure 2.9.).34

Figure 2.9. Schematic representation of LH-I from the purple bacterium


Rhodopseudomonas acidophila adapted from Cherezov et al.34

The overall process for the excitation energy transfer occurs from LH-II9 → LH-II18
→ LH-I → RC and the individual rates are in the order of femto- to picoseconds,
while this energy cascade takes place ~1000 times per second. These structures
undoubtedly confirmed the importance of porphyrinoids in the light-harvesting
process and inspired studies in diverse scientific fields.

In the past two decades synthetic chemists have attempted to mimic the light-
harvesting processes of plants and bacteria by designing various synthetic
multiporphyrin-containing entities. Naturally occurring chromophores are embedded
with weak hydrogen bonds in complex protein matrices. These chromophores have

18
Chapter 2

been challenging to synthesise and were often unstable when isolated from their
natural enviroment.35 Synthetic analogues therefore were constructed using strong
covalent or coordination bonds in order to enhance the efficiency of electron
transport and to stabilise the systems.

2.6. Synthetic multiporphyrin arrays

A detailed explanation of the numerous synthetic strategies developed to isolated


multiporphyrin-containing compounds is beyond the scope of this literature review.
Only an overview of the most recent and popular methods together with a summary
of the major types of assemblies will be presented.

2.6.1. Covalently linked multiporphyrin assemblies

The Osuka group studied extensively the formation and properties of directly linked
porphyrin systems. These oligomers were generated either chemically with Ag(I)
salts36 or electrochemically37 and by choosing DAPs with different central metal
ions, meso-meso (Zn, Mg) and meso-β (Cu, Pd, 2H, Ni) linked dimers were
prepared.38 In subsequent reports the same authors obtained linear arrays with 12839
and 51240 directly meso-meso linked chromophore units. The multiporphyrin systems
were analysed by various spectroscopic techniques41 and by quantum mechanics
calculations.42 The disadvantage of these systems was the orthogonal arrangement of
the chromophores which resulted in a weak ground state interaction between the
porphyrins. However there was still a strong interaction between the chromophores
in the excited state. The meso-meso linkage resulted in a larger interaction between
two macrocycles than the meso-β linked systems (Figure 2.10.).43

Ar

N
N
Ar
M
N N
N M N
N
N
Ar
Ar n

Figure 2.10. Porhyrin oligomers joined by meso-meso linkages.

19
Chapter 2

The same group also prepared doubly meso-β linked porphyrin dimers by treating the
respective monomers with tris(4-bromophenyl)aminium hexachloroantimonate
(BAHA), a typical one-electron oxidising agent.43 Larger electronic interactions
between the two macrocycles, due to the resulting planarity of the porphyrin dyads,
were observed by spectroscopic and electrochemical studies.

By reacting meso-meso linked porphyrin oligomers with two equivalents of BAHA,


the Osuka group also isolated triply fused porphyrin systems.44 The reaction was
performed by treating the respective singly linked dyads with 2,3-dichloro-5,6-
dicyano-1,4-benzoquininone (DDQ) and scandium trifluoromethanesulfonate
[Sc(OTf)3] as oxidising agents and triply fused porphyrin oligomers with up 128
chromophores were prepared.45 Interestingly Zn(II) porphyrins treated with DDQ
and Sc(OTf)3 yielded exclusively triply fused porphyrin dimers while Pd(II)
porphyrins formed only doubly linked meso-β dimers. Cu(II) and Ni(II) porphyrins
resulted in a mixture of the triply and doubly linked dyads.46 The fused
chromophores were again intensively studied,45a,47 revealing a much stronger
interaction between the macrocycles compared to the singly and doubly linked
analogues. Increasing numbers of porphyrins resulted in intensified and red shifted
Q-bands, a splitting of the B-band and a decreasing of the first one-electron oxidation
potential (Figure 2.11.).

Ar Ar

N N N N
M M
N N N N

Ar Ar n

Figure 2.11. Triply fused porphyrin dimers.

A directly fused tetrameric porphyrin sheet (5) is the most recent result in these
series of investigations. By treating a directly meso-meso linked porphyrin tetramer
with DDQ and Sc(OTf)3 the fused macrocycle 5 was prepared in an excellent yield
of 77% (Figure 2.12.).48

20
Chapter 2

Ar Ar
Ar

N N N N
N N Ar Zn Zn Ar
Ar
Zn N N N N
Ar
N N N
N Zn N DDQ
N
Ar Sc(OTf)3
Ar
N N N N N
N Zn N
N N N Ar Zn Zn Ar
Zn Ar N N N N
Ar
N N

Ar Ar
Ar
5

Figure 2.12. Formation of directly fused porphyrin tetramer.

Preliminary spectroscopic studies and quantum mechanics calculations showed large


differences in the electronic properties of 5 compared to singly, doubly and triply
linked porphyrin oligomers.48

Alkene, alkyne and azo linked porphyrins were studied in detail by the Arnold,
Anderson, Therien and other groups. A variety of meso-meso, meso-β and β-β ethene
linked porphyrin dimers were examined in a recent PhD thesis.49 In theory this type
of linkage should result in a large extended conjugation between two porphyrins.
Unfortunately, due to steric hindrance, the alkene bridge was not as efficient a linker
as simple conjugation arguments would predict.49,50

Butadiyne linked porphyrins, such as OEP dimer (6), were prepared first by Arnold
and co-workers.51 Since the initial report a variety of meso-meso, meso-β and β-β
alkyne linked porphyrin dimers were isolated and characterised by the same
group51,52 and by others.13,53,54

21
Chapter 2

N N N N
M M
N N N N

The macrocycles showed strong intramolecular interactions due to the linear and
sterically unhindered conjugated linkage and the meso-meso linked dimers
demonstrated a superior porphyrin π-orbital overlap compared to the β-β linked
dimers. Anderson et al. extended these dimeric systems to form hexameric53d and
polymeric53e-g,55 porphyrin entities, while the Sugiura group prepared a square shaped
porphyrin tetramer56 and a dodecamer (7)57 connected with alkyne bridges.
Ar Ar Ar Ar

N N N N N N N N
Ar Ni Ni Ni Ni Ar
N N N N N N N N

Ar Ar

N N N N
Ar Ni Ar Ar Ni Ar
N N N N

N N N N
Ar Ni Ar Ar Ni Ar
N N N N

Ar Ar

N N N N N N N N
Ar Ni Ni Ni Ni Ar
N N N N N N N N

Ar Ar Ar Ar
7

22
Chapter 2

In 2002 theoretical comparisons between alkene, alkyne and azo linked porphyrins
predicted the last to be the most efficient, due to good orbital matching and less steric
interference between the chromophores.58 Just recently the first examples of
azoporphyrins (8) were prepared by the Arnold group and the preliminary report
demonstrated the superiority of azo linkages in comparison to alkene and alkyne
linkages as stronger interaction between two macrocycles was observed.59
Ar
Ar

N N
N N N M Ph
Ph M N N N
N N

Ar
Ar
8

A variety of other linkages were utilised to form covalently linked porphyrin


oligomers. These included heteroatom, aromatic, saturated, and fused systems as
well as combinations of linkage types mentioned above.60 A covalently linked
porphyrin (Por) dyad attached to a carotenoid (Car) electron donor and two quinone
(Q) electron acceptor units (9) was designed to mimic the RC. 22,61 As in the natural
RC, a series of short range electron transfer steps limits the reverse charge
recombination processes, hence increasing the lifetime of the separated state. Thus
this kind of system is overall more efficient than one long-range electron transfer.
Upon excitation of the Zn porphyrin unit in 9 (Car-ZnPor*-H2Por-Q-Q), multiple fast
and efficient electron transfer steps occurred to yield the charge separated entity
(Car•+-ZnPor-H2Por-Q-Q•-). Although this system had an impressive quantum yield
of 83%, the lifetime of the charge-separated state was found to be only 50 μs.

O O
O O NH N
N N
N O
N Zn N H
N HN
H H O
N N

Car Por 9 Por Q Q

23
Chapter 2

Porphyrin wheels containing twelve62 or twenty-four (10)63 porphyrin subunits


respectively with a mixture of direct and phenyl linkages showed promising
properties to mimic LH-I. The excitation energy transfer rate between chromophores
in these systems of ~35 ps-1 is similar to rates encountered in Nature.
Ar Ar Ar Ar

N N N N N N N N
Zn Zn Zn Zn
Ar Ar
N N N N N N N N
N N

N Zn N N Zn N
Ar Ar Ar Ar
N N
Ar Ar
N Ar Ar N

N Zn N N Zn N

Ar N N Ar
N Ar Ar N

N Zn N N Zn N

N N
Ar Ar
N Ar Ar N

N Zn N N Zn N

N N
Ar Ar

Ar Ar
N N

N Zn N N Zn N

N N
Ar Ar
Ar Ar
N N

N Zn N N Zn N

N Ar Ar N
Ar N N Ar

N Zn N N Zn N

N N
Ar Ar
Ar Ar
N N
Ar Ar Ar Ar
N Zn N N Zn N

N N
N N N N N N N N
Ar Ar
Zn Zn Zn Zn
N N N N N N N N

Ar Ar Ar Ar

10

Covalently linked porphyrin oligomers have afforded extensive insights into electron
transfer mechanisms of the photosynthetic process and promising examples to mimic
these processes have also been developed in recent years. Unfortunately their
formation often included challenging and low-yielding multistep synthetic strategies,
thus limiting their practical applications.60 Hence other strategies were developed, as
described below.

2.6.2. Multiporphyrin assemblies linked by axial coordination

Coordination to the central metal ion of porphyrins to form porphyrin oligomers and
polymers has attracted considerable attention in recent years. By choosing the
appropriate peripheral functionality, the macrocycles act as ligands and coordinate to
other metalloporphyrins (Figure 2.13. A-D).64 If an appropriate metal is inserted into
the ligands, they can self-assemble to dimers, cyclic oligomers or polymers (Figure
2.13. E-G).65

24
Chapter 2

M' M' M'


L L L

M M M L M' M' L M L M' M

L L L L
M' M' M' M'

A B C D N N
M
M L M M L
N N
M L
M L
L L
L M
M L
M L M n

E F G

Figure 2.13. Complexation types of porphyrins which act as ligands.

Oligomers derived from such complexation schemes have the distinct advantage that
the only challenging synthetic step is the attachment of the ligand to the macrocycle.
The rich coordination chemistry of central metal ions and the large number of
possible organic ligands lead to a vast array of potential supramolecular systems. It is
however important that the ligand binds strongly to the metal centre and is in close
proximity to the macrocycle in order to ensure the presence of only a single
dominant product. Flexible systems result in a concentration-dependent mixture of
monomers and different oligomers.18e In order to achieve a π-orbital overlap of the
chromophores and therefore strong electronic communication, a cofacial geometry
between two porphyrins with a high degree of macrocycle overlap and a short inter-
porphyrin distance is preferred. This leads to systems with similar geometries of the
SP in the photosynthetic RC.66

In initial reports pyridine was often the ligand of choice. N-donor ligands are known
to bind strongly to a variety of metal centres and they can be easily incorporated into
the porphyrin framework by utilising pyridinecarboxaldehyde during the initial
porphyrin condensation reaction.67 Fleisher and Shachter reported the first such
porphyrin dimers (Type A, B and C) with ZnTPP and 5-pyridyl-, 5,10-dipyridyl and
5,15-dipyridylporphyrins respectively.68 By inserting Zn(II) into 5-pyridyl-10,15,20-

25
Chapter 2

triphenylporphyrin and recrystallisation from chloroform/methanol, a linear


coordination polymer (11) was obtained.68

Ph Ph
Ph
N
N Zn N
N Ph N
Ph N
N
Zn
N
N Ph

N
Ph Ph
N
N Zn N
N N
Ph N
N
Zn
N
N Ph

n
11

The Osuka group isolated a self-assembled porphyrin box (12) by inserting Zn(II)
into a 5-pyridyl-10-arylporphyrin with benzene as solvent and thus demonstrated the
importance of the choice of solvent in such reactions.69

Ar
N

N Zn N
Ar =
N
N

Ar N
N N
Zn
N N
N N
Zn
N N
N Ar

N
N
N Zn N
12
N

Ar

26
Chapter 2

In order to achieve a cofacial geometry, 2-pyridyl instead of 4-pyridyl units were


incorporated as ligands. Stibrany prepared the first Type E Zn(II)porphyrin dimer
utilising 2-pyridyl ligands,70 while an American group isolated one of the very few
examples of Mg(II) porphyrin oligomers (13). In this case the authors were not only
concerned about structural properties, but investigated also the aggregation chemistry
of the oligomers and only a weak binding of the nitrogen ligand with the Mg(II)
fragment was observed.71
R R

N
N Mg N
N

R N N R

N
N Mg N
R = Me, nPr or nBu N

R R

13

Kobuke and co-workers investigated the binding strength of various nitrogen ligands
to Zn(II) porphyrins in a variety of solvents. N-Methylimidazole had an order of
magnitude larger association constant than pyridine. The authors proposed a
combination of the electron donating ability of the ligand and the bond angle
between ligand and metal to be the rationale for this finding.67 Zn(II) (14) and Mg(II)
(15) OAP dimers of Type E with two imidazole ligands on each macrocycle were
prepared by the same authors and their structural properties compared.72

N N
N M N N N
N N N M N
N N N N
N N
N N
N N N N
N M N N N
N N N M N
N N

cis 14: M = Zn trans 14: M = Zn


cis 15: M = Mg trans 15: M = Mg

Two geometrical isomers were isolated, depending on the orientation of the N-


methyl groups. In the case of trans 15, a Type G polymer could also be obtained at
higher concentrations, while both isomers of 14 were observed only as dimers. Again
the Mg(II) porphyrin (15) coordinated weakly to the nitrogen ligand while a very

27
Chapter 2

large self association constant for 14 was estimated. UV-visible, fluorescence and
NMR spectra also indicated large porphyrin-porphyrin interactions.

A similar Zn(II) porphyrin dimer attached to a pyromellitdiimide electron acceptor


(16) was prepared by the same authors to mimic the SP of the RC. Detailed
electrochemical studies revealed that the radical cation was delocalised over the
whole dyad. Photophysical investigations on 16 in comparison to the monomer units
showed that the charge-separated lifetime in dimer 16 rises from 100 to 160 ps, the
charge recombination rate decreases from 1000 to 120 ps-1 and the fluorescence is
quenched by 95%. These facts showed for the first time the significance of the SP
arrangement in the RC and demonstrated the vital role of the SP in the initial charge
separation of the photosynthetic process.73
C7H15

N N N O O O
Zn
N N N NH N N OMe
C7H15 C H
7 15 O
O O
O O O N N N
Zn
HN N N N
16
MeO N N

O C7H15
O O

Examples of porphyrinyl coordination frameworks incorporating other metal centres


such as Mn(III), Fe(III), Co(III), Ru(II), Rh(III), Os (II), Ga (III), In (III), Sn (IV)
and P(V) porphyrins have also been reported, but interestingly hardly any non-
nitrogen containing ligands have been exploited for this type of supramolecular
chemistry.64,74 Multiporphyrin arrays in which the metal fragments are directly
attached to the carbon skeleton of the macrocycles will be discussed in Chapter 3.

Porphyrin oligomers have also attracted the attention of scientists concerned with
topics other than photosynthesis. Supramolecular arrays containing porphyrin units
have been successfully applied in related areas of research such as fluorescence
sensors,75 conducting polymers,76 catalysts77 and as reagents for photodynamic
therapy,78 however it is beyond the scope of this review to discuss these applications.

2.7 References

(1) Milgrom, L. R. The colours of life: an introduction to the chemistry of


porphyrins and related compounds; Oxford University Press: Oxford, 1997.

28
Chapter 2

(2) Bonnett, R. In The Porphyrins; Dolphin, D., Ed.; Academic Press: New York,
1978; Vol. 1, p 1-27.
(3) Buchler, J. W. In The Porphyrins; Dolphin, D., Ed.; Academic Press: New
York, 1978; Vol. 1, p 389-483.
(4) Dixon, M.; Webb, E. C. Enzymes; 3rd ed.; Longman: London, 1979.
(5) Stryer, L. Biochemistry; 4th ed.; W.H. Freeman: New York, 1995.
(6) (a) Fischer, H.; Orth, H. Die Chemie des Pyrrols; Akademische
Verlagsgesellschaft: Leipzig, 1934; Vol. 1. (b) Fischer, H.; Orth, H. Die
Chemie des Pyrrols; Akademische Verlagsgemeinschaft: Leipzig, 1937; Vol.
2A. (c) Fischer, H.; Stern, A. Die Chemie des Pyrrols; Akademische
Verlagsgemeinschaft: Leipzig, 1940; Vol. 2B.
(7) Porphyrins and Metalloporphyrins; Smith, K. M. Ed.; Elsevier Scientific
Publishing Company: Amsterdam, 1975.
(8) (a) The Porphyrins; Dolphin, D. Ed.; Academic Press Inc.: New York, 1978;
Vol. 1-5. (b) The Porphyrins; Dolphin, D. Ed.; Academic Press Inc.: New
York, 1979; Vol. 6-7.
(9) (a) The Porphyrin Handbook; Kadish, K. M., Smith, K. M., Guilard, R. Eds.;
Academic Press: New York, 2000; Vol. 1-10. (b) The Porphyrin Handbook;
Kadish, K. M., Smith, K. M., Guilard, R. Eds.; Academic Press: New York,
2004; Vol. 11-20.
(10) (a) Gouterman, M. J. Mol. Spectrosc. 1961, 6, 138-63. (b) Gouterman, M.;
Wagniere, G.; Snyder, L. C. J. Mol. Spectrosc. 1963, 11, 108-27. (c) Weiss,
C.; Kobayashi, H.; Gouterman, M. J. Mol. Spectrosc. 1965, 16, 415-50. (d)
Gouterman, M. In The Porphyrins; Dolphin, D., Ed.; Academic Press: New
York, 1978; Vol. 3, p 1-165.
(11) Gouterman, M. J. Chem. Phys. 1959, 30, 1139-61.
(12) Dorough, G. D.; Miller, J. R.; Huennekens, F. M. J. Am. Chem. Soc. 1951,
73, 4315-20.
(13) Anderson, H. L. J. Chem. Soc., Chem. Commun. 1999, 2323-2330.
(14) Michl, J. In Handbook of Photochemistry; 3rd ed.; Montalti, M., Credi, A.,
Prodi, L., Gandolfi, M. T., Eds.; CRC: Boca Raton, 2006, p 1-43.
(15) (a) Becker, R. S.; Allison, J. B. J. Phys. Chem. 1963, 67, 2662-9. (b) Bedford,
R. B.; Cazin, C. S. J.; Holder, D. Coord. Chem. Rev. 2004, 248, 2283-2321.
(c) Allison, J. B.; Becker, R. S. J. Phys. Chem. 1963, 67, 2675-9.

29
Chapter 2

(16) Hopf, F. R.; Whitten, D. G. In Porphyrins and Metalloporphyrins; Smith, K.


M., Ed.; Elsevier Scientific Publishing: Amsterdam, 1975, p 667-700.
(17) Gouterman, M.; Khalil, G. E. J. Mol. Spectrosc. 1974, 53, 88-100.
(18) (a) Sanders, J. K. M.; Bampos, N.; Clyde-Watson, Z.; Darling, S. L.; Hawley,
J. C.; Kim, H.-J.; Mak, C. C.; Webb, S. J. In The Porphyrin Handbook;
Kadish, K. M., Smith, K. M., Guilard, R., Eds.; Academic Press: San Diego,
2000; Vol. 3, p 1-48. (b) Barbe, J.-M.; Guilard, R. In The Porphyrin
Handbook; Kadish, K. M., Smith, K. M., Guilard, R., Eds.; Academic Press:
San Diego, 2000; Vol. 3, p 211-244. (c) Buchler, J. W.; Ng, D. K. P. In The
Porphyrin Handbook; Kadish, K. M., Smith, K. M., Guilard, R., Eds.;
Academic Press: San Diego, 2000; Vol. 3, p 245-294. (d) Guilard, R.; Tabard,
A.; Van Caemelbecke, E.; Kadish, K. M. In The Porphyrin Handbook;
Kadish, K. M., Smith, K. M., Guilard, R., Eds.; Academic Press: San Diego,
2000; Vol. 3, p 295-345. (e) Sanders, J. K. M. In The Porphyrin Handbook;
Kadish, K. M., Smith, K. M., Guilard, R., Eds.; Academic Press: San Diego,
2000; Vol. 3, p 347-368. (f) Aida, T.; Jiang, D.-L. In The Porphyrin
Handbook; Kadish, K. M., Smith, K. M., Guilard, R., Eds.; Academic: San
Diego, 2000; Vol. 3, p 369-384.
(19) Smith, K. M. In The Porphyrin Handbook; Kadish, K. M., Smith, K. M.,
Guilard, R., Eds.; Academic Press: San Diego, 2000; Vol. 1, p 119-148.
(20) (a) Kim, J. B.; Adler, A. D.; Longo, F. R. In The Porphyrins; Dolphin, D.,
Ed.; Academic Press Inc.: New York, 1978; Vol. 1, p 85-100. (b) Hatscher,
S.; Senge, M. O. Tetrahedron Lett. 2002, 44, 157-160.
(21) (a) Lindsey, J. S. In Catalysis by Metal Complexes 1994; Vol. 17, p 49-86.
(b) Lindsey, J. S. In The Porphyrin Handbook; Kadish, K. M., Smith, K. M.,
Guilard, R., Eds.; Academic Press: San Diego, 2000; Vol. 1, p 45-118.
(22) Gust, D.; Moore, T. A. In The Porphyrin Handbook; Kadish, K. M., Smith,
K. M., Guilard, R., Eds.; Academic Press: San Diego, 2000; Vol. 8, p 153-
190.
(23) Vicente, M. G. H. In The Porphyrin Handbook; Kadish, K. M., Smith, K. M.,
Guilard, R., Eds.; Academic Press: San Diego, 2000; Vol. 1, p 149-199.
(24) Jaquinod, L. In The Porphyrin Handbook; Kadish, K. M., Smith, K. M.,
Guilard, R., Eds.; Academic Press: San Diego, 2000; Vol. 1, p 201-237.

30
Chapter 2

(25) Laha, J. K.; Dhanalekshmi, S.; Taniguchi, M.; Ambroise, A.; Lindsey, J. S.
Org. Process Res. Dev. 2003, 7, 799-812.
(26) Dolphin, D.; Muljiani, Z.; Rousseau, K.; Borg, D. C.; Fajer, J.; Felton, R. H.
Ann. N. Y. Acad. Sci. 1973, 206, 177-200.
(27) Smith, K. M.; Barnett, G. H.; Evans, B.; Martynenko, Z. J. Am. Chem. Soc.
1979, 101, 5953-61.
(28) Kalisch, W. W.; Senge, M. O. Angew. Chem., Int. Ed. 1998, 37, 1107-1109.
(29) (a) Senge, M. O.; Bischoff, I. Tetrahedron Lett. 2004, 45, 1647-1650. (b)
Senge, M. O. Acc. Chem. Res. 2005, 38, 733-743.
(30) (a) Feng, X.; Senge, M. O. Tetrahedron 2000, 56, 587-590. (b) Senge, M. O.;
Feng, X. J. Chem. Soc., Perkin Trans. 1 2000, 3615-3621. (c) Feng, X.;
Senge, M. O. J. Chem. Soc., Perkin Trans. 1 2001, 1030-1038.
(31) Deisenhofer, J.; Epp, O.; Miki, K.; Huber, R.; Michel, H. J. Mol. Biol. 1984,
180, 385-98.
(32) (a) Deisenhofer, J.; Michel, H. Angew. Chem., Int. Ed. 1989, 28, 829-841. (b)
Deisenhofer, J.; Michel, H. Science 1989, 245, 1463-73. (c) Deisenhofer, J.;
Michel, H. EMBO J. 1989, 8, 2149-70. (d) Deisenhofer, J.; Epp, O.; Sinning,
I.; Michel, H. J. Mol. Biol. 1995, 246, 429-57. (e) Huber, R. Angew. Chem.,
Int. Ed. 1989, 28, 848-69.
(33) Roszak, A. W.; Howard, T. D.; Southall, J.; Gardiner, A. T.; Law, C. J.;
Isaacs, N. W.; Cogdell, R. J. Science 2003, 302, 1969-1972.
(34) Cherezov, V.; Clogston, J.; Papiz, M. Z.; Caffrey, M. J. Mol. Biol. 2006, 357,
1605-1618.
(35) Smith, K. M. In Porphyrins and Metalloporphyrins; Smith, K. M., Ed.;
Elsevier Scientific Publishing Company: Amsterdam, 1975, p 29-60.
(36) Osuka, A.; Shimidzu, H. Angew. Chem., Int. Ed. Eng. 1997, 36, 135-137.
(37) Ogawa, T.; Nishimoto, Y.; Yoshida, N.; Ono, N.; Osuka, A. J. Chem. Soc.,
Chem. Commun. 1998, 337-338.
(38) Ogawa, T.; Nishimoto, Y.; Yoshida, N.; Ono, N.; Osuka, A. Angew. Chem.,
Int. Ed. 1999, 38, 176-179.
(39) Aratani, N.; Osuka, A.; Kim, Y. H.; Jeong, D. H.; Kim, D. Angew. Chem.,
Int. Ed. 2000, 39, 1458-1462.
(40) Aratani, N.; Takagi, A.; Yanagawa, Y.; Matsumoto, T.; Kawai, T.; Yoon, Z.
S.; Kim, D.; Osuka, A. Chem.--Eur. J. 2005, 11, 3389-3404.

31
Chapter 2

(41) (a) Ohta, N.; Iwaki, Y.; Ito, T.; Yamazaki, I.; Osuka, A. J. Phys. Chem. B
1999, 103, 11242-11245. (b) Bhuiyan, A. A.; Seth, J.; Yoshida, N.; Osuka,
A.; Bocian, D. F. J. Phys. Chem. B 2000, 104, 10757-10764. (c) Yoshida, N.;
Osuka, A. Tetrahedron Lett. 2000, 41, 9287-9291. (d) Kim, Y. H.; Jeong, D.
H.; Kim, D.; Jeoung, S. C.; Cho, H. S.; Kim, S. K.; Aratani, N.; Osuka, A. J.
Am. Chem. Soc. 2001, 123, 76-86. (e) Jeong, D. H.; Yoon, M.-C.; Jang, S.
M.; Kim, D.; Cho, D. W.; Yoshida, N.; Aratani, N.; Osuka, A. J. Phys. Chem.
A 2002, 106, 2359-2368. (f) Aratani, N.; Cho, H. S.; Ahn, T. K.; Cho, S.;
Kim, D.; Sumi, H.; Osuka, A. J. Am. Chem. Soc. 2003, 125, 9668-9681.
(42) Matsuzaki, Y.; Nogami, A.; Iwaki, Y.; Ohta, N.; Yoshida, N.; Aratani, N.;
Osuka, A.; Tanaka, K. J. Phys. Chem. A 2005, 109, 703-713.
(43) Tsuda, A.; Nakano, A.; Furuta, H.; Yamochi, H.; Osuka, A. Angew. Chem.,
Int. Ed. 2000, 39, 558-561.
(44) (a) Tsuda, A.; Furuta, H.; Osuka, A. Angew. Chem., Int. Ed. 2000, 39, 2549-
2552. (b) Tsuda, A.; Furuta, H.; Osuka, A. J. Am. Chem. Soc. 2001, 123,
10304-10321.
(45) (a) Tsuda, A.; Osuka, A. Science 2001, 293, 79-82. (b) Tsuda, A.; Osuka, A.
J. Incl. Phenom. Macrocycl. Chem. 2001, 41, 77-81. (c) Hiroto, S.; Osuka, A.
J. Org. Chem. 2005, 70, 4054-4058.
(46) Kamo, M.; Tsuda, A.; Nakamura, Y.; Aratani, N.; Furukawa, K.; Kato, T.;
Osuka, A. Org. Lett. 2003, 5, 2079-2082.
(47) (a) Cho, H. S.; Jeong, D. H.; Cho, S.; Kim, D.; Matsuzaki, Y.; Tanaka, K.;
Tsuda, A.; Osuka, A. J. Am. Chem. Soc. 2002, 124, 14642-14654. (b) Ikeue,
T.; Furukawa, K.; Hata, H.; Aratani, N.; Shinokubo, H.; Kato, T.; Osuka, A.
Angew. Chem., Int. Ed. 2005, 44, 6899-6901.
(48) Nakamura, Y.; Aratani, N.; Shinokubo, H.; Takagi, A.; Kawai, T.;
Matsumoto, T.; Yoon, Z. S.; Kim, D. Y.; Ahn, T. K.; Kim, D.; Muranaka, A.;
Kobayashi, N.; Osuka, A. J. Am. Chem. Soc. 2006, 128, 4119-4127.
(49) Locos, O. B. PhD thesis, Queensland University of Technology, 2006.
(50) Locos, O. B.; Arnold, D. P. Org. Biomol. Chem. 2006, 4, 902-916. (b)
Frampton, M. J.; Akdas, H.; Cowley, A. R.; Rogers, J. E.; Slagle, J. E.; Fleitz,
P. A.; Drobizhev, M.; Rebane, A.; Anderson, H. L. Org. Lett. 2005, 7, 5365-
5368.

32
Chapter 2

(51) Arnold, D. P.; Johnson, A. W.; Mahendran, M. J. Chem. Soc., Perkin Trans.
1 1978, 366-70.
(52) (a) Arnold, D. P.; Nitschinsk, L. J. Tetrahedron 1992, 48, 8781-92. (b)
Arnold, D. P.; Nitschinsk, L. J. Tetrahedron Lett. 1993, 34, 693-6. (c)
Arnold, D. P.; Heath, G. A. J. Am. Chem. Soc. 1993, 115, 12197-8. (d)
Arnold, D. P.; James, D. A.; Kennard, C. H. L.; Smith, G. J. Chem. Soc.,
Chem. Commun. 1994, 2131-2. (e) Arnold, D. P.; Manno, D.; Micocci, G.;
Serra, A.; Tepore, A.; Valli, L. Langmuir 1997, 13, 5951-5956. (f) Arnold, D.
P.; James, D. A. J. Org. Chem. 1997, 62, 3460-3469. (g) Arnold, D. P.;
Heath, G. A.; James, D. A. New J. Chem. 1998, 22, 1377-1387. (h) Arnold,
D. P.; Manno, D.; Micocci, G.; Serra, A.; Tepore, A.; Valli, L. Thin Solid
Films 1998, 327-329, 341-344. (i) Arnold, D. P.; Heath, G. A.; James, D. A.
J. Porphyrins Phthalocyanines 1999, 3, 5-31. (j) Arnold, D. P. Synlett 2000,
296-305. (k) Arnold, D. P.; Hartnell, R. D. Tetrahedron 2001, 57, 1335-1345.
(l) Arnold, D. P.; Hartnell, R. D.; Heath, G. A.; Newby, L.; Webster, R. D. J.
Chem. Soc., Chem. Commun. 2002, 754-755. (m) Arnold, D. P.; Genga, A.;
Manno, D.; Micocci, G.; Serra, A.; Tepore, A.; Valli, L. Colloids Surf., A
2002, 198-200, 897-904.
(53) (a) Anderson, H. L. Inorg. Chem. 1994, 33, 972-81. (b) Anderson, H. L.;
Anderson, S.; Sanders, J. K. M. J. Chem. Soc., Perkin Trans. 1 1995, 2231-
45. (c) Wilson, G. S.; Anderson, H. L. Synlett 1996, 1039-1040. (d) Taylor, P.
N.; Huuskonen, J.; Aplin, R. T.; Anderson, H. L.; Rumbles, G.; Williams, E.
Chem. Commun. 1998, 909-910. (e) Taylor, P. N.; Anderson, H. L. J. Am.
Chem. Soc. 1999, 121, 11538-11545. (f) Screen, T. E. O.; Thorne, J. R. G.;
Denning, R. G.; Bucknall, D. G.; Anderson, H. L. J. Am. Chem. Soc. 2002,
124, 9712-3. (g) Drobizhev, M.; Stepanenko, Y.; Rebane, A.; Wilson, C. J.;
Screen, T. E. O.; Anderson, H. L. J. Am. Chem. Soc. 2006, 128, 12432-
12433.
(54) (a) DiMagno, S. G.; Lin, V. S. Y.; Therien, M. J. J. Org. Chem. 1993, 58,
5983-93. (b) Therien, M. J.; DiMagno, S. G. In International Patent;
(University of Pennsylvania, USA). USA, 1994, p 62 pp. (c) Lin, V. S. Y.;
DiMagno, S. G.; Therien, M. J. Science 1994, 264, 1105-11. (d) Shediac, R.;
Gray, M. H. B.; Uyeda, H. T.; Johnson, R. C.; Hupp, J. T.; Angiolillo, P. J.;
Therien, M. J. J. Am. Chem. Soc. 2000, 122, 7017-7033. (e) Therien, M. J. In

33
Chapter 2

International Patent; (The Trustees of the University of Pennsylvania, USA).


USA, 2002, p 76. (f) Susumu, K.; Duncan, T. V.; Therien, M. J. J. Am. Chem.
Soc. 2005, 127, 5186-5195.
(55) Screen, T. E. O.; Thorne, J. R. G.; Denning, R. G.; Bucknall, D. G.;
Anderson, H. L. J. Mater. Chem. 2003, 13, 2796-2808.
(56) Sugiura, K.-i.; Fujimoto, Y.; Sakata, Y. Chem. Commun. 2000, 1105-1106.
(57) Kato, A.; Sugiura, K.-i.; Miyasaka, H.; Tanaka, H.; Kawai, T.; Sugimoto, M.;
Yamashita, M. Chem. Lett. 2004, 33, 578-579.
(58) Screen, T. E. O.; Blake, I. M.; Rees, L. H.; Clegg, W.; Borwick, S. J.;
Anderson, H. L. J. Chem. Soc., Perkin Trans. 1 2002, 320-329.
(59) Esdaile, L. J.; Jensen, P.; McMurtrie, J. C.; Arnold, D. P. Angew. Chem., Int.
Ed. 2007, 46, 2090-2093.
(60) Burrell, A. K.; Officer, D. L.; Plieger, P. G.; Reid, D. C. W. Chem. Rev. 2001,
101, 2751-2796.
(61) (a) Gust, D.; Moore, T. A.; Moore, A. L.; Lee, S. J.; Bittersmann, E.; Luttrull,
D. K.; Rehms, A. A.; DeGraziano, J. M.; Ma, X. C.; et al. Science 1990, 248,
199-201. (b) Gust, D.; Moore, T. A.; Moore, A. L.; Macpherson, A. N.;
Lopez, A.; DeGraziano, J. M.; Gouni, I.; Bittersmann, E.; Seely, G. R.; et al.
J. Am. Chem. Soc. 1993, 115, 11141-52. (c) Gust, D.; Moore, T. A.; Moore,
A. L. Acc. Chem. Res. 2001, 34, 40-48. (d) Palacios, R. E.; Kodis, G.; Gould,
S. L.; de la Garza, L.; Brune, A.; Gust, D.; Moore, T. A.; Moore, A. L.
ChemPhysChem 2005, 6, 2359-2370.
(62) (a) Peng, X.; Aratani, N.; Takagi, A.; Matsumoto, T.; Kawai, T.; Hwang, I.-
W.; Ahn, T. K.; Kim, D.; Osuka, A. J. Am. Chem. Soc. 2004, 126, 4468-
4469. (b) Hwang, I.-W.; Ko, D. M.; Ahn, T. K.; Yoon, Z. S.; Kim, D.; Peng,
X.; Aratani, N.; Osuka, A. J. Phys. Chem. B 2005, 109, 8643-8651.
(63) (a) Hori, T.; Aratani, N.; Takagi, A.; Matsumoto, T.; Kawai, T.; Yoon, M.-C.;
Yoon, Z. S.; Cho, S.; Kim, D.; Osuka, A. Chem.--Eur. J. 2006, 12, 1319-
1327. (b) Park, M.; Yoon, M.-C.; Yoon, Z. S.; Hori, T.; Peng, X.; Aratani, N.;
Hotta, J.-i.; Uji-i, H.; Sliwa, M.; Hofkens, J.; Osuka, A.; Kim, D. J. Am.
Chem. Soc. 2007, 129, 3539-3544. (c) Hori, T.; Nakamura, Y.; Aratani, N.;
Osuka, A. J. Organomet. Chem. 2007, 692, 148-155.
(64) Bouamaied, I.; Coskun, T.; Stulz, E. Struct. and Bonding 2006, 121, 1-47.
(65) Kobuke, Y. Struct. and Bonding 2006, 121, 49-104.

34
Chapter 2

(66) Kobuke, Y. Eur. J. Inorg. Chem. 2006, 2333-2351.


(67) Satake, A.; Kobuke, Y. Tetrahedron 2005, 61, 13-41.
(68) Fleischer, E. B.; Shachter, A. M. Inorg. Chem. 1991, 30, 3763-9.
(69) Tsuda, A.; Nakamura, T.; Sakamoto, S.; Yamaguchi, K.; Osuka, A. Angew.
Chem., Int. Ed. 2002, 41, 2817-2821.
(70) Stibrany, R. T.; Vasudevan, J.; Knapp, S.; Potenza, J. A.; Emge, T.; Schugar,
H. J. J. Am. Chem. Soc. 1996, 118, 3980-1.
(71) Gerasimchuk, N. N.; Mokhir, A. A.; Rodgers, K. R. Inorg. Chem. 1998, 37,
5641-5650.
(72) (a) Kobuke, Y.; Miyaji, H. J. Am. Chem. Soc. 1994, 116, 4111-12. (b)
Kobuke, Y.; Miyaji, H. Bull. Chem. Soc. Jpn. 1996, 69, 3563-3569.
(73) Ozeki, H.; Nomoto, A.; Ogawa, K.; Kobuke, Y.; Murakami, M.; Hosoda, K.;
Ohtani, M.; Nakashima, S.; Miyasaka, H.; Okada, T. Chem.--Eur. J. 2004,
10, 6393-6401.
(74) (a) Wojaczyński, J.; Latos- Grażyński, L. Coord. Chem. Rev. 2000, 204, 113-
171. (b) Imamura, T.; Fukushima, K. Coord. Chem. Rev. 2000, 198, 133-156.
(75) De Silva, A. P.; Gunaratne, H. Q. N.; Gunnlaugsson, T.; Huxley, A. J. M.;
McCoy, C. P.; Rademacher, J. T.; Rice, T. E. Chem. Rev. 1997, 97, 1515-
1566.
(76) (a) Malinski, T. In The Porphyrin Handbook; Kadish, K. M., Smith, K. M.,
Guilard, R., Eds.; Academic Press: San Diego, 2000; Vol. 6, p 231-256. (b)
Chou, J.-h.; Kosal, M. E.; Nalwa, H. S.; Rakow, N. A.; Suslick, K. S. In The
Porphyrin Handbook; Kadish, K. M., Smith, K. M., Guilard, R., Eds.;
Academic Press: San Diego, 2000; Vol. 6, p 43-131.
(77) Aida, T.; Inoue, S. In The Porphyrin Handbook; Kadish, K. M., Smith, K.
M., Guilard, R., Eds.; Academic Press: San Diego, 2000; Vol. 6, p 133-156.
(78) (a) Silva, J. N.; Filipe, P.; Morliere, P.; Maziere, J.-C.; Freitas, J. P.; Cirne De
Castro, J. L.; Santus, R. Bio-Med. Mater. Eng. 2006, 16, S147-S154. (b)
Berg, K.; Selbo, P. K.; Weyergang, A.; Dietze, A.; Prasmickaite, L.; Bonsted,
A.; Engesaeter, B. O.; Angell-Petersen, E.; Warloe, T.; Frandsen, N.; Hogset,
A. J. Microsc. 2005, 218, 133-47. (c) Vicente, M. G. Curr. Med. Chem.
Anticancer Agents 2001, 1, 175-94. (d) Pandey, R. K.; Zheng, G. In The
Porphyrin Handbook; Kadish, K. M., Smith, K. M., Guilard, R., Eds.;
Academic Press: San Diego, 2000; Vol. 6, p 157-230.

35
CHAPTER 3

Porphyrins with metal, metalloid or


phosphorus atoms directly bonded to the
carbon periphery

36
Chapter 3

Statement of Contribution

The authors listed below have certified* that:


1) they meet the criteria for authorship in that they have participated in the
conception, execution, or interpretation, of at least that part of the publication in
their field of expertise;
2) they take public responsibility for their part of the publication, except for the
responsible author who accepts overall responsibility for the publication;
3) there are no other authors of the publication according to these criteria;
4) potential conflicts of interest have been disclosed to (a) granting bodies, (b) the
editor or publisher of journals or other publications, and (c) the head of the
responsible academic unit, and
5) they agree to the use of the publication in the student’s thesis and its publication
on the Australasian Digital Thesis database consistent with any limitations set by
publisher requirements.
In the case of this chapter:
Porphyrins with metal, metalloid or phosphorus atoms directly bonded to the
carbon periphery
Atefi F and Arnold DP. J. Porphyrins Phthalocyanines 2007, accepted for
publication
Contributor Statement of contribution*

Wrote manuscript

Dennis P. Arnold* Overall supervisor of the project, edited manuscript

Principal Supervisor Confirmation

I have sighted email or other correspondence from all Co-authors confirming their
certifying authorship and.

37
Chapter 3

3. Porphyrins with metal, metalloid or phosphorus atoms directly


bonded to the carbon periphery

Farzad Atefi and Dennis P. Arnold*

Synthesis and Molecular Recognition Program, School of Physical and Chemical


Sciences, Queensland University of Technology, G.P.O. Box 2434, Brisbane 4001,
Australia

*Correspondence to: Dennis P. Arnold, email: d.arnold@qut.edu.au, fax: +61 7 3138


1804

ABSTRACT: Organometallic porphyrins with a metal, metalloid or phosphorus


fragment directly attached to their carbon framework emerged for the first time in
1976, and these macrocycles have been intensively investigated in the past decade.
The present review summarises for the first time all reported examples as well as
applications of these systems.

KEYWORDS: Porphyrinoids, organometallic, mercury, palladium, platinum, boron,


phosphorus

38
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

INTRODUCTION
Porphyrins coordinate the majority of metals within the core to form
equatorial metal complexes. Traditionally the organometallic bond in porphyrin
chemistry is associated with this metal fragment [1] and in a review by Brothers a
classification system was used for these organometallic chromophores (Fig. 1) [1a].
R R R R MLn L'MLn MLn

N N N N N N N N N N
M M M M M Ar X MLn
N N N N N N N N N N

A B C D E

LnM'' N N N N
M M
N N N N N N
M M'Ln
N N
M'Ln M'Ln

F G H

Figure. 1. Classification of organometallic porphyrins.

According to this classification five types of organometallic porphyrins can be


discerned, namely complexes containing one or two metal-carbon single or multiple
bonds (Class A), complexes containing π-ligands (Class B), complexes containing
bridging μ-M,N ligands (Class C), complexes containing a metal-metal bond, where
the metal-carbon bond is not associated with the metal coordinated to inner N-
protons (Class D) and porphyrins in which the metal-carbon bond is associated with
a substituent attached to the porphyrin ring (Class E). The present review discusses
organometallic porphyrins in which the metal is either directly σ- (Class F) or π-
bonded (Class G) to the carbon framework of the porphyrin macrocycle. In the
previous reviews Class G porphyrins were included in Class E. However the metal-
carbon bond in a Class G porphyrin is associated with the carbon framework of the
porphyrin, and not with any substituents. Hence we suggest such π-bonded
organometallic porphyrins should be classified on their own. There are also a very
few examples of organometallic porphyrins with metal ions π-bonded to fused
exocyclic rings, now labelled Class H. The present review is restricted to examples
of Classes F, G and H. Although the first examples of Class F organometallic
porphyrins have been known for almost 30 years, the majority of reports discussed
here stem from the past decade. For completeness we have also included elements
such as B, P and Te in this review, but omitted ferrocenylporphyrins, as these

39
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

compounds are really examples of Class E porphyrins and the metal-carbon bond is
neither directly associated with the carbon framework of the porphyrin nor with
fused rings [2]. Core-modified porphyrins with at least one C atom in the core, such
as N-confused porphyrins and carbaporphyrins form various organometallic
derivatives with equatorially coordinated metal ions. These are also not considered in
this review, as they represent a different class of macrocycles altogether and should
be discussed on their own [3]. Readers concerned with ferrocenyl or core-modified
organometallic porphyrins should refer to some recent reports in this area. As our
emphasis is on structures and uses of the porphyrin derivatives, only brief mention is
made of catalysts, ligands, bases, solvents and conditions; readers are referred to the
original works for details.

CLASS F ORGANOMETALLIC PORPHYRINS


Mercury

The first example of a mercury fragment directly σ-bonded to the carbon


periphery of porphyrins was reported in 1980. Smith and co-workers prepared the
bis-chloromercurio derivative of Zn(II) deuteroporphyrin IX dimethyl ester (7) by
treating the parent porphyrin (1) with mercury(II) acetate in acetonitrile and
exchanging the anion using sodium chloride [4a]. The same authors successfully
mercurated (11-18) various other deuteroporphyrin IX analogues (2-6) under similar
reaction conditions [4b-4f]. By altering the exchanging ion to fluoride, bromide or
iodide, the corresponding halomercurials (8-10) were also obtained (Scheme 1) [4g].
R1 R2 R1 R2
1: M = Zn, R1 = R2 = R3 = R4 = R6 = H, R5 = R7 = P NaF
2: M = Cu, R1 = R2 = R4 = R6 = H, R3 = COMe, R5 = R7 = P R3 O R3
or
N N O N N
3: M = Cu, R1 = COMe, R2 = R3 = R4 = R6 = H, R5 = R7 = P NaCl
M R4 + Hg M R4
4: M = Zn, R1 = R2 = R4 = R6 = H, R3 = R5 = R7 = P or
N N O N N
5: M = Zn, R1 = R3 = R7 = Et, R2 = R4 = R5 = R6 = H NaBr
O
6: M = Zn, R1 = R3 = Q, R2 = R4 = R6 = R7 = H, R5 = P or
NaI R7 R5
R7 R6 R5 R6

7: M = Zn, R1 = R3 = HgCl, R2 = R4 = R6 = H, R5 = R7 = P >100% [4a,4c,4d,4f,4g]


8: M = Zn, R1 = R3 = HgF, R2 = R4 = R6 = H, R5 = R7 = P >100% [4f]
9: M = Zn, R1 = R3 = HgBr, R2 = R4 = R6 = H, R5 = R7 = P 81% [4f]
10: M = Zn, R1 = R3 = HgI, R2 = R4 = R6 = H, R5 = R7 = P 69% [4f]
11: M = Zn, R1 = R3 = R4 = HgCl, R2 = R6 = H, R5 = R7 = P [4d]
P = CH2CH2CO2Me
12: M = Cu, R1 = HgCl, R2 = R4 = R6 = H, R3 = COMe, R5 = R7 = P >100% [4a,4c]
Q = CH2CH2OAc
13: M = Cu, R1 = COMe, R2 = R4 = R6 = H, R3 = HgCl, R5 = R7 = P >100% [4a,4c]
14: M = Zn, R1 = HgCl, R2 = R4 = R6 = H, R3 = R5 = R7 = P >100% [4b-4d]
15: M = Zn, R1 = R2 = HgCl, R3 = R5 = R7 = P, R4 = R6 = H >100% [4d]
16: M = Zn, R1 = R3 = R7 = Et, R2 = R4 = R6 = H, R5 = HgCl 84% [4d]
17: M = Zn, R1 = R3 = R7 = Et, R2 = R4 = H, R5 = R6 = HgCl 84% [4d]
18: M = Zn, R1 = R3 = Q, R2 = R4 = R6 = H, R5 =P, R7 = HgCl 87% [4f]

Scheme 1. Formation of mercury porphyrins by Smith and co-workers [4].


40
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Initially the authors reported the substitution of the chloromercury fragments


only in the β- and not the meso-positions of the porphyrins [4a]. Further
investigations by demercurating the products with NaBD4 showed however that the
macrocycles were also partially mercurated in the meso-positions [4c]. High yields of
the bis-mercurial 7 could only be accomplished with an excess of Hg(II) acetate.
Therefore a mixture of the desired product 7 and a β,meso-trimercurated product 11
was always obtained. Thus the yields of some derivatives are recorded in Scheme 1
as “>100%” due to the presence of indeterminate small amounts of tri-mercurated
compounds. However the final products from reactions of the mercurated derivatives
were successfully separated by column chromatography.
Subsequently the authors utilised the mercurials for Heck-type C-C coupling
reactions. The reaction with ethylene and LiPdCl3 as catalyst gave the desired di-
vinyl porphyrin (19) in only 5% yield [4c]. Methyl acrylate on the other hand yielded
the corresponding acrylate (20) in up to 84% yield [4c-4e]. All mercurials shown in
Scheme 1 (7-18) were successfully substituted with methyl acrylate under similar
reaction conditions. The addition of bromine or iodine to the mercurials led to the
formation of the dihaloporphyrins in up to 96% yield, which were utilised for Stille-
type coupling reactions with tri-n-butylethenyl stannane [4f]. Interestingly, porphyrin
11 also formed a separable pair of cyclised products (30) when reacted with methyl
acrylate (Scheme 2) [4d]. In the next stage of their investigations the same authors
widened the scope of the Heck-type C-C coupling reaction reactions of the
mercurials by substituting the macrocycles with a variety of vinyl groups (e.g. 21-29,
Scheme 2) [4g]. Styrene derivatives formed the di-substituted porphyrins (23-27) in
higher yields than those from ethylene or acrylic substrates (19-22) under these
reaction conditions. The authors also reported that the coupling of bromoporphyrins
with vinylmercurials gave the desired products in higher yields than the inverse
coupling of porphyrinylmercurials with vinyl substrates [4g]. The dimercurated
deuteroporphyrin IX dimethyl ester was also successfully coupled with vinyl
carborane [5], while a Spanish group successfully iodinated dimercurated
deuteroporphyrin IX dioctyl ester [6].

41
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

R
19: R = H 5% [4c]

HgCl 20: R = CO2Me 31% [4a]


21: R = CHO 35% [4g]
HgCl 22: R = CN 30% [4g]
N N N N R
23: R = Ph 44% [4g]
Zn + R + LiPdCl3 M 24: R = p-MeOPh 43% [4g]
N N N N 25: R = m-NO2Ph 62% [4g]
26: R = p-NO2Ph 79% [4g]
P = CH2CH2CO2Me 27: R = PhCO2Me 57% [4g]
P P P P 28: R = 1-Naphthyl 52% [4g]
7
MeO2C 29: R = 9-Anthryl 17% [4g]
CO2Me

N N N N CO2Me
CO2Me
Zn H M
N N N N
30

P P P P

Scheme 2. Palladium-catalysed C-C coupling with mercurials by Smith and co-


workers [4a,4c,4g].

Buchler and Herget studied the effect of the central metal ion on the
mercuration reaction of a synthetic tetraarylporphyrin. Ni(II) (31), Pd(II) (32) and
Pt(II) (33) complexes of meso-tetra(p-tolyl)porphyrin were mercurated with mercuric
acetate or mercuric trifluoroacetate (Scheme 3) [7].
Ar Ar
R1
O O ClHg 34: M = Ni, R1 = R2 = H 35%
Hg
N N N N 35: M = Pd, R1 = R2 = H 20%
O O NaCl
Ar M Ar + or Ar M Ar 36: M = Pt, R1 = R2 = H 21%
F F
31: M = Ni N N F F N N 37: M = Pt, R1 = HgCl, R2 = H 13%
32: M = Pd O O 38: M = Pt, R1 = HgCl, R2 = HgCl 12%
F Hg F
33: M = Pt R2
O O
Ar Ar
Ar = p-tolyl

Scheme 3. Mercuration of meso-tetra(p-tolyl)porphyrin [7].

The yields of these reactions were highest for the Ni(II) complex (34, 35%), while
the Pd(II) (35, 20%) and Pt(II) porphyrins (36, 21%) gave similar but lower yields.
Di- and tri-mercurated porphyrins with Ni(II) and Pd(II) as the central metal ion
could not be isolated, as these derivatives degraded during column chromatography.
On the other hand, small amounts of the di- (37, 13%) and tri-mercurated (38, 12%)
Pt(II) porphyrins were stable during chromatography and could be successfully
characterised.
As vinylmercurials are accessible more easily than the porphyrin analogues
and no studies concerning the effect of the Hg(II) fragment on the porphyrin core

42
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

have been reported, the synthetic use of these macrocycles remained very limited for
almost two decades. Recently porphyrinylmercurials were resurrected by our group
and collaborators during studies on the mercuration of Ni(II) (39), Zn(II) (40) and
free base 5,15-diarylporphyrins (41) and Ni(II) 5,15-dialkylporphyrin (42) and
subsequent reactions of these mercurials (Scheme 4) [8].
βB

R βA R HgCl

F F
N N F F N N
O O NaCl
M + F Hg F M
N N O O N N

R R
39: R = 3,5-di-t-Bu-Ph, M = Ni 43: R = 3,5-di-t-Bu-Ph, M = Ni 33%
40: R = 3,5-di-t-Bu-Ph, M = Zn 44: R = 3,5-di-t-Bu-Ph, M = Zn 17%
41: R = 3,5-di-t-Bu-Ph, M = 2H 45: R = 3,5-di-t-Bu-Ph, M = 2H 21%
42: R = nC7H15, M = Ni 46: R = nC7H15, M = Ni 23%

Scheme 4. Mercuration of 5,15-diaryl- and 5,15-dialkylporphyrins [8].

To our surprise, in all reactions the mercuration took place in the more
hindered βB-position (43-46), which was established by spectroscopic and
preliminary single crystal X-ray analysis. A slight metal dependency was also
observed. While the Ni(II) mercurial (43) was isolated in 33% yield, the Zn(II) (44,
17%) and free base analogues (45, 21%) as well as the Ni(II) 5,15-dialkylporphyrin
(46, 23%) gave the desired mercurials in lower yields. Ni(II) porphyrin 43 was
incorporated in three subsequent reactions. Complex 43 was protiodemercurated with
trifluoroacetic acid (TFA) to yield the unsubstituted porphyrin, although mercurial 43
is stable to treatment with acetic acid or water. The reaction with TFA-d resulted in a
regioselective ipso-substitution to yield the corresponding deuterated porphyrin.
Coupling of methyl acrylate with catalytic amounts of Pd(OAc)2 and PPh3 as ligand
led to the unexpected formation of a rearrangement product 47, substituted in the βA-
position, as a major product. The expected βB-substituted porphyrin 48 was only
observed as a minor product. Exchanging PPh3 with a bulky phosphine ligand like
di-t-butyl-biphenylphosphine produced solely the rearrangement product 47. The
omission of phosphine ligands on the other hand yielded only the βB-substituted
product 48. The ipso-iodinated porphyrin 49 was isolated in quantitative yield after
treatment of 39 with iodine (Scheme 5).

43
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Ar HgCl Ar X

Y
N N N N
i, ii, iii
Ni M
N N N N

Ar Ar = 3,5-di-t-Bu-Ph Ar
39
i: Pd(OAc)2, PtBu2biphenyl, methyl acrylate; 47: X = H, Y = C2H2CO2Me 63%
ii: Pd(OAc)2, methyl acrylate; 48: X = C2H2CO2Me, Y = H 47%
iii: I2; 49: X = I, Y = H quant.

Scheme 5. Reactions of 5,15-diarylporphyrin mercurials [8].

Palladium and Platinum

Palladium-catalysed C-C coupling reactions are today one of the most


versatile and widely used substitution reactions in synthetic chemistry [9]. Since the
first examples reported by the Smith group (discussed above), this methodology has
also been widely utilised for substitution reactions on porphyrinyl macrocycles and
this field has been reviewed recently [10]. Therefore we will not generally cover
palladium-catalysed reactions of haloporphyrins except where organometallic
derivatives have been isolated. However, new applications involving C-H activation
are highlighted. Our studies of Class F palladioporphyrins stem from a failed attempt
at a catalytic reaction during which we serendipitously detected the formation of a
η1-palladioporphyrin. Subsequently we isolated the first example of such a transition
metal porphyrin complex (51) by the stoichiometric oxidative addition of a Pd(0)
fragment to bromoporphyrin 50 (Scheme 6) [11a].
Ph Ph

NH N Pd2dba3 + Ph3P NH N PPh3


Br + or Pd Br
N HN Pd(PPh3)4 N HN PPh3

dba = dibenzylideneacetone

Ph Ph

50 51

Scheme 6. Formation of η1-palladioporphyrin [11a].

Since our initial report we successfully prepared various mono- and bis-η1-
palladio- and η1-platinioporphyrins with different phosphine (51-53, 55-71, 77),
arsine (54), amine (72-76) and chiral phosphine ligands (78-83) under similar
reaction conditions (Fig. 2) [11].

44
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

51: M = 2H, X = H, Y = Pd(Ph3P)2Br [11a,11b]


52: M = Ni, X = H, Y = Pd(Ph3P)2Br [11a,11b]
53: M = 2H, X = H, Y = Pd(Ph3As)2Br [11a,11b]
54: M = Ni, X = Pd(Ph3P)2Br, Y = Pd(Ph3P)2Br [11a,11b]
55: M = 2H, X = H, Y = Pd(dppe)Br 80% [11a,11b]
56: M = 2H, X = H, Y = cis-Pt(Ph3P)2Br [11a,11b]
57: M = 2H, X = H, Y = trans-Pt(Ph3P)2Br 94% [11a,11b, 11e]
58: M = Ni, X = H, Y = cis-Pt(Ph3P)2Br [11a,11b]
Ph 59: M = Ni, X = H, Y = trans-Pt(Ph3P)2Br 76% [11a-11c]
60: M = Ni, X = Pd(Ph3P)2Br, Y = trans-Pt(Ph3P)2Br [11a,11b]
61: M = 2H, X = H, Y = Pd(dppf)Br 60% [11b]
N N 62: M = 2H, X = H, Y = Pd(dppp)Br 70% [11b]
X M Y 63: M = MnCl, X = H, Y = trans-Pt(Ph3P)2Br 64% [11c]
N N 64: M = Co, X = H, Y = trans-Pt(Ph3P)2Br 63% [11c]
65: M = Zn, X = H, Y = trans-Pt(Ph3P)2Br 53% [11c]
66: M = Ni, X = Br, Y = trans-Pt(Ph3P)2Br 37% [11c]
Ph 67: M = Ni, X = trans-Pt(Ph3P)2Br, Y = trans-Pt(Ph3P)2Br 32% [11c]
dppe = 1,2-bis(diphenylphosphino)ethane 68: M = Ni, X = butyl, Y = trans-Pt(Ph3P)2Br 28% [11d]
dppf = 1,1'-bis(diphenylphosphino)ferrocene 69: M = Ni, X = Ph, Y = trans-Pt(Ph3P)2Br 84% [11d]
dppp = 1,3-bis(diphenylphosphino)propane 70: M = 2H, X = Ph, Y = cis-Pt(Et3P)2Br 67% [11d]
tmeda = N,N,N',N'-tetramethylethylenediamine 71: M = 2H, X = Ph, Y = trans-Pt(Et3P)2Br 90% [11d]
bpy = 2,2'-bipyridyl 72: M = 2H, X = H, Y = Pd(tmeda)Br 93% [11e]
CHIRAPHOS = 2,2'-bis(diphenylphosphino)butane 73: M = 2H, X = H, Y = Pd(bpy)Br 91% [11e]
Tol-BINAP = 2,3-bis(di-p-tolylphosphino)-1,1'-binaphthyl 74: M = 2H, X = Pd(tmeda)I, Y = Pd(tmeda)Br [11e]
diphos = 1,2-bis(methylphenylphosphino)benzene 75: M = 2H, X = Ph, Y = Pt(tmeda)I 83% [11e]
76: M = 2H, X = Ph, Y = Pt(bpy)I 78% [11e]
77: M = 2H, X = Ph, Y = trans-Pt(Ph3P)2I 87% [11e]
78: M = 2H, X = H, Y = Pd[(R,R)CHIRAPHOS]Br 76% [11f]
79: M = 2H, X = H, Y = Pd[(S,S)CHIRAPHOS]Br 99% [11f]
80: M = 2H, X = H, Y = Pd[(R)Tol-BINAP]Br 80% [11f]
81: M = 2H, X = H, Y = Pd[(S)Tol-BINAP]Br [11f]
82: M = 2H, X = H, Y = Pd[(R,R)diphos]Br 57% [11f]
83: M = 2H, X = H, Y = Pd[(S,S)diphos]Br [11f]

Figure 2. η1-Palladio- and η1-platinioporphyrins [11].

The η1-metalloporphyrins were isolated in high yields of up to 99% and some


structures were studied in detail by single crystal X-ray analysis [11b-11d]. All
compounds were stable in air and in organic solvents, although in chlorinated
solvents a slow Br-Cl exchange was observed. Indeed, slow growth over three
months of a crystal from a CDCl3 solution of 57 unambiguously proved this fact, as a
single crystal of the chloro analogue was structurally characterised [12]. Recently
Dolphin and co-workers reported the single crystal X-ray structure of a η1-
palladioporphyrin zinc(II) chloro complex (the chloropalladium analogue of 65) and
studied its catalytic efficiency in the Heck reaction [13].
While the Pd(II) porphyrins were always obtained as trans-isomers (except
with chelating ligands), the cis-isomers of the platinated macrocycles could often be
observed and isolated as the initial products of the oxidative additions. Prolonged
heating however transformed all examples into the trans-isomers [11c]. The single
crystal X-ray structure of palladioporphyrin 55 (Figure 3, left) displays a slightly

45
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

distorted square planar coordination about the Pd(II) fragment with a slightly ruffled
porphyrin core, while one phenyl group of the diphosphine fragment shields the face
of the porphyrin [11a]. The single crystal X-ray structure for platinioporphyrin 58
(Figure 3, right) is consistence with the stability of the initially formed cis-adduct at
room temperature. Again one phenyl group of the phosphine ligand shields the face
of the porphyrin suggesting a recurring motif. The porphyrin ring is distorted in this
example, as expected for Ni(II) porphyrins. Interestingly macrocycle 58 adopts a
hybrid of the ruffled and saddled conformations [11b]. This pattern also applies to
the single crystal X-ray structures of other Ni(II) (59), Co(II) (64) and Zn(II) (65)
porphyrins with Pt(PPh3)2Br substituents [11c]. The smaller and more reactive PEt3
ligands led to a faster formation of the cis-product and transformation into the trans-
isomer could only be accomplished under harsher reaction conditions (xylene, 140°
C). In this case the smaller phosphine ligand does not shield the face of the
macrocycles, as shown in the single crystal X-ray structure of 71 [11d].

Figure. 3. X-ray single crystal structures of η1-palladioporphyrin 55 (left) and η1-


platinioporphyrin 58 (right). These figures were generated using data originally
reported by Arnold and co-workers [11a,b].

The Pt-porphyrins were stable during column chromatography while Pd(II)


analogues could only be purified by recrystallisation [11b]. Insertion of a metal
fragment into the porphyrin core could be achieved either before or after the
oxidative addition of the transition metal fragment [11c]. Cyclic and ac voltametry
studies of the η1-metalloporphyrins gave insight into the effects of the Pd(II) and
Pt(II) fragments on the redox properties of these macrocycles [11c]. Table 1
summarises the first reduction and oxidation potentials of Pd(II) (51, 52) and Pt(II)
porphyrins (57, 59, 67) together with the precursors 5,15-diphenylporphyrin
(H2DPP), 5,15-diphenylporphyrinatonickel (II) (NiDPP), 50 and 5-bromo-10,20-
diphenylporphyrinatonickel (II) (NiBrDPP). The voltametric data clearly confirmed

46
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

the electron-donating properties of both the Pt(II) and the Pd(II) fragment, although
the latter did not result in such a large cathodic shift. The effect of a second Pt(II)
fragment (67) is additive and the fact that both free base and Ni(II) complexes show
similar effects is evidence that the oxidation occurs at the porphyrin ring and not the
Ni(II) centre and that the changes in spectra are due to electronic effects and not due
to the distortion of the Ni(II) complexes. The first reduction for Pd(II) complexes 51
and 52 is not reversible, and loss of the Pd(II) fragment occurs. The electron-
donating substitution effect of both Pd(II) and Pt(II) fragments is also evident in the
remarkable decrease in reactivity towards a second oxidative addition. Both η1-bis-
platinio- and η1-bis-palladioporphyrins were only formed after longer reaction times
(e.g. four hours for Pt, forty minutes for Pd) as the metal fragments clearly decrease
the polarity of the opposite C-Br bond [11c].

Table 1. Electrochemical data for η1-metalloporphyrins and some precursors in


CH2Cl2-Bu4NPF6 vs. Ag/Ag+ using ferrocene as internal standard [E0(ox) = +0.55 V]
at 293K [11c].

Compound E0 (red1)/V E0 (ox1)/V

5,15-Diphenylporphyrin (H2DPP) -1.05 +1.16

5,15-Diphenylporphyrinatonickel(II) (NiDPP) -1.13 +1.17

50 -0.93 +1.19

5-Bromo-10,20-diphenylporphyrinatonickel (II) (NiBrDPP) -1.02 +1.20

51 -1.24a +0.89

52 -1.37a +0.95

57 -1.32 +0.83

59 -1.44 +0.90

67 -1.66 +0.61
a
Irreversible with loss of PdBr(PPh3) fragment

In order to enforce a cis arrangement of the platinum fragment, our group


pursued the formation of η1-platinioporphyrins with chelating phosphine ligands.
Unfortunately the desired organometallic porphyrins failed to form by the oxidative
addition of a haloporphyrin to zerovalent platinum species prepared in situ, as the
bis(chelating diphosphine)platinum fragments seem to be too stable to undergo such

47
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

a reaction. The same strategy utilising chelating amine ligands was successful and
the desired η1-platinioporphyrins (74, 75) were cleanly prepared in high yields. In
this case the zerovalent platinio fragment could be prepared in situ, which was a
distinct advantage over the previous method, as air- and moisture-sensitive Pt(0)
fragments could be avoided (Scheme 7). Under similar conditions η1-
palladioporphyrins with chelating amine ligands (72-74) were also isolated, which
were more stable than their analogues with chelating phosphine ligands, as they
could be purified by column chromatography [11e].
Ph Ph

NH N NH N N
Ph I + Pt(dba)2 + tmeda Ph Pt N
N HN N HN I

dba = dibenzylideneacetone
Ph tmeda = N,N,N',N'-tetramethylethylenediamine Ph
84 75 83%

Scheme 7. Formation of η1-platinioporphyrin with chelating amine ligand [11e].

The Arnold group has also prepared a suite of η1-palladioporphyrins with


chiral diphosphine ligands (78-83). Again the oxidative addition of a haloporphyrin
to the Pd(0) species prepared in situ could be utilised and the products formed in very
high yields of up to 99%. Enantiomeric pairs of three types of chiral diphosphines
were employed, out of which the Tol-BINAP [2,2’-bis(di-p-tolylphosphino)-1,1’-
binaphthyl] analogues (80, 81) were the most stable in solution and could be
recrystallised from chlorinated solvents. The same analogues induced the strongest
effect on the CD spectra (Figure 4) and therefore offer the most promise for further
applications such as chiral recognition and enantioselective catalysis [11f].

48
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Ph
P
N N
Ni Pd
N N P
Br
Ph

80

Figure 4. CD spectrum of 80 (upper, solid line), 81 (upper, dashed line) and visible
absorption spectrum (lower) in CH2Cl2 [11f].

We also studied subsequent reactions of these organometallic macrocycles.


Treatment of η1-platinioporphyrins with RLi led either to alkyl substitution on a
vacant meso-position, if present [14], or to alkylation at Pt and reductive elimination
of the Pt fragment, if no vacant meso-position was present [11d]. Alkynylation of the
Pt(II) fragment was cleanly achieved with CuI/Pd(PPh3)2Cl2 as catalyst to give a
phenylacetylene substituted Pt(II) porphyrin or even a porphyrin dimer with an
alkyne-Pt spacer (86, characterised by single crystal X-ray analysis) in 38% yield
(Scheme 8) [11d].
Ph Ph

N N PPh3 N N
Ni Pt Br + H Ni
N N PPh3 N N

Ph Ph
59 CuI/Pd(PPh3)2Cl2 85
Ph Ph

N N PPh3 N N
Ni Pt Ni
N N PPh3 N N

Ph 86 38% Ph

Scheme 8. Formation of alkynyl-Pt(II) linked porphyrin dimer [11d].

Bromide abstraction from the Pt(II) fragment to give a nitrato platinum


analogue was achieved by reacting the Pt(II) porphyrin (87) with silver nitrate. The

49
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

reaction with silver triflate yielded a triflatoplatinum porphyrin (88) which was
subsequently reacted with pyridine to form a cationic pyridyl complex in 97% yield.
Porphyrin 88 was also reacted with bipyridine to yield a dicationic bis-porphyrin in
93% yield. Similar reactions were carried out with a monopyridyl porphyrin to obtain
a cationic bis-porphyrin in 88% yield, with a di-pyridyl porphyrin to obtain a
dicationic tri-porphyrin in 89% yield and with a tetrapyridyl porphyrin 89, which
resulted in a tetracationic porphyrin pentamer 90 in 78% yield (Scheme 9) [15].
N

Ar Ar

N N PPh3 N N PPh3 O F NH N
Ph Ni Pt Br + AgOTf Ph Ni Pt O S F+N N
N N PPh3 N N PPh3 O F N HN

87 88 89
Ar Ar
Ar = 3,5-di-t-bu-Ph
Ph
N

N N
Ar Ni Ar
N N

Ph3P Pt+ PPh3

Ar Ar

N N PPh3 NH N PPh3 N N
Ph Ni Pt+ N N Pt+ Ni Ph
N N PPh3 N HN PPh3 N N

Ar Ar

Ph3P Pt+ PPh3 90 78%

N N
Ar Ni Ar
N N

Ph

Scheme 9. Self-assembly of porphyrin pentamer [15].

Bromination of 5,15-diarylporphyrins, although very rapid, is not selective


and the resulting monobromo- and di-bromoporphyrins are often not separable
without extensive chromatography. On the other hand, the monopalladation of
dibromoporphyrins has been achieved selectively, due to the much slower oxidative

50
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

addition of the second Pd(0) fragment. In collaboration with the Sugiura group, we
monobrominated various diarylporphyrins (91-94) via the respective
monopalladioporphyrins (95-98) and subsequent hydrodepalladation with NaOH in
MeOH. The reaction was performed either with the isolated η1-palladioporphyrins
95-98 or in a one-pot reaction with the respective dibromoporphyirins 91-94 and the
desired monobromoporphyrins 99-102 were isolated in good yields of 57-71%
(Scheme 10) [16].
Ar Ar

Ph Ph
95: R = H
NH N Pd2dba3 (1 mol) NH N P
96: R = Me
91: R = H Br Br + + Br Pd P
Ph 97: R = t-Bu
92: R = Me N HN dppe (2 mol) N HN Br Ph 98: R = O(CH2)2CH(CH3)2
93: R = t-Bu
94: R = O(CH2)2CH(CH3)2
Ar Ar Ar
R R
Ar =
NH N NaOH/MeOH
99: R = H 57% Br
dba =dibenzylideneacetone
100: R = Me 71% N HN
dppe = bis(diphenylphosphino)ethane
101: R = t-Bu 67%
102: R = O(CH2)2CH(CH3)2 61%
Ar

Scheme 10. Selective monobromination of 5,15-diarylporphyrins [16].

Recently we also reported the regiospecific halogen exchange reaction to


obtain meso-iodoporphyrins 84, 112 and 113 with η1-palladioporphyrins 51, 108 and
109 as intermediates. Again it was not necessary to isolate the transition metal
complexes and the products 84, 111 and 113 were prepared in high yields of up to
81%. Similarly meso-iodo-bromoporphyrins 114 and 115 were isolated from
dibromoporphyrins 106 and 107 via the respective η1-palladioporphyrins 110 and
111 (Scheme 11) [17]. The formation of the iodoporphyrins was unambiguously
confirmed by single crystal X-ray analysis and the same methodology was applied to
prepare meso-(2-iodovinyl)porphyrins [37]. The regioselective meso-mono- and bis-
halogenation of 5,15-substituted porphyrins can now be easily achieved using the
methods described above. η1-Palladioporphyrins were also utilised as starting
materials to isolate porphyrinylphosphine oxides, which will be discussed below.

51
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Ar Ar

103: R = H, Ar = Ph NH N Pd2dba3 + PPh3 NH N PPh3 51: R = H, Ar = Ph


104: R = Ph, Ar = Ph R Br + or Pd Br 108: R = Ph, Ar = Ph
R
105: R = Ph, Ar = 3,5-di-t-Bu-Ph Pd(PPh3)4 109: R = Ph, Ar = 3,5-di-t-Bu-Ph
N HN N HN PPh3
106: R = Br, Ar = Ph 110: R = Br, Ar = Ph
107: R = Br, Ar = 3,5-di-t-Bu-Ph 111: R = Br, Ar = 3,5-di-t-Bu-Ph
Ar Ar Ar

112: R = H, Ar = Ph 78%
84: R = Ph, Ar = Ph NH N I2/air
79%
113: R = Ph, Ar = 3,5-di-t-Bu-Ph 75% R I
114: R = Br, Ar = Ph 78% N HN dba =dibenzylideneacetone
115: R = Br, Ar = 3,5-di-t-Bu-Ph 81%

Ar

Scheme 11. Selective meso-iodination of 5,15-diarylporphyrins [17].

A new type of peripheral palladation reaction has emerged in the porphyrin


literature recently. Various groups have reported C-H activation reactions involving
proposed β-palladated porphyrin intermediates. The first such reaction is an
intramolecular cyclisation reaction of ortho-iodinated meso-phenyl porphyrins 116-
119 (Scheme 12) [18]. Both mono- (120-123, up to 44%) and double-cyclised
products (125, 126, up to 36%) were prepared in acceptable yields, however the
isomers 125 and 126 could not be separated. The same reactions with ortho-
bromophenylporphyrins were not successful.

N N N N
Pd(PPh3)4
R M R R M R
N N N N

R2 R2 R2 R2
R1 R1
116: R = H, R1 = COOMe, R2 = H, M = Ni 120: R = H, R1 = COOMe, R2 = H, M = Ni
117: R = R1 = H, R2 = OMe, M = Ni 121: R = R1 = H, R2 = OMe, M = Ni 44%
118: R = R1 = H, R2 = OMe, M = Cu 122: R = R1 = H, R2 = OMe, M = Cu
119: R = 4-(COOMe)Ph, R1 = COOMe, R2 = H, M = Ni 123: R = 4-(COOMe)Ph, R1 = COOMe, R2 = H, M = Ni

N N N N N N
Pd(PPh3)4
Ni Ni + M

N N N N N N

I
36%

124 125 126

Scheme 12. Intramolecular Pd(0) catalysed cyclisation [18].


52
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

In a similar case a naphthalene-triflate substituted porphyrin (127) has been


reported to cyclise. In an attempt to synthesise a cofacial ferrocene-porphyrin dyad
using a double Suzuki-Miyaura reaction, the desired target compound was not
formed. The only product of this reaction was an intramolecular cyclised product
(128) with a yield of 14%, and interestingly in the absence of the bis(boronate) the
cyclised porphyrin was observed only in trace amounts (Scheme 13) [19].
Ph Ph

N N B(OH)2 N N
PdCl2(dppf)
Ph Zn + Fe Ph Zn

N N B(OH)2 N N

TfO
Ph Ph
dppf = 1,1'-bis(diphenylphosphino)ferrocene
127 128 14%

Scheme 13. Formation of doubly linked naphthyl-porphyrins [19].

Our group reported recently the Heck-type coupling reaction of Ni(II) meso-
vinylporphyrin (131) with various meso-bromoporphyrins (104, 129, 130). In all
cases the only isolated product is a meso-β linked porphyrin dimer (132-134).
According to the proposed mechanism for this reaction, a β-substituted
palladioporphyrin must be formed as an intermediate, presumably also by C-H
activation. A plausible pathway involves a five-membered Pd(IV) palladcycle, which
forms the product by migratory reductive elimination (Scheme 14) [20]. Similar
migration products were also obtained from the reactions of meso-vinyl porphyrin
131 with iodobenzene and 9-bromoanthracene. However in these cases the normal
non-rearranged products were also isolated. Another example of migration to a
neighbouring β carbon in a palladium catalysed process was mentioned above in the
discussion of the Heck coupling of our mercurioporphyrins [8].

53
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Ph

N N
Ph M Br Ph
Ph
104: M = 2H N N
129: M = Ni
N N
130: M = Zn
N N Ni
Ph Pd(OAc)2
+ Ph M N N
PtBu2biphenyl
Ph N N

Ph
N N Ph 132: M = 2H 23%
Ni 133: M = Ni 33%
N N 134: M = Zn 15%

Ph 131 N
N
N N N H N
N
N N N N
N H
Pd
H Pd PdL2
L Br
Br
L

Scheme 14. Formation of meso-β linked porphyrin dimers [20].

Osuka and co-workers reported the formation of various cyclopentadiene


fused porphyrins (137-146) in the palladium-catalysed reactions of the respective
bromoporphyrins (39, 40, 135, 136) with symmetrical alkynes (Scheme 15) [21]. The
products (137-146) formed in high yields of 69-87% and two examples were
characterised by single crystal X-ray analysis. Again the proposed reaction
mechanism indicates intramolecular Pd attack on a β C-H bond with a β-substituted
palladioporphyrin as an intermediate. These last examples show that η1-
palladioporphyrins can be formed by C-H activation as well as via C-X oxidative
addition. Although these porphyrins were only proposed as possible intermediates,
the scope for applications of η1-transition metal porphyrins has been considerably
widened and this field is attractive for further development.
137: M = Ni, R = Ph 78%
Ar Ar 138: M = Ni, R = p-CF3-Ph 81%
dba = dibenzylideneacetone 139: M = Ni, R = p-MeO-Ph 78%
140: M = Ni, R = thienyl 82%
N N N N 141: M = Ni, R = nPr 85%
Pd2dba3 R
M Br + R R M 142: M = Cu, R = Ph 87%
o-tol3P
N N N N 143: M = Cu, R = nPr 86%
39: M = Ni 144: M = Zn, R = Ph 79%
R
40: M = Zn 145: M = Zn, R = p-CF3-Ph 78%
Ar = 3,5-di-t-Bu-Ph
135: M = Cu Ar Ar 146: Ar = dioctyloxyphenyl
136: Ar = dioctyloxyphenyl, M = Zn M = Zn, R = p-MeO-Ph 69%

Scheme 15. Pd(0) catalysed annulation of porphyrins [21].

54
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Boron

Organoboron compounds have been widely utilised in Suzuki-Miyaura C-C


coupling reactions [22]. As such, porphyrinylboronates are of immense value for
substitutions on the porphyrin carbon skeleton. Therien and co-workers first prepared
mono- (147) and bis-porphyrinylboronates (148) of Zn(II) 5,15-diphenylporphyrin
from the Pd(0) catalysed reaction of the respective bromoporphyrins with
pinacolborane in excellent yields of 79% and 86%, respectively, and characterised
both macrocycles by single X-ray crystallography [23a]. The same authors also
reported the formation of a similar boronate with alkoxyphenyl substituents on the
porphyrin (155, 156) under similar reaction conditions [23b], while the Osuka group
isolated boronates with bis-dioctyloxyphenyl (157) [24a], bis-
dioctyloxyphenylpyridyl (158) [24b] and bis-dodecyloxyphenyl (159) [24c]
substituents on the remaining meso-positions of the macrocycle. Again under similar
reaction conditions, Lindsey and co-workers prepared porphyrinylboronates with
mesityl groups on the porphyrin (149, 150) [25], while a Swiss group isolated
porphyrinylboronates with long chain ester functionality on the phenyl rings of the
porphyrin (160) [26]. Other porphyrinylboronates reported thus far include
[10,15,20-tris(mesityl)porphyrin-5-yl]boronate (151) [27], the free base [10,20-
bis(di-t-butylphenyl)porphyrin-5-yl]boronate (152) [28], a triarylporphyrinylboronate
with carboxyalkoxy functionality on the phenyl rings (161, 162) [29], a
diarylporphyrinylboronate with a cyanophenyl group on the remaining meso position
(153) [30], (10,15,20-tritolylporphyrin-5-yl)boronate (154) [31a] and [10,20-bis(4-
decyloxyphenyl)porphyrin-5-yl]boronate (163) [31b] (Scheme 16).

55
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Ar L = (PPh3)2 or 1,1'-bis(diphenylphosphino)ferrocene Ar

N N O N N O
LPdCl2
R M Br + HB R M B
N N O N N O

147: M = Zn, R = H, Ar = Ph 86% [23a]


Ar Ar
148: M = Zn, R = BO2C6H12, Ar = Ph 79% [23a]
149: M = Zn, R = mesityl, Ar = 4-t-Bu-Ph 89% [25a]
150: M = Zn, R = H, Ar = mesityl 93% [25b]
151: M = Zn, R = Ar = mesityl 96% [27b]
152: M = 2H, R = H, Ar = 3,5-di-t-Bu-Ph 32% [28]
153: M = Zn, R = p-CNPh, Ar = 3,5-di-t-Bu-Ph 84% [30]
154: M = Zn, R = Ar = tolyl 5%* [31a]

O O
155: M = Zn, R = H, Ar = [23b]

156: M = Zn, R = dimethylaminophenylethynyl]


O O
Ar = 88% [23b]

O O
157: M = Zn, R = H, Ar = H3C(H2C)7 (CH2)7CH3 90% [24a]

O O
158: M = 2H, R = H, Ar = 4-pyr, H3C(H2C)7 (CH2)7CH3 [24b]

O (CH2)11CH3
159: M = Zn, R = H, Ar = [24c]

160: M = Zn, R = H, Ar = EtO2C(H2C)3 (CH2)3CO2Et quant. [26]


O O

CO2K

161: M = Zn, R = H, Ar = KO2C(H2C)10O O(CH2)10CO2K 88% [29]

CO2K

162: M = Zn, R = Ar = KO2C(H2C)10O O(CH2)10CO2K 85% [29]

O(CH2)9CH3

163: M = Zn, R = H, Ar= 73% [31b]

* including initial porphyrin condensation

Scheme 16. Formation of porphyrinyl boronates [23-31] (the point of attachment of


the aryl groups to the meso carbon is from the lowermost ring carbon).

All porphyrinylboronates described above were utilised as synthons for


Suzuki-Miyaura C-C coupling reactions to attach a variety of functional groups to
the carbon framework of porphyrins. These include an alanine derivative (79%)
[23a], naphthalene derivatives (48-92%) [23a,32], thiophene derivatives (29-93%)
56
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

[23b], hydroxybenzaldehyde (60%) [25a], benzene derivatives (24-63%) [26,33],


dibenzofurancarboxylic acid (67%) [27a], xanthene derivatives (46-74%) [27a,34],
pyrenes with and without alkoxy substituents [35] and carbon nanotubes [36]. The
variety of these functional groups emphasises the almost unlimited possibility to
substitute porphyrinyl macrocycles with carbon groups by this methodology.
Multiporphyrin coupling reactions were also reported in which
porphyrinylboronates were the key synthons. Therien and co-workers successfully
isolated a carbazole linked porphyrin dimer (164) in a good yield of 76% (Scheme
17) [23a].
Ph

N N O
Ph Ph
Zn B
N N O
N N
N N
Zn Zn
N N
Ph 147 Pd(PPh3)4
+ N N
Br Br
Ph N Ph
H
N 164 76%
H

Scheme 17. Carbazole linked porphyrin dimer [23a].

The Lindsey group reported the formation of an unusual diporphyrin system


with a bis(phenyldipyrromethane)palladium linker (165, Scheme 18) [25b]. The
porphyrin dimer (165) was isolated in 50% yield. Subsequently the Pd(II) fragment
was removed from the linker by treatment with DTT (threo-1,4-dimercapto-2,3-
butanediol) and Zn(II) inserted into the linker to give a trinuclear Zn(II) porphyrin
dimer (166).

57
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

N N N N
Mes Zn Mes Mes Zn Mes
Mes N N N N

N N O
Zn B
N N O

N N N N
Mes 150
Pd(PPh3)4 DTT Zn(OAc)2
+ Pd 165 50% Zn 166 70%
N N N N
N N
I Pd I
N N

N N N N
Mes Zn Mes Mes Zn Mes
N N N N

DTT = threo-1,4-dimercapto-2,3-butanediol

Scheme 18. Bis(phenyldipyrromethane) linked porphyrin dimers [25b].

Cofacial bis(trimesitylporphyrins) (167, 168) and corresponding monomeric


models were synthesised from the reaction of a porphyrinylboronate 151 with
dibromoxanthene and dibromodibenzofuran (Scheme 19) [27a,b]. The proximity of
the two macrocycles in the bisporphyrins was controlled by varying the spacer
groups and the differing steric demands of the two systems explain the very different
yields. Zn(II) fragments were removed in 6N HCl and MnCl inserted into the
macrocycles subsequently.
t-Bu t-Bu
Mes
N N
Br Zn Mes
N N
Mes
+ O O 167 21%
Mes
Mes N N
Br Zn Mes
N N
Mes
N N t-Bu t-Bu
O
Pd(PPh3)4
Mes Zn B
O Mes Mes
N N N
N Zn N
Br N
Mes
Mes 151
+ O O 168 77%

Mes
Br N
N Zn N
N
Mes Mes

Scheme 19. Cofacial porphyrin dimers [27b].


58
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

A water soluble porphyrin dimer with phenyl spacers and long chain acid
salts on the phenyl groups (169) was prepared by a Japanese group (Scheme 20) [29].
The authors isolated a similar diporphyrin (50%) with a vacant meso-position on
each of the macrocycles under similar reaction conditions [29].
R1

R2 R2

R2
N N O I I

R1 Zn B +
N N O
R2

R2 R2 161

R1 = CO2K
R2 = O(CH2)10CO2K Pd(PPh3)4
R1
R1 R1

R2 R2
R1 R2 R2 R1
R2 R2
N N

R2 N N R2
Zn Zn
N N
N N
R2 R2

R2 R2
169 60%

R1 R1

Scheme 20. Phenyl linked porphyrin dimer [29].

The formation of two photoresponsive porphyrin dimers joined by ferrocene


linkers was recently reported by Aida and co-workers. In the first example the
ferrocene has an aminodiphenylethynyl unit attached on each cyclopentadienyl ring
(170, 22%) (Scheme 21) [31a]. UV light caused the molecule 170 to rotate and the
dyad locked internally as each aniline unit coordinated to the central Zn(II) metal of
one porphyrin (170, lower). Addition of dipyridylethylene to the dimer formed an
externally locked diporphyrin [31a].

59
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Tol

Br NH2
N N O
Tol Zn B + Fe
N N O
Br NH2
154

Tol Tol Pd(PPh3)4

Tol N
N
Zn
N
N
NH2
Tol Fe
Tol
NH2
N
N
Zn
N
Tol N Tol
170
N Tol
Tol N
Zn
N
N
H2N Tol Tol

N N NH2
Fe
Tol Zn
N N

Tol

Scheme 21. Aniline-ferrocene linked porphyrin dimer [31a].

In the second example the ferrocene spacer had an azobenzene unit attached
to the side remote from the porphyrin (171, 20%) (Scheme 22) [31b]. In this example
the two porphyrins were strapped together by bis(isoquinoline) coordinating to the
central zinc ions. UV irradiation caused the molecule to rotate around the ferrocene
group in a pedal-like motion, by isomerisation to the cis-azobenzene configuration,
while this motion was reversed by visible light, which restored the azobenzene to the
trans configuration.

60
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Ar

Br
N N O
N
Zn B + Fe
N
N N O
Br
163

Ar Ar Pd(PPh3)4

Ar = O(CH2)9CH3
N
N
Zn
N
N

Ar Fe N
Ar
N
N
N
Zn
N 171 20%
N

Ar

Scheme 22. Azobenzene-ferrocene linked porphyrin dimer [31b].

The Osuka group reported the formation of various meso-meso directly (174,
175) and phenyl linked porphyrin arrays (176). The coupling of boronate 157 to
bromoporphyrins 172 and 173 resulted in porphyrin dimers 174 and 175. A
porphyrinyl heptamer and octamer were also isolated by repeating the reaction with a
bromo-terminated hexamer [24a]. Similarly the coupling with diiodobenzene yielded
phenylene bridged porphyrin dimer (176), tetramer, hexamer and octamer (Scheme
23) [24c]. The Ag+ promoted meso-meso coupling of Zn(II) porphyrins enabled the
formation of porphyrin wheels with 12 [37a] and 24 porphyrin subunits [37b],
respectively, by combining the directly (174, 175) and phenylene linked (176)
porphyrin oligomers and these supramolecules exhibited very interesting light-
harvesting properties.

61
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Ar Ar

N N O N N
Zn B + M Br
N N O N N
172: M = 2H
173: M = Ni
157
Ar Pd(PPh3)4 Ar
Ar Ar Ar = dioctyloxyphenyl

N N N N
Zn M 174: M = 2H 62% [24a]
175: M = Ni 57% [24a]
N N N N

Ar Ar
Ar

N N O
Zn B + I I
N N O

PdCl2(PPh3)2/AsPh3
Ar 157
Ar Ar = p-dodecyloxyphenyl Ar

N N N N
Zn Zn
N N N N

176 64% [24c]


Ar Ar

Scheme 23. Directly linked and phenylene bridged porphyrin dimers [24a,24c].

Porphyrin oligomers with up to four porphyrin units have been prepared


under similar conditions by the Fukuzumi group. In this case one macrocycle also
had ester functionality on one aryl ring (178, Scheme 24) [28,38]. The macrocycles
were subsequently attached to indium tin oxide electrodes and the macrocycles
exhibit photocurrent generation.

62
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Ar Ar

NH N O NH N
B + Br CO2CH3
N HN O N HN

Ar 151 Ar 177
Pd(PPh3)4
Ar Ar
Ar = 3,5-di-t-Bu-Ph

NH N NH N
CO2CH3
N HN N HN

Ar 178 37% Ar

Scheme 24. Directly linked porphyrin dimers with ester functionality [28,38].

The palladium-catalysed coupling of dibromoporphyrin 179 with boronate


153 resulted in the formation of a directly meso-meso linked trimer with two
cyanophenyl groups (180). The authors subsequently oxidized it to a triply-linked
trimer and studied its interaction with carbon nanotubes. (Scheme 25) [30].
Ar Ar

N N O N N
NC Zn B + Br Zn Br
N N O N N

Pd(PPh3)4
Ar = 3,5-di-t-Bu-Ph Ar 153 Ar 179
Ar Ar Ar

N N N N N N
NC Zn Zn Zn CN
N N N N N N

Ar Ar 180 17% Ar

Scheme 25. Directly linked porphyrin trimer with cyanide functionality [30].

β-Boronated porphyrins have also been prepared in recent years. The first
example (182) was obtained from the palladium-catalysed boronation of
bromoporphyrin 181 and subsequently utilised in various Suzuki-type C-C bond
forming reactions to isolate phenyl-substituted porphyrins (characterised by single
crystal X-ray analysis) and a β-β linked porphyrin dimer 183 in a good yield of 62%
(Scheme 26) [39].

63
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

N N O O N N
PdCl2(dppf) 76%
Zn + B B Zn + 181
O O N N
N N O
Br B
O
181 dppf = 1,1'-bis(diphenylphosphino)ferrocene 182

PdCl2(dppf)

N N N N
Zn Zn

N N N N

183 62%

Scheme 26. Formation of β-β linked porphyrin dimer [39].

A similar reaction of β-boronated tetraarylporphyrins 186 and 187 (65-70%)


with β-brominated tetraarylporphyrins 184 and 185 yielded chiral β-β linked
porphyrin dimers (188, 199) in good yields of 54-59%. The axial chirality of the
dimers stemmed from the hindered rotation around the β-β bond and was assigned by
CD measurements (Scheme 27) [40].
Ar Ar

NH N O NH N 184
Pd(PPh3)4
Ar Ar + HB Ar Ar + or
N HN O N HN O 185
Br B
O
Ar
Ar Ar
184: Ar = Ph 186: Ar = Ph 70%
185: Ar = p-tolyl 187: Ar = p-tolyl 65%
NH N
Ar Pd(PPh3)4
Ar Ar
N HN

NH N
Ar Ar Ar
188: Ar = Ph 59% N HN
189: Ar = p-tolyl 54%

Ar

Scheme 27. Formation of chiral β-β linked porphyrin dimers [40].

Osuka and co-workers pioneered the regioselective iridium-catalysed


boronation of various unsubstituted diarylporphyrins (39, 41, 135). In all cases the β
position adjacent to the meso-carbon was selectively boronated (190-192) in yields
up to 47% and surprisingly no meso boronation was observed (Scheme 28) [41]. A

64
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

diboronated triarylporphyrin (80%) and a tetraboronated diarylporphyrin (73%) were


prepared under similar reaction conditions and the substitution again took place only
in the β position adjacent to the vacant meso-carbon [41]. The latter compound was
unambiguously characterised by single crystal X-ray analysis. The authors also
demonstrated the usefulness of the porphyrinylboronates as building blocks for
multiporphyrin arrays. While the monoboronated porphyrin formed in a palladium-
catalysed homocoupling reaction a β-β linked porphyrin dimer (65%), the reaction of
the same compound with meso-bromoporphyrin resulted in the formation of a meso-
β linked porphyrin dimer (80%) [41]. This selective βA substitution complements the
βB substitutions obtained in the mercuration reactions described above [8].
Ar Ar
O
B
39: M = Ni N N N N O
O O
[Ir(cod)OMe]2
41: M = 2H M + B B M
dtbpy 190: M = 2H 43%
135: M = Cu N N O O N N
191: M = Ni 47%
Ar = 3,5-di-t-Bu-Ph 192: M = Cu 44%
cod = cyclooctadiene
Ar Ar
dtbpy = di-t-Bu-2,2'-bipyridyl

Scheme 28. β-Boronated diarylporphyrins [41].

Porphyrins with directly attached boronic acid substituents have also been
reported recently. The β-boronic acid (195) was prepared from the deprotection of a
β-substituted boronate (194) with NaIO4 in acidic solution followed by metalation. A
new meso-porphyrinylboronate (197) had to be employed for the formation of the
meso-boronic acid (198), as the already discussed boronate 147 could not be
deprotected. Therefore the neopentylglycolatoborane-substituted porphyrin (197)
was isolated first and subsequently deprotected with N-methyldiethanolamine under
mild conditions followed by metalation to yield the desired boronic acid 198. The
macrocycles (195, 198) were subsequently tested as sensors for sugars and
carbohydrates (Scheme 29) [42].

65
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

O O
Ph Br Ph B

N N O O N N
B Pd(t-Bu3P)2
Ph Zn Ph + Ph Zn Ph
B
N N O O N N
quant.
HO OH
Ph 193 Ph B Ph 194

NaIO4
N N HCl/MeOH
Ph Zn Ph then Zn2+
N N
195 74%

Ph
Ph Ph

N N O O N N O
B Pd(t-Bu3P)2
Zn Br + Zn B
B O
N N O O N N
197 63%

Ph 196 Ph
THF/H2O
Ph + HO OH
N

N N OH then Zn2+
Zn B
N N OH

Ph 198 61%

Scheme 29. Boronic acid substituted porphyrins [42].

Phosphorus

The first report of a porphyrin with phosphorus directly attached to the meso
carbon of a porphyrin was by the Smith group in 1977. The authors generated a
Zn(II) octaethylporphyrin (199) cation radical by treatment with tris(p-
bromophenyl)ammoniumyl hexachloroantimonate and reacted the radical species
with triphenylphosphine to obtain a meso-octaethylporphyrinylphosphonium salt
(200, Scheme 30) [43].

66
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

N N NH N
HCl/H2O P+Ph3 Cl-
Zn + [(p-BrPh)3N]+[SbCl6]- + PPh3
N N N HN

199 200 26%

Scheme 30. meso-Porphyrinylphosphonium salt [43].

Porphyrinyl β-phosphonium (and arsonium) salts generated from Zn(II)


tetraphenylporphyrin (201) radical cation were first reported by Shine et al. [44]. The
same phosphonium salt (202) was reported in 1991 by Giraudeau and El Kahef, who
prepared it from electrochemically generated cation radical. (Scheme 31) [45].
Ph Ph

P+Ph3
N N N N
-e-
Ph Zn Ph + PPh3 Ph Zn Ph ClO4-
N N N N

35% [44]
50% [45]
Ph Ph
201 202

Scheme 31. β-Porphyrinylphosphonium salt [44,45].

Similarly a chemically formed Fe(III)tetraphenylporphyrin cation radical


treated with PPh3 formed the corresponding β-phosphonium salt. This compound
was unambiguously characterised by single X-ray crystallography and studied in
detail by NMR analysis [46].
The Giraudeau group isolated a variety of di- (203-219, 211, 212) and
triporphyrins (210) linked by phosphonium bridges [47]. Electrochemically-
generated porphyrin cation radicals were reacted with the respective di- and
triphosphines to yield porphyrin oligomers 203-212, including a mixed
phosphonium-arsonium linked dimer 208. The electron-withdrawing effect of the
phosphonium cations in the oligomers resulted in a red shift of the absorption spectra
in comparison to the monomeric units. The Soret bands of the dimers were also
considerably broadened, suggesting exciton coupling between the macrocycles.
Electrochemical studies on the benzene-diphosphonium bridged dimer 211 revealed
that the chromophores were reduced before the phosphonium spacer and the red shift
in the UV-visible spectrum was ~50 nm in comparison to the monomer. Therefore
the authors concluded that the positive charge of the P atoms is delocalised over both
67
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

macrocycles and propose a bis-chlorin-like structure (212) for this dimer (Scheme
32) [47c].
Ph
N N
Ph Zn Ph
N N
PPh2 P+Ph2
Ph
+
203 72% [47a] Ph
PPh2 Ph2+P
N N
Ph Zn Ph
N N
Ph Ph
N N
Ph Zn Ph
N N
PPh2 P+Ph2
+ n Ph n Ph
PPh2 204: n = 1 81% [47b] Ph2P+
N N
n =1-4 205: n = 2 65% [47b] Ph Zn Ph
N N
206: n = 3 70% [47b]
207: n = 4 93% [47b] Ph

Ph
Ph
N N
Ph Zn Ph
N N
PPh2 P+Ph2
N N
Ph
-e- Ph
Ph Zn Ph +
Ph2As+
N N AsPh2 N N
208 60% [47b] Ph Zn Ph
N N

Ph Ph
Ph 201 Ph
Ph N Ph
CH3 N
N N Zn
CH3 N Zn + N
N P Ph2
Ph N
PPh2 Ph
Ph Ph
+ +
P Ph2 P(O)Ph2
PPh2 P(O)Ph2 209 75% [47b]
Ph Ph
Ph N Ph
N
N N Zn
N Zn P+Ph2 N
N Ph N
PPh2 Ph Ph
+ Ph
P+Ph2 Ph2P Ph
+ PPh2 Ph
210 70% [47b] N
PPh2 N
Zn
N
Ph N
Ph Ph
Ph
N N N N
Ph Zn Ph Ph Zn Ph
N N N N
P+Ph2 + PPh2
Ph Ph
Ph2P
212
+
Ph Ph
Ph2P Ph2P+ Ph2P +
N N N N
211 70% [47c] Ph Zn Ph Ph Zn Ph
N N N N

Ph Ph

Scheme 32. β-β Linked porphyrinyl phosphonium salt dimers and trimers [47].

Our group prepared the first porphyrins with a pentavalent phosphorus


fragment directly attached to the carbon framework. The palladium-catalysed
reaction of bromoporphyrin 213 with Ph2PSiMe3 yielded a porphyrinylphosphine

68
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

oxide (214) in 2% yield, while the same reaction with iodoporphyrin 113 led to the
formation of phosphine oxide 214 in 20% yield (Scheme 33) [17].
Ar Ar
Ar = 3,5-di-t-Bu-Ph

NH N NH N Ph Ph
Ph (CH3CN)2PdCl2 from 213 2%
Ph X+ P Si Ph P
Ph from 113 20%
N HN N HN O

Ar 213: X = Br Ar 214
113: X = I

Scheme 33. Formation of diarylporphyrinylphosphine oxide [17].

In order to optimise the reaction condition we utilised η1-palladioporphyrin


55 discussed above. The transition metal porphyrin (55), which is the intermediate of
the proposed palladium-catalysed cycle, was treated with a variety of Ph2P sources
under different conditions (Scheme 34) [48].
Ph Ph

Ph Ph
NH N P NH N Ph Ph
base
Pd P + Ph2P source P
Ph
N HN Br Ph N HN O

Ph2P source base yield


55 Ph2PH - <10% 215
Ph Ph
Ph2PSiMe3 - <10%
Ph2P(BH3)H - 35%
Ph2P(BH3)H DBU <5%
Ph2P(BH3)H Cs2CO3 65%
Ph2P(O)H - 70%
Ph2P(O)H Cs2CO3 85%
DBU = 1,5-diazabicyclo[5.4.0]undec-7-ene

Scheme 34. Reactions of η1-palladioporphyrin with various phosphines and


diphenylphosphine oxide [48].

After determining the optimum reaction conditions, we successfully repeated the


reaction under catalytic conditions. Various bromoporphyrins (50, 91, 216-219) were
reacted with Ph2P(O)H, Cs2CO3 as base and a Pd(dppe)2 catalyst. The resulting
phosphine oxides (215, 220-224) were isolated in good to excellent yields and Zn(II)
porphyrin 223 was also prepared in almost quantitative yield by standard Zn(II)
insertion into the free base analogue 215 (Scheme 35) [48].

69
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Ph Ph

NH N NH N Ph Ph
Pd(dppe)2
Y Br + Ph2P(O)H + Cs2CO3 R P
N HN N HN O
dppe = bis(diphenylphosphino)ethane

Ph Ph
50: M = 2H, Y = H 215: M = 2H, Y = H >95%
91: M = 2H, Y = Br 220: M = 2H, Y = Ph2PO >95%
216: M = Ni, Y = H 221: M = Ni, Y = H >95%
217: M = Ni, Y = Br 222: M = Ni, Y = Ph2PO >95%
218: M = Zn, Y = H 223: M = Zn, Y = H 61%
219: M = Zn, Y = Br 224: M = Zn, Y = Ph2PO 85%

Scheme 35. Palladium-catalysed formation of mono- and bis-porphyrinylphosphine


oxides [48].

The X-ray structure for 222 (Figure 5, left) shows a typical Ni(II) porphyrin ruffled
conformation, and both diphenylphosphine oxide groups project on the same side of
the macrocycle in a cis-like geometry. In the case of the Zn(II)-centred compounds
223 and 224 a phosphine oxide group from one macrocycle coordinated to the Zn(II)
centre of a neighbouring porphyrin, forming a chain-like polymer, which was
confirmed by single crystal X-ray analysis of 224 (Figure 5, right). In this case the
two phosphine oxide groups protrude on opposite sides of the macrocycles in the
trans-like arrangement and the P-O-Zn unit is almost linear (178º). The electron-
withdrawing properties of the phosphine oxide substituent(s) were quantified by
cyclic and square wave voltametry [48].

Figure. 5. X-ray single crystal structure of Ni(II) porphyrinylphosphine oxide 222


(left) and Zn(II) porphyrinylphosphine oxide polymer 224 (right). Reproduced with
permission from Inorg. Chem. 2006; 45: 6479-6489. Copyright 2006 American
Chemical Society.

70
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Matano et al. prepared independently similar porphyrinylphosphine oxides


(228, 230, 232) and porphyrinyl phosphonates (226, 227). In their methodology
Cu(I) iodide was utilised as catalyst and Zn(II) iodoporphyrin as starting material,
because free base porphyrins would form the Cu(II) porphyrins. In order to obtain
the free base analogues (227, 230, 231), the Zn(II) porphyrins (226, 228, 229) were
demetalated under standard acidic conditions (Scheme 36) [49]. Single crystal X-ray
analysis revealed that phosphine oxide 228 existed in this case also as a linear chain-
like polymer while phosphonate 226 has a cofacial, partially overlapping, dimeric
structure in the solid state. The authors proposed that 228 exists as a dyad in solution.
The self-association constant for complex 228 of 5.9 × 106 M-1 is significantly
smaller than values reported for similar porphyrin dimers, for example Zn(II)
imidazolylporphyrins [49].
Ar Ar

N N BuO OBu NH N BuO OBu


+ HP(O)(OBu)2 H+
Ar Zn P Ar P
N N O N HN O
Ar

N N Ar 226 81% Ar 227 85%


CuI
Ar Zn I
MeNH(CH2)2NHMe
N N Ar Ar

O Ph
Ar 225 N N Ph Ph N N P Ph
Ar Zn P + Ar Zn O
+ Ph2P(O)H
N N O N N
Ar = 3,5-di-t-Bu-Ph

Ar 228 72% Ar 229 5%


H+
Ar Ar Ar

O Ph
N N Ph Ph NH N Ph Ph NH N P Ph
Pd(OAc)2
Ar Pd P Ar P Ar O
N N O N HN O N HN

Ar 232 87% Ar 230 Ar 231

Scheme 36. CuI catalysed formation of porphyrinylphosphine oxides, phosphonates


and phosphonic esters [49].

Very recently the coordination properties of porphyrinylphosphine oxides


towards Mg(II) porphyrins were investigated by us. We prepared and characterised
di- (234, 235) and triporphyrin (236, 237) complexes consisting of Mg(II)
diphenylporphyrin 233 coordinated to Ni(II) and free base porphyrinyl mono- (215,
221) and bis- (220, 222) phosphine oxides (Scheme 37) [50]. All complexes were

71
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

isolated in excellent yields (>92%) and were remarkably robust. The phosphine
oxides also stabilised the coordination of the Mg(II) in the porphyrin and the
complexes did not decompose in organic solvents. The solubility of the Mg(II)
macrocycles (234-237) was also significantly increased in comparison to precursor
233.
Ph
Ph
Ph Ph
Ph
H H P
O N N N N
Ph O N
N N N
Ph Mg Ph + M P O M Mg
N N N 234: M = 2H 96%
N
N N Ph N N 235: M = Ni 92%
233
Ph
215: M = 2H Ph
Ph 221: M = Ni
Ph Ph Ph
Ph Ph
Ph
H H P
O N N N
Ph N N Ph N O
N N
N N
2 Ph Mg Ph + O P M P O Mg M Mg
N N N
Ph Ph N O N
N N N
233 N N
P
Ph
220: M = 2H Ph Ph
Ph Ph Ph
222: M = Ni 236: M = 2H 95%
237: M = Ni 95%

Scheme 37. Coordination arrays of Mg(II) diphenylporphyrin with


porphyrinylphosphine oxides [50].

We also isolated the Mg(II) porphyrinyl phosphine oxide 238 by insertion of


Mg(II) into the free base analogue 215 (Scheme 38) [50]. The UV-visible and
fluorescence spectra suggested that the porphyrin exists in a dimeric form (238) in
solution. The self-association constant of the dimer was found to be 5.5 × 108 M-1,
the largest binding constant reported for Mg(II) porphyrins. Thus this complex is the
first strong binding synthetic porphyrin analogue of the “special pair” of the
photosynthetic reaction centre.
Ph Ph

N N
Mg Ph
N N P
NH N Ph
Ph MgBr2
Ph O Ph
P O O

N HN Ph N N
Ph P
Mg
Ph N N

Ph 215 Ph 238 75%

Scheme 38. Mg(II) porphyrinylphosphine oxide [50].

72
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Reported intermediates: Te, Mg, Zn, Ni

Porphyrins of Class F with four other elements have been proposed as


intermediates, but not isolated and characterised. Sugiura et al. prepared a doubly
linked porphyrin dimer 240 via a meso-tellurium porphyrin intermediate 239
(Scheme 39) [51].
Ar

Ar
N N Cl
Ar
Ni TeCl2
N N
N N
2 Ni + 2 TeCl4
N N
N N
Cl2Te Ni
Ar
Cl N N
Ar 39

Ar = 3,5-di-t-Bu-Ph Ar
239
-2 HCl
Ar
Ar

N N Ar
Ar
Ni N N
TeCl2
Ni
N N
N N
N N N N
Ni
Cl2Te Ni
Ar N N
N N
Ar

Ar
Ar 240

Scheme 39. Formation of doubly-fused porphyrin dimer via Te-porphyrin


intermediate [51].

Just recently the nickel-catalysed etheration, amination and C-substitution


reactions on the meso-positions of Ni(II) porphyrins was reported. The authors
proposed a catalytic cycle during which a σ-nickelioporphyrin is formed. Zn(II) and
Cu(II) porphyrins formed the products in lower yields while free base porphyrins did
not undergo the same reaction [52]. The Therien group reported in two patents the
formation and in situ use of bromozinc- and bromomagnesium organometallic
porphyrins. The compounds were utilised for subsequent palladium-catalysed
reactions and a carboxylation reaction, respectively [53].

73
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

CLASS G ORGANOMETALLIC PORPHYRINS

Two reports by the same group describe organometallic fragments directly π-


bonded to the carbon framework of porphyrins [54]. The results were also discussed
by Senge [55]. Both Ni(II) and Zn(II) octaethylporphyrins (199, 241) were treated
with Ru(II), Os(II) and Ir(II) fragments to form cationic η5-organometallic
porphyrins (242-246), while the free base analogues did not form such products
(Scheme 40). The Zn species coordinated one triflate anion to the central zinc. Ni(II)
and Zn(II) η5-Ru(II) porphyrins (242, 243) were unambiguously characterised by
single crystal X-ray analysis. The complexes 242-246 were stable in air but
decomposed slowly in coordinating solvents like acetonitrile. Fluorescence
quenching for the Zn(II) porphyrins (242, 245) demonstrated the large effect on the
metalloporphyrin properties, as did the existence of a long-wavelength absorption
band (680-720 nm).

N N 242: M = Zn(OTf), M' = Ru 67% [54b]


M
N N
243: M = Ni, M' = Ru 60% [54a]
+ [(cymene)M'(F3CSO3)2] 244: M = Ni, M' = Os

M'

N N
cymene = p-isopropyltoluene
M
Cp* = C5Me5
N N N N
M 245: M = Zn(OTf) 64% [54b]
N N
246: M = Ni 60% [54b]

199: M = Zn + [(Cp*)Ir(F3CSO3)2] Ir
241: M = Ni

Scheme 40. Formation of π-organometallic porphyrins (charges, counterions not


shown) [54].

CLASS H ORGANOMETALLIC PORPHYRINS

Smith and co-workers reported the formation of β,β’-fused mono- (253-258)


and bis(metallocene) (261, 262) porphyrins (Scheme 41) [56]. The structures were
confirmed by single crystal X-ray analysis for porphyrin 253 and dimer 258. UV-
visible spectroscopy showed severely broadened and red-shifted absorption bands for
these η5-organometallic porphyrins (253-258) in comparison to the respective
precursors (247-252) and electrochemical studies suggested significant electronic
interaction between the attached organometallic fragment and the porphyrin core.
74
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Diruthenium species 261 and 262 were only detected by MS and attempts to obtain
free base analogues of these metallocenes failed.
Ar
Ar
N N
Ar M Ar
+ Cp*RuCl2 N N
N N
Ru Ar 253: M = Ni, Ar = 3,5-di-t-Bu-Ph 25% [56a,56b]
Ar M Ar
254: M = Cu, Ar = 3,5-di-t-Bu-Ph 14% [56b]
N N
255: M = Zn, Ar = 3,5-di-t-Bu-Ph 2% [56b]
Cp* = C5Me5
Ar
247: M = Ni, Ar = Ph Ar

248: M = Ni, Ar = 3,5-di-t-Bu-Ph N N


Ar M Ar
249: M = Cu, Ar = Ph N N
250: M = Cu, Ar = 3,5-di-t-Bu-Ph + FeCl2
Fe Ar
251: M = Zn, Ar = Ph Ar 256: M = Ni, Ar = Ph 30% [56a,56b]
252: M = Zn, Ar = 3,5-di-t-Bu-Ph N N 257: M = Cu, Ar = Ph 8% [56b]
Ar M Ar
N N 258: M = Zn, Ar = Ph 35% [56b]

Ar
Ar
Ar Ar Ar N
M N Ru
Ru N
N Ar
Ar
N N N N 261
+
Ar M Ar + Ar M Ar + Cp*RuCl2 Ar
N N N N Ar N
M N Ru
Ar = 3,5-di-t-Bu-Ph N Ar
N
Ru Ar
Ar Ar 262
259 260

Scheme 41. Formation of β,β’-fused metalloceneporphyrins [56].

CONCLUSION AND FUTURE OUTLOOK

The majority of the work summarised in the present review stems from
reports during the past decade. We can therefore conclude that investigations into the
formation and reactions of these types of organometallic porphyrins are only in their
infant stage. Nonetheless σ-organometallic porphyrins already play a significant role
in synthetic porphyrin chemistry. Substitution reactions which seemed impossible 15
years ago can easily be carried out today due to the wider knowledge of palladium-
catalysed coupling reactions. The possibility to isolate the intermediate η1-
palladioporphyrins simplifies the understanding and optimisation of these processes.
Porphyrinylboronates are also very valuable tools for such substitution reactions as
can be seen by the variety of functional groups that have been introduced into the
porphyrin moiety with their aid. It is possible today to substitute regiospecifically the
meso- and the β-position of diarylporphyrins via the respective
porphyrinylmercurials, porphyrinylboronates and η1-palladioporphyrins. This leads
to an almost unlimited opportunity of fast functionalisation of the porphyrin carbon

75
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

framework. But organometallic porphyrins are also of significant interest on their


own as more and more interesting and imaginative multiporphyrin arrays
incorporating such fragments emerge. Nonetheless the presented work can only be
seen as the beginning of what we believe is a fascinating and rewarding area of
research. We hope that in the future a greater variety of metal fragments can be
directly attached to the carbon framework of porphyrins, which will undoubtedly
lead to even more stimulating reports in this field. For example, after initial
submission of this review, Osuka and co-workers published a novel application of
their β-borylated porphyrins [41], that led to isolation and complete characterisation
of Class F meso-metallated porphyrin NCN pincer complexes of Pd(II) and Pt(II),
which exhibited interesting catalytic behaviour in the Heck alkenylation reaction
[57].

Acknowledgements

F.A. thanks the Department of Education, Science and Training (DEST) and the
Faculty of Science, Queensland University of Technology for Post-Graduate
Scholarships. We thank all past and present members of our group and our
collaborators as well as all other researchers who have contributed to the knowledge
in this area. Their names can be found in the following list of references.

REFERENCES

1. a) Brothers PJ. Adv. Organomet. Chem. 2001; 46: 223-321. b) Brothers PJ.
Adv. Organomet. Chem. 2001; 48: 289-342. c) Guilard R, Tabard A, van
Caemelbecke, E and Kadish, KM. In The Porphyrin Handbook, Vol. 3,
Kadish KM, Smith KM, Guilard R. (Eds.) Academic Press: San Diego, 2000;
pp 295-345. d) Barbe J-M and Guilard R. In The Porphyrin Handbook, Vol.
3, Kadish KM, Smith KM, Guilard R. (Eds.) Academic Press: San Diego,
2000; pp 211-244.
2. a) Kalita D, Morisue M and Kobuke Y. New J. Chem. 2006; 30: 77-92. b)
Shoji O, Okada S, Satake A and Kobuke Y. J. Am. Chem. Soc. 2005; 127:
2201-2210. c) Koszarna B, Butenschoen H and Gryko DT. Org. Biomol.
Chem. 2005; 3: 2640-2645.
3. a) Latos- Grażyński L. In Porphyrin Handbook, Vol. 2, Kadish KM, Smith
KM, Guilard R. (Eds.) Academic Press: San Diego, 2000; pp 361-416. b)

76
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

Furuta H, Maeda H and Osuka A. Chem. Commun. 2002: 1795-1804. c)


Harvey JD and Ziegler CJ. Coord. Chem. Rev. 2003; 247: 1-19. d) Srinivasan
A and Furuta H. Acc. Chem. Res. 2005; 38: 10-20. e) Gupta I and Ravikanth
M. Coord. Chem. Rev. 2006; 250: 468-518. f) Chmielewski PJ and Latos-
Grażyński L. Coord. Chem. Rev. 2005; 249: 2510-2533.
4. a) Smith KM and Langry KC. J. Chem. Soc., Chem. Commun. 1980: 217-
218. b) Smith KM and Langry KC. J. Chem. Soc., Chem. Commun. 1981:
283-284. c) Smith KM and Langry KC. J. Org. Chem. 1983; 48: 500-506. d)
Smith KM, Langry KC and Minnetian OM. J. Org. Chem. 1984; 49: 4602-
4609. e) Smith KM, Miura M and Morris IK. J. Org. Chem. 1986; 51: 4660-
4667. f) Minnetian OM, Morris IK, Snow KM and Smith KM. J. Org. Chem.
1989; 54: 5567-5574. g) Morris IK, Snow KM, Smith NW and Smith KM. J.
Org. Chem. 1990; 55: 1231-1236.
5. Miura M, Gabel D, Oenbrink G and Fairchild RG. Tetrahedron Lett. 1990;
31: 2247-2250.
6. Castella M, Trull FR, Lopez Calahorra F, Velasco D and Gonzalez MM.
Tetrahedron 2000; 56: 4017-4025.
7. Buchler JW and Herget G. Z. Naturforsch., B: Chem. Sci. 1987; 42: 1003-
1008.
8. Sugiura K-i, Kato A, Iwasaki K, Miyasaka H, Yamashita M, Hino S and
Arnold DP. Chem. Commun. 2007: DOI: 10.1039/b618464b.
9. Handbook of Organopalladium Chemistry for Organic Synthesis, Vols. 1, 2;
Negishi E-i. (Ed.) John Wiley & Sons, Inc., 2002.
10. a) Sharman WM and van Lier JE. J. Porphyrins Phthalocyanines 2000; 4:
441-453. b) Setsune J-i. J. Porphyrins Phthalocyanines 2004; 8: 93-102.
11. a) Arnold DP, Sakata Y, Sugiura K-i and Worthington EI. J. Chem. Soc.,
Chem. Commun. 1998: 2331-2332. b) Arnold DP, Healy PC, Hodgson MJ
and Williams ML. J. Organomet. Chem. 2000; 607: 41-50. c) Hodgson MJ,
Healy PC, Williams ML and Arnold DP. J. Chem. Soc., Dalton Trans. 2002:
4497-4504. d) Hartnell RD, Edwards AJ and Arnold DP. J. Porphyrins
Phthalocyanines 2002; 6: 695-707. e) Hartnell RD and Arnold DP. Eur. J.
Inorg. Chem. 2004: 1262-1269. f) Hodgson MJ, Borovkov VV, Inoue Y and
Arnold DP. J. Organomet. Chem. 2006; 691: 2162-2170.

77
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

12. Hodgson MJ. M.App.Sc.(Research); Queensland University of Technology,


2005, p 128.
13. Jiang Y, Patrick B and Dolphin D. J. Porphyrins Phthalocyanines 2006; 10:
692.
14. Senge MO. Acc. Chem. Res. 2005; 38: 733-743.
15. Hartnell RD and Arnold DP. Organometallics 2004; 23: 391-399.
16. Kato A, Hartnell RD, Yamashita M, Miyasaka H, Sugiura K-i and Arnold
DP. J. Porphyrins Phthalocyanines 2004; 8: 1222-1227.
17. Atefi F, Locos OB, Senge MO and Arnold DP. J. Porphyrins
Phthalocyanines 2006; 10: 176-185.
18. Fox S and Boyle RW. Chem. Commun. 2004: 1322-1323.
19. Cammidge AN, Scaife PJ, Berber G and Hughes DL. Org. Lett. 2005; 7:
3413-3416.
20. Locos OB and Arnold DP. Org. Biomol. Chem. 2006; 4: 902-916.
21. Sahoo AK, Mori S, Shinokubo H and Osuka A. Angew. Chem., Int. Ed. 2006;
45: 7972-7975.
22. Suzuki A. In Handbook of Organopalladium Chemistry for Organic
Synthesis, Vol. 1, Negishi E-i. (Ed.) John Wiley & Sons, Inc., 2002; pp 249-
262.
23. a) Hyslop AG, Kellett MA, Iovine PM and Therien MJ. J. Am. Chem. Soc.
1998; 120: 12676-12677. b) Zhang T-G, Zhao Y, Asselberghs I, Persoons A,
Clays K and Therien MJ. J. Am. Chem. Soc. 2005; 127: 9710-9720.
24. a) Aratani N and Osuka A. Org. Lett. 2001; 3: 4213-4216. b) Tsuda A,
Nakamura T, Sakamoto S, Yamaguchi K and Osuka A. Angew. Chem., Int.
Ed. 2002; 41: 2817-2821. c) Peng X, Nakamura Y, Aratani N, Kim D and
Osuka A. Tetrahedron Lett. 2004; 45: 4981-4984.
25. a) Sazanovich IV, Balakumar A, Muthukumaran K, Hindin E, Kirmaier C,
Diers JR, Lindsey JS, Bocian DF and Holten D. Inorg. Chem. 2003; 42:
6616-6628. b) Yu L, Muthukumaran K, Sazanovich IV, Kirmaier C, Hindin
E, Diers JR, Boyle PD, Bocian DF, Holten D and Lindsey JS. Inorg. Chem.
2003; 42: 6629-6647.
26. Felber B, Calle C, Seiler P, Schweiger A and Diederich F. Org. Biomol.
Chem. 2003; 1: 1090-1093.

78
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

27. a) Chang CJ, Chng LL and Nocera DG. J. Am. Chem. Soc. 2003; 125: 1866-
1876. b) Chng LL, Chang CJ and Nocera DG. J. Org. Chem. 2003; 68: 4075-
4078.
28. Hasobe T, Imahori H, Yamada H, Sato T, Ohkubo K and Fukuzumi S. Nano
Letters 2003; 3: 409-412.
29. Wada K, Mizutani T, Matsuoka H and Kitagawa S. Chem.--Eur. J. 2003; 9:
2368-2380.
30. Cheng F, Zhang S, Adronov A, Echegoyen L and Diederich F. Chem.--Eur. J.
2006; 12: 6062-6070.
31. a) Muraoka T, Kinbara K and Aida T. J. Am. Chem. Soc. 2006; 128: 11600-
11605. b) Muraoka T, Kinbara K and Aida T. Nature 2006; 440: 512-515.
32. a) Iovine PM, Kellett MA, Redmore NP and Therien MJ. J. Am. Chem. Soc.
2000; 122: 8717-8727. b) Kang YK, Rubtsov IV, Iovine PM, Chen J and
Therien MJ. J. Am. Chem. Soc. 2002; 124: 8275-8279.
33. Felber B and Diederich F. Helv. Chim. Acta 2005; 88: 120-153.
34. Chng LL, Chang CJ and Nocera DG. Org. Lett. 2003; 5: 2421-2424.
35. Yamane O, Sugiura K-i, Miyasaka H, Nakamura K, Fujimoto T, Nakamura
K, Kaneda T, Sakata Y and Yamashita M. Chem. Lett. 2004; 33: 40-41.
36. Cheng F and Adronov A. Chem. Mater. 2006; 18: 5389-5391.
37. a) Peng X, Aratani N, Takagi A, Matsumoto T, Kawai T, Hwang I-W, Ahn
TK, Kim D and Osuka A. J. Am. Chem. Soc. 2004; 126: 4468-4469. b) Hori
T, Aratani N, Takagi A, Matsumoto T, Kawai T, Yoon M-C, Yoon ZS, Cho
S, Kim D and Osuka A. Chem.--Eur. J. 2006; 12: 1319-1327.
38. Hasobe T, Imahori H, Ohkubo K, Yamada H, Sato T, Nishimura Y,
Yamazaki I and Fukuzumi S. J. Porphyrins Phthalocyanines 2003; 7: 296-
312.
39. Deng Y, Chang CK and Nocera DD. Angew. Chem., Int. Ed. 2000; 39: 1066-
1068.
40. Bringmann G, Ruedenauer S, Goetz DCG, Gulder TAM and Reichert M.
Org. Lett. 2006; 8: 4743-4746.
41. Hata H, Shinokubo H and Osuka A. J. Am. Chem. Soc. 2005; 127: 8264-
8265.
42. Zhang C and Suslick KS. J. Porphyrins Phthalocyanines 2005; 9: 659-666.

79
Copyright © 2007 Society of Porphyrins & Phthalocyanines
Chapter 3

43. a) Evans B and Smith KM. Tetrahedron Lett. 1977: 3079-3082. b) Smith
KM, Barnett GH, Evans B and Martynenko Z. J. Am. Chem. Soc. 1979; 101:
5953-5961.
44. a) Padilla AG, Wu S-M and Shine HJ. J. Chem. Soc., Chem. Commun. 1976:
236-237. b) Shine HJ, Padilla AG and Wu S-M. J. Org. Chem. 1979; 44:
4069-4075.
45. Giraudeau A and El Kahef L. Can. J. Chem. 1991; 69: 1161-1165.
46. Małek A, Latos- Grażyński L, Bartczak TJ and Żądło A. Inorg. Chem. 1991;
30: 3222-3230.
47. a) Ruhlmann L and Giraudeau A. J. Chem. Soc., Chem. Commun. 1996:
2007-2008. b) Ruhlmann L and Giraudeau A. Eur. J. Inorg. Chem. 2001:
659-668. c) Ruhlmann L, Gross M and Giraudeau A. Chem.--Eur. J. 2003; 9:
5085-5096.
48. Atefi F, McMurtrie JC, Turner P, Duriska M and Arnold DP. Inorg. Chem.
2006; 45: 6479-6489.
49. Matano Y, Matsumoto K, Terasaka Y, Hotta H, Araki Y, Ito O, Shiro M,
Sasamori T, Tokitoh N and Imahori H. Chem.--Eur. J. 2007; 13: 891-901.
50. Atefi F, McMurtrie JC and Arnold DP. J. Chem. Soc., Dalton Trans. 2007:
2163-2170.
51. Sugiura K-i, Matsumoto T, Ohkouchi S, Naitoh Y, Kawai T, Takai Y,
Ushiroda K and Sakata Y. J. Chem. Soc., Chem. Commun. 1999: 1957-1958.
52. Liu C, Shen D-M and Chen Q-Y. J. Org. Chem. 2007; 72: 2732-2736.
53. a) Therien MJ and DiMagno SG. USA Patent WO 94/04614, 1994. b)
Therien MJ. USA Patent WO 02/104072, 2002.
54. a) Koczaja Dailey K, Yap GPA, Rheingold AL and Rauchfuss TB. Angew.
Chem., Int. Ed. Eng. 1996; 35: 1833-1835. b) Dailey KK and Rauchfuss TB.
Polyhedron 1997; 16: 3129-3136.
55. Senge MO. Angew. Chem., Int. Ed. Eng. 1996; 35: 1923-1925.
56. a) Wang HJH, Jaquinod L, Nurco DJ, Vicente MGH and Smith KM. Chem.
Commun. 2001: 2646-2647. b) Wang HJH, Jaquinod L, Olmstead MM,
Vicente MGH, Kadish KM, Ou Z and Smith KM. Inorg. Chem. 2007; 46:
2898-2913.
57. Yamaguchi S, Katoh T, Shinokobu H and Osuka A. J. Am. Chem. Soc. 2007;
129: 6392-6393.

80
Copyright © 2007 Society of Porphyrins & Phthalocyanines
CHAPTER 4

meso-Iodo- and meso-iodovinylporphyrins via


organopalladium porphyrins and the crystal
structure of 5-iodo-10,20-diphenylporphyrin

81
Chapter 4

Statement of Contribution

The authors listed below have certified* that:


6) they meet the criteria for authorship in that they have participated in the
conception, execution, or interpretation, of at least that part of the publication in
their field of expertise;
7) they take public responsibility for their part of the publication, except for the
responsible author who accepts overall responsibility for the publication;
8) there are no other authors of the publication according to these criteria;
9) potential conflicts of interest have been disclosed to (a) granting bodies, (b) the
editor or publisher of journals or other publications, and (c) the head of the
responsible academic unit, and
10) they agree to the use of the publication in the student’s thesis and its publication
on the Australasian Digital Thesis database consistent with any limitations set by
publisher requirements.
In the case of this chapter:
meso-Iodo- and meso-iodovinylporphyrins via organopalladium porphyrins and
the crystal structure of 5-iodo-10,20-diphenylporphyrin
Atefi F, Locos OB, Senge MO and Arnold DP. J. Porphyrins Phthalocyanines 2006;
10: 176-185.
Contributor Statement of contribution*
Synthesised compounds 3,9-13; all characterisation of these
compounds except: single crystal X-ray analysis (see below), mass
spectra and microanalysis (independent providers); grew X-ray
quality single crystals; analysed data; wrote majority of manuscript
Oliver B. Locos* Synthesised compounds 14-18 following methodology developed
by Farzad Atefi, characterised these compounds except: mass
spectra and microanalysis (independent providers); wrote part of
the manuscript about these compounds
Mathias O. Senge* Collected data for singe crystal X-ray analysis and wrote part of
the manuscript about crystal structure
Dennis P. Arnold* Overall supervisor of the project, edited manuscript

Principal Supervisor Confirmation

I have sighted email or other correspondence from all Co-authors confirming their
certifying authorship and.

82
Chapter 4

meso-Iodo- and meso-iodovinylporphyrins via organopalladium porphyrins and


the crystal structure of 5-iodo-10,20-diphenylporphyrin

Farzad Atefia◊◊, Oliver B. Locosa, Mathias O. Sengeb◊ and Dennis P. Arnold*a◊

a
Synthesis and Molecular Recognition Program, School of Physical and Chemical
Sciences, Queensland University of Technology, G.P.O. Box 2434, Brisbane 4001,
Australia
b
School of Chemistry, SFI Tetrapyrrole Laboratory, Trinity College Dublin, Dublin
2, Ireland

Received 28 April 2006


Accepted 7 June 2006


SPP full member or ◊◊ student member in good standing

*Correspondence to: Dennis P. Arnold, email: d.arnold@qut.edu.au, fax: +61 7 3864


1804

ABSTRACT: The regiospecific halogen exchange reactions of various easily


accessible meso-bromoporphyrins to obtain meso-iodoporphyrins, using η1-
palladioporphyrins as intermediates, have been investigated. This one-pot
methodology allows the isolation of meso-iodoporphyrins in excellent yields with
short reaction times. Similarly meso-(2-bromoethenyl)porphyrins can be converted to
their iodoethenyl analogues. These iodoporphyrins show great potential as starting
materials for various palladium catalyzed reactions. The X-ray crystal structure of 5-
iodo-10,20-diphenylporphyrin has been determined.

KEYWORDS: porphyrinoids, iodination, palladium, organometallic, crystal


structure.

83
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

INTRODUCTION
In the past decade the palladium catalyzed formation of new C-C and C-X
bonds has emerged as a versatile method for the meso-functionalization of various
porphyrins. Due to the selectivity and mild reaction conditions a variety of otherwise
challenging reactions can now be accomplished with relative ease, leading to novel
substitution patterns and interesting multiporphyrin arrays [1]. The intermediates of
these reactions are meso-η1-palladioporphyrins, which were first isolated in 1998 [2].
Since then we have reported numerous examples of these novel organopalladium
porphyrins and their more robust platinum(II) analogues [3]. Bromoporphyrins have
been the preferred starting materials for palladium catalyzed reactions, as the NBS
bromination of unsubstituted porphyrins can be accomplished with short reaction
times [4]. In the case of porphyrins substrates with just one unsubstituted meso-
position, mono-bromoporphyrins can be isolated with almost quantitative yield [3b],
while with a slight excess of NBS, porphyrins with two vacant meso-positions afford
the 5,15-dibromoporphyrins quantitatively [4]. The lack of selectivity of the NBS
bromination of porphyrins with two vacant meso-positions has been overcome
recently by our group with the aid of organopalladium porphyrins [5].
Iodide on the other hand is known to be a better leaving group than bromide
in transition metal catalyzed reactions [6]. The oxidative addition of the zerovalent
palladium fragment, often the first step of the catalytic cycle, occurs much more
readily with iodides than with bromides, leading to faster reactions times, higher
yields and selective reactions at the iodinated site, if both halogens are present. In
some cases the desired substitution reaction cannot be accomplished at all, due to the
less polarisable C-Br bond [7]. It is therefore no surprise that often iodoporphyrins
have to be utilized in order to accomplish palladium catalyzed substitution reactions.
Examples of alkynes [8], organoboranes [8d, 8g, 9], organotins [8l, 10], alkenes [8d,
8l] and organozincs [11] reacting with meso-iodoporphyrins as starting materials can
be found in the literature.
Unfortunately the selective meso-iodination of porphyrins cannot be as
rapidly and selectively accomplished as the meso-bromination. Only two practical
methods for the meso-iodination of porphyrins are described in the literature, both
with their respective limitations. Boyle and co-workers [8a, 8b, 8h] introduced a
hypervalent I(III) reagent to accomplish the iodination of porphyrins, and most other

84
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

users of iodoporphyrins have used this method. Although excellent yields can be
achieved, long reaction times of up to 48 hours are required and unselective
additional iodination in β-positions can occur. The porphyrin cation radical method
reported by Osuka’s group [8c] on the other hand yields selectively meso-iodinated
porphyrins with very short reaction times. Unfortunately lower yields and the
necessity of zinc porphyrins as starting materials limit this method.
We were seeking to accomplish the selective meso-iodination of 5,15-diaryl-
and 5,10,15-triarylporphyrins with short reaction times, in high yields and with the
possibility to utilize free-base porphyrins as starting materials. Here we wish to
report such a methodology using the stoichiometric amounts of organopalladium
complexes as intermediates for the selective meso-iodination of porphyrins. This
method has been extended to the formation of meso-(2-iodoethenyl)porphyrins,
useful starting materials that have previously been prepared only for Ni(II) octaethyl-
and tetraphenylporphyrin [12].

EXPERIMENTAL

General

All synthetic reactions involving zerovalent palladium metal reagents were


carried out in an atmosphere of high-purity argon using standard Schlenk techniques
[13]. Chemical reagents and ligands were of laboratory reagent (LR) or analytical
reagent (AR) grade, received from Sigma-Aldrich and used without further
purification. Solvents used were AR grade where available. Solvents used in
palladium catalyzed reactions were degassed by freeze/thawing under vacuum three
times prior to their use. Toluene and diethyl ether were stored over sodium wire.
CH2Cl2 and CHCl3 were stored over anhydrous sodium carbonate. Tetrahydrofuran
(THF) was distilled immediately before use from a dark blue sodium/benzophenone
solution under an atmosphere of high purity argon. Bromoporphyrins 2 [8h], 5 [3b],
6 [3b], 7 [3c] and 8 [14], η1-palladioporphyrin 1 [2], ethenylporphyrin 14 [15] and
Pd(PPh3)4 [16] were synthesized according to the literature. Recrystallization from
two solvents was accomplished by dissolving the product in a minimum amount of
the first solvent and carefully layering the solution with a tenfold excess of the
second solvent. Analytical TLC was performed using aluminium backed Merck silica
gel 60 F254 plates and column chromatography was carried out using silica gel (230-

85
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

400 mesh) from Qingdao Haiyang Chemical Co. Ltd., Qingdao, China. NMR spectra
were recorded on a Bruker Avance 400 MHz instrument with CHCl3 as internal
standard at 7.26 ppm for 1H spectra and 85% aqueous H3PO4 as external standard for
protondecoupled 31P spectra. Quantitative UV-vis spectra were recorded on a Varian
Cary 3 spectrometer. High accuracy ESI mass spectra were recorded on a Bruker
BioApex 47e FTMS fitted with an Analytica Electrospray source. Dichloromethane
was used as a solvent and the samples were diluted either with
dichloromethane/methanol or just methanol. The samples were introduced into the
source by direct infusion (syringe pump) at 60 μL/h with a capillary voltage of 80 V.
Sodium iodide clusters were used as internal standard for the calibration of the
instrument. Laser Desorption/Ionization (LDI) MS analysis was performed with an
Applied Biosystems Voyager-DESTR BioSpectrometry Workstation. The instrument
was operated in positive polarity in reflectron mode. Samples were spotted on a
stainless steel sample plate and allowed to air dry. Data from 500 laser shots (337 nm
nitrogen laser) were collected, signal averaged, and processed with the instrument
manufacturer's Data Explorer software. Isotopic modelling was performed using
MassLynx V3.5 Software by Micromass Limited. Elemental analyses were carried
out by the Microanalytical Service, The University of Queensland.

Synthesis

5-iodo-10,20-diphenylporphyrin (3) from η1-palladioporphyrin (1).


Palladioporphyrin 1 (20 mg, 0.019 mmol) was dissolved in toluene (6 mL) and a
solution of iodine in toluene (5 mg, 0.020 mmol, 2 mL) was added dropwise to the
reaction mixture. After stirring for 10 min at room temperature, the solvent was
removed under vacuum and the product was purified on a short column packed with
silica gel using CH2Cl2/hexane/triethylamine (400:200:3) as eluent and recrystallized
from CH2Cl2/MeOH to give dark purple crystals (10.2 mg, 91%). By changing the
eluent to CH2Cl2, the palladium(II) complexes 4 can be isolated. 1H NMR (400 MHz,
CDCl3, 25 °C): δ, ppm 10.18 (s, 1H, meso-H), 9.76, 9.29, 8.95, 8.92 (d, each 2H, J =
4.7 Hz, β-H), 8.22-8.20 (m, 4H, o-H), 7.78-7.76 (m, 6H, m,p-H), -3.01 (br s, 2H,
NH). UV-vis (CH2Cl2): λmax, nm (ε, 103 M-1.cm-1) 417 (317), 513 (14.9), 546 (5.1),
588 (4.4), 644 (2.2). High-resolution ESI MS: m/z 589.0884 ([C32H21IN4 + H]+
requires 589.0889).

86
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

General procedure for the one-pot iodination

The bromoporphyrin was added to a Schlenk vessel, dried under high vacuum
and subsequently the vessel was charged with argon. Toluene was added and the
solvent was deoxygenated by freeze/thawing three times before heating to 105 °C.
The appropriate amounts of Pd2dba3/PPh3 or freshly prepared Pd(PPh3)4 were added
in portions and the reaction was monitored by TLC. After the disappearance for the
spot representing the haloporphyrin, the mixture was cooled to room temperature and
diluted with toluene. At this stage the Schlenk vessel was opened to air and the
appropriate amount of iodine was added as a solid. The solvent was removed under
vacuum and the product was purified on a short silica gel column using
CH2Cl2/hexane/triethylamine (100:300:2) as eluent and recrystallized from
CH2Cl2/MeOH.
5-iodo-10,20-diphenylporphyrin (3) from bromoporphyrin (2).
Bromoporphyrin 2 (10 mg, 0.018 mmol), Pd2dba3 (22 mg, 0.024 mmol) and PPh3 (25
mg, 0.095 mmol) gave after stirring in 5 mL toluene at 105 °C for 2 h, addition of
toluene (1.25 mL) and iodine (6.4 mg, 0.025 mmol), the desired product as a dark
purple crystalline powder (8.5 mg, 78%).
5-iodo-10,15,20-triphenylporphyrin (9). Bromoporphyrin 5 (25 mg, 0.040
mmol), Pd2dba3 (37 mg, 0.041 mmol) and PPh3 (43 mg, 0.164 mmol) gave after
stirring in 10 mL toluene at 105 °C for 2 h, addition of toluene (7 mL) and iodine (20
mg, 0.079 mmol), the desired product as a dark purple crystalline powder (21 mg,
79%). The NMR data agreed with those of an authentic sample previously prepared
by our group [3b].
5-iodo-10,20-bis(3’,5’-di-tert-butylphenyl)-15-phenylporphyrin (10).
Bromoporphyrin 6 (25 mg, 0.030 mmol), Pd2dba3 (28 mg, 0.031 mmol) and PPh3 (32
mg, 0.122 mmol) gave after stirring in 15 mL toluene at 105 °C for 2.5 h, addition of
toluene (10 mL) and iodine (19 mg, 0.075 mmol), the desired product as a dark
purple crystalline powder (21 mg, 75%). 1H NMR (400 MHz, CDCl3, 25 °C): δ, ppm
9.68, 8.91, 8.82, 8.80 (d, each 2H, J = 4.6 Hz, β-H), 8.22-8.18 (m, 2H, o-H on 15-
phenyl), 8.05 (d, 4H, J = 1.7 Hz, 2,6-H on 5,15-aryl), 7.82 (t, 2H, J = 1.7 Hz, 4-H on
5,15-aryl), 7.77-7.71 (m, 3H, m,p-H on 15-phenyl), 1.56 (s, 36H, t-butyl), -3.01 (br s,
2H, NH). UV-vis (CH2Cl2): λmax, nm (ε, 103 M-1.cm-1) 423 (311), 522 (14.6), 557
(8.3), 597 (4.7), 654 (4.4). High-resolution ESI MS: m/z 889.3705 ([C54H57IN4 + H]+

87
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

requires 889.3706). Anal. calcd., %: C, 72.96; H, 6.46; N, 6.30. Found: C, 73.34; H,


6.43; N, 6.00.
5-bromo-15-iodo-10,20-diphenylporphyrin (11). Dibromoporphyrin 7 (31
mg, 0.050 mmol) and Pd(PPh3)4 (60 mg, 0.052 mmol) gave after stirring in 30 mL
toluene at 105 °C for 5 h, addition of toluene (30 mL) and iodine (63 mg, 0.248
mmol), the desired product as a dark purple crystalline powder (26 mg, 78%). The
NMR data agreed with those in the literature [8h].
5-bromo-15-iodo-10,20-bis(3’,5’-di-tert-butyl)-phenylporphyrin (12).
Dibromoporphyrin 8 (42 mg, 0.050 mmol) and Pd(PPh3)4 (58 mg, 0.050 mmol) gave
after stirring in 30 mL toluene at 105 °C for 4h, addition of toluene (30 mL) and
iodine (64 mg, 0.252 mmol), the desired product as a dark purple crystalline powder
(36 mg, 81%). 1H NMR (400 MHz, CDCl3, 25 °C): δ, ppm 9.64-9.61 (m, 4H, β-H),
8.88-8.83 (m, 4H, β-H), 8.02 (d, 4H, J = 1.9 Hz, 2,6-H on 10,20-aryl), 7.83 (t, 2H, J
= 1.8 Hz, 4-H on 10,20-aryl), 1.55 (s, 36H, t-butyl), -2.63 (br s, 2H, NH). UV-vis
(CH2Cl2): λmax, nm (ε, 103 M-1.cm-1) 425 (252), 524 (10.8), 560 (8.7), 604 (3.1), 661
(4.3). High-resolution ESI MS: m/z 891.2492 ([C48H52BrIN4 + H]+ requires
891.2498). Anal. calcd., %: C, 64.65; H, 5.88; N, 6.28. Found: C, 64.63; H, 5.91; N
5.98.
[10,20-bis(3’,5’-di-tert-butylphenyl)-15-phenylporphyrin-5-
yl]diphenylphosphine oxide (13). Iodoporphyrin 10 (10 mg, 0.0112 mmol) and
(CH3CN)2PdCl2 (0.3 mg, 0.0012 mmol) were dried under high vacuum in a Schlenk
vessel which was subsequently charged with argon. After the addition of toluene (1
mL), the reaction mixture was heated to 80 °C, before adding
diphenyl(trimethylsilyl)-phosphine (6 μL, 0.0235 mmol) by syringe. The solvent was
removed under vacuum after 72 h, and the residue purified on a silica gel column
using CHCl3/triethylamine (200:1) as eluent. The major less polar fraction was
identified as a mixture of iodoporphyrin 10 and [10,20-bis(3’,5’-di-tert-butylphenyl)-
15-phenylporphyrin. After changing the eluent to a mixture of
CHCl3/MeOH/triethylamine 196:4:1) the crude target compound was collected. This
product was dissolved successively in a series of solvents (MeOH, diethyl ether,
CHCl3, pentane) every time filtering off unwanted solids, then recovering by
evaporation. The porphyrinylphosphine oxide 13 was finally isolated as dark red
powder (2.2 mg, 20%). The desired target compound can also be isolated under
similar reaction conditions in only 2% yield if starting from bromoporphyrin 6. 1H

88
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

NMR (400 MHz, CDCl3, 25 °C): δ, ppm 9.42 (br d, 2H, J = 4.7 Hz, β-H), 8.77, 8.71,
8.63 (d, each 2H, J = 4.7 Hz, β-H), 8.19-8.17 (m, 2H, o-H on 15-phenyl), 7.97 (d,
4H, J = 1.7 Hz, 2,6-H on 10,20-aryl), 7.95-7.90 (m, 4H, o-H PPh2), 7.76 (t, 2H, J =
1.7 Hz, 4-H on 10,20-aryl), 7.77-7.70 (m, 3H, m,p-H 15-phenyl), 7.55-7.50 (m, 2H,
p-H PPh2), 7.45-7.40 (m, 4H, m-H PPh2), 1.49 (s, 36H, t-butyl), -2.04 (br s, 2H, NH);
31
P NMR (162 MHz, CDCl3, 25 °C): δ, ppm 34.6. UV-vis (CH2Cl2): λmax, nm (ε, 103
M-1.cm-1) 423 (217), 521 (10.2), 557 (6.1), 592 (3.8), 649 (3.5). High-resolution ESI
MS: m/z 963.5141 ([C66H67N4OP + H]+ requires 963.5131).
(E)-5-(2-bromoethenyl)-10,15,20-triphenylporphyrin (15). Porphyrin 14
(150 mg, 0.27 mmol) and pyridinium tribromide (91 mg, 0.28 mmol) were added to a
50 mL flask, dissolved in THF (20 mL) and allowed to stir at room temperature
overnight. The solution turned bright green, due to protonation of the free-base
porphyrin. After 14 h, triethylamine (1 mL) was added, changing the color of the
solution to purple. The solvent was removed under vacuum and the residue was then
loaded onto a plug of silica gel using CH2Cl2/hexane/triethylamine (100:100:1) as
eluent. The porphyrin band was collected and the residue recrystallized from
CH2Cl2/MeOH to give the product as a purple crystalline powder (141 mg; 81%). 1H
NMR (400 MHz, CDCl3, 25 °C): δ, ppm 9.61 (d, 1 H, J = 13.9 Hz, alkene-H), 9.43,
8.95 (d, each 2 H, J = 4.9 Hz, β-H), 8.83 (br s, 4 H, β-H), 8.22 (m, 2 H, o-H ), 7.79
(m, 9 H, m, p-H), 7.10 (d, 1 H, J = 13.9 Hz, alkene-H), -2.70 (br s, 2 H, NH). UV-vis
(CH2Cl2): λmax, nm (ε, 103 M-1.cm-1) 420 (151), 520 (7.2), 555 (4.4), 594 (1.6), 651
(1.2). LDI MS: m/z 642.14 ([C40H27BrN4]+ requires 642.14). Anal. calcd., %: C,
74.65; H, 4.23; N, 8.71; Found C, 74.71; H, 4.17; N, 8.56.
(E)-5-(2-iodoethenyl)-10,15,20-triphenylporphyrin (16).
Bromoethenylporphyrin 15 (50 mg, 0.078 mmol) was dissolved in dry toluene (10
mL) in a Schlenk vessel. The solvent was deoxygenated by freeze/thawing three
times and purged with argon. Freshly prepared Pd(PPh3)4 (100 mg, 0.087 mmol) was
added and the reaction mixture was allowed to stir for 1.5 hours at room temperature.
Iodine (50 mg, 0.20 mmol), triethylamine (25 μL, 0.2 mmol) and toluene (10 mL)
were added and the reaction was allowed to stir for 14 hours overnight, open to air.
The reaction was transferred to a separating funnel and washed with Na2S2O3
solution (1 M, 50 mL × 3) and dried over Na2SO4. The solvent was removed under
vacuum and the residue was purified on a silica gel column using
CH2Cl2/hexane/triethylamine (300:100:2) as eluent. The porphyrin band was

89
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

collected, dried and recrystallized to give a purple crystalline powder (50 mg, 93%).
1
H NMR (400 MHz, CDCl3, 25 °C): δ, ppm 9.92 (d, 1 H, J = 14.7 Hz, alkene-H),
9.39, 8.92 (d, each 2 H, J = 4.9 Hz, β-H), 8.81 (br s, 4 H, β-H), 8.18 (m, 2 H, o-H),
8.20 (m, 9 H, m,p-H), 7.19 (d, 1 H, J = 14.7 Hz, alkene-H), -2.70 (br s, 2 H, NH);
UV-vis (CH2Cl2): λmax, nm (ε, 103 M-1.cm-1) 422 (151), 519 (7.6), 557 (4.9), 595
(3.0), 651 (2.6). LDI MS: m/z 691.14 ([C40H27IN4 + H]+ requires 691.14).
(E)-5-(2-iodoethenyl)-10,15,20-triphenylporphyrinatonickel(II) (17). Free
base iodoethenylporphyrin 16 (50 mg, 0.066 mmol) and Ni(acac)2 (18 mg, 0.071
mmol) were added to a 50 mL flask, dissolved in toluene (10 mL) and brought to
reflux. The progress of the reaction was monitored by TLC using CH2Cl2/hexane
(1:1) as eluent, which indicated after 3 h that the starting porphyrin was consumed.
The solvent was removed under vacuum and the product was isolated after passage
through a plug of silica gel using CHCl3 as the eluent. Recrystallization of the
product from CHCl3/MeOH yielded the product as a red crystalline powder (109 mg,
99%). 1H NMR (400 MHz, CDCl3, 25 °C): δ, ppm 9.61 (d, 1 H, J = 14.7 Hz, alkene-
H), 9.25, 8.81, 8.69, 8.66 (d, each 2 H, J = 4.9 Hz, β-H), 7.98 (m, 2 H, o-H), 7.68 (m,
9 H, m,p-H), 6.71 (d, 1 H, J = 14.7 Hz, alkene-H). UV-vis (CH2Cl2): λmax, nm (ε, 103
M-1.cm-1) 419 (145), 535 (11.2), 571 (4.2). LDI MS: m/z 769.04 ([C40H25IN4Ni +
Na]+ requires 769.04).
(E)-5-(2-iodoethenyl)-10,15,20-triphenylporphyrinatozinc(II) (18). Free
base iodoethenylporphyrin 16 (20 mg, 0.029 mmol) was dissolved in CHCl3 (5 mL)
and brought to reflux. Zinc(II) acetate (20 mg, 0.1 mmol) was dissolved in MeOH
(0.5 mL) and added dropwise through the condenser to the refluxing mixture. The
progress of the reaction was monitored by TLC using CHCl3/hexane (1:1) as eluent.
The heat source was removed after 15 minutes and the solution was allowed to cool.
The solvents were removed under vacuum and the product was isolated after passage
through a plug of silica gel using CHCl3 as the eluent. Recrystallization of the purple
residue using CHCl3 (1% pyridine)/MeOH yielded the product as a bright purple
crystalline powder (21 mg, 90%). 1H NMR (400 MHz, CDCl3 + 1% d5-pyridine, 25
°C): δ, ppm 10.00 (d, 1 H, J = 14.7 Hz, alkene-H), 9.42, 8.93 (d, each 2 H, J = 4.6
Hz, β-H), 8.81 (br s, 4 H, β-H), 8.17 (m, 6 H, o-H), 7.8 (m, 9 H, m,p-H), 7.05 (d, 1 H,
J = 14.7 Hz, alkene-H). UV-vis (CH2Cl2): λmax, nm (ε, 103 M-1.cm-1) 425 (151), 555
(9.1), 608 (4.8). Highresolution ESI MS: m/z 775.0304 ([C40H25IN4Zn + Na]+
requires 775.0313).

90
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

X-ray single crystal structure determination of 5-iodo-10,20-diphenylporphyrin


(3)

The crystals were obtained by recrystallizing the product from CHCl3/MeOH.


Growth and handling of crystals followed the concept developed by Hope [17].
Intensity data were collected at 90 K with a Bruker SMART Apex2 system complete
with 3-circle goniometer and CCD detector utilizing Mo-Kα radiation (λ = 0.71073
Å). The intensities were corrected for Lorentz, polarization, absorption and
extinction effects. The structure was solved with direct methods using the SHELXTL
PLUS program system [18] and refined against |F2| with the program XL from
SHELX-97 using all data [19]. Nonhydrogen atoms were refined with anisotropic
thermal parameters. Hydrogen atoms were generally placed into geometrically
calculated positions and refined using a riding model. Residual electron density was
located in the middle of the porphyrin ring system indicating the presence of a
metalloporphyrin impurity. Refinement with a free variable of occupancy as a
palladium atom gave an occupancy of 0.014. Due to the small amount of
metalloporphyrin present this peak was not included in the final refinement. CCDC-
605794 contains the supplementary crystallographic data for this paper. These data
can be obtained free of charge from the Cambridge Crystallographic Data Centre via
www.ccdc.cam.ac.uk/data_request/cif.
Crystal data: C32H21IN4. FW = 588.43, black parallelepiped, crystal size 0.3 ×
0.08 × 0.01 mm, monoclinic, P21/c, a = 13.349(3) Å, b = 18.981(4) Å, c = 9.841(2)
Å, β = 104.87(3)°, V = 2410.0(10) Å3, Z = 4, dcalcd = 1.622 Mg.m-3, μ(Mo-Kα) =
1.359 mm-1, Tmin = 0.92, Tmax = 0.99, θmax = 31.53°, 20390 reflections collected, 7981
independent reflections, Rint = 0.0235, 6844 reflections with I > 2.0σ(I), 334
parameters, R1 (I > 2.0σ(I)) = 0.0291, R1 (all data) = 0.0364, wR2 (all data) =
0.0782, S = 1.024, ρmax = 2.902 e.Å-3.

RESULTS AND DISCUSSION

The oxidative addition of molecular I2 to organo organoplatinum(II)


complexes to yield Pt(IV) complexes is well documented in the literature [20]. On
the other hand the same reaction with organopalladium(II) complexes is very rarely
investigated, presumably because the resulting Pd(IV) complexes are very unstable.
Only two examples of Pd(IV) complexes resulting from the oxidative addition of I2

91
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

to a Pd(II) complex have been reported so far and in both examples, these
compounds were only detected by NMR and could not be isolated in the solid state
[21]. There are however some reports describing the iodination of an organic
fragment resulting from the organopalladium(II) intermediate [21b, 22]. Although
none of these reports investigates a halogen exchange of an organic species, we
decided to pursue this avenue in order to achieve the meso-iodination of porphyrins.
In our first experiment, isolated η1-palladioporphyrin 1, which can be
prepared from the reaction of 5-bromo-10,20-diphenylporphyrin (H2DPPBr, 2) with
Pd2dba3 (dba = dibenzylideneacetone) and dppe (dppe = 1,2-
bis(diphenylphosphino)ethane) [2], was reacted with I2 (Scheme 1).
Ph Ph

Ph Ph
NH N NH N P
Pd2dba3
Br + + Pd
P
dppe Ph
N HN N HN Br
Ph

Ph Ph
2 1
Ph I2

Ph
Ph
P NH N
Ph P Pd X + I
Ph X N HN

X = I,Br or Cl
Ph
4 3 91%

Scheme 1. Iodination of η1-palladioporphyrin.

The reaction was carried out at room temperature under aerobic conditions
and was complete in 10 minutes to give the desired iodoporphyrin in 91% yield. The
purification was achieved on a short silica gel column using a mixture of CH2Cl2 and
hexane as eluent. After collecting the porphyrin fraction, byproduct palladium(II)
complexes 4 were eluted using CH2Cl2.
Encouraged by this result, we decided to carry out the halogen exchange
reaction in situ starting from bromoporphyrins, omitting the isolation of the η1-

92
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

palladioporphyrin intermediates. A variety of meso-iodoporphyrins can be isolated


by this method in excellent yields (Scheme 2).
Ar

NH N
R Br
N HN

2: R=H, Ar=Ph
Ar 5: R=Ph, Ar=Ph
+ 6: R=Ph, Ar=3,5-di-t-Bu-phenyl
Pd2dba3 + PPh3 7: R=Br, Ar=Ph
or 8: R=Br, Ar=3,5-di-t-Bu-phenyl
Pd(PPh3)4

3: R=H, Ar=Ph; 78%


I2 air 9: R=Ph, Ar=Ph; 79%
10: R=Ph, Ar=3,5-di-t-Bu-phenyl; 75%
11: R=Br, Ar=Ph; 78%
Ar 12: R=Br, Ar=3,5-di-t-Bu-phenyl; 81%

NH N
R I
N HN

Ar
Scheme 2. One-pot meso-iodination.

The oxidative addition of the Pd(0) fragment is best carried out under an
atmosphere of argon at 105 °C. After the total consumption of the starting
bromoporphyrin the argon atmosphere is not necessary. The reaction mixture should
be diluted and cooled to room temperature before the addition of iodine, in order to
avoid the formation of insoluble materials, whose nature is presently unknown. The
slightly lower yields of these reactions in comparison to the iodination of η1-
palladioporphyrin 1 can be explained by the expected partial dehalogenation of the
bromoporphyrins [3a]. In the case of dibromoporphyrins 7 and 8, we used freshly
prepared Pd(PPh3)4 as the source of Pd(0) in order to control the rate and hence
selectivity of the reaction. As the addition of only one molar equivalent of Pd(0) was
desired, these reactions were also carried out in dilute solutions. The clean formation

93
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

of the 5-η1-palladio(II)-15-bromo-10,20-diarylporphyrins was initially monitored by


1
H NMR of the reaction mixtures and the results agreed well with the data for the
isolated compounds [5]. In order to achieve the iodination, solid I2 pieces were added
to the reaction mixtures. The slow solubility of I2 in toluene at room temperature
ensures that no insoluble materials are formed. The products can be easily purified
on a short silica gel column and the dihalobis(phosphine) Pd(II) by-products can also
be recovered in the cases where Pd(PPh3)4 is utilized. Again the only observed side
products were the mono-dehalogenated porphyrins, i.e. 5-bromo-10,20-
diarylporphyrins. The overall yields and short reaction times for the formation of
mixed haloporphyrins 11 and 12 suggest this is an improvement to the reported
method for the preparation of 7 [8h], as the yield limiting monobromination of the
5,15-diarylporphyrins is not necessary and the iodination can be carried out in 4-5 h
instead of the previously reported 48 h. Moreover, the palladium could, in principle,
be recycled, as any dihalobis(triphenylphosphine)palladium(II) complex (or mixed
complexes) can be reduced in the presence of excess triphenylphosphine to
Pd(PPh3)4 [16].
Our initial reason for the development of this iodination method was the
unsuccessful palladium catalyzed phosphination of porphyrins (Scheme 3).

94
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

Ar

NH N
Ph X
N HN

Ar = 3,5-di-t-Bu-phenyl
Ph 6: X = Br
10: X = I
+
Ph
P Si
Ph

(CH3CN)2PdCl2

Ar
13 from 6: 2%
13 from 10: 20%

NH N Ph Ph
Ph P
N HN O

Ar

Scheme 3. Palladium catalyzed phosphination.

Due apparently to the less reactive C-Br bond, bromoporphyrin 6 yielded the
phosphine oxide in a very poor yield of only 2% after 72 hours at 80 °C. Use of
iodoporphyrin 10 on the other hand resulted in the formation of
porphyrinylphosphine oxide 13 in 20% yield under the same reaction conditions. The
desired tertiary phosphine itself is extremely easily oxidised and only the oxide has
been isolated from these reactions. Prolonged reaction times and temperatures above
85 °C lead only to the decomposition of the target compound. The successful
generation of a suite of porphyrinylphosphine oxides by a different palladium
catalyzed route will be reported elsewhere [23].
Another reason for the development of this halogen exchange methodology is
our interest in the palladium catalyzed formation of meso-meso ethene-linked
porphyrin dyads. As seen in our previous studies of the Heck alkenation reaction, the
desired coupling does not occur with 5-ethenyl-10,15,20-triphenylporphyrin 14 and
bromoporphyrins, but Ni(II) analogues result in novel meso-β ethene-linked

95
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

bisporphyrins. Consequently, we tried a different methodology for the synthesis of


the meso-meso linked dyads, namely Suzuki coupling [15]. For this route, (2-
haloethenyl)porphyrins are potential synthons for the formation of these dyads. Ni(II)
bromoethenyloctaethylporphyrins were first synthesized by Arnold et al. from the
parent ethenylporphyrin and pyridinium tribromide [24]. Later, this methodology
was adapted by Shi et al. for free-base porphyrins [25]. Using the same brominating
reagent in a THF solution at room temperature overnight with ethenylporphyrin 14,
the bromoethenylporphyrin 15 was synthesized as the sole isomer and in good yield
(81%). Interestingly, attempts at coupling 15 with porphyrinyl boronates were
unsuccessful [26]. Thus a method for the formation of iodoethenylporphyrins was
developed (Scheme 4).
Ph Ph

NH N NH N
i
Ph Ph
N HN N HN Br

Ph ii
Ph Ph
14 15 81%

NH N
Ph
N HN I

Ph iii iv Ph
Ph
16 93%

N N N N
Ph Ni Ph Zn
N N I N N I

Ph Ph
17 99% 18 90%

Scheme 4. Formation of iodoethenylporphyrins; conditions: i) pyridinium


tribromide; ii) 1. Pd(PPh3)4, 2. I2; iii) Ni(acac)2; iv) Zn(OAc)2.

96
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

Ni(II) iodoethenyloctaethyl- and tetraphenylporphyrins have been previously


synthesized by Arnold and Hartnell using the CrCl2 promoted Takai iodoalkenation
[12], however, other metallated and free-base analogues have not been reported. This
reaction is unsuitable for free base porphyrins due to the use of the transition metal
salt. The reaction conditions we normally use for inserting Pd(0) into meso-
bromoporphyrins proved too vigorous for insertion into 15, the result being
decomposition of the starting material into multiple products that have not yet been
characterized. Presumably this result is due to the alkenyl-Pd bond being more labile
than a Pd-porphyrin direct bond. The insertion of Pd(0) onto arylalkenyl bromides at
0 °C was reported by Loar and Stille [27]. Inserting Pd(0) into 15 at room
temperature, followed by the addition of I2 in a one-pot procedure, led to the desired
free-base iodoethenylporphyrin in very good yield. Fortunately, the high reactivity of
the alkene-Br bond allows a high yield to be obtained under very mild conditions.
The palladium catalyzed formation of meso-meso ethenelinked diporphyrins from
iodoethenylporphyrins 16, 17 and 18 and various porphyrinyl boronates will be
reported elsewhere [26].
The X-ray structure of 3 is shown in Fig. 1. The compound crystallized in the
monoclinic space group P21/c and overall shows the typical structural features for a
meso-substituted porphyrin of the A2B–type [28]. Only very few crystal structures of
this type have been reported [29]. A related structure is the planar Ni(II) complex of
the related 5-bromoporphyrin [30]. The macrocycle exhibits a moderate degree of
nonplanarity with an overall Δ24 (deviation of the 24 macrocycle atoms from their
least-squares-plane) of 0.078 Å. The conformation is characterized by out-of-plane
deformations with contributions from sad, ruf, and wav(x) distortion modes [31].
Additionally, minor contributions from in-plane distortions are present. The largest
deviations are observed for the β-positions C2 (0.17 Å) and C3 (0.24 Å), i.e. the
pyrrole ring neighboured both by the meso-iodo and a meso-phenyl substituent. The
C5-I1 bond is of typical length [2.119(2) Å] and the iodinated meso-quadrant shows
a widening of the Ca-Cm-Ca bond angle [128.42° (16)] compared to the two meso-
phenyl quadrants [average Ca-Cm-Ca bond angle 123.85° (17)]. The unsubstituted
C15 quadrant also exhibits a widened Ca-Cm-Ca bond angle [127.60° (17)], indicating
the presence of in-plane distortion modes as a result of the bulky iodine residue.

97
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

Figure 1. View of the molecular structure of 3 in the crystal. Thermal ellipsoids are
drawn for 50% occupancy and hydrogen atoms have been omitted for clarity.

In summary, we have developed a rapid, selective and high yielding


methodology for the convenient one-pot meso-iodination of various 5,15-diaryl- and
5,10,15-triarylporphyrins from bromoporphyrins via η1-palladioporphyrin
intermediates. Due to the mild reaction conditions, we propose that this method is
applicable to any diarylporphyrin. This sequence is also very convenient for
converting meso-(2-bromoethenyl)porphyrins to their iodoethenyl analogues.
Subsequent use of the iodo starting materials has confirmed their utility in palladium
catalyzed couplings to form novel substituted porphyrins and diporphyrins.

Acknowledgements

F.A. thanks the Department of Education, Science and Training (DEST) and
the Faculty of Science, Queensland University of Technology for Post-Graduate
Scholarships. O.L. thanks the Queensland University of Technology for a Post-
Graduate Scholarship. M.O.S. gratefully acknowledges support by Science
Foundation Ireland (SFI Research Professorship 04/RP1/B482).

REFERENCES

1. a) Sharman WM and van Lier JE. J. Porphyrins Phthalocyanines 2000; 4:


441-453. b) Setsune J-I. J. Porphyrins Phthalocyanines 2004; 8: 93-102. c)
Senge MO and Richter J. J. Porphyrins Phthalocyanines 2004; 8: 934-953.
2. Arnold DP, Sakata Y, Sugiura K-I and Worthington EI. J. Chem. Soc., Chem.
Commun. 1998; 2331-2332.

98
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

3. a) Arnold DP, Healy PC, Hodgson MJ and Williams ML. J. Organomet.


Chem. 2000; 607: 41-50. b) Hartnell RD, Edwards AJ and Arnold DP. J.
Porphyrins Phthalocyanines 2002; 6: 695-707. c) Hodgson MJ, Healy PC,
Williams ML and Arnold DP. J. Chem. Soc., Dalton Trans. 2002; 4497-4504.
d) Hartnell RD and Arnold DP. Organometallics 2004; 23: 391-399. e)
Hartnell RD and Arnold DP. Eur. J. Inorg. Chem. 2004; 1262-1269.
4. DiMagno SG, Lin VSY and Therien MJ. J. Org. Chem. 1993; 58: 5983-5993.
5. Kato A, Hartnell RD, Yamashita M, Miyasaka H, Sugiura K-I and Arnold
DP. J. Porphyrins Phthalocyanines 2005; 8: 1222-1227.
6. Tsuji J. Palladium Reagents and Catalysts: Innovations in Organic Synthesis,
Wiley & Sons: Chichester New York, 1996.
7. Negishi E-I, Handbook of Organopalladium Chemistry for Organic
Synthesis; Ed., John Wiley & Sons, Inc., 2002; Vols. 1, 2.
8. a) Boyle RW, Johnson CK and Dolphin D. J. Chem. Soc., Chem. Commun.
1995; 527-528. b) Boyle RW, Dolphin D and Johnson CK. PCT Int. Appl.,
9616966, 1996. c) Nakano A, Shimidzu H and Osuka A. Tetrahedron Lett.
1998; 39: 9489-9492. d) Shultz DA, Gwaltney KP and Lee H. J. Org. Chem.
1998; 63: 769-774. e) Shultz DA, Gwaltney KP and Lee H. J. Org. Chem.
1998; 63: 4034-4038. f) Wytko J, Berl V, McLaughlin M, Tykwinski RR,
Schreiber M, Diederich F, Boudon C, Gisselbrecht J-P and Gross M. Helv.
Chim. Acta 1998; 81: 1964-1977. g) Shultz DA, Lee H, Kumar RK and
Gwaltney KP. J. Org. Chem. 1999; 64: 9124-9136. h) Shanmugathasan S,
Johnson CK, Edwards C, Matthews EK, Dolphin D and Boyle RW. J.
Porphyrins Phthalocyanines 2000; 4: 228-232. i) Nakano A, Yasuda Y,
Yamazaki T, Akimoto S, Yamazaki I, Miyasaka H, Itaya A, Murakami M and
Osuka A. J. Phys. Chem. A 2001; 105: 4822-4833. j) Sutton JM and Boyle
RW. Chem. Commun. 2001; 2014-2015. k) Hirata O, Yamamoto M, Sugiyasu
K, Kubo Y, Ikeda M, Takeuchi M and Shinkai S. J. Supramol. Chem. 2003;
2: 133-142. l) Odobel F, Suresh S, Blart E, Nicolas Y, Quintard J-P, Janvier
P, Le Questel JY, Illien B, Rondeau D, Richomme P, Haupl T, Wallin S and
Hammarstrom L. Chem.--Eur. J. 2002; 8: 3027-3046. m) Kubo Y, Yamamoto
M, Ikeda M, Takeuchi M, Shinkai S, Yamaguchi S and Tamao K. Angew.
Chem., Int. Ed. 2003; 42: 2036-2040. n) Kato A, Sugiura K-I, Miyasaka H,
Tanaka H, Kawai T, Sugimoto M and Yamashita M. Chem. Lett. 2004; 33:

99
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

578-579. o) Yahioglu G and Ibanez Garcia D. PCT Int. Appl., 2004046151,


Wo, 2004. p) Susumu K, Duncan TV and Therien MJ. J. Am. Chem. Soc.
2005; 127: 5186-5195. q) Chen Y-J, Lee G-H, Peng S-M and Yeh C-Y.
Tetrahedron Lett. 2005; 46: 1541-1544. r) Wiehe A, Shaker YM, Brandt JC,
Mebs S and Senge MO. Tetrahedron 2005; 61: 5535-5564.
9. a) Yu L and Lindsey JS. Tetrahedron 2001; 57: 9285-9298. b) Bonifazi D
and Diederich F. Chem. Commun. 2002; 2178-2179. c) Bonifazi D, Scholl M,
Song F, Echegoyen L, Accorsi G, Armaroli N and Diederich F. Angew.
Chem., Int. Ed. 2003; 42: 4966-4970.
10. a) Odobel F, Suzenet F, Blart E and Quintard J-P. Org. Lett. 2000; 2: 131-
133. b) Blart E, Suzenet F, Quintard J-P and Odobel F. J. Porphyrins
Phthalocyanines 2003; 7: 207-213. c) Liu Z, Yasseri AA, Loewe RS,
Lysenko AB, Malinovskii VL, Zhao Q, Surthi S, Li Q, Misra V, Lindsey JS
and Bocian DF. J. Org. Chem. 2004; 69: 5568-5577.
11. Fang Z and Breslow R. Bioorg. Med. Chem. Lett. 2005; 15: 5463-5466.
12. Arnold DP and Hartnell RD. Tetrahedron 2001; 57: 1335-1345.
13. Shriver DF and Drezdzon MA. The manipulation of air-sensitive compounds,
Wiley: New York, 1986.
14. Balaban TS, Goddard R, Linke-Schaetzel M and Lehn J-M. J. Am. Chem.
Soc. 2003; 125: 4233-4239.
15. Locos OB and Arnold DP. Org. Biomol. Chem. 2006; 4: 902-916.
16. Coulson DR. Inorg. Synth. 1971; 13: 121-124.
17. Hope H. Prog. Inorg. Chem. 1994; 41: 1-19.
18. Sheldrick GM. SHELXS-93, Program for the Solution of Crystal Structures,
Universität Göttingen, Göttingen (Germany), 1993.
19. Sheldrick GM. SHELXL-97, Program for the Refinement of Crystal
Structures, Universität Göttingen, Göttingen (Germany), 1997.
20. a) Hopgood D and Jenkins RA. J. Am. Chem. Soc. 1973; 95: 4461-4463. b)
Hodges KD and Rund JV. Inorg. Chem. 1975; 14: 525-528. c) Terheijden J,
Van Koten G, De Booys JL, Ubbels HJC and Stam CH. Organometallics
1983; 2: 1882-1883. d) van Beek JAM, van Koten G, Smeets WJJ and Spek
AL. J. Am. Chem. Soc. 1986; 108: 5010-5011. e) van Koten G, van Beek
JAM, Wehman-Ooyevaar ICM, Muller F, Stam CH and Terheijden J.
Organometallics 1990; 9: 903-912. f) Gossage RA, Ryabov AD, Spek AL,

100
Copyright © 2006 Society of Porphyrins & Phthalocyanines
Chapter 4

Stufkens DJ, van Beek JAM, van Eldik R and van Koten G. J. Am. Chem.
Soc. 1999; 121: 2488-2497. g) Yahav A, Goldberg I and Vigalok A.
Organometallics 2005; 24: 5654-5659.
21. a) van Asselt R, Rijnberg E and Elsevier CJ. Organometallics 1994; 13: 706-
720. b) van Belzen R, Elsevier CJ, Dedieu A, Veldman N and Spek AL.
Organometallics 2003; 22: 722-736.
22. a) Domenech J and Sales J. An. Quim. B-Inorg. Anal. 1984; 80: 182-188. b)
Vedernikov AN, Kuramshin AI and Solomonov BN. Zh. Org. Khim. 1993;
29: 2129-2140.
23. Atefi F, McMurtrie JC, Turner P, Duriska M and Arnold DP. submitted for
publication 2006.
24. Arnold DP, Johnson AW and Mahendran M. J. Chem. Soc., Perkin Trans. 1
1978; 366-370.
25. Shi X, Amin SR and Liebeskind LS. J. Org. Chem. 2000; 65: 1650-1664.
26. Locos OB and Arnold DP. manuscript in preparation.
27. Loar MK and Stille JK. J. Am. Chem. Soc. 1981; 103: 4174-4181.
28. Senge MO. Acc. Chem. Res. 2005; 38: 733-743.
29. Senge MO. In Porphyrin Handbook; Kadish KM, Smith KM and Guilard R,
Eds.; Academic Press: San Diego, 2000; Vol. 1, pp 239-347.
30. Arnold DP, Bott RC, Eldridge H, Elms FM, Smith G and Zojaji M. Aust. J.
Chem. 1997; 50: 495-503.
31. Senge MO. Chem. Commun. 2006: 243-256.

101
Copyright © 2006 Society of Porphyrins & Phthalocyanines
CHAPTER 5

meso-Porphyrinylphosphine Oxides: Mono- and


Bidentate Ligands for Supramolecular
Chemistry and the Crystal Structures of
Monomeric {[10,20-
Diphenylporphyrinatonickel(II)-5,15-diyl]-bis-
[P(O)Ph2]} and Polymeric Self-Coordinated

{[10,20-Diphenylporphyrinatozinc(II)-5,15-
diyl]-bis-[P(O)Ph2]}

102
Chapter 5

Statement of Contribution

The authors listed below have certified* that:


11) they meet the criteria for authorship in that they have participated in the
conception, execution, or interpretation, of at least that part of the publication in
their field of expertise;
12) they take public responsibility for their part of the publication, except for the
responsible author who accepts overall responsibility for the publication;
13) there are no other authors of the publication according to these criteria;
14) potential conflicts of interest have been disclosed to (a) granting bodies, (b) the
editor or publisher of journals or other publications, and (c) the head of the
responsible academic unit, and
15) they agree to the use of the publication in the student’s thesis and its publication
on the Australasian Digital Thesis database consistent with any limitations set by
publisher requirements.
In the case of this chapter:
meso-Porphyrinylphosphine Oxides: Mono- and Bidentate Ligands for
Supramolecular Chemistry and the Crystal Structures of Monomeric {[10,20-
Diphenylporphyrinatonickel(II)-5,15-diyl]-bis-[P(O)Ph2]} and Polymeric Self-
Coordinated {[10,20-Diphenylporphyrinatozinc(II)-5,15-diyl]-bis-[P(O)Ph2]}
Atefi F, McMurtrie JC, Turner P, Duriska M and Arnold DP. Inorg. Chem. 2006; 45:
6479-6489.
Contributor Statement of contribution*
Synthesised compounds; all characterisation of the compounds
except: single crystal X-ray analysis (see below), mass spectra and
microanalysis (independent providers); grew X-ray quality single
crystals; analysed data; wrote complete manuscript except part for
single crystal X-ray analysis
John C. Analysed crystal structure, wrote manuscript part about crystal
McMurtrie* structure
Peter Turner* Collected data for singe crystal X-ray analysis
Martin Duriska* Collected data for singe crystal X-ray analysis
Dennis P. Arnold* Overall supervisor of the project, edited manuscript

Principal Supervisor Confirmation

I have sighted email or other correspondence from all Co-authors confirming their
certifying authorship and.

103
Chapter 5

meso-Porphyrinylphosphine Oxides: Mono- and Bidentate Ligands for


Supramolecular Chemistry and the Crystal Structures of Monomeric {[10,20-
Diphenylporphyrinatonickel(II)-5,15-diyl]-bis-[P(O)Ph2]} and Polymeric Self-
Coordinated {[10,20-Diphenylporphyrinatozinc(II)-5,15-diyl]-bis-[P(O)Ph2]}

Farzad Atefi,† John C. McMurtrie,† Peter Turner,‡ Martin Duriska,§ and Dennis P.
Arnold*,†

Synthesis and Molecular Recognition Program, School of Physical and Chemical


Sciences, Queensland University of Technology, GPO Box 2434, Brisbane 4001,
Australia, Crystal Structure Analysis Facility, School of Chemistry, University of
Sydney, NSW 2006, Australia and School of Chemistry, Monash University, Victoria
3800, Australia

Received March 6, 2006


Published on Web 07/14/2006

* To whom correspondence should be addressed. Email: d.arnold@qut.edu.au.


† Queensland University of Technology.
‡ University of Sydney.
§ Monash University.

ABSTRACT
A series of porphyrins substituted in one or two meso positions by
diphenylphosphine oxide groups has been repared by the palladium-catalyzed
reaction of diphenylphosphine or its oxide with the corresponding
bromoporphyrins. Compounds {MDPP-[P(O)Ph2]n} (M = H2, Ni, Zn; H2DPP
= 5,15-diphenylporphyrin; n = 1, 2) were isolated in yields of 60-95%. The
reaction is believed to proceed via the conventional oxidative addition,
phosphination, and reductive elimination steps, as the stoichiometric reaction
of η1-palladio(II) porphyrin [PdBr(H2DPP)(dppe)] (H2DPP = 5,15-

104
Copyright © 2006 American Chemical Society
Chapter 5

diphenylporphyrin; dppe = 1,2-bis(diphenylphosphino)ethane) with


diphenylphosphine oxide also results in the desired mono-
porphyrinylphosphine oxide [H2DPP-P(O)Ph2]. Attempts to isolate the
tertiary phosphines failed due to their extreme air-sensitivity. Variable-
temperature 1H NMR studies of [H2DPP-P(O)Ph2] revealed an intrinsic lack
of symmetry, while fluorescence spectroscopy showed that the phosphine
oxide group does not behave as a “heavy atom” quencher. The electron-
withdrawing effect of the phosphine oxide group was confirmed by
voltammetry. The ligands were characterized by multinuclear NMR and UV-
visible spectroscopy, as well as mass spectrometry. Single-crystal X-ray
crystallography showed that the bis(phosphine oxide) nickel(II) complex
{[NiDPP-[P(O)Ph2]2} is monomeric in the solid state, with a ruffled
porphyrin core and the two P=O fragments on the same side of the average
plane of the molecule. On the other hand, the corresponding zinc(II) complex
formed infinite chains through coordination of one Ph2PO substituent to the
neighboring zinc porphyrin through an almost linear P=O···Zn unit, leaving
the other Ph2PO group facing into a parallel channel filled with disordered
water molecules. These new phosphine oxides are attractive ligands for
supramolecular porphyrin chemistry.

KEYWORDS: Porphyrinoids, palladium, ligand design, phosphine oxides,


crystal structures.

105
Copyright © 2006 American Chemical Society
Chapter 5

This article is not available here.


Please consult the hardcopy thesis
available from QUT Library

106
Copyright © 2006 American Chemical Society
CHAPTER 6

Multiporphyrin coordination arrays based on


complexation of magnesium(II) porphyrins
with porphyrinylphosphine oxides

138
Chapter 6

Statement of Contribution

The authors listed below have certified* that:


16) they meet the criteria for authorship in that they have participated in the
conception, execution, or interpretation, of at least that part of the publication in
their field of expertise;
17) they take public responsibility for their part of the publication, except for the
responsible author who accepts overall responsibility for the publication;
18) there are no other authors of the publication according to these criteria;
19) potential conflicts of interest have been disclosed to (a) granting bodies, (b) the
editor or publisher of journals or other publications, and (c) the head of the
responsible academic unit, and
20) they agree to the use of the publication in the student’s thesis and its publication
on the Australasian Digital Thesis database consistent with any limitations set by
publisher requirements.
In the case of this chapter:
Multiporphyrin coordination arrays based on complexation of magnesium(II)
porphyrins with porphyrinylphosphine oxides
Atefi, F.; McMurtrie, J. C.; Arnold, D. P. J. Chem. Soc., Dalton Trans. 2007, 2163-
2170.
Contributor Statement of contribution*

Synthesised all compounds; all characterisation of these


compounds except: mass spectra (independent providers); analysed
data; wrote complete manuscript

John C.
Input on data interpretation, edited manuscript
McMurtrie*
Dennis P. Arnold* Overall supervisor of the project, edited manuscript

Principal Supervisor Confirmation

I have sighted email or other correspondence from all Co-authors confirming their
certifying authorship and.

139
Chapter 6

Multiporphyrin coordination arrays based on complexation of magnesium(II)


porphyrins with porphyrinylphosphine oxides†

Farzad Atefi, John C. McMurtrie and Dennis P. Arnold*

Received 9th March 2007, Accepted 10th April 2007


First published as an Advance Article on the web 19th April 2007
DOI: 10.1039/b703589f

Synthesis and Molecular Recognition Program, School of Physical and Chemical


Sciences, Queensland University of Technology, GPO Box 2434, Brisbane, 4001,
Australia. E-mail: d.arnold@qut.edu.au; Fax: +61 (0)7 3138 1804; Tel: +61 (0)7
3138 2482

† Electronic supplementary information (ESI) available: 1D 1H NMR spectrum of


porphyrin 2 and 1D and 2D 1H NMR spectra of complexes 3, 8–12. See DOI:
10.1039/b703589f

ABSTRACT
Di- and triporphyrin arrays consisting of 5,15-diphenylporphyrinatomagnesium(II)
(MgDPP) coordinated to free-base and Ni(II) porphyrinyl mono- and bis-phosphine
oxides, as well as the self-coordinating diphenyl[10,20-
diphenylporphyrinatomagnesium(II)-5-yl]phosphine oxide [MgDPP(Ph2PO)], were
synthesised in excellent yields and characterised by various spectroscopic
techniques. Phosphine oxides stabilise Mg(II) coordination to porphyrins and the
resulting complexes have convenient solubilities, while the Ni(II) complexes exhibit
interesting intramolecular fluorescence quenching behaviour. The binding constant
of MgDPP to triphenylphosphine oxide (5.3 ± 0.1 × 105 M−1) and the very high self-
association constant of [MgDPP(Ph2PO)] (5.5 ± 0.5 × 108 M−1) demonstrate the
strong affinity of phosphine oxides towards Mg(II) porphyrins. These complexes are
the first strongly bound synthetic Mg(II) multiporphyrin complexes and could
potentially mimic the “special pair” in the photosynthetic reaction centre.

140
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

INTRODUCTION
The formation of multiporphyrin arrays has received considerable attention in
recent years, not only due to their critical role in the biological light harvesting
process of plants and bacteria, but also because of their potential applications in
diverse fields such as molecular wires, non-linear optics and photodynamic therapy.1
As the majority of metals form equatorial complexes with porphyrins, the study of
multiporphyrinic systems obtained through the coordination of this central metal ion
with an axial ligand attached to a second porphyrinic macrocycle is a very fast
growing area of research.2 Such molecules have been developed in recent years to
mimic the primary electron donor in photosynthetic reaction centres of bacteria.3
This so-called “special pair” consists of two bacteriochlorophylls, which are
noncovalently held apart by 3.2A˚ and the central metal atom of these porphyrinic
macrocycles is Mg(II).4
Magnesium porphyrins have long singlet excited state lifetimes, large
fluorescence quantum yields and low oxidation potentials.5 However most examples
of Mg(II) porphyrins incorporated in self-assembled synthetic multiporphyrin
systems are concerned with aggregation behaviour of chlorophyll based dimers
which are weakly bound to each other.6 Recently some theoretical studies of
chlorophyll and its aggregation behaviour have also been reported.7 There is one
example of a synthetic Mg(II) porphyrin dimer in which an imidazole moiety
attached to the porphyrin periphery coordinates to the central Mg(II) centre of a
neighbouring macrocycle.8 Although the authors did not report any self-association
constant for their dimer, it has been established that N-ligands bind very weakly to
Mg(II) porphyrins.9 The vast majority of self-complexed metalloporphyrin oligomers
involve zinc porphyrins, which have a shorter singlet state excited lifetime than their
Mg(II) analogues but are synthetically more accessible.1a,2,3,10 Until the mid 1990’s
magnesium(II) porphyrins have been avoided by synthetic chemists due to the
problems associated with the insertion of Mg(II) into the inner core of porphyrinic
macrocycles. This obstacle has been removed by the heterogeneous and
homogeneous Mg(II) insertion methods of Lindsey and co-workers.5,11 However,
suitable ligands that show a high affinity towards magnesium porphyrins have not
been reported so far. There is only one example of triphenylphosphine oxide as a
ligand for chlorophyll a and b.12 This IR spectroscopic study showed that the

141
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

phosphine oxide forms a strong complex with the chlorophylls and thus breaks the
self-aggregation that depends on the keto group of a neighbouring chlorophyll
molecule.4 Triphenylphosphine oxide has also been used as a ligand for other
metalloporphyrins such as Os(II),13 Cr(III),14 Fe(III),15 Ru(IV)16 and Sn(IV)17
porphyrins.
During our systematic studies of the formation and reactions of η1-
palladioporphyrins,18 we recently reported the synthesis and characterisation of
meso-porphyrinylphosphine oxides.19 These pentavalent phosphorus compounds can
be isolated in high yields under mild reaction conditions either from the direct
reaction of η1-palladioporphyrins or from a Pd(0) catalysed reaction of
haloporphyrins with diphenylphosphine oxide. Our results showed a weak affinity of
the phosphorus-oxo ligand towards Zn(II) porphyrins. A Japanese group prepared
similar phosphine oxides and quantified the self-coordination of the Zn(II) species.20
Both reports19b,20 show only a weak reciprocal interaction between the phosphine
oxide ligand and the neighbouring zinc(II) porphyrin, which is not surprising due to
the low binding strength of oxo ligands towards zinc(II). In earlier reports it has been
well established that zinc(II) porphyrins show a larger affinity towards N-donor
ligands than hydroxy or carbonyl ligands.2a Magnesium porphyrins on the other hand
show very weak coordination capabilities towards nitrogen bases.
We report here the formation and characterisation of di- and triporphyrin
complexes consisting of porphyrinylphosphine oxides and magnesium(II)
porphyrins. We have studied these complexes with various techniques, namely 1H
31
and PNMR, UV-visible and fluorescence spectroscopies and cyclic voltammetry.
The phosphorus-oxo ligands indeed show a strong affinity towards the Mg(II)
porphyrins. These novel complexes have potential as synthetic mimics for the
“special pair” in photosynthetic reaction centres.

RESULTS AND DISCUSSION

Synthesis and NMR characterisation


Due to the known low solubility of Mg(II) porphyrins in organic solvents, we
decided at first to utilise 5,15-bis(3,5-di-tbutylphenyl) porphyrin (H2DAP) as
substrate for the formation of a Mg(II) porphyrin. To our surprise we discovered that
the resulting 5,15-bis(3,5-di-t-butylphenyl)porphyrinatomagnesium(II) (MgDAP) has

142
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

a very low solubility in organic solvents (<10−5 M). 5,5-


Diphenylporphyrinatomagnesium(II) (MgDPP, 2) on the other hand can be dissolved
in concentrations of up to 10−4 M in CDCl3, CH2Cl2 and toluene without the aid of
sonication, which is not the case for MgDAP. Porphyrin 2 can be isolated in almost
quantitative yield from the reaction of 5,15-diphenylporphyrin (H2DPP) 1 with either
MgBr2·O(Et)2 (homogeneous method) or with freshly prepared anhydrous MgBr2
(heterogeneous method) as shown in Scheme 1.5,11 The advantage of the
heterogeneous method in this case is the absence of any axial ligand except water in
2.
Ph

H H
NH N O
MgBr2.O(Et)2 N N
Ph
N
Mg
N
Ph 2
or MgBr2
N HN

K = 5.2 x 105 M-1 Ph PO


3
Ph (lg 5.7)

1
Ph
Ph Ph
P
O
N N
3 Ph
N
Mg
N
Ph

Scheme 1. Formation of MgDPP 2 and MgDPP-PPh3PO complex 3.

Next we isolated the Ph3PO complex 3 of Mg(II) porphyrin 2 as shown in


Scheme 1. During NMR studies, no six-coordinate Mg(II) species could be observed
as the peak representing water coordinated to Mg(II) (~0.1 ppm) disappears and free
phosphine oxide can always be observed in the spectra upon addition of excess
Ph3PO.Water displacement by Ph3PO was measured by UV-visible titration as
described by Connors.21 The titration spectra (Fig. 1) show an isosbestic point at 415
nm which indicates the formation of a single complex. A Job’s plot confirms the
formation of a 1 : 1 complex between Ph3PO and porphyrin 2.21 The displacement
constant of 5.3 ± 0.1 × 105 M−1 calculated by the Benesi–Hildebrand method is two
orders of magnitude larger than any other Mg(II) porphyrin-ligand binding constant
reported thus far.22

143
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

Figure 1. UV-visible titration of MgDPP 2 (1.4 × 10−6 M) in toluene at 25 ºC upon


successive addition of Ph3PO (7.9 × 10−7-1.5 × 10−4 M). Inset: Benesi–Hildebrand
plot (solid line: fitted curve).21

Encouraged by this result, we then isolated di- and triporphyrin Mg(II)


complexes with free-base and Ni(II) porphyrinyl phosphine oxides as ligands
(Scheme 2). Complexes 8–11 were isolated in a similar fashion. Porphyrin 2 and the
appropriate ligand (4–7) were first dissolved in CDCl3 in a 1 : 1 stoichiometry and
the 1H NMR spectra of the complexes were recorded. Afterwards the solvent was
removed and the products recrystallised from toluene–pentane. The complexes all
form in quantitative yield and losses are due only to the recrystallisation process. We
discovered that these complexes are remarkably robust. The solid samples were
subjected to high resolution ESI, LDI or LSI mass spectrometry and molecular ions
were observed for the respective complexes. The solubility of the complexes 8–11
also drastically increases in comparison to uncoordinated 2, as concentrations of 10−3
M in CDCl3, CH2Cl2 and toluene can be easily achieved. The phosphine oxide
ligands also stabilise the basal coordination of the Mg(II) centre to DPP. While
partial demetalation of 2 in CDCl3 was observed after only 8 h, complexes 8–11 are
stable in CDCl3 for months. All products were analysed by various 1D and 2D 1H
31
NMR techniques (see ESI†) as well as P NMR, and similar features could be
observed for all complexes. Again no peaks representing water coordinated to the
Mg(II) fragment were observed. If an excess of ligand 4–7 is present, three different

144
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

kinds of porphyrinyl units can be discerned, namely excess uncomplexed ligand 4–7
and both the metalloporphyrin 2 and the phosphine oxide components of the di- or
tri-porphyrin complexes 8–11 in the appropriate integral ratios. Likewise when
excess metalloporphyrin 2 is present, the peaks representing both complexed and
uncomplexed MgDPP are apparent. This indicates that exchange between complexed
and uncomplexed species is slow enough to enable the observation of sharp spectra.
Exchange between complexed and uncomplexed species was only observed by
elevated temperature experiments.

145
Copyright © 2007 The Royal Society of Chemistry
146
Chapter 6

Copyright © 2007 The Royal Society of Chemistry


Scheme 2. Formation of di-porphyrin complexes 8, 9 (above) and tri-porphyrin complexes 10, 11 (below).
Chapter 6

Fig. 2 shows the 1H NMR spectra of complexes 8 (left) and 11 (right) in


CDCl3 at 23 ºC. The arrows indicate some of the differences between the spectra of
the complexes (8, 11) and uncomplexed porphyrins 2, 4 and 7. One can readily
appreciate the large effect the two macrocycles have on each other upon
complexation. The most evident features are the differences in chemical shifts of the
peaks representing the o-Ph protons on the phosphorus (H3: shifted upfield by more
than 2.5 ppm) and the peaks representing the b-protons next to the P–C bond (H2;
shifted upfield by more than 3 ppm). These results indicate that the two porphyrin
macrocycles are in close proximity, as the influence of the respective ring currents
can be observed in a large upfield shift of the 1H NMR peaks. Even the peak
associated with the meso-proton on the metalloporphyrin (H1) and the inner nitrogen
protons of the free-base phosphine oxide (H4) are shifted upfield.
31
The P NMR spectra of the complexes also show peaks in the expected
region (3: 25.5 ppm, 8–11: 26.4–31.5 ppm) and these differ only slightly from the
values of the uncomplexed ligands (Ph3PO: 29.6 ppm; 4–7: 29.2–34.2 ppm).19b This
stems from the fact that the P–O–Mg coordination leads to a downfield shift while
the ring current from an adjacent porphyrin causes an upfield shift of similar
magnitude.
We also isolated Mg(II) phosphine oxide 12 as shown in Scheme 3 and in this
case only the heterogeneous Mg(II) insertion method was successful. The complex
adheres very strongly to glassware and accumulates between the organic and the
water phases during the work up. Nevertheless we could isolate 12 in a satisfactory
yield of 75%. Due to its low solubility (10−5 M) in organic solvents (CDCl3, CH2Cl2
31
and toluene) the P NMR and the 1H NOESY NMR of porphyrin 12 were of poor
quality. The UV-visible and fluorescence spectra (discussed below) show that the
complex exists in its dimeric form 12a in solution. The previously reported single
crystal X-ray structures of the polymeric Zn(II) phosphine oxides19b,20 suggest that a
polymeric structure 12b is likely for complex 12 in the solid state. The 1H NMR
spectrum is similar to the spectra already discussed above. Even the addition of large
amounts of coordinating solvents (MeOH, pyridine) does not improve the solubility,
indicating a very stable self-coordinated complex.

147
Copyright © 2007 The Royal Society of Chemistry
148
Chapter 6

Figure 2. (Left) 1H NMR spectra of MgDPP 2 (bottom), complex 8 (middle) and ligand 4 (top); (right) 1H NMR spectra of MgDPP 2

Copyright © 2007 The Royal Society of Chemistry


(bottom), complex 11 (middle) and ligand 7 (top).
149
Chapter 6

Scheme 3. Formation of Mg(II) phosphine oxide 12 and possible polymeric form.

Copyright © 2007 The Royal Society of Chemistry


Chapter 6

Electronic absorption spectra, emission spectra and redox properties


The UV-visible spectra of the complexes (3, 8–11) were recorded in toluene
at 25 ºC at various concentrations (1 × 10−4 to 5 × 10−6 M) in order to gain some
insight into the complexation properties in solution. High concentrations (10−4–10−5
M) were measured in 1 mm pathlength cells while lower concentrations were
recorded in a 10 mm cell. As one example, Fig. 3 shows the spectral changes of
complex 8. High concentrations show the same spectral profiles, while at 10−6 M a
small deviation of the spectral properties can be observed. At these low
concentrations the complexes show some dissociation into their monomeric forms. A
splitting of the Soret bands due to excitonic coupling can be observed for the multi-
porphyrin complexes. This feature is most prominent in complex 12, as in the other
cases (8–11) the Soret bands of the monomers overlap with those of the complexes.
The self-association constant of 12 was measured by UV-visible displacement
titration by successive addition of Ph3PO, to obtain the data displayed in Fig. 4 that
indicate the expected 1 : 2 complex formation.23 With the reasonable assumption that
the binding constants for the complexation of Ph3PO to MgDPP 2 and
MgDPP(Ph2PO) 12 are the same, the self-association constant was calculated to be
5.5 ± 0.5 × 108 M−1 (lg 8.7), showing that very little dissociation will occur at 10−5 M
monomer concentration. The stability and solubility of complexes such as 8–12 in
organic solvents are encouraging for further studies of these aggregates.

150
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

Figure 3. UV-visible spectra of 8 (1.1 × 10−4 to 5.4 × 10−6 M; 10 mm and 1 mm


cells).

Figure 4. UV-visible titration of 12 (3.1 × 10−6 M) in toluene at 25 ºC upon


successive addition of Ph3PO (5.7 × 10−5 to 2.9 × 10−2 M). Inset: Binding isotherm
(solid line: fitted curve for 1 : 2 complex).

The fluorescence spectra of the complexes 8–12 were recorded in toluene at


25 ºC (10−5 M, except for 12: 10−6 M). We excited at 485 nm (minimum between
Soret and Q-bands) in order to avoid any effects due to internal absorption. The

151
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

spectrum of 9 was compared to the spectra of free MgDPP 2 and a solution


containing 2 and 5,15-diphenylporphyrinatonickel(II) (NiDPP, 13). Nickel(II)
porphyrins are non-fluorescent due to radiationless decay pathways associated with
the unfilled d orbitals.24 As shown in Fig. 5, the fluorescence profiles for MgDPP 2
and a mixture of 2 with NiDPP 13 differ only slightly. Therefore no significant
intermolecular fluorescence quenching due to the presence of the Ni(II) porphyrin
can be detected. However, the spectrum of complex 9 shows a large quenching
(~80%) due to intramolecular energy transfer between the Ni(II) andMg(II)
macrocycles, which indicates effective electronic communication in the excited state.
The fluorescence for complex 11 is less efficiently quenched, as there is only one
Ni(II) for two Mg(II) centres present. Due to the low solubility of dimer 12, the
spectrum for this complex was recorded at 10−6 M. Both maxima for 12 are red-
shifted by ~25 nm compared with the other species. This indicates a more rigid face-
to-face geometry in comparison to the other complexes.25 These preliminary
empirical studies indicate that measurements of excited state lifetimes and
intramolecular energy transfer rates are warranted in the future.

Figure 5. Fluorescence spectra of 2, 2 + 13 and 9 (left); 8, 10-12 (right) excited at


485 nm in toluene at 25 ºC and normalised to 10−5 M (except for 12: measured at
10−6 M on a more sensitive detector setting).

The oxidation properties of MgDPP 2 and complexes 3, 8 and 10 were briefly


studied by cyclic and square wave voltametry and compared to values for H2DPP 1

152
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

and free-base ligands 4 and 6. The measurements were conducted in CH2Cl2 (10−3
M) containing 0.1 M Bu4NPF6 at a Pt working electrode (vs Ag/Ag+;
ferrocene/ferrocinium at +0.55 V). The values for the first oxidation potential (Table
1) show, as expected, the electron donating properties of Mg(II) in 2 (+0.78 V) in
comparison to free-base porphyrin 1 (+1.16 V). The first oxidation potentials of the
complexes show only a slight further cathodic shift in comparison to MgDPP 2,
which is largest for complex 3 (+0.73 V). The second oxidation wave of complexes 8
and 10 comprises overlapped waves due to the second oxidation of 2 and the first
oxidation of the free-base phosphine oxide ligands 4 and 6. Relative areas from
square wave measurements support this being an overall two-electron process.

Table 1. Oxidation potentials for 1–4, 6, 8 and 10 (10−3 M) measured in CH2Cl2


containing 0.1 M Bu4NPF6.

Compound Ox(1) Ox(2)


118c +1.16
2 +0.78 1.12
3 +0.73 1.20
19b
4 +1.24
619b +1.33
8 +0.76 +1.2 a
10 +0.77 +1.2 a
a
Ox(2) of Mg(II) and Ox(1) of free-base part of complex overlap.

CONCLUSION

We have prepared several representatives of a new class of multiporphyrin


complexes containing Mg(II) porphyrins and our recently reported
porphyrinylphosphine oxides.19 These compounds can be obtained in excellent yield
by complexation in organic solvents. Inter-porphyrin electronic interaction can be
observed with various spectroscopic techniques and the complexes are remarkably
robust in solution. The large self-association constant (5.5 ± 0.5 × 108 M−1) of
complex 12 is a very good indication of this behaviour. In comparison the self-
association constant of the Zn(II) dimer reported by the Japanese group is two orders
of magnitude lower (5.9 × 106 M−1).20 The presence of Ni(II) in the phosphine oxide
ligands leads to intramolecular fluorescence quenching of the complexes. Although
no X-ray quality single crystal for any of the complexes has been obtained so far, the
comparison with zinc(II) porphyrinylphosphine oxide polymers19b,20 leads us to
expect very interesting porphyrin–porphyrin interactions in the solid state. We have

153
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

shown that Mg(II) porphyrins have a very high affinity towards oxo-ligands,
especially phosphine oxides, and our complexes are the first strongly bound synthetic
Mg(II) porphyrin analogues of the “special pair” of the photosynthetic reaction
centre.

EXPERIMENTAL

General remarks
Chemical reagents were of laboratory reagent (LR) or analytical reagent (AR) grade,
received from Sigma-Aldrich and used without further purification. Toluene was
stored over sodium wire and CH2Cl2 was stored over anhydrous sodium carbonate.
MgBr2,26 H2DPP27 and phosphine oxides 4–719b were synthesised according to
literature procedures. Recrystallisation from two solvents was accomplished by
dissolving the product in a minimum amount of the first solvent (or solvent mixture)
and carefully layering the solution with a tenfold excess of the solvent in which the
product is less soluble. Analytical TLC was performed using aluminium backed
Merck silica gel 60 F254 plates. NMR spectra were recorded on a Bruker Avance
400 MHz instrument and J values are given in Hz. Quantitative UV-visible spectra
were recorded on a Cary 3 spectrometer in toluene solutions. UV-visible titrations
were performed by adding aliquots of a solution containing excess Ph3PO and the
Mg(II) porphyrin to a solution of the same concentration containing only the Mg(II)
porphyrin. Fluorescence spectra were recorded on a Varian Cary Eclipse
fluorescence spectrophotometer equipped with a standard multicell Peltier
thermostatted sample holder in toluene solutions. Accurate mass electrospray
ionisation (ESI) mass spectra were recorded at Monash University, Australia on a
Bruker BioApex 47e FTMS fitted with an Analytical Electrospray source.
Dichloromethane was used as a solvent and the samples were diluted either with
methanol or dichloromethane–methanol. The samples were introduced into the
source by direct infusion (syringe pump) at 60 μL h−1 with a capillary voltage of 80
V. Sodium iodide clusters were used as internal standard for mass calibration. Laser
desorption/ionization (LDI) MS analysis was performed at Monash University,
Australia with an Applied Biosystems Voyager-DE STR BioSpectrometry
Workstation. The instrument was operated in positive polarity in reflecton mode.
Samples were spotted on a stainless steel sample plate and allowed to air dry. Data

154
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

from 500 laser shots (337 nm nitrogen laser) were collected, signal averaged, and
processed with the instrument manufacturer’s Data Explorer software. Liquid
secondary ion (LSI) MS measurements were collected at the University of Tasmania,
Australia using m-nitrobenzyl alcohol as proton donor and CsI/TEG mixture for
external reference on a Kratos Concept ISQ double focussing Magnetic/Electrostatic
Mass Spectrometer. Isotopic modelling was performed by MassLynx V3.5 Software
by Micromass Limited.
5,15-Diphenylporphyrinatomagnesium(II) 2
Homogeneous method. H2DPP 1 (50 mg, 0.11 mmol) was dried under high
vacuum. After the addition of freshly distilled CH2Cl2 (10 cm3) and triethylamine
(TEA) (0.3 cm3, 2.2 mmol), freshly ground MgBr2·OEt2 (280 mg, 1.1 mmol) was
added and the mixture was stirred for 2 h. The reaction was monitored by TLC
(CH2Cl2–TEA=200 : 1) and a new more polar spot could be observed. The reaction
mixture was diluted with 25 cm3 of CH2Cl2, washed with 2 × 25 cm3 NaHCO3, 5 ×
50 cm3 water, dried over sodium sulfate, filtered and the residue was recrystallised
from CH2Cl2–TEA–hexane to obtain 52 mg (0.11 mmol, 99%) of a bright red
crystalline powder.
Heterogeneous method. The desired Mg(II) porphyrin can also be obtained
under the same reaction conditions and with the same reactants as above by using
freshly prepared MgBr2 (203 mg, 1.1 mmol) instead of MgBr2·OEt2. δH (400 MHz,
CDCl3, CHCl3) 10.24 (2H, s, meso-H), 9.35 (4H, br d, β-H), 9.05 (4H, br d, β -H),
8.29–8.21 (4H, br, o-H 5,15-phenyl), 7.79–7.77 (6H, m, m, 30 p-H 5,15-phenyl),
0.05 (br s, H2O coordinated to Mg, variable according to amount of water in solvent);
UV/Vis (toluene) λmax/nm 416 (ε/103 dm3 mol−1 cm−1 336), 514 (2.2), 551 (19), 588
(3.1); m/z (High-resolution ESI) 507.1430 (C32H20MgN4 + Na+ requires 507.1436).
5,15-Diphenylporphyrinatomagnesium(II)–triphenylphosphine oxide complex 3
Mg(II)DPP 2 (15 mg, 0.030 mmol) and triphenylphosphine oxide (8 mg,
0.030 mmol) were dissolved in 5 cm3 of CDCl3 and the 1H NMR spectrum of the
mixture was recorded. The solvent was removed under vacuum and the complex was
recrystallised from toluene–pentane to obtain 22 mg (0.029 mmol, 97%) of a purple
1
crystalline powder. The HNMR spectrum of the redissolved solid after
recrystallisation agreed with the one recorded of the initial solution. δH (400 MHz,
CDCl3, CHCl3) 10.13 (2H, s, meso-H), 9.28 (4H, d, 3J 4.0, β-H), 8.98 (4H, d, 3J 4.1,
β-H), 8.35–8.30 (2H, br, o-H 5,15 Ph), 7.93 (2H, br d, 3J 6.4, o-H 5,15 Ph), 7.78–

155
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

7.72 (4H, br, m, p-H 5,15 Ph), 7.70–7.66 (2H, br, m-H 5,15 Ph), 7.06 (2H, t, 3J 6.3,
p-H PPh3), 6.77–6.70 (4H, br, m-H PPh3), 4.65–4.60 (4H, m, o-H PPh3); δP (162
MHz, CDCl3, 85% H3PO4 in H2O) 25.5; UV/Vis (toluene) λmax/nm 416 (ε /103 dm3
mol−1 cm−1 524), 552 (19), 589 (3.6);m/z (High-resolution ESI) 785.2277
(C50H35MgN4OP+Na+ requires 785.2297).
Free-base di-porphyrin Complex 8
Magnesium porphyrin 2 (10 mg, 0.020 mmol) and ligand 4 (13 mg, 0.020
mmol) were dissolved in 7 cm3 of CDCl3 and the 1H NMR spectrum of the mixture
was recorded. The solvent was removed under vacuum and the complex was
recrystallised from toluene/pentane to obtain 22 mg (0.019 mmol, 96%) of a purple
1
crystalline powder. The H NMR spectrum of the redissolved solid after
recrystallisation agreed with the one recorded of the initial solution. δH (400 MHz,
CDCl3, CHCl3) 10.15 (1H, s, meso-H P=O), 10.02 (2H, s, meso-H Mg), 9.22 (2H, d,
3
J 4.4, β-H P=O), 9.14 (4H, d, 3J 4.3, β-H Mg), 8.78 (6H, d, 3J 4.2, β-H P=O, β-H
Mg), 8.20 (2H, br d, 3J 6.8, o-H 5,15 Ph Mg), 7.97 (4H, d, 3J 6.4, m-H 10,20 Ph
P=O), 7.93 (2H, d, 3J 4.4, β-H P=O), 7.80 (2H, d, 3J 6.4, p-H 10,20 Ph P=O), 7.77-
7.73 (4H, m, o-H 10,20 Ph P=O), 7.65 (2H, t, 3J 6.8, m-H 5,15 Ph Mg), 7.54 (2H, t,
3
J 7.3, p-H 5,15 Ph Mg), 7.09 (2H, t, 3J 6.4, p-H PPh2), 7.01-6.99 (2H, br, o-H 5,15
Ph Mg), 6.91 (2H, t, 3J 6.3, m-H 5,15 Ph Mg), 6.74-6.72 (4H, br, m-H PPh2), 5.84-
5.82 (2H, br, β-H P=O), 4.75-4.70 (4H, m, o-H PPh2), -3.07 (2H, br, NH); δP (162
MHz, CDCl3, 85% H3PO4 in H2O) 31.5; UV/Vis (toluene) λmax/nm 416 (ε/103 dm3
mol-1 cm-1 655), 514 (14), 551 (23), 589 (7.8), 640 (2.8); m/z (High-resolution ESI)
1147.3825 (C76H51MgN8OP+H+ requires 1147.3852).
Nickel(II) di-porphyrin Complex 9
Magnesium porphyrin 2 (12 mg, 0.025 mmol) and ligand 5 (18 mg, 0.025
mmol) were dissolved in 7 cm3 of CDCl3 and the 1H NMR spectrum of the mixture
was recorded. The solvent was removed under vacuum and the complex was
recrystallised from toluene/pentane to obtain 29 mg (0.023 mmol, 92%) of a purple
1
crystalline powder. The H NMR spectrum of the redissolved solid after
recrystallisation agreed with the one recorded of the initial solution. δH (400 MHz,
CDCl3, CHCl3) 10.03 (2H, s, meso-H Mg), 9.68 (1H, s, meso-H P=O), 9.15 (4H, d,
3
J 4.4, β-H Mg), 9.04 (2H, d, 3J 4.5, β-H P=O), 8.77 (4H, d, 3J 4.2, β-H Mg), 8.69
(2H, d, 3J 452, β-H P=O), 8.24 (2H, br d, 3J 6.8, o-H 5,15 Ph Mg), 7.84 (2H, d, 3J

156
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

4.6, β-H P=O), 7.78-7.64 (12H, br m, o,m,p-H, 10,20 Ph P=O; m-H 5,15 Ph Mg),
7.54 (2H, br, p-H 5,15 Ph Mg), 7.00-6.97 (2H, m, p-H PPh2), 6.85 (4H, br, m,o-H
5,15 Ph Mg), 6.63-6.57 (4H, br, m-H PPh2), 5.60-5.55 (2H, br, β-H P=O), 4.53-4.47
(4H, m, o-H PPh2); δP (162 MHz, CDCl3, 85% H3PO4 in H2O) 27.2; UV/Vis
(toluene) λmax/nm 416 (ε/103 dm3 mol-1 cm-1 501), 552 (24), 586 (10); m/z (High-
resolution LSI) 1202.2920 (C76H49MgN8NiOP+ requires 1202.2971.
Free-base tri-porphyrin Complex 10
Magnesium porphyrin 2 (10 mg, 0.020 mmol) and ligand 6 (9 mg, 0.010
mmol) were dissolved in 7 cm3 of CDCl3 and the 1H NMR spectrum of the mixture
was recorded. The solvent was removed under vacuum and the complex was
recrystallised from toluene/pentane to obtain 18 mg (0.0096 mmol, 95%) of a purple
1
crystalline powder. The H NMR spectrum of the redissolved solid after
recrystallisation agreed with the one recorded of the initial solution. δH (400 MHz,
CDCl3, CHCl3) 9.99 (4H, s, meso-H Mg), 9.11 (8H, d, 3J 4.1, β-H Mg), 8.75 (8H, d,
3
J 4.1, β-H Mg), 8.19 (4H, br d, 3J 6.9, o-H 5,15 Ph Mg), 7.78 (4H, d, 3J 4.6, β-H
P=O), 7.72-7.67 (10H, m, o,m,p-H 10,20 Ph P=O), 7.64 (4H, br t, 3J 6.8, m-H 5,15
Ph Mg), 7.49 (4H, br t, 3J 7.7, p-H 5,15 Ph Mg), 7.13 (4H, br t, 3J 6.4, p-H PPh2),
6.89-6.85 (4H, br, o-H 5,15 Ph Mg), 6.80-6.70 (12H, m, m-H 5,15 Ph Mg, m-H
PPh2), 5.72 (4H, br d, 3J 4.1, β-H P=O), 4.74-4.68 (8H, m, o-H PPh2), -3.16 (2H, br,
NH); δP (162 MHz, CDCl3, 85% H3PO4 in H2O) 29.9; UV/Vis (toluene) λmax/nm 416
(ε/103 dm3 mol-1 cm-1 1040), 515 (11), 552 (45), 588 (14), 661 (7.1); m/z (High-
resolution ESI) 938.2732 (C120H80Mg2N12O2P2+Na22+ requires 938.2750).
Nickel(II) tri-porphyrin Complex 11
Magnesium porphyrin 2 (10 mg, 0.020 mmol) and ligand 7 (9 mg, 0.010
mmol) were dissolved in 7 cm3 of CDCl3 and the 1H NMR spectrum of the mixture
was recorded. The solvent was removed under vacuum and the complex was
recrystallised from toluene/pentane to obtain 18 mg (0.0095 mmol, 95%) of a purple
1
crystalline powder. The H NMR spectrum of the redissolved solid after
recrystallisation agreed with the one recorded of the initial solution. δH (400 MHz,
CDCl3, CHCl3) 10.05 (4H, s, meso-H Mg), 9.15 (8H, d, 3J 4.1, β-H Mg), 8.76 (8H, d,
3
J 4.0, β-H Mg), 8.24 (4H, br d, 3J 7.0, o-H 5,15 Ph Mg), 7.71 (4H, d, 3J 4.6, β-H
P=O), 7.69-7.61 (12H, m, o,m-H 10,20 Ph P=O, m-H 5,15 Ph Mg), 7.55-7.45 (6H, m,
p-H 10,20 Ph P=O, p-H 5,15 Ph Mg), 7.01 (4H, br t, 3J 6.7, p-H PPh2), 6.81-6.73

157
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

(8H, br, o,m-H 5,15 Ph Mg), 6.67-6.60 (8H, br, m-H PPh2), 5.50-5.46 (4H, br, β-H
P=O), 4.56-4.48 (8H, m, o-H PPh2); δP (162 MHz, CDCl3, 85% H3PO4 in H2O) 26.4;
UV/Vis (toluene) λmax/nm 416 (ε/103 dm3 mol-1 cm-1 856), 515 (8.3) 552 (41), 590
(14) 615 (13); m/z (High-resolution ESI) 1886.4851 (C120H78Mg2N12NiO2P2+
requires 1886.4900).
Bis{diphenyl[10,20-diphenylporphyrinatomagnesium(II)-5-yl]phosphine oxide}
12
Phosphine oxide 4 (13 mg, 0.020 mmol) was dried under high vacuum. After
the addition of freshly distilled CH2Cl2 (5 cm3) and TEA (0.220 cm3, 1.6 mmol),
freshly ground MgBr2.OEt2 (204 mg, 0.790 mmol) was added and the mixture was
stirred for 7 h. Due to the polarity and insolubility of the resulting product, the
reaction could not be monitored by TLC. The reaction mixture was diluted with 50
cm3 of CH2Cl2. Even with the aid of sonication the product could not be fully
dissolved. The mixture was washed with 50 cm3 of water in the flask. The majority
of the water was decanted off and the remaining solvents were removed under
vacuum. The solid was dissolved in 30 cm3 of a toluene/MeOH/TEA (100:100:1)
mixture with the aid of sonication and washed with 5 × 50 cm3 water. The solvents
were removed under vacuum and the product was dried thoroughly under high
vacuum. After recrystallisation from toluene/MeOH/TEA/pentane, 10 mg (0.015
mmol, 75%) of the desired product was obtained as a purple crystalline powder. δH
(400 MHz, CDCl3, CHCl3) 10.27 (1H, s, meso-H), 9.26 (2H, d, 3J 4.2, β-H), 8.68
(2H, d, 3J 4.2, β-H), 8.40-8.36 (2H, br, m-H 10,20 Ph), 9.05 (2H, d, 3J 4.1, β-H),
7.74-7.59 (4H, m, m,p-H 10,20 Ph), 7.55-7.51 (2H, m, m-H 10,20 Ph), 7.48-7.43
(2H, br, o-H 10,20 Ph), 7.23-7.19 (2H, m, p-H PPh2), 6.86-6.80 (4H, m, m-H PPh2),
6.34 (2H, d, 3J 4.4, β-H), 6.53-6.43 (4H,br, o-H PPh2); UV/Vis (toluene) λmax/nm
414sh (ε/103 dm3 mol-1 cm-1 54), 426 (74), 568 (7), 595 (5); m/z (LDI) 685.2
(C44H29MgN4PO+H+ requires 685.2); dimer m/z (LDI) 1369.4
+
(C88H30Mg2N8P2O2+H requires 1369.4)

Acknowledgements

F.A. would like to thank the Department of Education, Science and Training (DEST)
and the Faculty of Science, Queensland University of Technology for Post-Graduate
Scholarships.

158
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

REFERENCES

1 (a) J.-C. Chambron, V. Heitz and J.-P. Sauvage, in Porphyrin Handbook, 60


ed. K. M. Kadish, K. M. Smith and R. Guilard, Academic Press, San Diego,
2000, vol. 6, pp. 1–42; (b) C. M. Drain, I. Goldberg, I. Sylvain and A. Falber,
Top. Curr. Chem., 2005, 245, 55–88.
2 (a) J. K.M. Sanders, N. Bampos, Z. Clyde-Watson, S. L. Darling, J. C.
Hawley, H.-J. Kim, C. C. Mak and S. J.Webb, in Porphyrin Handbook, ed.
K. M. Kadish, K. M. Smith and R. Guilard, Academic Press, San Diego,
2000, vol. 3, pp. 1–48; (b) J. K. M. Sanders, in Porphyrin Handbook, ed. K.
M. Kadish, K.M. Smith and R. Guilard, Academic Press, San Diego, 2000,
vol. 3, pp. 347–368.
3 T. S. Balaban, Acc. Chem. Res., 2005, 38, 612–623; Y. Kobuke, Eur. J.
Inorg. Chem., 2006, 2333–2351.
4 J. Deisenhofer and H. Michel, Science, 1989, 245, 1463–1473.
5 J. S. Lindsey and J. N. Woodford, Inorg. Chem., 1995, 34, 1063–1069.
6 R. R. Bucks and S. G. Boxer, J. Am. Chem. Soc., 1982, 104, 340–343; R. J.
Abraham, A. E. Rowan, N. W. Smith and K. M. Smith, J. Chem. Soc., Perkin
Trans. 2, 1993, 1047–1059; R. G. Brereton and J. K. M. Sanders, J. Chem.
Soc., Perkin Trans. 1, 1983, 423–430; R. G. Brereton and J. K. M. Sanders, J.
Chem. Soc., Perkin Trans. 1, 1983, 431–434; I. S. Denniss and J. K. M.
Sanders, Tetrahedron Lett., 1978, 295–298; K. N. Ganesh, J. K. M. Sanders
and J. C. Waterton, J. Chem. Soc., Perkin Trans. 1, 1982, 1617–1624; K. M.
Smith, F.W. Bobe, D. A. Goff and R. J. Abraham, J. Am. Chem. Soc., 1986,
108, 1111–1120.
7 M. Nsangou, A. Ben Fredj, N. Jaidane, M. G. Kwato Njock and Z. Ben
Lakhdar, THEOCHEM, 2004, 681, 213–224; M. Nsangou, A. Ben Fredj, N.
Jaidane, M. G. Kwato Njock and Z. Ben Lakhdar, THEOCHEM, 2005, 726,
245–251.
8 Y. Kobuke and H. Miyaji, Bull. Chem. Soc. Jpn., 1996, 69, 3563–3569.
9 N. N. Gerasimchuk, A. A. Mokhir and K. R. Rodgers, Inorg. Chem., 1998,
37, 5641–5650.
10 J. Wojaczyński and L. Latos- Grażyński, Coord. Chem. Rev., 2000, 204, 113–
171; T. Imamura and K. Fukushima, Coord. Chem. Rev., 2000, 198, 133–

159
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

156; A. Satake and Y. Kobuke, Tetrahedron, 2004, 61, 13–41; E. Iengo, E.


Zangrando and E. Alessio, Acc. Chem. Res., 2006, 39, 841–851; Y. Kobuke,
Struct. Bonding, 2006, 121, 49–104.
11 D. F. O’Shea, M. A. Miller, H. Matsueda and J. S. Lindsey, Inorg. Chem.,
1996, 35, 7325–7338.
12 S. G. Otavin, Dopov. Akad. Nauk Ukr. RSR, Ser. B: Geol., Khim. Biol. Nauki,
1978, 165–169.
13 C. M. Che, T. F. Lai, W. C. Chung, W. P. Schaefer and H. B. Gray, Inorg.
Chem., 1987, 26, 3907–3911.
14 J. W. Buchler, C. Dreher and K. L. Lay, Chem. Ber., 1984, 117, 2261–2274;
M. Inamo,N.Matsubara, K. Nakajima, T. S. Iwayama, H. Okimi and M.
Hoshino, Inorg. Chem., 2005, 44, 6445–6455.
15 T. Mashiko, M. E. Kastner, K. Spartalian, W. R. Scheidt and C. A. Reed, J.
Am. Chem. Soc., 1978, 100, 6354–6362; A. Hoshino and M. Nakamura,
Chem. Lett., 2004, 33, 1234–1235.
16 Y. Li, J.-S. Huang, G.-B. Xu, N. Zhu, Z.-Y. Zhou, C.-M. Che and K.-Y.
Wong, Chem.–Eur. J., 2004, 10, 3486–3502.
17 H. Inoue, K. Chandrasekaran and D. G.Whitten, J. Photochem., 1985, 30,
269–284.
18 (a)D. P. Arnold, Y. Sakata,K.-i. Sugiura and E. I.Worthington, Chem.
Commun., 1998, 2331–2332; (b) R. D. Hartnell, A. J. Edwards and D. P.
Arnold, J. Porphyrins Phthalocyanines, 2002, 6, 695–707; (c) M. J. Hodgson,
P. C. Healy, M. L. Williams and D. P. Arnold, J. Chem. Soc., Dalton Trans.,
2002, 4497–4504; (d) R. D. Hartnell and D. P. Arnold, Organometallics,
2004, 23, 391–399; (e) R. D. Hartnell and D. P. Arnold, Eur. J. Inorg. Chem.,
2004, 1262–1269; (f)M. J. Hodgson, V. V. Borovkov, Y. Inoue and D. P.
Arnold, J. Organomet. Chem., 2006, 691.
19 (a) F. Atefi, O. B. Locos, M. O. Senge and D. P. Arnold, J. Porphyrins
Phthalocyanines, 2006, 10, 176–185; (b) F. Atefi, J. C. McMurtrie, P. Turner,
M. Duriska and D. P. Arnold, Inorg. Chem., 2006, 45, 6479–6489.
20 Y. Matano, K.Matsumoto, Y. Terasaka, H. Hotta, Y. Araki, O. Ito,M. Shiro,
T. Sasamori, N. Tokitoh and H. Imahori, Chem.–Eur. J., 2007, 13, 891–901.
21 K. A. Connors, Binding Constants: The Measurements of Molecular Complex
Stability, John Wiley & Sons, New York, 1987.

160
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

22 M. Tabata and J. Nishimoto, inPorphyrin Handbook, ed. K. M.Kadish, K. M.


Smith and R. Guilard, Academic Press, San Diego, 2000, vol. 9, pp. 221–419.
23 SigmaPlot 10.0, Systat Software, Inc., 2006, San Jose.
24 M. Gouterman, in The Porphyrins, ed. D. Dolphin, Academic Press,
Vancouver, 1978, vol. 3, pp. 1–165.
25 Y. Kobuke and H. Miyaji, J. Am. Chem. Soc., 1994, 116, 4111–4112; C. J.
Chang, E. A. Baker, B. J. Pistorio, Y. Deng, Z.-H. Loh, S. E. Miller, S.D.
Carpenter andD.G. Nocera, Inorg. Chem., 2002, 41, 3102–3109. 15
26 D. Bryce-Smith, Inorg. Synth., 1960, 6, 9–11.
27 J. K. Laha, S. Dhanalekshmi, M. Taniguchi, A. Ambroise and J. S. Lindsey,
Org. Process Res. Dev., 2003, 7, 799–812.

161
Copyright © 2007 The Royal Society of Chemistry
Chapter 6

ELECTRONIC SUPPLEMENTARY INFORMATION

Multiporphyrin coordination arrays based on complexation of magnesium(II)


porphyrins with porphyrinylphosphine oxides
by
Farzad Atefi, John C. McMurtrie and Dennis P. Arnold

S2 Fig. S2 1H NMR of 2
S3 Fig. S3 1H NMR of 3
S4 Fig. S4 1H COSY NMR of 3
S5 Fig. S5 1H NOESY NMR of 3
S6 Fig. S6 1H COSY NMR of 8
S7 Fig. S7 1H NOESY NMR of 8
S8 Fig. S8 1H NMR of 9
S9 Fig. S9 1H COSY NMR of 9
S10 Fig. S10 1H NOESY NMR of 9
S11 Fig. S11 1H NMR of 10
S12 Fig. S12 1H COSY NMR of 10
S13 Fig. S13 1H NOESY NMR of 10
S14 Fig. S14 1H COSY NMR of 11
S15 Fig. S15 1H NOESY NMR of 11
S16 Fig. S16 1H NMR of 12
S17 Fig. S17 1H COSY NMR of 12

162
Chapter 6

Figure S2. 1H NMR of 2.

163
Chapter 6

164
Chapter 6

Figure S3. 1H NMR of 3.

Figure S4. 1H COSY NMR of 3.

165
Chapter 6

Figure S5. 1H NOESY NMR of 3.

166
Chapter 6

Figure S6. 1H COSY NMR of 8.

167
Chapter 6

Figure S7. 1H NOESY NMR of 8.

168
Chapter 6

Figure. S8 1H NMR of 9.

169
Chapter 6

Figure S9. 1H COSY NMR of 9.

170
Chapter 6

Figure S10. 1H NOESY NMR of 9.

171
Chapter 6

Figure S11. 1H NMR of 10.

172
Chapter 6

Figure S12. 1H COSY NMR of 10.

173
Chapter 6

Figure S13. 1H NOESY NMR of 10.

174
Chapter 6

Figure S14. 1H COSY NMR of 11.

175
Chapter 6

Figure S15. 1H NOESY NMR of 11.

176
Chapter 6

Figure S16. 1H NMR of 12.

177
Chapter 6

Figure S17. 1H COSY NMR of 12.

178
CHAPTER 7

Discussion, conclusion and future work

179
Chapter 7

7. Discussion, conclusion and future work

7.1. Discussion and conclusion

This PhD project has shown that organometallic porphyrins are attractive systems for
scientists from various backgrounds. Strategies and materials described in this thesis
are the basis for future research which should be undertaken in order to assess fully
the scope of this field. Essential biological processes, in which porphyrins play key
roles, are the driving forces behind a broad spectrum of investigations and
organometallic porphyrins are important components in the quest for simpler
synthetic mimics of these biological systems.

Iodoporphyrins 4.3, 4.9-4.12 were prepared from the respective η1-


palladioporphyrins or bromoporphyrins (chapter 4) and are valuable starting
materials for further functionalisations of porphyrins. The high yielding, rapid and
selective methodology is a great improvement to previous synthetic strategies and
due to the mild reaction conditions it should be applicable to a variety of diaryl- and
triarylporphyrins. This halogen exchange reaction to form iodinated organic
fragments could be potentially utilised for other synthetic compounds as well. The
necessity of stoichiometric amounts of expensive organopalladium reagents in this
methodology might seem to limit the full potential of this transformation reaction.
However palladium was recovered from the reaction mixture and could be in
principle recycled, making this strategy also very attractive for large scale synthetic
applications. The utility of these iodoporphyrins was confirmed in a palladium-
catalysed coupling reaction to prepare a porphyrinylphosphine oxide 4.13 and the X-
ray single crystal structure of iodoporphyrin 4.3 confirmed the structure of the
chromophore unambiguously.

Porphyrinylphosphine oxide 5.6 was prepared from η1-palladioporphyrin 5.5 as


described in chapter 5. The reaction conditions were optimised by utilising 5.5
instead of improving the catalytic reaction itself. This methodology could in theory
be applied to any palladium-catalysed reaction. These substitution reactions are today
one of the most versatile and widely used functionalisation pathways in porphyrin
chemistry, and indeed in general organic chemistry. However the optimisation

180
Chapter 7

process is often challenging, as various ligands, solvents and bases have to be


evaluated in order to obtain the most favourable requirements. By using η1-
palladioporphyrins, the first intermediate in proposed catalytic cycles, this
challenging process can be simplified.

These findings were subsequently successfully applied to isolate a suite of


porphyrinylphosphine oxides (5.6, 5.12-5.16) under catalytic conditions. Chapter 5
shows that these macrocycles, which represent a new class of porphyrin, were
cleanly prepared in very high yields once the optimum reaction conditions were
obtained. Lower solubility for Zn(II) phosphine oxides 5.15 and 5.16 without
coordinating solvents indicated the self-coordination of these macrocycles. All
porphyrinylphosphine oxides were studied in detail by various spectroscopic and
electrochemical techniques. Mass spectra confirmed the formation of phosphine
oxides and strong peaks representing the P=O vibrations were observed in the IR
spectra of the macrocycles. A decrease of ~25 cm-1 in the P=O vibration frequencies
of 5.15 and 5.16 in comparison to the other phosphine oxide analogues again
31
demonstrated the O-Zn coordination. P NMR analysis showed the expected peaks
for the phosphorus fragments. Variable–temperature 1H NMR revealed an intrinsic
lack of symmetry for free base phosphine oxide 5.6, due to the tautomerism of the
inner NH protons, a feature which was observed at room temperature for the first
time in our experience. The fluorescence quantum efficiency of 5.6 and 5.12 was
slightly increased in comparison to unsubstituted porphyrin, thus revealing that the
diphenylphosphine oxide does not behave as a “heavy atom” quencher. This feature
is very important for further possible applications as synthetic mimic of the
photosynthetic process. Brief electrochemical investigations revealed the electron
withdrawing properties of phosphine oxide substituents.

The X-ray single crystal structure of Ni(II) bis(phosphine oxide) 5.14 (chapter 5) is a
typical example for the ruffled conformation of four-coordinate Ni(II) porphyrins,
unlike the crystal structure of η1-palladioporphyrin 3.58, which adopts a hybrid of
the ruffled and saddled conformations. Interestingly both diphenylphosphine oxide
groups project on the same side of the macrocycles in a cis-like geometry. The X-ray
single crystal structure of Zn(II) bis(phosphine oxide) 5.16 showed a five coordinate
Zn(II) centre with approximate square pyramidal stereochemistry. The apical

181
Chapter 7

coordination site of the Zn(II) centre is occupied by an oxygen atom from one of the
diphenylphosphine oxide groups of a neighbouring macrocycle. This configuration
produces a chainlike, one-dimensional coordination polymer. In this case the two
diphenylphosphine oxide groups protrude on opposite sides of the chromophore in a
trans-like arrangement.

The findings of chapter 5 demonstrated the possibility to prepare a variety of


multiporphyrin assemblies through axial coordination with porphyrinylphosphine
oxides which were further investigated in chapter 6. At first the binding strength of
triphenylphosphine oxide towards Mg(II) porphyrin 6.2 was evaluated. The 1H NMR
and UV-visible spectra of 6.3 indicated a strong affinity of the phosphorus-oxo
ligand towards the Mg(II) porphyrin. The calculated displacement constant of 5.3 ×
105 M-1 was two orders of magnitude larger than any other Mg(II) porphyrin-ligand
binding constant reported thus far. Next a variety of di- and triporphyrin arrays (6.8-
6.11) consisting of Mg(II) porphyrin 6.2 coordinated to free base and Ni(II)
porphyrinyl mono- and bis-phosphine oxides (6.4-6.7) were prepared. These
compounds represent a new class of multiporphyrin complexes. All products were
isolated in very high yields and were stable in organic solvents for extended periods
and in the solid state for at least several months. The phosphine oxide ligands also
increased the solubility of the Mg(II) porphyrin in organic solvents. The large upfield
shifts of the 1H NMR peaks of the complexes in comparison to the uncomplexed
macrocycles indicated that the two macrocycles were in close proximity. The
excitonic couplings observed in the UV-visible spectra were a good indication for the
inter-porphyrin electronic interaction. Ni(II) complexes 6.9 and 6.11 exhibit
intramolecular fluorescence quenching, which showed efficient electronic
communication in the excited state.

Mg(II) porphyrinylphosphine oxide dimer 6.12 was isolated in a satisfactory yield as


discussed in chapter 6. UV-visible spectroscopy again indicated inter-porphyrin
electronic interaction and the fluorescence spectra showed a more rigid face-to-face
geometry compared to the other complexes (6.8-6.12). The large self-association
constant of 5.5 × 108 M-1 was again proof for the strong affinity of phosphine oxide
ligands towards Mg(II) porphyrins establishing these complexes as the first strongly

182
Chapter 7

bound synthetic Mg(II) porphyrin analogues of the “special pair” of the


photosynthetic reaction centre.

7.2 Future Work

Investigations into the formation and reactions of Class F-H organometallic


porphyrins are almost in their infant stage. As shown in chapter 3, only very few
metal and metalloid fragments have been directly attached to the carbon framework
of porphyrins and more investigations are needed if we hope to unveil the full
potential of this interesting area of research.

The oxidative-addition methodology to form η1-palladio- and platinioporphyrins


(chapter 3) is to date the most successful strategy to prepare organometallic
porphyrins. Recent reports have proposed the formation of η1-Ni(II)-1 and η1-Ir(III)
porphyrin intermediates2 under catalytic conditions. It is therefore very likely that
these organometallic porphyrins can be isolated from the stoichiometric reaction of a
haloporphyrin or unsubstituted porphyrin position with the respective metal
fragment. Various additional strategies have also been pursued in the present work in
order to try to prepare other organometallic porphyrins and these are presented in
Appendix A.

The transmetallation of η1-palladioporphyrins with other metal fragments such as


Sn(IV) and Si(IV) should lead to the formation of the respective organometallic
porphyrins. This strategy has been briefly investigated during the course of this
project (Appendix A). Due to time constraints the desired macrocycles have not been
isolated yet. However the initial results suggest that porphyrinylstannanes can be
prepared by this methodology.

The formation and reactions of electrophilic porphyrin cation radicals have been
discussed in chapter 2. Similarly, nucleophilic porphyrin anions would be of
immense value for subsequent functionalisation reaction. Recent reports of Grignard
porphyrin analogues3 were encouraging to pursue such a synthetic strategy.
Unfortunately all attempts to form such macrocyclic anion synthons chemically were
unsuccessful (Appendix A). However porphyrin anion radicals have been prepared
electrochemically4 and these species should react with electrophilic metal fragments

183
Chapter 7

to yield the desired organometallic porphyrins (Figure 7.1.). There can be no doubt
that different organometallic porphyrins will emerge in the future which will lead to
even more potential applications for these macrocycles.
-
N N N N N N
e-
M M + "M'Ln" - "H" M M'Ln
N N N N N N

M = Mg, Ni, Cu, Zn


M' = Si, Sn, P, Au, Pd, Pt, Fe, Ru, Os, Rh, Ir, Zr

Figure 7.1. Concept of the preparation of organometallic porphyrins from a


porphyrin anion radical intermediate (charges are not shown and exact reaction
sequence is not implied).

Porphyrin anion radicals might also be applicable to form metal bridged porphyrin
dimers utilising already existing organometallic porphyrins such as the robust
porphyrinylmercurials (1) and η1-platinioporphyrins (2). The former have thus far
only been utilised as synthons for subsequent functionalisations (chapter 3). Detailed
spectroscopic and electrochemical studies on the monomeric mercurials should
therefore commence first. The effect of such direct metal bridges is at this stage only
speculative, however the physical and electrochemical characteristic of other
organometallic multiporphyrin arrays suggest that the metal bridge should alter the
properties of the macrocycles considerably and might also lead to interesting inter-
porphyrin interactions. The analogues of 2 with ethynylporphyrin units on Pt have
been prepared and indeed show electronic coupling across the Pt centre.5

N N
Hg M N N L N N
N N M' Pt M
N N
M' N N L N N
N N

1 2

This project has shown that porphyrinylphosphine oxides are very interesting new
ligands for supramolecular chemistry. Further experiments have to be carried out in
order to exploit the full potential of these novel ligands. Simpler phosphine oxides

184
Chapter 7

have been utilised as ligands for other metalloporphyrins, such as Os(II),6 Cr(III),7
Fe(III),8 Ru(IV)9 and Sn(IV).10 It would be of great interest to assess the bridging
properties of porphyrinylphosphine oxides towards these metalloporphyrins.
Especially the complexation with six-coordinate metal fragments should lead to
attractive coordination frameworks.

The effect of the substituents on the phosphorus on the macrocycle should also be
evaluated. Smaller alkyl substituents, rather than the phenyl groups used so far,
might lead to different electronic properties and different complexation behaviour of
the phosphine oxides. The formation of a triporphyrinylphosphine oxide by a similar
synthetic strategy as discussed in chapter 5 should produce a multiporphyrin ligand
with considerably altered physical, chemical and structural properties (Figure 7.2.),
although the steric challenges of such a situation may be difficult to overcome. The
corresponding phosphite or phosphate could be accessible from the
hydroxyporphyrin with a PX3 equivalent.

N
N
M
N
N
N N N N
M Br +"PH3 equivalent" M P O
N N N N
N
N
M
N
N

Figure 7.2. Preparation of triporphyrinylphosphine oxide.

Porphyrins attached to surfaces through anchoring groups such as carboxylates, have


been investigated for molecular information storage11 and as components for
porphyrin-sensitised solar cells.12 Attaching porphyrinylphosphine oxides or the
similarly accessible porphyrinylphosphonates to surfaces such as TiO2 could result in
attractive materials for such applications. The porphyrins could be anchored to the
surface in a perpendicular sense by using mono(phosphonates), while
bis(phosponates) should result in parallel anchoring (Figure 7.3.).

185
Chapter 7

O O
N N
P M P
N N
O O O O

Figure 7.3. Attachment of porphyrinylphosphonic acids to TiO2 surface.

Porphyrin oligomers consisting of porphyrinyl phosphine oxides and robust


organometallic porphyrins such as η1-platinioporphyrins would be of great interest.
A triply linked porphyrin dyad (3) incorporating the electron withdrawing properties
of phosphine oxides and the electron donating properties of η1-platinioporphyrins
might improve the electron transfer properties of such oligomers.

R R N N N N L
P M M' Pt Br
O N N N N L

Further evaluation of dimer 6.12 as synthetic mimic for the special pair of the
photosynthetic reaction centre should also be pursued. In the first stage the solubility
of the dimer in organic solvents has to be increased in order to simplify further
spectroscopic and electrochemical measurements. This might be achieved by using
different aryl groups on the macrocycles, such as tolyl-, mesityl-, di-t-butyl-phenyl or
long-chain alkyl groups. The substituents on the phosphorus might also alter the
solution properties of the macrocycles. The dimer should in a subsequent step be
attached to electron acceptor units (4) similar to the porphyrin dyad prepared by
Kobuke and co-workers (2.16).13 Detailed studies of the electrochemical,
photophysical and structural properties will be necessary in order to assess possible
further applications of such porphyrin dyads.
Ar
R R O
N N O O
P Mg
N N NH N N OMe
O
Ar Ar O
O O
N O
O O O N
Mg P
N N
MeO N N HN R R
Ar
4
O O O

186
Chapter 7

7.3. References

(1) Liu, C.; Shen, D.-M.; Chen, Q.-Y. J. Org. Chem. 2007, 72, 2732-2736.
(2) Hata, H.; Shinokubo, H.; Osuka, A. J. Am. Chem. Soc. 2005, 127, 8264-8265.
(3) Therien, M. J.; DiMagno, S. G. In International Patent; (University of
Pennsylvania, USA). USA, 1994, p 62 pp. (b) Therien, M. J. In International
Patent; (The Trustees of the University of Pennsylvania, USA). USA, 2002, p
76.
(4) (a) Mack, J.; Stillman, M. J. J. Porphyrins Phthalocyanines 2001, 5, 67-76.
(b) Kadish, K. M.; Van Caemelbecke, E.; Royal, G. In The Porphyrin
Handbook; Kadish, K. M., Smith, K. M., Guilard, R., Eds.; Academic Pres:
San Diego, 2000; Vol. 8, p 1-114.
(5) Chen, Y.-J.; Chen, S.-S.; Lo, S.-S.; Huang, T.-H.; Wu, C.-C.; Lee, G.-H.;
Peng, S.-M.; Yeh, C.-Y. Chem. Commun. 2006, 1015-1017.
(6) Che, C. M.; Lai, T. F.; Chung, W. C.; Schaefer, W. P.; Gray, H. B. Inorg.
Chem. 1987, 26, 3907-11.
(7) (a) Buchler, J. W.; Dreher, C.; Lay, K. L. Chem. Ber. 1984, 117, 2261-74. (b)
Inamo, M.; Matsubara, N.; Nakajima, K.; Iwayama, T. S.; Okimi, H.;
Hoshino, M. Inorg. Chem. 2005, 44, 6445-6455.
(8) (a) Mashiko, T.; Kastner, M. E.; Spartalian, K.; Scheidt, W. R.; Reed, C. A. J.
Am. Chem. Soc. 1978, 100, 6354-62. (b) Hoshino, A.; Nakamura, M. Chem.
Lett. 2004, 33, 1234-1235.
(9) Li, Y.; Huang, J.-S.; Xu, G.-B.; Zhu, N.; Zhou, Z.-Y.; Che, C.-M.; Wong, K.-
Y. Chem.--Eur. J. 2004, 10, 3486-3502.
(10) Inoue, H.; Chandrasekaran, K.; Whitten, D. G. J. Photochem. 1985, 30, 269-
84.
(11) Jiao, J.; Anariba, F.; Tiznado, H.; Schmidt, I.; Lindsey, J. S.; Zaera, F.;
Bocian, D. F. J. Am. Chem. Soc. 2006, 128, 6965-6974.
(12) Imahori, H.; Hayashi, S.; Umeyama, T.; Eu, S.; Oguro, A.; Kang, S.; Matano,
Y.; Shishido, T.; Ngamsinlapasathian, S.; Yoshikawa, S. Langmuir 2006, 22,
11405-11411.
(13) Ozeki, H.; Nomoto, A.; Ogawa, K.; Kobuke, Y.; Murakami, M.; Hosoda, K.;
Ohtani, M.; Nakashima, S.; Miyasaka, H.; Okada, T. Chem.--Eur. J. 2004,
10, 6393-6401.

187
APPENDIX A

Further investigations into the preparation of


novel σ-bonded organometallic porphyrins

I
Appendix A

A. Further investigations into the preparation of novel σ-bonded


organometallic porphyrins

During the course of this PhD project various synthetic strategies were pursued in
order to prepare novel σ-bonded organometallic porphyrins. These results were often
inconclusive or did not lead to the formation of the desired product. Nonetheless it is
important to document these experiments for future investigations.

A.1. Transmetallation: Sn, Si, Pt, Au

The palladium-catalysed C-C bond forming reaction of organic electrophiles, such as


halides or triflates, with organotin reagents is often referred to as the Stille protocol
and has been well documented.1 This methodology has also been widely used to
functionalise porphyrins.2 Similarly, organosilyl reagents have been successfully
utilised as cross-coupling reagents in palladium-catalysed substitution reactions.3
Therefore it was proposed that the transmetallation of a η1-palladioporphyrin with
hexamethylditin or hexamethyldisilane should lead to the formation of the desired
porphyrinylstannane or porphyrinylsilane, respectively (Scheme A.1.).
Ph Ph

Ph Ph
NH N P NH N
Pd P + M M M
Ph
N HN Br Ph N HN

M = Sn or Si
Ph Ph
A2: M = Sn
A1 A3: M = Si

Scheme A.1. Proposed formation of porphyrinylstannane and porphyrinylsilane by


transmetallation.

While reactions with the silyl reagent formed only 5-bromo-10,20-diphenylporphyrin


and unsubstituted 5,15-diphenylporphyin, reactions with hexamethylditin led to the
formation of a new product as indicated by 1H NMR (Figure A.1.).

II
Appendix A

Figure A.1. 1H NMR spectrum of the reaction mixture after stannylation in CDCl3.

A typical experimental procedure is described here: Porphyrin A1 (20 mg, 0.019


mmol) was dried in a Schlenk vessel and subsequently flushed three times with ultra-
high purity (UHP) argon. Hexamethylditin (45 mg, 1.37 mmol), which was weighed
out under an argon shower, was dissolved in a separate Schlenk flask in 6 ml freshly
distilled toluene. The mixture was degassed three times by freeze-thawing and 2.5 ml
of the tin solution (0.057 mmol) was added to the vessel containing the porphyrin by
syringe. The reaction mixture was stirred at 85º C. During the reaction time, aliquots
of the mixture were removed, dried under high vacuum and subjected to 1H NMR.
The peaks representing the starting material completely disappeared after 26 hours
and two major porphyrin fractions were present in the 1H NMR spectrum (Figure
A.1.).

The peaks marked with “+” represent unsubstituted 5,15-diphenylporphyrin, while


the peaks marked with “*” represent an unknown porphyrin. The chemical shifts of
these peaks do not match the starting material A1 nor 5-bromo-10,20-
diphenylporphyrin. Due to the presence of four doublets, each representing two β-
protons, it can be concluded that three meso-positions of the macrocycles are

III
Appendix A

substituted. A trimethylstannyl peak with two tin-satellites at ~1.1 ppm


(hexamethylditin: ~0.2 ppm) indicates the presence of a Sn(IV) fragment in the
periphery of the porphyrin. The peaks representing the phosphorus-phenyl groups on
the palladium fragment were also not present, thus eliminating the formation of a Pd-
Sn substituted porphyrin. It can therefore be proposed that the desired
porphyrinylstannane A3 was formed under these reaction conditions. The integrals of
the peaks indicate the presence of the two porphyrins in a ~ 1:1 ratio. Unfortunately
the porphyrin decomposed during purification by column chromatography and due to
time constraints this methodology could not be further investigated. Nonetheless the
desired porphyrin A3 should be isolable using this strategy. Different solvents,
porphyrin starting materials and the use of an appropriate base might lead to the
formation of A3 in higher yields. This should result in a simplified purification
process (e.g. recrystallisation). Further optimisation reactions should be carried out at
80-85º C, as higher temperatures led only to the formation of unsubstituted 5,15-
diphenylporphyin and no reactions were observed at lower temperatures. The fact
that the product is stable in air and in chlorinated solvents is also promising for future
studies.

As discussed in chapter 3, η1-platinioporphyrins with chelating phosphine ligands


have not yet been isolated. It was proposed that the transmetallation of
porphyrinylboronates with platinum salts might yield the desired product. Similarly
porphyrinylboronates were reacted with gold salts to try to prepare
porphyrinylgold(I) complexes (Scheme A.2.).

IV
Appendix A

Ph

[Pt(dppe)]2+ NH N
Pt(dppe)Cl2 O
base
AgOTf + or or + B
Au(PEt3)Cl [AuPEt3]+ O
N HN

Ph
dppe = 1,2-bis(diphenylphosphino)ethane
OTf = triflate (trifluoromethanesulfonate) A4

Ph Ph

Ph Ph
NH N P NH N

Pt P or Au PEt3
Ph
N HN Cl N HN
Ph

Ph Ph
A5 A6

Scheme A.2. Proposed formation of porphyrinylplatinum(II) and porphyrinylgold(I)


complexes by transmetallation.

The following typical experimental procedure was carried out: Silver triflate (9 mg,
0.035 mmol) and the metal salt [Pt(dppe)Cl2, 11 mg, 0.017 mmol or Au(PEt3)Cl 12
mg, 0.034 mmol) were dried under high vacuum in a two neck round bottomed flask
equipped with a septum. The flask was subsequently flushed three times with UHP
argon before adding 5 ml of freshly distilled THF. In a separate Schlenk vessel
porphyrin A44 (10 mg, 0.017 mmol) was dried under high vacuum before the vessel
was flushed three times with UHP argon. After stirring for 30 minutes, the mixture
containing the metal reagent was filtered by cannula into the Schlenk vessel
containing the porphyrin. A white precipitate could always be observed at this stage
confirming the formation of AgCl. Aliquots of the reaction mixture were removed,
which at this stage had turned from purple to green-brown. The mixture was
neutralised with triethylamine, after which the colour turned back to purple, dried
under high vacuum and dissolved in CDCl3, before the 1H NMR spectra were
recorded. Even after prolonged reaction times (up to three weeks) only porphyrin A4
together with unsubstituted 5,15-diphenylporphyrin were recovered as indicated by

V
Appendix A

1
H NMR spectroscopy. η1-Platinioporphyrins are known to be stable in air and
chlorinated solvents. Therefore it can be only concluded that no reaction took place
under these conditions.

A.2. Formation of porphyrin anion radicals and Grignard porphyrin analogues:


Pt, Au, Sn

The formation and reactions of electrophilic porphyrin cation radicals have been
discussed in chapter 3. Similarly it has been proposed that nucleophilic porphyrin
anion radicals would react with metal salts to form σ-bonded organometallic
porphyrins. Two synthetic strategies were investigated in order to prepare such an
anionic species. In the first, Ni(II) porphyrin A7 was reacted with sodium
naphthalenide. This powerful organic reductant, which has been often utilised to
generate various radical hydrocarbons,5 was prepared in situ and reacted with Ni(II)
porphyrin A7 to generate the desired radical species A8 according to a typical
experimental procedure described bellow (Scheme A.3.).6
Ph

- N N
Na + Na+ + Ph Ni
N N
-
Ph

Ph
N N A7

Ph Ni
N N

Ph
A8

Scheme A.3. Proposed formation of porphyrin anion radical by treatment with


sodium naphthalenide.

All glassware used for these experiments was dried in the oven for at least twelve
hours and subsequently flame-dried under high vacuum. A two-neck round bottomed
flask equipped with a septum and a glass coated stirrer bar was charged with freshly
cut Na (33 mg, 1.4 mmol). Subsequently the Na was dried under high vacuum before

VI
Appendix A

the vessel was flushed three times with UHP argon. After the addition of freshly
sublimed LR grade naphthalene (167 mg, 1.3 mmol) the flask was cooled in an
ice/water/NaCl mixture before adding 20 ml freshly distilled THF. The flask was
removed from the ice bath before the mixture was sonicated for one hour at room
temperature. At this stage the clear solution turned dark green indicating the
formation of the naphthalenide and no solid sodium could be observed anymore. The
mixture was stirred for additional three hours to ensure the completion of the
reaction. In a separate Schlenk vessel porphyrin A7 (20 mg, 0.034 mmol) was dried
under high vacuum before the vessel was flushed three times with UHP argon.
Afterwards the porphyrin was dissolved in 5 ml freshly distilled THF and degassed
by three cycles of freeze-thawing before 6 ml of the sodium naphthalenide solution
(0.39 mmol) was added by syringe. At this stage the colour of the reaction mixture
turned from orange to brown. This mixture was stirred for an additional three hours
at room temperature before adding Au(Et3P)Cl (15 mg, 0.043 mmol) dissolved in 1
ml freshly distilled and freeze-thawed THF. After an aqueous workup the 1H NMR
spectrum of the reaction mixture was recorded, which indicated the formation of
demetallated 5,10,15-triphenylporphyrin. The demetallation of A7 under these
reaction conditions was not unexpected, as the demetallation of porphyrins under
very strong basic or reducing conditions has been well documented.7

In a similar experiment the formation of the porphyrin anion radical was attempted
utilising freshly prepared activated magnesium (Mg*).8 In this case the formation of
small amounts of 5,10,15-triphenylporphyrinatomagnesium(II) in addition to
demetallated A7 was also observed. The Mg(II) insertion into porphyrins using
Grignard reagents7a,9 is a well established procedure and it can be therefore
concluded that under these reaction conditions A7 was first demetallated and
subsequently small amounts of Mg(II) were inserted into the free base porphyrin..

In the second synthetic strategy it was attempted to prepare a Grignard porphyrin


analogue as reported by the Therien group in two patents.10 In this methodology
freshly prepared activated magnesium (Mg*)9 was reacted with 5-iodo-10,15,20-
triphenylporphyrinato zinc(II) (A9) in situ (Scheme A.4.). Subsequently the
porphyrin (A10) was treated with gold, tin or platinum salts in order to obtain the

VII
Appendix A

desired σ-bonded organometallic porphyrins. A control reaction with benzaldehyde


was also carried out in order to assess the reactivity of A10.
Ph Ph

N N N N
Ph Zn I + Mg * Ph Zn MgI
N N N N

Mg* = activated Mg
Ph Ph
A9 A10

Scheme A.4. Proposed formation of Grignard porphyrin analogue.

A typical experimental procedure is described here: All glassware used for these
experiments was dried in the oven for at least twelve hours and subsequently flame-
dried under high vacuum. A two-neck round bottomed flask equipped with a
condenser, a septum and a glass coated stirrer bar was charged with freshly cut Mg
(225 mg, 8.5 mmol). After drying the Mg under high vacuum the flask was flushed
three times with UHP argon and 20 ml freshly distilled and three times freeze-thawed
THF was added. After the addition of 1,2-dibromoethane (0.810 ml, 4.7 mmol) the
reaction was stirred at room temperature. Vigorous bubbling was observed after 5
min. which ceased after further 10 min. of stirring. Subsequently the mixture was
refluxed for one hour, the solvents removed and the solids dried under high vacuum
for one hour at 150º C to obtain anhydrous MgBr2. Freshly cut Li (68 mg, 8.6 mmol)
was added to the same flask and the solids were dried under high vacuum (30 min.)
before the vessel was flushed three times with UHP argon. The flask was then
charged with AR grade naphthalene (109 mg, 8.50 mmol) and 10 ml of freshly
distilled and freeze-thawed THF. The mixture was stirred for 24 hours after which a
black-grey suspension of highly activated Mg was formed. In a separate Schlenk
flask, which was equipped with a glass coated stirrer, porphyrin A9 (50 mg, 0.069
mmol) was dried under high vacuum before the flask was flushed three times with
UHP argon. The porphyrin was dissolved in 3 ml freshly distilled THF and the
mixture was degassed by three cycles of freeze-thawing. Afterwards the Mg*
suspension (~1ml, 0.085 mmol) was transferred by cannula into the Schlenk vessel
and the reaction mixture was stirred for 24 hours.

VIII
Appendix A

The porphyrin reaction mixture was treated subsequently either with Pt(dppe)Cl2,
(Bu)3SnCl, AuCl(Et3)P or with benzaldehyde. All reactions gave after aqueous
workup only unsubstituted 5,10,15-triphenylporphyrinato zinc(II). It seems unlikely
that the desired Grignard porphyrin analogue A10 was formed under these reaction
conditions as treatment with benzaldehyde, which is known to be very reactive under
Grignard conditions, should have resulted in an addition reaction. Future experiments
should therefore investigate the electrochemical formation of porphyrin anion
radicals and their subsequent reactions with metal salts.

A.3. Formation of porphyrinylgold(I) complexes

The auration of heteroarenes with gold salts has been successfully achieved.11 It was
therefore proposed that porphyrins might react in a similar fashion with a gold salt to
give the desired porphyrinylgold(I) complexes (Scheme A.5.).
Ar Ar

N N N N
Zn + (Et3P)AuPF6 Zn Au(PEt3)
N N N N

Ar = di-t-Bu-Ph
Ar Ar
A11 A12

Scheme A.5. Proposed formation of porphyrinylgold(I) complex.

In a typical experimental procedure, (Et3P)AuCl (27 mg, 0.077 mmol) and AgPF6
(20 mg, 0.079 mmol) were dried in a round bottomed flask equipped with a septum
and a stirrer bar under high vacuum, before the vessel was flushed three times with
UHP argon. After addition of 7 ml freshly distilled and freeze-thawed THF the
reaction mixture was stirred for 10 min. at room temperature. In a separate Schlenk
vessel porphyrin A11 (25 mg, 0.033mmol) was dried under high vacuum before the
flask was flushed three times with UHP argon. The porphyrin was dissolved in 1 ml
freshly distilled and freeze-thawed THF and the Au-solution was filtered by cannula
into the Schlenk flask. A white precipitate could be observed at this stage in the
round bottomed flask indicating the formation of AgCl. Even after prolonged
reaction times (up to 72 hours) and heating of the reaction mixture (up to 70º C in a

IX
Appendix A

sealed Schlenk flask) no reactions were observed and only porphyrin A11 was be
recovered. It can be concluded that the formation of porphyrinylgold(I) complexes
by this synthetic strategy seems unlikely, as reactions with Ni(II)- and free base
porphyrins gave similar results.

A.4. References

(1) (a) Stille, J. K. Pure Appl. Chem. 1985, 57, 1771-80. (b) Stille, J. K. Angew.
Chem. 1986, 98, 504-19. (c) Kosugi, M.; Fugami, K. J. Organomet. Chem.
2002, 653, 50-53. (d) Kosugi, M.; Fugami, K. In Handbook of
Organopalladium Chemistry for Organic Synthesis; Negishi, E.-I., Ed.; John
Wiley & Sons, Inc.: 2002; Vol. 1, p 263-283. (e) Espinet, P.; Echavarren, A.
M. Angew. Chem., Int. Ed. 2004, 43, 4704-4734.
(2) (a) Setsune, J.-i. J. Porphyrins Phthalocyanines 2004, 8, 93-102. (b)
Sharman, W. M.; van Lier, J. E. J. Porphyrins Phthalocyanines 2000, 4, 441-
453.
(3) Hayashi, T. In Handbook of Organopalladium Chemistry for Organic
Synthesis; Negishi, E.-i., Ed.; John Wiley & Sons, Inc.: 2002; Vol. 1, p 791-
806.
(4) Hyslop, A. G.; Kellett, M. A.; Iovine, P. M.; Therien, M. J. J. Am. Chem. Soc.
1998, 120, 12676-12677.
(5) (a) Holy, N. L. Chem. Rev. 1974, 74, 243-77. (b) Connelly, N. G.; Geiger,
W. E. Chem. Rev. 1996, 96, 877-910.
(6) Chi, K. M.; Frerichs, S. R.; Stein, B. K.; Blackburn, D. W.; Ellis, J. E. J. Am.
Chem. Soc. 1988, 110, 163-71.
(7) (a) Buchler, J. W. In The Porphyrins; Dolphin, D., Ed.; Academic Press: New
York, 1978; Vol. 1, p 389-483. (b) Sanders, J. K. M.; Bampos, N.; Clyde-
Watson, Z.; Darling, S. L.; Hawley, J. C.; Kim, H.-J.; Mak, C. C.; Webb, S. J.
In The Porphyrin Handbook; Kadish, K. M., Smith, K. M., Guilard, R., Eds.;
Academic Press: San Diego, 2000; Vol. 3, p 1-48.
(8) (a) Rieke, R. D.; Li, P. T.-J.; Burns, T. P.; Uhm, S. T. J. Org. Chem. 1981,
46, 4323-4. (b) Burns, T. P.; Rieke, R. D. J. Org. Chem. 1987, 52, 3674-80.
(9) (a) Lindsey, J. S.; Woodford, J. N. Inorg. Chem. 1995, 34, 1063-9. (b)O'Shea,
D. F.; Miller, M. A.; Matsueda, H.; Lindsey, J. S. Inorg. Chem. 1996, 35,
7325-7338.

X
Appendix A

(10) (a) Therien, M. J.; DiMagno, S. G. In International Patent; (University of


Pennsylvania, USA). USA, 1994, p 62. (b) Therien, M. J. In International
Patent; (The Trustees of the University of Pennsylvania, USA). USA, 2002, p
76.
(11) Porter, K. A.; Schier, A.; Schmidbaur, H. Organometallics 2003, 22, 4922-
4927.

XI

You might also like