Advances in Catalytic Activation of Dioxygen Metal Complexes [1st Edition.] 1441952381

The catalytic activation of dioxygen continues to attract interest both due to its biological importance and synthetic p

234 1 15MB

English Pages 348 Year 2010

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
528.pdf......Page 1
528_001.pdf......Page 14
528_002.pdf......Page 91
528_003.pdf......Page 134
528_004.pdf......Page 167
528_005.pdf......Page 237
528_006.pdf......Page 275
528_007.pdf......Page 339
Recommend Papers

Advances in Catalytic Activation of Dioxygen Metal Complexes [1st Edition.]
 1441952381

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

ADVANCES IN CATALYTIC ACTIVATION OF DIOXYGEN BY METAL COMPLEXES

Catalysis by Metal Complexes Volume 26

Editors: Brian James, University of British Columbia, Vancouver, Canada Piet W. N. M. van Leeuwen, University of Amsterdam, The Netherlands Advisory Board: Albert S.C. Chan, The Hong Kong Polytechnic University, Hong Kong Robert Crabtee, Yale University, U.S.A. David Cole-Hamilton, University of St Andrews, Scotland István Horváth, Eotvos Lorand University, Hungary Kyoko Nozaki, University of Tokyo, Japan Robert Waymouth, Stanford University, U.S.A.

The titles published in this series are listed at the end of this volume.

ADVANCES IN CATALYTIC ACTIVATION OF DIOXYGEN BY METAL COMPLEXES edited by

László I. Simándi Chemical Research Center, Hungarian Academy of Sciences, Budapest

KLUWER ACADEMIC PUBLISHERS NEW YORK, BOSTON, DORDRECHT, LONDON, MOSCOW

eBook ISBN: Print ISBN:

0-306-47816-1 1-4020-1074-5

©2002 Kluwer Academic Publishers New York, Boston, Dordrecht, London, Moscow Print ©2003 Kluwer Academic Publishers Dordrecht All rights reserved

No part of this eBook may be reproduced or transmitted in any form or by any means, electronic, mechanical, recording, or otherwise, without written consent from the Publisher

Created in the United States of America

Visit Kluwer Online at: and Kluwer's eBookstore at:

http://kluweronline.com http://ebooks.kluweronline.com

v

Preface The subject of dioxygen activation and homogeneous catalytic oxidation by metal complexes has been in the focus of attention over the last 20 years. The widespread interest is illustrated by its recurring presence among the sessions and subject areas of important international conferences on various aspects of bioinorganic and coordination chemistry as well as catalysis. The most prominent examples are ICCC, ICBIC, EUROBIC, ISHC, and of course the ADHOC series of meetings focusing on the subject itself. Similarly, the number of original and review papers devoted to various aspects of dioxygen activation are on the rise. This trend is due obviously to the relevance of catalytic oxidation to biological processes such as dioxygen transport, and the action of oxygenase and oxidase enzymes related to metabolism. The structural and functional modeling of metalloenzymes, particularly of those containing iron and copper, by means of low-molecular complexes of iron, copper, ruthenium, cobalt, manganese, etc., have provided a wealth of indirect information helping to understand how the active centers of metalloenzymes may operate. The knowledge gained from the study of metalloenzyme models is also applicable in the design of transition metal complexes as catalytsts for specific reactions. This approach has come to be known as biomimetic or bioinspired catalysis and continues to be a fruitful and expanding area of research. This book is the sequel of the monograph Catalytic Activation of Dioxygen by Metal Complexes by the editor of the present volume, published by Kluwer Academic Publishers in 1992 as Volume 13 of the Series Catalysis by Metal Complexes. Ten years later it is appropriate to cover the developments in selected areas of the field, which has been the objective of this edited volume in the same series. It is my great pleasure to thank the authors of individual chapters for their excellent contributions. The following prominent scientists have accepted invitations to review major areas of dioxygen activation: Brian R.. James (Catalytic oxidations using ruthenium porphyrins), Kenneth D. Karlin (Copper-dioxygen complexes and their roles in biomimetic oxidation reactions), Roger A. Sheldon (Catalytic oxidations of alcohols), Takuzo Funabiki (Functional model oxygenations by nonheme iron complexes) and Craig L. Hill (Catalysis for selective aerobic oxidation under ambient conditions). A chapter on Catalytic oxidations using cobalt(II) complexes has been contributed by myself.

vi I am grateful to my wife Tatiana for her expert help at various stages of the work and critical reading of Chapter 6. I thank Mary Egresi for her assistance in the manuscript prepration. László I. Simándi Budapest August 2002

vii

Contents Preface Contributors About the Editor

Chapter 1 Catalytic oxidations using ruthenium porphyrins

v xiii xiv

1

Maria B. Ezhova and Brian R. James Department of Chemistry, University of British Columbia, 2036 Main Mall, Vancouver, BC, V6T1Z1, Canada

1. INTRODUCTION: OXYGENASE AND OXIDASE ACTIVITY 2. REACTIONS OF RUTHENIUM PORPHYRIN COMPLEXES WITH AND OTHER OXIDANTS 3. OXIDATION OF ORGANIC SUBSTRATES 3.1 Oxidation of phosphines, phosphites, arsines and stibines 3.2 Oxidation of thioethers 3.3 Epoxidation of olefins 3.4 Oxidation of saturated hydrocarbons 3.5 Oxidative-dehydrogenation of phenols and other arenes 3.6 Oxidative-dehydrogenation of alcohols 3.7 Oxidative dehydrogenation of amines 4. CONCLUSIONS 5. ABBREVIATIONS 6. REFERENCES

3 12 16 16 19 21 40 44 47 51 61 64 66

Chapter 2 Copper-dioxygen complexes and their roles in biomimetic oxidation reactions 79 Christiana Xin Zhang, Hong-Chang Liang, Kristi J. Humphreys and Kenneth D. Karlin Department of Chemistry, Johns Hopkins University, Baltimore, MD 21218, USA

1. INTRODUCTION 1.1 Practical Copper Oxidative Processes 1.2 Copper in Biology 1.2.1 Hemocyanin, Tyrosinase, and Catechol Oxidase 1.2.2 Amine Oxidases; Galactose Oxidase

79 80 80 81 84

viii

1.2.3 Cytochrome c Oxidases 2. COPPER-DIOXYGEN ADDUCTS 2.1 Copper-Dioxygen Complex Generation; 1984-1999 2.2 Recent Further Advances in Copper-Dioxygen Complex Generation 3. COPPER OXYGENASE CHEMISTRY 3.1 Aromatic Hydroxylation 3.2 Recent Tyrosinase Models 3.3 TPQ Biogenesis 3.4 Aliphatic Hydroxylation 4. COPPER OXIDASE MODELS; CATALYTIC ALCOHOL OXIDATION 5. COPPER-PHENANTHROLINE DNA OXIDATION 6. REFERENCES

Chapter 3 Catalytic oxidations of alcohols

86 86 87 93 97 97 99 101 103 107 111 116

123

R.A. Sheldon and I.W.C.E. Arends Biocatalysis and Organic Chemistry, Delft University of Technology, Julianalaan 136, 2628 BL Delft, The Netherlands

1. 2. 3. 4. 5. 6. 7.

123 124 126 138 144 146

INTRODUCTION MECHANISMS RUTHENIUM-CATALYZED OXIDATIONS WITH PALLADIUM-CATALYZED OXIDATIONS WITH COPPER-CATALYZED OXIDATIONS WITH OTHER METALS AS CATALYSTS FOR OXIDATION WITH CATALYTIC OXIDATION OF ALCOHOLS WITH HYDROGEN PEROXIDE AND ALKYL HYDROPEROXIDES 8. CONCLUDING REMARKS 9. REFERENCES

148 151 152

Chapter 4 Functional model oxygenations by nonheme iron complexes

157

Takuzo Funabiki Department of Molecular Engineering, Graduate School of Engineering, Kyoto University, Sakyo-ku, Kyoto 606-8501, Japan

1. INTRODUCTION

158

ix 2. HEME AND NONHEME OXYGENASES 3. FUNCTIONAL MODEL STUDIES ON NONHEME IRON DIOXYGENASES 3.1 Catechol Dioxygenases 3.1.1 Intradiol Cleavage Oxygenations 3.1.2 Extradiol Cleavage Oxygenations 3.1.3 Mechanisms of Oxygenations 3.2 Dioxygenases other than Catechol Dioxygenases 4. FUNCTIONAL MODEL SYSTEMS FOR NONHEME IRON MONOOXYGENASES 4.1 Functional Model Oxygenations by Diiron Complexes 4.1.1 Monooxygenation by Diiron Complexes with Peroxides 4.1.2 Monooxygenation by Diiron Complexes with Molecular Oxygen 4.2 Monooxygenation by Diiron Complexes 4.3 Functional Model Oxygenations by Iron Species in the Polyoxometalate and Heterogeneous Matrix 4.3.1 Oxygenation by Iron Species in the Homogeneous System 4.3.2 Oxygenation by Iron Species in the Heterogeneous System 4.4 Functional Model Oxygenations by Mono-Iron Species 4.4.1 Oxygenation by Mono-Iron Complex/Activated Oxygen System 4.4.2 Oxygenation by Mono-Iron System 4.4.3 Mono-Iron Oxygen Species 5. FROM FUNCTIONAL MODEL TO CATALYSIS 6. REFERENCES

Chapter 5 Catalysis for selective aerobic oxidation under ambient conditions. Thioether sulfoxidation catalyzed by gold complexes

159 161 161 162 170 173 179 181 183 183 187 187 189 189 190 191 191 198 202 204 207

227

Eric Boring, Yurii V. Geletii and Craig L. Hill Department of Chemistry, Emory University, Atlanta, Georgia 30322, USA

1. INTRODUCTION CATALYTIC 2. DISCOVERY OF OXIDATION SYSTEM 3. STOICHIOMETRIC Au(III) REDUCTION BY THIOETHERS

228 230 231

x 4. IN SITU CATALYST PREPARATION 5. REACTION STOICHIOMETRY 6. EMPIRICAL REACTION RATE LAW 7. RATE LIMITING STEP 8. PROPOSED REACTION MECHANISM 9. MECHANISMS RULED OUT 10. ORIGIN OF OXYGEN IN SULFOXIDE PRODUCT; ROLE OF IN SULFOXIDATION 11. REOXIDATION OF Au(I) BY DIOXYGEN. CATALYST PREPARATION FROM Au(I) COMPLEX 12.EFFECT OF LIGANDS ON REACTIVITY 13.PRODUCT INHIBITION (DMSO EFFECT) 14. CO-CATALYSIS BY TRANSITION METAL IONS 15. SOLVENT EFFECTS 16.HETEROGENEOUS SYSTEMS 17.EFFECT OF AMINO ACIDS 18. OXIDATION OF THIOETHERS OTHER THAN CEES 19.EXPERIMENTAL DETAILS 20.CONCLUSIONS

Chapter 6 Catalytic oxidations using cobalt(II) complexes

232 233 235 237 238 242 244 245 247 249 251 252 255 256 257 259 261

265

László I. Simándi Chemical Research Center, Institute of Chemistry, Hungarian Academy of Sciences, H-1525 Budapest, P.O. Box 17, Hungary

1. INTRODUCTION 2. COBALT DIOXYGEN COMPLEXES 3. OXIDATIONS CATALYZED BY Co(salen) COMPLEXES 3.1 Oxidation of substituted phenols 3.1.1 2,6-di-tert-butylphenol 3.2 Oxidation of 3,5-di-tert-butylcatechol 3.3 Oxidation of lignin phenolics 3.4 Nitrogen monoxide 3.5 Oxidative dehydrogenation 3.6 Oxidation of quercetin 3.7 Mercaptoethanol 3.8 Alkenes and alcohols 3.9 Alkene epoxidation 3.10 Primary amines 4. OXIDATIONS CATALYZED BY COBALOXMES

266 267 269 269 269 270 270 274 275 276 276 276 278 280 280

xi 4.1 4.2 4.3 4.4 4.5

Oxidation of o-phenylenediamine Oxidation of 2-aminophenol Oxidation of 3,5-di-tert-butylcatechol Oxidative cleavage of a stilbene derivative Oxygen insertions 4.5.1 Oxygen insertion to terminal P, C and N atoms 4.6.2 Insertion of into alkylcobaloximes 5. OXIDATIONS CATALYZED BY COBALT(II) PORPHYRINS 6. OXIDATION WITH COBALT(II) PHTHALOCYANINES 7. OXIDATIONS CATALYZED BY COBALT(II) AMINE COMPLEXES 7.1 Catalytic oxidation of substituted anilines 7.2 Oxidation of miscellaneous substrates 8. OXIDATIONS CATALYZED BY COBALT(II) PYRIDINE COMPLEXES epoxidation 8.1 9. COBALT-FENTON SYSTEMS 10. CATALYZED OXIDATIONS 10.1 10.2 Substituted phenols 10.3 Pinanediols 11. OXIDATIONS VIA ALKYLPEROXOCOBALT COMPLEXES 12.OXIDATIONS WITH Co-CYCLIDENE COMPLEXES 13. OXIDATIONS WITH COBALT PEPTIDE COMPLEXES 15.CARBOXYLATOCOBALT COMPLEXES AND SALTS 16.MISCELLANEOUS COBALT CATALYSTS 17.CONCLUSIONS 18. REFERENCES

281 282 283 287 289 289 290 292 298

Subject Index

329

301 301 306 307 307 308 309 309 309 310 311 313 316 317 320 322 323

xiii

Contributors I.W.C.E. Arenda, Organic Chemisty and Catalysis, Delft University of Technology, Julianalaan 136, 2628 BL Delft, The Netherlands Eric Boring, Department of Chemistry, Emory University, Atlanta, Georgia 30322, USA Maria B. Ezhova, Department of Chemistry, University of British Columbia, 2036 Main Mall, Vancouver, BC, V6T1Z1, Canada Takuzo Funabiki, Professor, Department of Molecular Engineering, Graduate School of Engineering, Kyoto University, Sakyo-ku, Kyoto 6068501, Japan Yurii V. Geletii, Department of Chemistry, Emory University, Atlanta, Georgia 30322, USA Craig L. Hill, Professor, Department of Chemistry, Emory University, Atlanta, Georgia 30322, USA Kristi J. Humphreys, Department of Chemistry, Johns Hopkins University, Baltimore, MD 21218, USA Brian R. James, Professor, Department of Chemistry, University of British Columbia, 2036 Main Mall, Vancouver, BC, V6T1Z1, Canada Kenneth D. Karlin, Professor, Department of Chemistry, Johns Hopkins University, Baltimore, MD 21218, USA Hong-Chang Liang, Department of Chemistry, Johns Hopkins University, Baltimore, MD 21218, USA Roger A. Sheldon, Professor, Organic Chemisty and Catalysis, Delft University of Technology, Julianalaan 136, 2628 BL Delft, The Netherlands László I. Simándi, Professor, Institute of Chemistry, Chemical Research Center, Hungarian Academy of Sciences, H-1525 Budapest, P.O. Box 17, Hungary Christiana Xin Zhang, Department University, Baltimore, MD 21218, USA

of Chemistry,

Johns

Hopkins

xiv

About the Editor László I. Simándi is Head of the Department of Coordination Chemistry at the Institute of Chemistry, Chemical Research Center, Hungarian Academy of Sciences and Professor of Chemistry at L. Eötvös University (both in Budapest). His research interests include catalysis by cobalt, iron and manganese complexes (biomimetic dioxygen activation, catalytic oxidation and carbonylation), as well as kinetics and mechanisms of inorganic reactions in solution (fast redox and electron transfer). He was visiting professor at the University of Texas (Arlington) and lectured widely as invited speaker at conferences and seminars in the US, Europe and Japan. He was the organizer of the Fourth International Symposium on the Activation of Dioxygen and Homogeneous Catalytic Oxidation (Balatonfüred, Hungary, 1990), and co-organizer of the XXII International Conference on Coordination Chemistry (Budapest, 1982). He is member of the Hungarian Chemical Society and Fellow of the Royal Society of Chemistry. He is the author of the monograph Catalytic Activation of Dioxygen by Metal Complexes (Kluwer, 1992) and the editor of Dioxygen Activation and Homogeneous Catalytic Oxidation (Elsevier, 1991).

Chapter 1 Catalytic oxidations using ruthenium porphyrins

Maria B. Ezhova and Brian R. James Department of Chemistry, University of British Columbia, 2036 Main Mall, Vancouver, BC, V6T 1Z1, Canada

Abstract: The major goal of the Chapter is to review developments in the use of Ru-porphyrin complexes as homogeneous (or matrix-supported) catalysts for oxygenation and oxidation processes. The subject was given impetus with the discovery of a remarkable reaction in which a Ru(II) porphyrin complex reacted with to give a trans-dioxo-Ru(VI) species. Such species, which can be formed from a wide range of O-atom donors, were shown subsequently to be capable of acting as a bis(monooxygenase) in transferring both the coordinated oxo ligands (as O-atoms) to olefinic substrates, saturated hydrocarbons, phosphines, and thioethers, and the processes become catalytic in the presence of excess of the O-atom donor. Further, the dioxo species can also exhibit oxidase-like activity, and effect stoichiometric or catalytic oxidative-dehydrogenation of phenols, alkoxyarenes, alcohols, and amines. Use of chiral porphyrins has led to catalytic, asymmetric epoxidation and hydroxylations, even though radical intermediates are invoked, as well as oxygenation of racemic substrates (phosphines and more interestingly tertiary alkanes) to yield chiral products by kinetic resolution processes. The reaction mechanisms invoked range from genuine O-atom transfer (from or species, where the disproportionation reaction is important), to free-radical induced processes, particularly when the porphyrin ligands are extensively halogenated, as Ru complexes generally of such porphyrins are extremely active in radical-type decomposition of hydroperoxides, often present as trace impurities in hydrocarbon substrates. 1 L.I. Simándi (ed.), Advances in Catalytic Activation of Dioxygen by Metal Complexes, 1-77. © 2003 Kluwer Academic Publishers. Printed in the Netherlands.

2

Maria B. Ezhova and Brian R. James

The amine dehydrogenation chemistry has led to characterization of Ru-imine, -imido (i.e. nitrene), -oxo(imido), -amido, as well as -amine complexes, the types of species that are intermediates in some reductive enzyme processes, and in reactions such as oxyamination, amination, and aziridination; the last mentioned process involves transfer of the imido moiety (isoelectronic with the oxo group), and corresponding catalytic aziridination of alkenes and insertion of the imido group into alkanes to give amides can be catalyzed by Ru-porphyrin complexes; with chiral porphyrins, chiral recognition in binding of racemic aminoesters has been realized and enantioeselective amidation of hydrocarbon substrates has been demonstrated. The beginning of the Chapter gives a general introduction to catalyzed oxygenation/oxidation reactions involving and briefly describes enzymatic oxygenase and oxidase systems, particularly the monooxygenase, cytochrome P-450, as the attempted modelling of this system using Ruporphyrin complexes led to the discovery of the trans-dioxo-Ru(VI) species and their chemistry. Key words:

amidation reactions, asymmetric hydroxylations, asymmetric epoxidation of alkenes, aziridination,

oxidation, chiral porphyrins, C=C

cleavage, Cu-containing proteins, cytochrome P-450, dioxoruthenium(VI) species, dioxygenases, epoxidation, galactose oxidase, Haber-Weiss reactions, halogenated porphyrins, hemoglobin, metalloporphyrins, monooxygenases, mustard gas, myoglobin, nitrene complexes, nitrene transfer, nitrous oxide, Oatom donors, oxidases, oxidation of olefins, oxidation of phosphines, oxidation of saturated hydrocarbons, oxidation of thioethers, oxidative dehydrogenation, oxidative dehydrogenation of alcohols, oxidative dehydrogenation of amines, oxidative dehydrogenation of phenols, porphyrins, pyridine-oxides, pyrocatechase, rebound mechanism, reductive activation of

ruthenium

amido complexes, ruthenium(IV) disproportionation, ruthenium imido complexes, ruthenium imine/amine complexes, ruthenium porphyrins, steroid epoxidation, supported catalysts, tryptophan dioxygenase.

1. Catalytic oxidations using ruthenium porphyrins

1.

3

INTRODUCTION: OXYGENASE AND OXIDASE ACTIVITY

Molecular oxygen is the most abundant and inexpensive oxygenating/oxidizing agent, which can effect in the presence of an appropriate catalyst a variety of useful oxidation reactions. These range from so-called di- or monooxygenase systems where two or one O-atoms of the are incorporated into the organic substrate (eqs. 1 and 2, respectively) to oxidative-dehydrogenation systems, where H-atoms are removed from organics as or (eqs. 3 and 4, respectively). In terms of strict, modern nomenclature, “oxygenase activity” (or O-atom incorporation) is effected by enzymatic di- or monooxygenases, and is represented by eqs. 1 and 2, while “oxidase activity” (H-atom abstraction) is effected by oxidases, and is represented by eqs. 3 and 4. Many studies with models (i.e. nonprotein systems) aim to mimic the enzyme systems, particularly their high selectivity (formation of a single product) and operation under ambient conditions.

Dioxygen is clearly an attractive oxidant (whether as an oxygenation or oxidizing agent), and is highly desirable when environmental requirements are considered as any inorganic co-product is typically water (eqs. 2 and 4). This may be contrasted with a classical stoichiometric oxidant such as dichromate (e.g. for alcohol oxidations), where undesirable Cr(III) is the coproduct. However, because of its biradical nature, non-coordinated reacts with organic substrates preferably according to a radical-chain mechanism, which generally operates with low selectivity. A typical example is of cyclohexene to give ene-one, ene-ol and epoxide via the intermediate cyclohexyl hydroperoxide, formed by attack of on the allylic radical, itself produced by hydrogen-atom abstraction (eq. 5): such reactions are catalyzed by transition metal salts and complexes which initiate

4

Maria B. Ezhova and Brian R. James

decomposition of hydroperoxides via so-called Haber-Weiss pathways (eqs. 6 and 7)1.

Such radical reaction pathways do not involve directly at the metal. Much more selective oxidation processes can be realized by activation of the through coordination to a metal center; the coordination is generally followed by transfer of electrons from the metal to the oxygen moiety (see below). The reactions shown in eqs. (1) - (4) can be exemplified by specific enzyme systems that incorporate such and these will be discussed briefly, but with greater emphasis on the mono-oxygenase type (eq. 2) because: (a) this type of activity is better understood, and (b) the majority of reported studies on model oxygenation systems has involved attempts at mimicking mono-oxygenase activity. A volume entitled “Oxygenases and Model Systems” has appeared within this series2, and discusses in detail reactions of the types shown in eqs. 1 and 2. As noted below, Fe-porphyrin complexes embedded in a protein environment are responsible in human systems for binding and transport of dioxygen, as well as for incorporation of one or two O-oxygen atoms into organic substrates via mono- or dioxygenase activity. Interest in the second-row analogues, Ruporphyrins, stems from this general “oxygen chemistry” but, before discussing the Ru systems in subsequent sections, a brief introduction to the enzyme chemistry of eqs. 1-4 is appropriate. Dioxygenase activity (eq. 1). Two examples of dioxygenase systems are provided by a pyrocatechase, which effects oxygenation of catechol to cis-muconic acid (eq. 8), and tryptophan dioxygenase, which catalyzes ring cleavage of tryptophan derivatives to the N-formyl products, called formylkynurenines (eq. 9). The active site within protocatechuate 3,4-dioxygenase (the enzyme that effects reaction 8, with as determined crystallographically, is high spin, 5-coordinate Fe(III) containing 2 tyrosine and 2 histidine ligands, and a coordinated water or hydroxide ((1), see Fig. 1) , and a postulated mechanism, based largely on EXAFS and EPR data, for the

1. Catalytic oxidations using ruthenium porphyrins

5

dioxygenase activity is shown in Figure 12,3. The catechol coordinates to

give (2) with the substrate phenolic protons used to remove the coordinated as water and one coordinated tyrosine; with (2) being written as an Fe(II)/o-semiquinone, subsequent binding of generates an intermediate

6

Maria B. Ezhova and Brian R. James

(3) set up to activate the substrate for attack by a dioxygen moiety, although details of the oxygen insertion and cleavage reactions remain obscure. No crystallographic data are available for L-tryptophan 2,3-dioxygenase (TPO), but the active site is now an Fe-heme centre2,4; that is, there is coordination to a porphyrin ligand, specifically protoporphyrin IX (PpIX), the same porphyrin present in myoglobin (Mb) and hemoglobin (Hb), the Fe(II)-containing proteins used by humans for storage and transportation of respectively. TPO is isolated as the Fe(III) form and spectroscopic data suggest that there is an axial histidine with water as a possible 6th ligand; at least in acid pH, the system is high spin. The system has to be reduced to the Fe(II) form before activity is realized and, although there is no direct evidence for formation of an (or intermediate, mechanisms involving prior binding of have been suggested, as well as mechanisms showing prior activation of tryptophan before reaction with particularly as L-tryptophan does combine with the Fe(III) and Fe(II) forms of the enzyme in the absence of Figure 2 shows suggested pathways involving reaction of a preformed species interacting with a deprotonated tryptophan to form eventually a dioxetane (4) that leads to the formyl products; as written here, the function of the heme is to localize and activate the Of note, ignoring the protein environment, deoxy-Mb and deoxy-Hb have active sites essentially identical to that of reduced TPO, and there are no tryptophan residues near the pocket5, and so the ‘simple’ reversible occurs efficiently unaccompanied by potentially deleterious ‘oxidation’ processes.

Monooxygenase activity (eq. 2). Monooxygenase activity is exemplified best by the well understood enzyme cytochrome P-450; this contains

1. Catalytic oxidations using ruthenium porphyrins

7

Fe(PpIX) as the prosthetic group with an axial thiolate ligand provided by a cysteinyl amino acid residue of the protein. Investigations on the structure and oxidizing ability of cytochrome P-450, as well as on the related model (biomimetic) systems, are extensive and the topic is well developed6-16. The protein and model studies have brought about detailed understanding of the activation of by P-450 and related systems; the scheme depicted in Figure 3 outlines the essential features. The scheme incorporates the so-called “reductive activation of dioxygen”, and demonstrates the counter-intuitive concept of requiring the presence of a reducing agent in order to execute transfer of a single O-atom in to an organic. The process is accomplished by coordination of dioxygen, a net transfer of two electrons (via the Fe center) and two protons, concomitant with the heterolytic cleavage of the oxygen-oxygen bond; one O-atom is

incorporated into the substrate via a high-valent metal-oxo species, while the second O-atom is reduced to water. The following steps of the process have been distinguished within the protein or model systems: (a) Addition of the substrate (S) to the resting state of the enzyme, a low spin Fe(III) state (5), gives a more easily reducible, high spin, Fe(III) enzyme-substrate complex

8

Maria B. Ezhova and Brian R. James

(6); crystallographic data for a enzyme that hydroxylates camphor have revealed that the camphor is held by the protein in close proximity to the Fe(III) center16. (b) A 1-electron reduction of (6) gives a high spin, 5coordinate Fe(II) complex (7). (c) (7) binds to give a low spin complex (8). (d) A second 1-electron reduction of the dioxygenated adduct (8) gives what is probably a high spin Fe(III)-peroxide (9)8,17. (e). Heterolytic cleavage of the oxygen-oxygen bond within (9), with consumption of protons, generates water and a high-valent oxo-iron species (10), written here as Fe(V)=O, although model studies favor a O=Fe(IV)(porp+.), porphyrin cation-radical species7-11. (f) Incorporation of the O-atom into the substrate gives the oxygenated product with concomitant regeneration of the initial Fe(III) state (5). The net reaction is that shown in eq. (2); the 2electrons are supplied by NADPH, and involve coupling to flavin reductase and putidaredoxin systems. The addition of the O-atom to S is usually written as a “rebound” mechanism, involving H-atom abstraction from the substrate by the Fe=O moiety and subsequent rebound of the OH group7. The protein pocket, in which cytochrome P-450 is embedded together with the amino-acid cysteine axial ligand, provides favorable and presumably optimal conditions for monooxygenase activity. In vitro (i.e. outside of a protein) metalloporphyrins, especially those containing naturally occurring porphyrins such as PpIX, can undergo undesirable side-reactions such as dimerization and aggregation and, under can be irreversibly oxidized (e.g. Fe and Ru-porphyrins give species - see below) with resulting loss of their oxygen-activating abilities. Efforts have thus been made to minimize these undesirable effects by introducing substituent alkyl or halogen substituents into, for example, the phenyl rings of the well known, easily synthesized meso-tetraphenylporphyrin ligand such modifications create steric hindrance against formation of species (loosely called “dimerization”), but also, of course, change the electrophilicity of the metal (or metal oxo) centre. The majority of biomimetic studies, including the use of a plethora of metal complexes, and not just metalloporphyrins, can be categorized and rationalized in terms of the pathways illustrated in Figure 3, especially when the so-called “shunt pathways” A and B are included. Basically, and of key importance, the P-450 studies show that generation of a high-valent metaloxo species such as (10) is needed to mimic monooxygenase activity. Clearly one way to achieve formation of model species akin to (10) via (9) is to follow pathways (b) - (d), i.e. utilize a metalloporphyrin (or other metal complex), and a reducing agent to supply the electrons11-15,18-31; a metal (M) in oxidation state (III) could generate a species. This type of system represents genuine monooxygenase-like character, i. e. is employed; however, a problem arises in that both the substrate (S) and the

1. Catalytic oxidations using ruthenium porphyrins

9

reducing agent will compete with the active form of the catalyst, the highvalent oxo species, and this renders such systems catalytically less efficient the competing reaction, catalytic of the reductant instead of (S) takes place! Many reducing agents, including borohydride, ascorbic acid, Zn dust (the GIF-type systems), aldehydes, CO and as well as electrons, have been tested with varying degrees of success. Such competition is avoided, of course, in the enzyme systems by membrane separation of the key components. Another general way to form species like (10) is to treat a precursor such as (6) with a net O-atom donor (XO, see Figure 3, pathway B); examples include iodosylbenzene, hypochlorite, chlorate, periodate, amine oxides, and Closely related are systems providing the dioxygen already reduced by 2-electrons, i.e. at the peroxide level using reagents such as organic hydroperoxides, peroxy acids, and magnesium monoperoxyphthalate, which again are net O-atom donors (see Figure 3, pathways A and B). A huge number of papers have been devoted to studies using the general “shunt pathways” exemplified by A and B (for reviews see refs. 10, 11, 14, and 32). Extensive investigations on metalloporphyrin-catalyzed oxidations of organic substrates using O-atom donors have shown that high-valent oxo-metal species generated in situ are the active species in the oxygenation of alkenes and alkanes6-15. However, in some cases, it has also been suggested that a key step of the catalytic reaction is the formation of an intermediate ternary complex composed of the metalloporphyrin, oxidizing agent and substrate33. Reactivity of metalloporphyrins with As both certain mono- and dioxygenase systems mentioned above utilize metalloporphyrins for activation (by binding) of it is worth reflecting on such systems that are known to react with Of the 1st-row transition metalloporphyrins, only those of Cr(II), Fe(II), Mn(II) and Co(II) bind dioxygen34,35. The products are type complexes with end-on geometry for the coordinated superoxide, except in the case of manganese which yields a Mn(IV) side-on peroxide type species. The Fe and Mn dioxygenated adducts of “non hindered” porphyrins are stable only at low temperatures, and at room temperature the species are too transient to be of catalytic use31. However, in the presence of a reducing agent and protons, bound and stabilized (as by the Fe(II)-picket fence porphyrin can be cleaved heterolytically and the system under showed some catalytic activity in the oxidation of olefins24. In contrast to the Fe and Mn systems, well defined adducts can be detectable in solutions at room temperature, although “dimerization” to bridged peroxo species is facile for non-sterically hindered porphyrins36; in any case, the bound is not a strong enough electrophile or nucleophile to effect epoxidation of olefins or hydroxylation of alkanes31.

10

Maria B. Ezhova and Brian R. James

The 2nd-row transition metal porphyrin complexes of Rh(II) and Ru(II) also react with but the resulting species are usually quite different in character. Rhodium octaethyl- and tetraphenylporphyrin oxygen adducts are like the Co analogues with an end-on superoxide; however, the octaethylporphyrin (OEP) species is only stable below -80°C and on warming to 20° C converts to a Rh(III) complex, analogous to the Co systems37,38. The interaction of oxidants, including with Ru porphyrins, and subsequent oxygenation/oxidation catalysis, constitute the topics of the remaining sections of this chapter; this review updates earlier 1992 and 1994 reviews from this laboratory on this topic15, while related reviews from the group of Che39 and Groves40 appeared in 1999 and 2000, respectively. Under certain conditions, “non-hindered” Ru(II) porphyrins can bind reversibly to give species that contain either coordinated superoxide or peroxide, although irreversible oxidation to a dinuclear species is more common (Section 2). Ru(II) porphyrins, initially containing bulky substituents at ortho-positions of meso-phenyl rings present in the porphyrin ligand, were shown to have a remarkable and unique reactivity with yielding trans-dioxo species which are stable in solutions at room temperature; an example is shown in eq. 10, where TMP = the dianion of 5,10,15,20-tetramesitylporphyrin, see Figure 4 in Section 2). Some nonhindered Ru(II)-porphyrins were later shown to form, in alcohol solutions, less stable trans-dioxo species (Section 2).

Over the last 17 years or so, studies based on these trans-Ru(VI)-dioxo complexes have led to the development of new, catalytic and selective oxygenation and oxidation systems that generally operate via non-radical pathways and that involve direct reaction between a metal complex and as illustrated in eq. 10. The Ru(VI)-dioxo species, depending on experimental conditions and substrates, can exhibit: (i) monooxygenase activity in terms of effecting addition of one O-atom of to a molecule of substrate, although both O-atoms of the are utilized for two molecules of substrate (cf. eq. 1), and so the systems are in effect illustrating dioxygenase activity based on the stoichiometric use of the molecule (Sections 3.13.5). Perhaps the term bis(monooxygenase) activity is appropriate; (ii) oxidase activity of the type illustrated in eq. 4 for certain dehydrogenation reactions (Section 4); and (iii) free-radical type activation of (eqs. 5-7). A very wide range of organic oxidation reactions is catalyzed by nonporphyrin complexes of Ru, including a vast literature on the use of Ru-oxo

1. Catalytic oxidations using ruthenium porphyrins

11

species; these systems are more commonly based on the use of O-atom donors. There are systems; however, these are typically radicalinduced processes and do not involve formation of oxo species. This extensive literature on Ru/non-porphyrin systems can be traced through refs. 41-43. Oxidase activity (eqs. 3 and 4). As the complexes (porp = porphyrin dianion) introduced above do exhibit oxidase activity (Sections 3.5-3.7), some brief elaboration of the enzymatic oxidase systems seems appropriate. To our knowledge, there are no oxidases that utilize metalloporphyrin centres as the active catalytic site. The majority of oxidases are Cu-containing protein systems44,45, and those exemplified in eq. 3 typically convert a primary alcohol moiety in a sugar residue to an aldehyde group. The “simplest” is probably galactose oxidase, which is one of a growing class of “free radical” enzymes, and it is only recently that the nature of the active site has been elucidated46,47. There is only a single redoxactive metal centre, which, in the Cu(II) state, is a so-called “normal, non blue, type-2 Cu”; the second required redox site needed to mediate the overall 2e-redox reaction is provided by a modified tyrosine radical, present as an equatorial ligand within an overall square pyramidal Cu site that contains (at pH 7) 2 histidines and water as the other equatorial ligands and a tyrosine as the axial ligand44,47. The axial tyrosine is thought to play a mechanistic role by facilitating abstraction of a proton from the substrate alcohol47. An oxidase system as outlined in eq. 4 typically transforms polyphenols/diols to quinone/diketones, as exemplified by lactase and ascorbate oxidase (eq. 11). Both these enzymes in the oxidized state contain, as well as the normal type-2 Cu, the “blue, type-1 Cu” (which exhibits an

intense blue colour, and an EPR signal with small Cu hyperfine coupling due to delocalization of the spin density towards a cysteine S-atom), and “type-3 Cu” (a centre that is EPR silent because of anti-ferrromagnetic exchange coupling between the Cu-atoms via a bridging ligand)48. The structure of ascorbate oxidase shows the 4 Cu centres quite close together: the type-1 Cu is coordinated by 2 histidines, 1 methionine, and 1 cysteine residue, the type-2 Cu has 3 histidines and a water and the type-3,

12

Maria B. Ezhova and Brian R. James

dimetallic centre contains a single bridging-OH and 3 terminal histidines on each Cu48. Oxidase mechanisms are not well understood in detail, but those presented here generally function via an electron transfer chain from the substrate to the various Cu centres through to which may or may not be coordinated to the metal. Studies on models for ascorbate oxidase have invoked reactivity via a ternary species, as well as oxidation by “free” essentially, the oxidizes the reduced metal and the ascorbate via its anionic, semiquinone radical1c.

2.

REACTIONS OF RUTHENIUM PORPHYRIN COMPLEXES WITH

AND OTHER

OXIDANTS Before detailing the catalytic oxygenations/oxidations that have been effected by Ru-porphyrin species, it is instructive to consider the reactions of these complexes with various oxidants in the absence of the substrates to be oxidized; clearly any such Ru products must be considered as catalysts or, at least, catalyst precursors. Studies on Ru-porphyrins in the laboratories here were initiated in the mid-1970s, with the aim of mimicking biological oxygenation/oxidation processes, especially those effected by their 1st-row, often naturally occurring Fe-based analogues8,15,49-53. In polar, aprotic solvents such as DMF, DMA (N,N’dimethylacetamide) or pyrrole, complexes of the type (L = solvent; porp = an anion of OEP or TPP) bind reversibly at ambient conditions to yield complexes50, but such species are readily converted in the presence of trace water to dinuclear Ru(IV) species of the type In toluene, under 50 undergoes a slow irreversible oxidation to a species , which is the usual oxidation product of Ru(II) “non hindered” porphyrin complexes using or an O-atom donor as oxidant8,53. The reaction of TPP complexes with in DMF at ~0°C is 1st-order in both metalloporphyrin and and the findings, particularly a Hammett plot, were rationalized in terms of formation of a Ru(IV), peroxo species54. Of note, myoglobin, which has been reconstituted with Ru(II)mesoporphyrin IX, is 6-coordinate low spin and undergoes, like

1. Catalytic oxidations using ruthenium porphyrins

13

(Im = imidazole), to the so-called met form via the outer-sphere process exemplified in eq. 12 (one L = histidine, the other L being unknown); the oxidation rate is faster than that of an axial ligand dissociation, which, for example, is the initial step in formation of Ru(porp)L(CO) via reaction with CO51.

A complex derived from has been stabilized within the protected cap on one side of a “picnic basket” porphyrin55, and at low temperature within a species56. An unusual demetallation reaction of a Ru(II) porphyrin complex with to give the insoluble, black and the free-base porphyrin is illustrated in eq. 13, trace again being the key ingredient57. The phosphine ligands were considered to be “burned off as phosphine oxides,

with a Ru(IV)=O species acting as oxidant the remaining ‘Ru(OEP)’, known in the form of the dimer then giving the known that was shown to react with to give (eq. 15). In the presence of excess the phosphine dissociation required

to generate the vacant site for coordination of (eq. 14) is prevented, and superoxide is formed via an outer-sphere electron transfer, according to eq. 12. Introduction of substituents at the ortho positions of the phenyl rings of creates steric hindrance against ‘dimerization’of the Ru derivatives (cf. eq. 15) and allows for an initial binding of that finally yields readily isolable, stable trans-Ru(VI)-dioxo species; e.g. in benzene, (L = MeCN15,59 or THF60,61) reacts with (or air) to yield (eq. 10, Figure 4). Its formation is thought to occur via steps of the type shown in eq. 14, followed by disproportionation of the Ru(IV)-monooxo

14

Maria B. Ezhova and Brian R. James

species, eq. 168,15,40,60,61. This equilibrium is evidenced by formation of on mixing samples of and in benzene under Ar at ambient conditions40,61; see also Section 3.3.

The species has been detected by as a monomeric, intermediate in the reaction of with air in benzene at ambient conditions en route to The monooxo species has 2 unpaired electrons, typical of paramagnetic, high spin, porphyrin species that obey a Curie plot49,62. A corresponding species was also evident from the reaction of with 1 mole equivalent of in benzene that generates which was initially thought to be noncoordinated40,61, although later kinetic data imply that the product is see Section 3.1; the v(RuO) value of for the dioxo64 was shifted to for the monooxo species, and correct isotope shifts were observed on using species. The oxidation state was also supported by an oxidation state marker at a narrow range in this IR region has been found to be very sensitive to the oxidation state of the metal in Ru-porphyrin complexes61,65-68.

The trans-Ru(VI)-dioxo species can effect stoichiometric and/or catalytic oxygenations/oxidations of organic substrates such as phosphines, thioethers,

1. Catalytic oxidations using ruthenium porphyrins

15

olefins, alkanes, alcohols, phenols and amines (see Section 3), via reaction pathways that illustrate the whole gamut of reactivity patterns”: oxygenase, oxidase, and free-radical activity, see Section 1. More details illustrating the importance of the disproportionation step (eq. 16) in the catalytic oxygenations are given later (especially in Section 3.3). Single oxygen atom donors, such as iodosylbenzene, mchloroperbenzoic acid (m-CPBA), and amine oxides, have also been reacted with Ru-porphyrins. In terms of mimicking the Fe(III) resting state of cyto P-450 (Fig. 3), a Ru(III) precursor is an obvious candidate; for example, reacts with PhIO according to the stoichiometry of eq. 17 to yield a green complex, isolable at room temperature and tentatively formulated as the oxo-Ru(IV) cation radical species, (11)8,69; an EPR signal at g = 2.00, the visible spectrum, and the

demonstrated stoichiometry of eq. 18, supported the formulation of (11)70, which showed some catalytic oxidation activity8,69 (Section 3.3). The Ru(III) precursors Ru(OEP)X (X = Cl, Br), synthesized from the corresponding dihalo-Ru(IV) complexes49,62 according to the chemistry outlined in eq. 1971, react with PhIO or m-CPBA again to generate a green solution, but work-up

procedures yield only species such as (12)71. Of note, corresponding Fe(IV)-oxoporphyrin cation-radical species are also green72,73. Addition of a 2-fold excess of m-CPBA, PhIO or to or benzene solutions of the sterically hindered Ru(porp)CO complexes (porp = TMP or TDCPP, Figure 4, cf. eq. 10) gives the species, which are well characterized15,40,61,64,74, including a crystal structure of the TDCPP complex40. The trans-dioxo complexes are sensitive to decomposition by traces of acids64 (see Section 3.4). Corresponding oxidation of the non-hindered Ru(TTP)CO with m-CPBA gave the species where L = OH or mchlorobenzoate64; similarly, Ru(OEP)CO with gives a product75. The derivatives with the non-hindered TPP and OEP can be made by reaction of the appropriate Ru(porp)CO with m-CPBA in alcohol, the formation of species being inhibited by coordination of

16

Maria B. Ezhova and Brian R. James

alcohol at the vacant site of the initially formed Ru(IV)(porp)O; the nonhindered dioxo complexes are reasonably stable in the solid state and in organic solvents, although they react readily with to form Useful precursors for a myriad of derivatives are the ‘bare” species Ru(TMP)59,76, and the dimeric and species58. For example, Ru(TMP), a 14-electron species, reacts readily with donor ligands ranging from (which gives CO, olefins, acetylenes, MeCN and ethers59,76,77. Figure 5 summarizes synthetic routes to the trans-dioxo species using or an O-atom donor59-61,67 .

3.

OXIDATION OF ORGANIC SUBSTRATES

3.1

Oxidation of phosphines, phosphites, arsines and stibines

of can be catalyzed by Ru(II) non-hindered porphyrin species57. The mechanism in benzene solution involves an initial outer-sphere oxidation of a 6-coordinate bis(phosphine) species which generates superoxide according to eq. 12 (porp = OEP; ). Although readily reduces species, the presence of protons (needed as a cocatalyst) forces the equilibrium of eq. 12 to the right via stabilization and subsequent disproportionation of to and (eq. 20). The oxidizes free while is regenerated using

1. Catalytic oxidations using ruthenium porphyrins

17

the 2-equiv. reducing power of phosphines in the presence of (eqs. 21, 22); the overall oxidation (reactions 12, 20-22) is catalytic in Ru, and That the oxidation proceeds faster than substitution of (e.g. by CO) supports the outer-sphere process, while EPR data on the system (eq. 12) showed formation of a mixture of the hydrated superoxide (g = 2.00, 2.10) and a low spin Ru(III) species (g = 1.98, 2.30)57.

In principle, any substrate that is oxidizable by can be oxidized stoichiometrically by a metal complex that allows for outer-sphere generation of and a catalytic process results if an appropriate reducing agent, preferably the substrate itself, is present to generate the lower-valent metal complex - in this case a Ru(II)-porphyrin. A related system (see Section 3.2) using with leads to catalytic oxidation of thioethers to sulfoxides, where the reaction corresponding to eq. 12 is photo-assisted78. The system gave non-reproducible kinetic data57, and it is possible that a photochemical pathway contributes to the reaction. The species (cf. eq. 18) can be reduced by excess to generate and thus can be used as a catalyst precursor for catalytic of at least under the outer-sphere conditions. can oxidize 2 moles of (or stoichiometrically, and a 3rd mole of can subsequently coordinate, according to eq. 23 in the presence of catalytic oxidation of

the phosphine can be readily effected at ambient conditions61,63,64,80. Initially, the detectable (13) was written without the axial ligand61,64, but kinetic data for the stoichiometric reaction point to the oxide being coordinated63. and were determined for the first O-atom transfer to (step a) by stopped-flow kinetics, as well as corresponding data for other phosphines; X = OMe, Me, F, Cl, values increase with increasing electron-withdrawing power of the p- substituent position, consistent with electrophilic attack of a Ru=O moiety on the lone-pair of the phosphine. The negative values

18

Maria B. Ezhova and Brian R. James

show an interesting trend becoming more favorable with increasing mass of the phosphines (and of and that were similarly oxidized), and this was rationalized in terms of O-atom transfer occurring via strong Ru=O vibrational coupling63,81. The kinetics for the slower, subsequent stages of the stoichiometric reaction, which monitored directly the conversion of to (again in the stopped-flow regime), were 1 st-order in the Ru complex, 1 st- to zero-order in and inverse on consistent with the steps shown in eqs. 24 -26. There was no direct evidence for the disproportionation step (eq. 25), but the kinetic data were most consistent with its involvement; some exhange data40 using imply that Ru(porp)O undergoes rapid exchange with water, while exchanges slowly, and the findings suggest that the disproportionation reaction is generally faster than, for example, Ru(porp)O reacting directly with a 2nd mole of oxidizable substrate, in this case (see also Section 3.3). Under 1 atm (or, less efficiently, air) at 24°C in benzene, in the presence of excess PPh3, does effect catalytic oxidation of but at a rate much lower than the stoichiometric reaction; the catalysis is governed by the rate of conversion by of the product of eq. 26, back to the dioxo species63.

Use of

containing a chiral picket-fence type porphyrin see Fig. 10, Section 3.3) in or under at 25°C can stoichiometrically oxidize the racemic phosphine to chiral phosphine oxide with 41% ee82,83. This reaction effects a kinetic resolution of the racemate, with the O-atom transfer occurring with retention of the configuration at the P-atom. The other reaction product is Ru(porpcontaining the phosphines as a mixture of SS : RS : RR isomers in the ratios 38 : 54 : 9. The stereoselectivity was explained by differences in reactivity of and Ru(porp)(O) toward the R- and S-isomers of the phosphine; this implies that Ru(porp)(O) is also an effective O-atom donor to phosphines, which conflicts with the kinetic interpretation discussed above. The relative capabilities of the dioxo versus the mono species for O-atom transfer is a question that will arise again in later sections. Equilibrium titration data for the addition of to were interpreted in

1. Catalytic oxidations using ruthenium porphyrins

19

terms of the dioxo species being a more powerful O-atom donor61; a key factor will be the nature of the ligand coordinated trans to the oxo ligand, and to draw a general conclusion is likely untenable. The mechanistic details of net O-atom transfer processes from metal-oxo species generally are of intense current interest14,15,32,80,84,85.

3.2

Oxidation of thioethers

Selective of thioethers to sulfoxides is of industrial importance74,86-88 and has been accomplished for dialkyl sulfides using complexes (porp = TMP, TDCPP)74,89. Kinetic and data in benzene solution near ambient conditions were first interpreted in terms of the chemistry shown in eq. 27. Under Ar, a stoichiometric reaction gives the bis(O-bonded sulfoxide) product (14), while under or air the labile sulfoxides are displaced to regenerate the trans-dioxo species and the

process becomes catalytic. The k values for the TMP system at 20°C are 0.0075, 0.012 and 0.11 respectively, for and the differences resulting more from differences in than values74; as with the phosphine systems, the determined negative values are consistent with O-atom transfer occurring via electronic coupling induced by strong Ru=O vibrational motion81, the data implying an easier Oatom transfer process with bulkier substrates. Alkylaryl and diaryl sulfides do not react with the trans-dioxo species74,89,90 presumably because of their decreased nucleophilicity. Of note, an Fe(TPP)Cl/PhIO system effects oxidation of dialkyl, alkylaryl and diaryl sulfides, via a proposed ClFe(TPP)O intermediate91; the axial ligand trans to the oxo ligand is likely critical. The chemistry of eq. 27 would imply that is a more efficient O-atom donor than while data for phosphine oxidation (see Section 3.1) imply that the Rumonooxo species disproportionates more quickly than reacting with phosphine. The kinetic data for thioether systems could be consistent with the type of mechanism described for the phosphine systems, and a careful reevaluation of the thioether systems is needed. With Fe systems, of course, disproportionation to a dioxo species is unknown. Relevant to this discussion are the findings that LRu(porp)(O) species are better catalyst precursors than the corresponding for oxidation of saturated hydrocarbons

20

Maria B. Ezhova and Brian R. James

(Section 3.4.) and some observations for the Ru(III)(OEP) systems (Section 3.3.), which again demonstrate the likely critical role of the axial ligand trans to the oxo ligand. That this is a key factor in biologically important7-9,11 and biomimetic Fe-porphyrin systems92 is well documented. The catalytic thioether oxidation (eq. 27) is limited by isomerization of the bis(O-bonded sulfoxide) to the more substitution inert bis(S-bonded) species, and by degradation of the TMP ligand74,89. The catalyst is more effective, the k value for oxidation being 10 times greater; complete conversion of 0.035 M to the sulfoxide and sulfone (4:1 mixture) can be accomplished at 100°C using mM catalyst with no degradation of the catalyst74. Substitution of halogens in to a porphyrin favours O-atom transfer by increasing the electrophilicity of the oxo ligand, and the decreased electron density in the ring also makes any metal-oxo catalyst less susceptible to attack by itself (self-destruction); this is related to an increase of reduction potentials within the systems92-94 (see also Section 3.3). In the thioether systems, the chlorine substituents perhaps reduce the extent of and slow the 15 isomerization process . The presence of chlorine in the thioether decreases the nucleophilicity of the S-atom to the extent that no O-atom transfer occurs from trans-Ru-dioxo species; such studies on bis(2-chloroethyl)sulfide (Mustard Gas) have led to isolation of the first (and possibly last!) structurally characterized Mustard Gas complex, Phenyl methyl and benzyl sulfides have been catalytically oxidized to corresponding sulfoxides (>80%) and sulfones ( > cyclooctene > while epoxidation of cis- and proceeds with retention of configuration60. The suggested catalytic cycle shown in Figure 6 implies the key disproportionation of a Ru(IV)=O intermediate to the and Ru(II) species (see also Section 2, eq. 16, and Section 3.1, eq. 25)60,

while a plausible route for reoxidation of Ru(II) is shown in eq. 1440,74. In catalyzed epoxidations of cis-but-2-ene in benzene, using in the presence of excess a 10:1 ratio of to was produced, and the 2 different pathways outlined in Figure 7 were considered in order to rationalize the data40: the dioxo species could be formed by disproportionation, or by cleavage of a dimer. Separate experiments had shown that the dioxo species undergoes oxygen exchange slowly with (see Fig. 7), and that epoxides do not undergo exchange; thus the pathway would be expected to generate a 1:1 mixture of the and epoxides, while the disproportionation pathway would yield in principle 100% The conclusion was that the observed 10:1 ratio thus favored the disproportionation pathway40, which is reasonable, with the presumption that undergoes rapid exchange, and that the does not. Although these data suggest that the fastest step in the epoxidations (Fig. 6) is the disproportionation reaction (ref. 40, see Fig. 7), incorporation from added to a reaction mixture does not necessarily imply the intermediacy of a metal-oxo species; other intermediates such as M-OOH or M(O-atom donor) could also have exchangeable O-atoms103. These selective of olefins, a commercially important process104, are based on the somewhat exotic hindered porphyrins, but are highly significant. Such O-atom transfer via an or air-generated metal-

1. Catalytic oxidations using ruthenium porphyrins

23

dioxo species represents unique chemistry, even though turnovers are only up to at 10 mM Ru and 0.5 M olefin. Stoichiometric epoxidation of para-substituted styrenes is 1st-order in both Ru-dioxo and styrene, at least for [styrene] up to 30 mM, and a Hammett plot shows increasing rate constants with electron-donating substituents, the data being consistent with a concerted process involving an electrophilic metal-oxo species40. Two important factors for applications of these Ru(VI)-dioxo/olefin systems using the TMP, TDCPP or TBCPP systems, and that could contribute to the low epoxidation rates, include: (a) a competitive binding of the olefin and the epoxide product at the Ru40,77,84,105 and (b) a build-up of catalytically inactive Ru(porp)CO species40,106. Ru(II) complexes generally have a strong affinity for a carbonyl moiety and, indeed, Ru-porphyrins under appropriate conditions can catalyze decarbonylation of aldehydes107, which are detected in trace amounts during epoxidation of styrenes40. Even a coordinated methyl group at a Ru-porphyrin centre has been transformed to a coordinated carbonyl108. After a catalyzed of oct-1-ene and propene to give 1,2-epoxyoctane and propene oxide, respectively, Ru(TMP)CO was isolated106; aldehyde was formed during the propene oxidation, while there was no reaction between aldehydes (or ketones) with Use of -oct-1-ene led to formation of and in a 3:1 ratio, showing that most of the CO was derived from the terminal C-atom of the olefin106. With cyclooctene at concentrations of ~0.5 M), the epoxidation rate becomes independent of [cyclooctene], implying a rate-determining step other than O-atom transfer60; this could result if either of factors (a) or (b) play a role, when regeneration

24

Maria B. Ezhova and Brian R. James

of the Ru-dioxo could become the rate-determining step and a dependence on would be expected; alternatively, genuine saturation kinetics might be operative, implying the build-up of a kinetically important ‘Ru-dioxoolefin’ ternary intermediate (cf. the benzhydrol system in Section 3.6.). Irradiation of the cyclooctene system with visible light from a tungsten lamp gives a five-fold increase in the epoxidation rate40, and this could result from photolytic decarbonylation of Ru(TMP)CO109. The isolable ethylene complex is stable to torr at 20°C, but the cyclohexene analogue readily dissociates the coordinated alkene77. The competition of alkene binding versus reactivity to give dioxo species is reminiscent of mechanisms in catalytic homogeneous hydrogenation, where hydride and unsaturated routes have been identified110; the former operates by olefin attack on a metal-hydride, and the latter by attack of on a metal-alkene species. Epoxides coordinate to Ru(II)-porphyrins, and Ru(TDCPP)(CO)(styrene oxide) has been characterized by X-ray analysis84. The bent geometry of the coordinated epoxide ring (Fig. 8) may be similar to the transition state geometry for olefin epoxidation, with a side-on approach of the alkene allowing for favourable interactions between its filled and the metal-oxygen orbitals; solution data show that the coordinated epoxide rotates about the Ru-O axis84,105. The corresponding thioepoxide and aziridine complexes also have bent geometries analogous to that of the epoxide40. Several other mechanistic possibilities, invoking a metallaoxetane, carbon radical, carbocation, ion-pair, or charge transfer species as an intermediate or transition state, have been proposed for epoxidations catalyzed by 1st-row metalloporphyrin monooxo species13,32,111.

A further complication in the catalyzed olefin epoxidation is a catalyzed cis-trans isomerization of the coordinated epoxide84,105. For example, in a reaction that is 1st-order in metal complex, catalyzes isomerization of cis- or oxide in benzene to give a 1:5 equilibrium mixture of the cis and trans forms. The non-hindered tetra-p-

1. Catalytic oxidations using ruthenium porphyrins

25

tolylporphyrin analogue exhibited similar reactivity, while Ru(TMP)CO was catalytically inactive. Coordinating olefins such as styrene, norbornene, cyclooctene and inhibited the isomerization of methylstyrene oxide, while the more weakly coordinating methylstyrene was a poorer inhibitor. The relative binding constants for olefin and epoxide are critical in evaluating any catalytic epoxidation processes. The binding constant of for the replacement of MeOH within Ru(TMP)(CO)(MeOH) is > at -50°C105, while that for styrene oxide coordination to Ru(TDCPP)CO is at -40°C84. Studies on chiral oxide suggested that the isomerization involved coordination of the epoxide, homolytic cleavage of the bond to give a benzylic radical, and then rotation about the single bond and subsequent reclosure to give coordinated epoxide105. The (air) catalyst has found an application at ambient conditions for epoxidation of steroids containing C=C bonds (Fig. 9)112-117. The catalyst (~5 mM), generated in situ in benzene from Ru(TMP)CO and

m-CPBA, was first applied to (at ~0.1 M) bearing a C5-C6 double bond, the reactivity depending very much on the nature of the C3 substituent 112-114 . Thus, there is no epoxidation if C3 has an OH group, perhaps because of reactivity of the OH with the Ru-dioxo species (Section 3.6). Protection of the OH as an ester or with a silyl group allows for up to 90% epoxidation with up to 99% initially surprising as the is more crowded and as epoxidation with peroxy acids takes place on the The best results were obtained with an acetate ester at C3; longer chain aliphatic esters and the benzoate ester required longer reaction

26

Maria B. Ezhova and Brian R. James

times of up to 3-6 days113,114. The cholesteryl acetate epoxidation revealed a 2 h induction period, with subsequent maximum rates roughly proportional to the initial [Ru]; complete epoxidations after 5 h were realized for a substrate : catalyst ratio of ~25. Reuse of the catalyst gave much lower rates but the product selectivities were retained; catalyst deactivation was attributed to the formation of hydroxylic species via protonation of the oxo ligands113. Although slow, the epoxidations are synthetically useful. With steroids having additional C=C bonds, either elsewhere in the nucleus or in the C17 side-chain, epoxidation still predominates at the 5,6position; for conjugated 5,7-dienes, the epoxidation is regioselective at C5C6 but with loss of stereoselectivity115. There is no effective epoxidation of cholest-4-ene-3-one, which has a carbonyl conjugated with the olefin bond, but ketalization of the conjugated carbonyl shifts the double bond to the 5,6-position and epoxidation occurs as described above114. The non-conjugated cholest-5-ene-3-one yields a mixture of epimeric 6-hydroxy-4-ene-3-ones, where the C=C bond has been shifted, and a 4-ene-3,5-dione116; this reaction was insensitive to the addition of a radical inhibitor, indicating a non-radical process. Ru(TMP)CO also catalyzes equally well this same reaction, but the true catalyst was again the trans-dioxo species formed from the carbonyl via reaction with a 6hydroperoxy-4-ene-3-one (cf. Fig. 5), formed by radical-initiated, incipient autoxidation of the cholest-5-ene-3-one. The reactivity and high of the epoxidations have been rationalized generally in terms of steric interactions between the catalyst and substrate117. Steroids containing a Me group on C6, or a double bond in ring C or D, are not epoxidized because of non-bonded interactions between the steroid and the porphyrin ring for a side-on approach of the alkene moiety, with the mean plane of the steroid orthogonal and not planar to the Ru=O bond (Fig. 9; cf. Fig. 8). A rationale for the has emerged from the structures of cholesteryl ethyl carbonate and its epoxide, and is based on conformational differences between the two structures along the C5-C10 bond117. Molecular modeling indicates that epoxidation on the ‘folds’ the A-B junction and allows for an easier approach of the substrate by releasing steric strain that results from interactions between the C3-ester and porphyrin mesityl; epoxidation on the has no effect on the A-B junction and is relatively disfavoured. NMR data confirm that only the interacts with the metal in 117 Ru(TMP)CO . The activity and stereoselectivity of other related metalloporphyrin systems depend on the metal and on the oxidant used; Fe(III)- or Mn(III)TMP complexes are weakly active or inactive for the aerobic oxidations9,14,40,118-120.

1. Catalytic oxidations using ruthenium porphyrins

27

in effects the stoichiometric epoxidation of olefins such as norbornene, styrene and cis- and trans-stilbene according to eq. 29; cis-stilbene yields cis- and trans-epoxides in a 1:2.7 ratio, and transstilbene gives mainly trans-epoxide67. The supposed Ru(IV)=O co-product reacts with present to form inactive and no catalysis occurs even up to 10 atm The 2nd-order rate constant for reaction 29 with norbornene was at 25°C67. In pyridine solution, a reaction similar to 29 generates and two equivalents of epoxide.

In EtOH, the monooxo species is thought to be stabilized as Ru(OEP)O(EtOH) and a very slow catalytic epoxidation, observed at 1 atm was attributed to slow of Ru(IV)=O to Ru(VI)-dioxo species67; whether disproportionation of the Ru(IV)=O occurs under these conditions (eq. 16) is unclear. As noted for the thioether oxidations, introduction of halogen substituents into the porphyrin, especially in positions to prevent sterically formation of Ru species, can increase and/or prolong catalytic activity. More generally, extensive halogenation of TPP-type porphyrins at the mesophenyl rings and positions gives rise to the so-called third generation metalloporphyrins9,92,94, and excitement became intense when such Fe-porphyrins were found to effect catalytic of light alkanes with high activity, and a mechanism involving high-valent Fe=O intermediates was proposed92; the mechanism was shown subsequently to be free-radical in nature, involving hydroperoxide decomposition pathways (eqs. 6, 7)121. Such substituents on the porphyrin ring can also modify the conformation of the porphyrin, and out-of-plane distortion can enhance catalytic activity, as first suggested for some Fe(III)-porphyrin systems.121 One rationale for increased activity via out-of-plane distortion is an imposed, so-called “unidirectionality of electron transfer”122. The highly distorted (X-Ray data revealed that its precursor, Ru(DPP)(CO)(py), also exhibits both saddle and ruffle distortion)123 catalyzes in benzene-MeCN (9:1) some epoxidation of norbornene, styrene, cyclohexene, cyclooctene and cis-stilbene under 1 atm with turnovers of 8 - 40 in 4 h before deactivation of catalyst123; the numbers are comparable with those for and greater than for and There were also co-oxidation products: benzaldehyde (7 turnovers) and trace phenylacetylaldehyde were formed from styrene, cyclohex-2-en-1-ol (20 turnovers) and cyclohex-2-en-1-one (11 turnovers) from cyclohexene, and trace benzaldehyde and trans-stilbene

28

Maria B. Ezhova and Brian R. James

oxide (3 turnovers) from cis-stilbene. did not oxidize trans-stilbene under these conditions. Stoichiometric alkene oxidation by in in the presence of pyrazole was also studied; allylic C-H oxidation was seen for cyclohexene, while styrene and cis-stilbene gave solely styrene oxide and cis-stilbene oxide, respectively. the final Ru product, was characterized by X-Ray123. The Stoichiometric alkene oxidations obeyed a rate law of the form = k[Ru(porp)][alkene], with k values for norbornene and styrene (in at 25.9°C) of 3.8 and respectively, but whether this referred to a radical process or an Oatom transfer process was uncertain. Use of perhalogenated Ru porphyrins certainly leads to radical autoxidation processes65,124. This is reflected in the fact that is significantly less active catalytically than the carbonyl for of cyclohexene124; a total turnover of 20 to radical products (epoxide, 2-cyclohexen-1-ol, 2cyclohexen-1-one) was achieved using M in at ambient conditions. In contrast, can catalyze similar radical of cyclohexene (and styrene) with turnovers up to 300; cyclooctene gave only epoxide, but such high product selectivity is not unusual for this substrate even in radical, autoxidation processes125. Of note, these systems were photo-initiated with visible light. Closely related are the autoxidations catalyzed by or the 65 carbonyl precursor . Use of M of these complexes in neat hydrocarbon substrate at ~90°C gives extremely efficient catalyzed autoxidations; turnovers of up to are found for non-selective cyclohexene oxidation (trace cyclohexene hydroperoxide was also detected), while cyclooctene gives >80% selectivity to epoxide65. Corresponding autoxidations using TDCPP systems were completely inhibited by addition of a radical inhibitor such as BHT65. The TDCPP-Cl8 systems similarly catalyze autoxidations of saturated hydrocarbons; see Section 3.4 for further details on radical processes, where the presence of the porphyrin ligand is not always essential. Reports have appeared on the rates of decomposition of cyclohexyl hydroperoxide (an intermediate in the industrial oxidation of cyclohexane126,127) to cyclohexanol and cyclohexanone catalyzed by Ru(porp)CO and systems (porp = tCPP , mCtPP , TDCPP, TMCPP, TMP, TPP) either in solution or anchored to polystyrene or silica128-130. The systems were studied in 20 : 1 at 25°C, when decompositions in the 28-66% range were observed after 2 h, and close to 100% after 48 h129,130. Several, plausible reaction pathways were presented for decomposition of the alkyl hydroperoxides130. Introduction of bulky and chiral substituents at the 5,10,15,20 (meso)positions of the porphyrin ring allows for aerobic, enantioselective epoxidation of olefins 131 . Use of chiral Fe(III)- and especially Mn(III)-

1. Catalytic oxidations using ruthenium porphyrins

29

porphyrin catalysts (using PhIO and LiOCl as oxidants), respectively, for regio- and enantioselective olefin epoxidation had been demonstrated earlier132, but use of chiral Ru-porphyrins for enantioselective epoxidation has been explored only recently131,133-139. Thus, (see Fig. 10) can catalyze aerobic oxidation of several olefins. Under the conditions given in eq. 30, only small turnovers were achieved, e.g. 10 for styrene oxide with 70% ee; in toluene, 21 turnovers were seen for conversion of

methylstyrene to a 7:1 cis/trans ratio of oxide (73% ee), 11 turnovers for the epoxidation of p-chlorostyrene (52% ee), and 14 turnovers for oxidation of 2-vinylnaphthalene The low turnovers were attributed to deactivation of the catalyst after ~24 h through formation of diamagnetic Ru(II) complexes containing epoxide (cf. Fig. 8); in the case of oxidation, the epoxide complex was isolated and characterized by FAB MS131. Application of the chiral dioxoruthenium(VI) porphyrin (Fig. 10) for aerobic oxidation of in benzene (with 9 atm 40 h) gave trans-epoxide 139 with 59% ee after 7 turnovers . The same types of chiral porphyrins have been used in conjunction with O-atom donors for asymmetric epoxidation (see below)133-139. The systems were developed largely from a non-chiral one that utilized in conjunction with pyridine N-oxides and other heteroatomic N-oxides that had been used for selective epoxidation of olefins96,140,141; these had proceeded efficiently under mild conditions under Ar and, in some cases (e.g. using 2,3,5,6-tetramethylpyrazine oxide), styrene was epoxidized over one day in 100% yield based on both styrene and N-oxide. Amine oxides with large substituents, such as phenyls, ortho to the nitrogen-oxo group were inert, presumably because of no interaction between the amine O-atom and the Ru. Strong coordination of the reduced amine oxide at the Ru (e.g. 4,6-dimethyltriazine formed from the N-oxide) also destroys the catalysis96. High selectivities, comparable to those obtained utilizing were observed using 2,6-lutidine N-oxide with cis- and methylstyrene, and with cis- and trans-stilbene; the cis moiety of trans,cis,trans-1,5,9-cyclododecatriene was the most reactive, while the 6,7-

30

Maria B. Ezhova and Brian R. James

double bond in some terpene acetates was also selectively epoxidized141. The catalyzed amine-oxide oxidation of 2-vinylnaphthalene was faster than the stoichiometric oxidation using and thus Ru(TMP)(O)(amine oxide) was considered a viable active intermediate; some data using labeled water were interpreted in terms of a reaction pathway other than via Some representative data showing the use of chiral Ru-porphyrins for asymmetric epoxidation using amine oxides or PhIO as O-donors are given in Table 1. The system134 effects epoxidation with 5-88% yields and 28-77% ee values. Either or

1. Catalytic oxidations using ruthenium porphyrins

31

32

Maria B. Ezhova and Brian R. James

in combination with PhIO are also active; however, yields do not exceed 71% with ee values from 7 to 63%136. Comparison of data for stoichiometric and catalytic epoxidations using and in the presence and absence of PhIO led to the conclusion that the dioxo-species was the key oxidizing intermediate using this oxidant136. Formation of was suggested as an intermediate, based on stoichiometric styrene oxidation seen in the presence of pyrazole (Hpz), eq. 31, where was detected by UV-vis spectroscopy (by comparison with data for an authentic sample). The pyrazolate complex

was similarly isolated from the reaction of with an alkene in the presence of pyrazole123. Kinetic studies on the stoichiometric oxidation of alkenes by in the presence of pyrazole revealed the rate law: with k values of and at 25°C for styrene, cisstilbene and trans-stilbene, respectively136. The rate-determining step for the O-atom transfer for epoxidation of was given as formation of the radical intermediate that explained generation of some oxide (Fig. 11). Such long-lived radical intermediates have been invoked in earlier catalytic epoxidation work using homochiral Fe- and Mn-porphyrins135,142-144. The system (see Fig. 10) was also thought to involve the Ru(VI)-dioxo species as the most active catalyst133, but with an N-oxide as O-donor an alternative mechanism has been suggested (see Fig. 12) 135 . With such oxidants, there is a strong dependence of the reactivity and selectivity of the epoxidations on the solvent and nature of the oxidant135,137, 145-147 . A general increase in enantioselectivity for epoxidations in benzene vs. those in (e.g. with for epoxidation of styrene 42 vs. 5% 135, and with 68 vs. 58%137 (see Table 1)) has been considered to result from a strong interaction of Ru with the aromatic solvent137, and there is, for example, NMR evidence for formation of species59. Association of benzene with species (or perhaps some other active Ru(IV) or Ru(V)

1. Catalytic oxidations using ruthenium porphyrins

33

intermediates) has been suggested to affect the solution structure of the chiral cavity associated with chiral induction135. Further, the dependence of enantioselectivity on the nature of the oxidant has been considered to indicate that the structure of the active oxidant includes an axially coordinated molecule of the O-atom donor; for example, a Ru(porp)(O)(pyNO) species has been proposed as an active intermediate in 2,3-dimethyl2-butene epoxidation catalyzed by (Fig. 12)137. Such possible intermediates have been discussed prior to reports on the asymmetric epoxidation (see Section 3.1.). The initial cycle starts with slow oxidation of the olefin by to give and epoxide (step a), the monooxo complex then reacting rapidly with N-oxide to form complex (15) (step b). (15) is then considered to effect very rapid epoxidation with liberation of the pyridine (step c). Regeneration of from (15) (step d) was ruled out, as this step was slower than epoxidation via (15). The role of coordinated pyNO was rationalized in terms of its effect on the chiral environment on the trans oxo-coordination site137. Much improved chiral induction resulted from the introduction of Cl-atoms into the meta-positions of the bridged phenyl rings in (Fig. 10): thus, under the catalysis conditions listed in Table 1, styrene was epoxidised with 79% ee at 551 turnover numbers, while cis- and yielded epoxides with 57 and 69% ee at 244 and 487 turnovers, respectively. Trans-stilbene was also oxidized to a mixture of epoxides with 38% ee at 242 turnovers. The authors concluded that overall, the observed chiral induction was higher for terminal and trans-olefins, versus cis-olefins138.

34

Maria B. Ezhova and Brian R. James

Attractive from an environmental point of view, nitrous oxide has been employed as an oxidising agent. acyclic and steroidal substrates have been epoxidized using Ru(TMP)-based systems148,149. Stoichiometric epoxidation of is effected by generated via reaction of with but catalytic oxidation of cholesteryl esters has been reported149. Although the conditions are relatively severe: 0.2 mmol of olefins, 5.0 mol % at 10 atm in chlorobenzene at 140°C, the yields and selectivities to the steroidal epoxidation are good; only were formed with yields from 48 to 99%. Use of polar solvents (THF, EtOH, or use of other complexes (porp = TPP, OEP, 4MeO-TPP) resulted in no epoxidation149. Application of Ru-porphyrin complexes for catalytic, homogeneous alkene epoxidation is limited because of small turnover numbers, only moderate enantioselectivities in the case of chiral systems, the cost of porphyrin and the O-atom donor and often limited stability of the porphyrin ring under the oxidizing conditions. In attempts to improve possible applicability, Ru-porphyrins have been “heterogenized” on various supports150-152, following methodology developed previously for Fe- and Mn-

1. Catalytic oxidations using ruthenium porphyrins

35

porphyrin systems14,153. The Ru(porp)(CO)(EtOH) complexes (porp = TCPP150 and TDCPP151) were successfully encapsulated into the channels of mesoporous material MCM-41154, modified by treatment with 3(aminopropyl)triethoxysilane M-41(m); the ligand exchange reaction utilized is shown in Figure 13 150,151 . Ru(TCPP)(CO)/M-41(m) containing 8.3 wt.% Ru showed good catalytic activity for alkene epoxidation using as

oxidant in under at r.t.150. A higher catalytic activity at lower Ru content was attributed to efficient site isolation, coupled with diffusion reaction pathways, conditions that precluded formation of stable, catalytically inactive species. Total turnovers of 9000 were achieved with 0.1 wt.% of Ru for oxidation of norbornene to give 53% exo-epoxide (yield based on consumption of TBHP), in comparison with only 216 turnovers using the 8.3 wt.% Ru catalyst; 230 turnovers were achieved with Ru(TCPP)(CO)(EtOH) as a homogeneous catalyst in In general, the turnovers were 20-40 times higher with the supported catalyst, which was stable under the oxidation conditions, with activities not changing after 2 days of reaction. Although the turnover numbers are reasonable, the chemoselectivity for the supported catalyst is poor. Thus, styrene gave only 11.0% of styrene oxide with benzaldehyde being the co-product (26.5%); cyclohexene gave 2.2% epoxide, cyclohex-2-en-l-ol (7%) and cyclohex-2en-l-one (12.3%); cis- and trans-stilbene gave trans-stilbene oxide as the major product for both isomers (8.6 and 9.35%, respectively), with cis-oxide (1.4 and 0.18%) and benzaldehyde (0.94 and 0.27%) being the other coproducts150. Ru(TDCPP)(CO)/M-41(m) also catalyzed selective alkene epoxidation using as an O-donor in the presence of HC1151. Aromatic and aliphatic alkenes gave epoxides in good yields (up to 98%, based on the amount of substrate consumed) with complete selectivities, styrene, cisstilbene, norbornene and octene1 giving only cis-products; trans-stilbene was not oxidized. Of note, the formation of cis-oxides from cis-alkenes here is quite different from the preferred formation of trans-oxides from cis- or trans-stilbene catalyzed by

36

Maria B. Ezhova and Brian R. James

the Ru(TDCPP)CO/TBHP system150. The catalyst retains 67% activity after 11,700 turnovers under the selected reaction conditions: 1 mmol of alkene, 1.1 mmol of ~ 0.3 mmol of HC1, 50 mg of 0.4 wt.% Ru/M-41(m) in 5 ml of The loss of activity was attributed to decomposition of the catalyst. The possibility of being an active intermediate was ruled out based on product distribution studies for and cyclohexene oxidation, and on Hammett correlation studies for the stoichiometric, epoxidation of p- and msubstituted styrenes; a species was proposed as a key intermediate, particularly as a paramagnetic species, assumed to be Ru(III) (presumably formed after O-atom transfer from a moiety) was detected in the reaction mixture after the catalysis (see also Fig. 20, Section 3.4)151. The longevity of the catalyst 150,151 is not the only advantage of using Ru-porphyrins supported on mesoporous materials. Differences in selectivity for the heterogeneous and homogeneous catalysis may also be an asset. Thus oxidation of (+)-limonene catalyzed by gives the 8,9epoxide as major product (61%) and 27 and 10% of the cis- and trans-1,2epoxides, respectively (Fig. 14). Under homogeneous conditions, in gives preferential formation of

1,2-epoxides (55 and 19% of cis- and trans-l,2-epoxides) and 22% 8,9epoxide151. Restricted space in the M-41(m) channels likely makes the more sterically hindered, trisubstituted C=C bond less accessible to the active metal center, compared with the exocyclic methylene. Another example of the influence of steric constraint of M-41(m) on selectivity is oxidation of the bulky 3,4,6-tri-O-acetyl-D-glucal. Under homogeneous conditions, the catalyst system affords stereoselectively the correspondent glycal epoxide (this undergoes methanolysis with an inversion of configuration, leading to methyl (77%))151. In contrast, under heterogeneous conditions, a 3 : 1 mixture of was formed

1. Catalytic oxidations using ruthenium porphyrins

37

(Figure 15) 151 . The more bulky benzyl-D-glycal was not oxidized under the heterogeneous conditions. Further developments in supported Ru-porphyrin catalysts include attachment to a polymer through a covalent linkage152. Ruthenium complexes with the unsymmetrical 5,10,15-tris(4-R-phenyl)-20-(4-hydroxy-

phenyl)porphyrins (R = Cl, Me) (abbreviated, for example, have been attached through the hydroxy group to Merrifield’s peptide resin (MPR) bearing benzyl chloride groups according to Figure 16. The resulting catalysts and

(4-MPR)TPP)CO epoxidized a variety of alkenes with as oxidant, the chloro system giving higher yields (56-98%) with Ru (8.6 wt.%.) : oxidant: substrate = 1 : 1400 : 1000, in benzene, at r. t. for 24 h. Cis-stilbene and norbornene yield exclusively cis-epoxides. In contrast, with the M-41(m) system, the polymer oxidized trans-stilbene and to the corresponding trans epoxides in 90 and 86 % yields, respectively. Several alkenes were oxidized catalytically for the first time with high selectivity: 3,4-dihydronaphthalene yielded 62% epoxide,

38

Maria B. Ezhova and Brian R. James

1,5-cyclooctadiene gave only bis-epoxide (76%), and cis-1-phenyl-3-penten1-yne only gave cis-epoxide (91%). The 3,4,6-tri-O-acetyl-D-glucal was epoxidized to only (56%), while the system gave a 3 : 1 mixture of and (see Fig. 15). Of note, were selectively oxidized to threo-(amino)epoxides in 87-89% yields. The system was consecutively reused four times without detectable leaching of activity, giving for example epoxide yields of 96, 93, 90, 92 and 91% from styrene. Lowering the Ru concentration to 0.17 wt.% decreased the yields (e.g. 68% for styrene oxidation), but the turnover numbers remained high ( for styrene). The remarkable activity of was attributed to the ability of MPR to solvate and swell in some organic solvents152. That Ru(II)-porphyrins do not effect deoxygenation of epoxides has been ascribed to the relatively low oxophilicity of the incipiently formed Ru(IV)-monooxo species105; thus such species are able to transfer the oxo ligand to olefins, and data on the amine-oxides donor systems described above and for related Ru non-porphyrin systems41,42,155 support such an inference. Ruthenium porphyrins in the presence of O-donors complexes are also capable of cleaving C=C double bonds. Thus a Ru(TDFPP)/TBHP system can cleave the C=C bond in to give acetophenone156, while cleavage of trans,trans-1,4-diphenyl-1,3-butadiene by Ru(TDCPP)/TBHP yields benzaldehyde and cinnamaldehyde157. Epoxidation is proposed as the first step, followed by a TBHP/Ru-porphyrin-mediated fragmentation of the epoxide. Another example is reaction of a Ru supramolecular system ((16), essentially Ru(TPP) with two units linked to two trans-phenyl rings through metheneoxo linkages, Fig. 17) with In a

1. Catalytic oxidations using ruthenium porphyrins

39

40

Maria B. Ezhova and Brian R. James

biphasic system ((hexane and (9:1)) : water = 10 : 1), preference for the cleavage of E-double bonds yields a mixture of aldehydes (15’- , 12’and l0’-apocarotenals (1:0.95:0.5)) in 30% yield. The “simpler” system also can mediate oxidative cleavage of to give a mixture of and Isomerization of trans- to cis-isomers of first takes place, followed by fast epoxidation of the trans-isomers. Simultaneously oxidation of the initially present trans-C=C double bonds yields the corresponding 5,6and 5,8-epoxides, with subsequent cleavage leading to the mixture of products (Fig. 18)158.

3.4

Oxidation of saturated hydrocarbons A

marginally

catalytic oxidation of cyclohexane using a system was mentioned above (Section 3.3.). The species is thermally unreactive toward cyclooctane at 1 atm at 70°C, but some O-atom insertion into saturated C-H bonds has been achieved under photolytic conditions and with the electrochemically generated species40; few details are available, but the latter system showed for adamantane oxidation the usual radical selectivity (tertiary > secondary carbon). Work in this laboratory has shown also that the complexes (porp = TMP, TDCPP, and are “practically” inactive for thermal of saturated hydrocarbons65. Some activity data for 0.2 mM Ru solutions in benzene under air at ~25°C for “optimum” substrates such as adamantane and triphenylmethane at 6 mM did show selective formation of 1-adamantol and trityl alcohol, respectively, but with turnover numbers of only ~0.2 per day; the maximum turnover realized was ~15 after 40 days for the TDCPP system! Nevertheless, this was a non-radical catalytic processes; there was < 10% decomposition of the and a genuine O-atom transfer process was envisaged65. Quite remarkably (and as mentioned briefly in Section 3.3), at the much lower concentration of 0.05 mM, in neat cyclooctene gave effective oxidation. For example, at 90°C under 1 atm an essentially linear oxidation rate over 55 h gave about ~70% conversion of the olefin with ~ 80% selectivity to the epoxide; however, the system was completely bleached after ~ 20 h and, as the activity was completely inhibited by addition of the radical inhibitor BHT, the catalysis is operating by a radical process, but in any case the conversion corresponds to a turnover of 110,00065! As in related Fe(porp) systems (Section 3.3, ref. 121), the Ru(porp) species are considered to be very effective catalysts for the decomposition of hydroperoxides (eqs. 6, 7) 65,124,128-130 . The radical nature was more obvious in a corresponding

1. Catalytic oxidations using ruthenium porphyrins

41

oxidation of cyclohexene using 0.03 mM Ru, when the usual mixture of enol, enone, epoxide, and other products were observed, with again a turnover in the region of ~105 being realized65. At much higher 1.0 mM Ru concentration with cyclohexene in benzene at 50°C and 1 atm a selective O-atom transfer oxygenation to epoxide is evident, but at only ~ 1 turnover per day! Similarly, at 0.01 - 0.03 mM Ru in neat methylcyclohexane, using either or Ru(PCP)CO, at 90°C and 1 atm turnovers up to 20,000 were attained in non-selective, radical oxidations giving mainly tertiary alcohols, secondary alcohols, and several ketones65. Of note, several other Ru non-porphyrin complexes initiate (as effectively as the “exotic” porphyrin complexes) the reasonably selective but radical oxidation of neat cyclooctene at 90°C and 1 atm even the essentially insoluble hydrate, and show moderate activity (~16% conversion in 16 h, with ~85% selectivity to epoxide), while the soluble at (!) gives 90% conversion with 69% selectivity (a turnover of at 1 mM, the phosphine complex gives only 50% conversion (at 79% selectivity) implying a contribution from a less effective pathway, possibly O-atom transfer, which has been invoked for Ruphosphine/O-atom donor systems159. Three important points are evident: (i) The introduction of halogen substituents into the porphyrin ring does stabilize Ru(VI)-dioxo species for executing O-atom transfer catalysis, but the halogenated species are more effective than the non-halogenated species for catalyzing decomposition of hydroperoxides; and these, if present in trace amounts, can lead to efficient radical oxidation pathways, (ii) At lower concentrations of a Ru(porp) complex (or apparently many other Ru non-porphyrin complexes) in neat hydrocarbon (typically ~10 M) under where there will be trace amounts of hydroperoxides, the favored reaction appears to catalyzed decomposition of the ROOH to generate radicals with accompanying bleaching (i.e. destruction) of the porphyrin. (iii) At higher Ru concentrations (~1 mM) in, for example, benzene solutions of the substrate (typically ~ 0.1 M) under where there will be less hydroperoxide, the radical chemistry is less evident and conditions for O-atom transfer from a Ru(VI)-dioxo species are improved. The dioxo species can, of course, be generated by reaction of a Ru-precursor and a hydroperoxide, the latter acting an O-atom donor (Section 2, Fig. 5), and so which pathway dominates will depend critically on concentrations of Ru and hydroperoxide (particularly within a second-order process) and the relative activation energies of the reaction pathways. The Ru(porp)-based oxidizing systems using as oxidant (see Section 3.3 for use in olefin epoxidation) have been reported to be highly efficient for oxidation of alkanes138,160-163. or

42

Maria B. Ezhova and Brian R. James

was initially used as a catalyst precursor in benzene containing “2-3 drops of concentrated aqueous HCl or HBr” under Ar at 20-40°C, with molecular sieves added to maintain anhydrous conditions160. Essentially complete conversion of adamantane yielded adamantan-1-ol (up to 68%), adamantan1,3-diol (25%) and adamantan-2-one (1%), while methylcyclohexane was oxidized to 1-methylcyclohexanol (77%) and methylcyclohexanones (6%); ethylbenzene and cyclohexanol were converted in 88% yield to acetophenone and cyclohexanone, respectively. The acids were essential, and NMR data suggest that these convert the dioxo species into the known, paramagnetic species (X = halide)49,71. The dihalo complexes, as well as Ru(porp)CO complexes (porp = TMP, TDCPP, TPFPP, and even the non-sterically hindered TPP) were also effective catalyst precursors161,162. For example, the (or HBr) system effected hydroxylation of adamantane to 1-adamantanol (76.2%) and adamantan-1,3diol (13%) with turnovers of depending on the while cis-decalin was converted to (Z)-9-decalol (79.6%) and (Z)-decal-9,10diol (4.2%) at up to 64 turnovers . The TPP, TMP, and TDCPP systems also effected oxidation of steroids, with retention of configuration for those containing chiral centers; for example, in the presence of Ru(TPP)CO gave the 25-hydroxy product (11%) (Fig. 19), while Ru(TMP)CO and Ru(TDCPP)CO generated other alcohols as well161.

Analysis of product evolution during adamantane and cis-decalin hydroxylation by the Ru(TPFPP)CO system, coupled with observation of intermediates (18 – 20) by UV-vis, IR and ESR spectroscopies, led to the mechanism outlined in Figure 20162. An observed induction time was attributed to formation of (20), and this induction period could be reduced by initial treatment of the precursor (17) with O-atom transfer occurs from (21), a Ru(V)-oxo (or a Ru(IV)-oxo-porphyrin cation radical160, which had been suggested much earlier8,69 - see above), and its formation was

1. Catalytic oxidations using ruthenium porphyrins

43

considered rate-determining under conditions when there was a zero-order dependence on substrate. At lower [substrate], the reaction of (21) with the hydrocarbon became rate-determining, when kinetic deuterium isotopic effects for hydroxylation of adamantane and cis162 decalin were measured . It is clear that the dioxo species, does not participate in these systems using pyridine oxides as an O-atom donor. Some rate data had suggested that the HCl/HBr acids may also accelerate deoxygenation of the N-oxide by the Ruporphyrin Total turnovers up to 18,800 were reached for adamantane oxidation and, for the TDCPP system, a high turnover of was attained160. An example of enantioselective hydroxylation of a benzylic C-H bond

using or complexes (see Fig. 10) has been reported by Che’s group163. The stoichiometric oxidation of substituted ethylbenzenes, 2-ethylnaphthalene, indane and dehydronaphthalene by in containing pyrazole at r.t., led to the corresponding alcohols (27-48% yield) and ketones (24-34% yield), with ee values of 9-58% (S) for the alcohols. The second-order rate constant determined for oxidation, for example, of ethylbenzene was at 25°C with a kinetic isotope effect for of 8.9 at 40°C and 11.2 at 25°C. A dual-parameter Hammett correlation with data for 4-substituted ethylbenzenes was considered consistent with a ratelimiting step involving C-H bond cleavage (Fig. 21). Preferential formation of the S-isomer was explained by preferential collapse of the benzylic radical

44

Maria B. Ezhova and Brian R. James

on the pro-S face versus pro-R face during an O-atom rebound step, resulting from the fit of the substrate in the chiral cavity (see Fig. 10). Chiral induction via radical intermediates is a remarkable result. Under catalytic conditions (5 ml benzene, 25°C, alkane : : catalyst = 0.5 mmol : 0.55 mmol : ), effectively oxidizes the substrates to alcohols in 28 to 72% yields, with ee values of 12-76% (S). Use of the system138 (see Fig. 10) for hydroxylation of the tertiary alkanes (rac-2-phenylbutane and -2-phenylhexane) under catalytic conditions ( of Ru, oxidant and 1 mmol alkane, 48% aq. HBr, 50 mg of 4A molecular sieves in 1 ml benzene under Ar at 25°C) results in tertiary alcohols with modest enantioselectivity and yields, again remarkable considering the possible radical nature of the process (see above). Hydroxylation of the alkanes gave

PhC(Me)(R)OH (16% ee, 41% yield, 103 TON for R = Bu; 27% ee, 54% yield, 135 TON for R = Et). The highest ee (38%) was obtained for hydroxylation of the 2-phenylbutane at 10°C; some kinetic resolution of the unreacted alkane (up to 8%) was evident138. This is the first report of catalytic enantioselective hydroxylation of tertiary alkanes.

3.5

Oxidative-dehydrogenation of phenols and other arenes

Preliminary studies had suggested that benzene solutions of react stoichiometrically with phenol under 1 atm to give the paramagnetic species (23), via a detected species, according to eq. 32, and data were given for the second-order rate constant Evidence for (23) included data,

1. Catalytic oxidations using ruthenium porphyrins

45

elemental analysis and a magnetic moment showing 2 unpaired electrons.

More detailed studies have shown subsequently that the product is, in fact, (24 in Fig. 22 below)65. (24) was synthesized independently from the metathesis reaction between with excess phenol under Ar, and mass spectral data distinguished the product from the supposed (23), which was probably the dihydrate of (24). A revised mechanism for the stoichiometric phenol oxidation is presented in Figure 22. The reaction initially produces p-hydroquinone, which is further oxidized to p-benzoquinone at a rate faster than the rate of phenol oxidation, and this was confirmed by direct oxidation of p-hydroquinone with The generated species is then thought to react reversibly with water to give the dihydroxy species (the X-ray structure for 164 has been reported ) that then undergoes metathesis with

46

Maria B. Ezhova and Brian R. James

phenol to give (24). The disproportionation of to and (see Sections 2 and 3.3) is thought to be a competing pathway for the reaction with water, and this would give rise to the observed which was also formed during the anaerobic reaction of with excess phenol. Under anaerobic conditions, the kinetic dependence on phenol for loss of the reactant becomes > 1 because of the faster reaction of the p-hydroquinone than phenol. The selective attack of at the para-position of the phenol is presumably imposed by the steric restraints of the TMP ligand. The mechanism of the net O-atom insertion into C-H bond could involve either sequential electron and H-atom transfer processes as suggested for the reaction between phenol and transor something akin to the oxygen-rebound mechanism76,101 with initial H-atom abstraction from the phenol at a Ru=O site. Effective catalytic hydroxylation/oxidation of phenol to give phydroquinone and or p-quinone has yet to realized using these Ru porphyrin systems. N,N-Dimethylaniline is marginally catalytically oxidized by 3.3 x 10-3 mM benzene solutions of under aerobic conditions at 50°C perhaps to the p-hydroxy-derivative; the turnover of 2.1 after 30 h, was limited by decomposition of the catalyst to Ru(TMP)CO via a decarbonylation process, presumably of the phenolic product65. The preliminary findings could be consistent with initial attack of an electrophilic oxo moiety at the para-position of the aniline, as suggested for the phenol oxidation (see above), or attack at the moiety (see Section 3.7). Of note, methoxybenzene, toluene, chlorobenzene, bromobenzene and nitrobenzene were not oxidized by The Ru(porp)CO/HX/2,6-Cl2pyNO systems (porp = TMP, TPP; X = Cl, Br; see Sections 3.3 and 3.4) can, however, convert alkoxyarenes eventually to the p-benzoquinone derivatives selectively, as exemplified in Fig. 23 l66 . The yields of 11-97% varied with the structure of the substrates, but the most electron-rich C-atom was preferentially attacked by an

1. Catalytic oxidations using ruthenium porphyrins

47

electrophilic oxygen; thus 1,3,5-trimethoxybenzene was oxidized to 2,6dimethoxy-p-quinone with 97% yield, the total turnover reaching 33,000. Oxidation of m-dimethoxybenzene gave 2-methoxy-p-quinone (74%), while o-dimethoxybenzene yielded the corresponding p-quinone in only 11% yield, and p-dimethoxybenzene was not oxidized at all; the chemoselectivity of these reactions is clear. Similarly, oxidation of 2,3’,4,4’- or 2,2’,4,5’tetramethoxy-1,1’-biphenyl resulted in 2-methoxy-5-(3,4-dimethoxyphenyl)and 2-methoxy-5-(2,5-dimethoxyphenyl)-p-quinones in 77 and 46% yields, respectively (Fig. 24). Oxidation of naphthalene and phenanthrene yielded 1,4-naphthoquinone (29%) and 9,10-phenanthroquinone (40%), respectively; benzene was not oxidized. The nature of the catalytic Ru-oxo species involved was considered uncertain (see Section 3.3.). The intermediate phenol formed in the first step was isolated in the m-dimethoxy system, and the second stage involved loss of MeOH. An experiment showed that at least one ketonic O-atom of the p-quinone originated from the pyridine oxide.

3.6

Oxidative-dehydrogenation of alcohols

at 1 mM concentrations in benzene catalyzes aerobic oxidative-dehydrogenation of 2-propanol to acetone65,74,167,168, benzylalcohols to the corresponding benzaldehydes65,167,168, and various benzhydrols (solid substrates, where X = F, H, MeO) to the corresponding benzophenones169, and in each system water is the coproduct, but the turnovers are small. After 24 h under 1 atm air at 50°C, for substrates typically at ~0.2 M, 6 turnovers were observed for 2-propanol, and 50 for a maximum 23 was seen (after 45 h) for (p-

48

Maria B. Ezhova and Brian R. James

Addition of increases the turnover numbers by factors of 2-6, depending on the substrate, while addition of aq. KOH or NaOH increases turnovers by a further factor of up to 6. Thus addition of KOH increases turnovers for isopropanol and benzylalcohol by factors of 56. Maximum turnovers over 24 h of ~200 (100% yield) were achieved for 2-, 3-, or 4-MeO substituted benzyl alcohols, and the catalyst still remains active. Oxidation of primary alcohols (such as 1-butanol) to the aldehyde was less efficient, with turnovers of up to 40 being noted over 24 h, with no formation of carboxylic acids. The use of rather than air did not increase the oxidation rates 65,167,168 . In general, use of instead of for the aerobic oxidations made little difference, i.e. the activities were similar over 24 h, but the TDCPP analogue was more susceptible to deactivation via a decarbonylation process that generated the Ru(CO) derivative168. For the oxidation of benzhydrols, the perchloro derivative exhibited activity similar to that of The stoichiometric reaction between and the alcohols (2propanol, and under Ar conforms to that shown in eq. 33 or Ph(H)), as evidenced by NMR. The

paramagnetic bis(alkoxy)ruthenium(IV) complexes were isolated and fully characterized by NMR, UV-vis and IR spectroscopies, magnetic moment data (2 unpaired electrons), and X-ray structures for and The NMR spectra show upfield, paramagnetic shifts for the hydrogens to the to -34 region (vs. 8 - 9 for diamagnetic or complexes), consistent with Ru(IV) species possessing symmetry49,170. Detailed kinetic studies on the stoichiometric oxidation of 2-PrOH, benzyl alcohol and several benzhydrols by have led to the mechanism outlined in Figure 25; the rapid pre-equilibrium (K) to give a Ru-dioxo-alcohol intermediate (25), followed by a rate-determining (k) hydride transfer from the conforms to the measured saturation kinetic behaviour and the rate-law given in eq. 34169. The reactions were

monitored by NMR and UV-Vis spectroscopies but even at higher [ROH], when the rate is zero-order in ROH and (25) should be fully formed,

1. Catalytic oxidations using ruthenium porphyrins

49

there was no spectroscopic evidence for this intermediate. Such species are commonly proposed for alcohol oxidation using other Ru complexes171-175. For the porphyrin dioxo system, the “more sensitive” v(Ru=O) IR region was investigated, but the data were inconclusive169. A primary kinetic isotope effect at 20°C of ~15 (e.g., for and for with no isotope effect found for oxidation of is consistent with the rate-determining hydride transfer. Large kinetic isotope effects of between 8176 and 50177 have been reported for cleavage of bonds in alcohol oxidation by non-porphyrin transand complexes, respectively. A Hammett plot of log against for oxidation the p-substituted MeO-, H- and F-benzhydrols gave a linear relationship with a slope (the F-system being least active), implying that transfer of electron density from the to the Ru oxidant occurs in the formation of the transition state; the Ru=O...H bonding within (25) is akin to H-bonding but involving a hydride H-atom and an electrophilic O-atom169. The presence of a small and reasonably constant quantity of (see Section 3.3.) was detected by NMR during the stoichiometric oxidation of and its formation was suggested to take place concomitantly with that of the ketone via a net loss of from the intermediate and protonated ketone (Fig. 25). Production of the bis(alkoxide) was written as involving a ligand exchange reaction

50

Maria B. Ezhova and Brian R. James

between the alcohol and (formed reversibly from the oxo species). There is no direct evidence for the existence of the isolation of this compound has been claimed (from crystal structure data on an alcohol solvated species)66, but other work has suggested that the structure reported is that of a species167. Nevertheless, crystals of have been isolated during a synthesis of and analyzed by X-ray crystallography (see Section 3.5) 164 . Isomerization of to species via proton migration has been invoked within some non-porphyrin systems176. Studies on the catalytic alcohol oxidation process involved investigation of the above systems under an atmosphere of air or The initial rate of production of the ketone product is governed by the rate of formation of the bis(alkoxy) species, while subsequently the catalysis rate is likely determined by regeneration of the dioxo species via disproportionation of the species (see Sections 2 and 3.3.); however, this will depend on the concentration of which increases as the catalysis proceeds, and indeed water has been considered to accelerate the disproportionation reaction by increasing the rate of dioxo formation (eq. 35)65. Of note, the Ru-

bis(alkoxide) species are stable in dry benzene under an atmosphere of dry while in air in wet benzene the species are slowly regenerated over several hours, with eventual formation of 2 mole equivalents of ketone (eq. 36); such a reaction suggests the likely involvement of the bis(alkoxy)species as intermediates in the catalysis65. No radical

intermediates were detected during the above alcohol oxidations, and addition of radical inhibitors such as BHT in small amounts did not significantly effect the rate of ketone production, although there was some NMR evidence for reaction of BHT with possibly to form 169 a bis(phenoxy) species . The O-atom donor, lutidine-N-oxide, has been used with to catalyze the room temperature oxidation of alcohols to the corresponding aldehydes or ketones in ~80%96,145. Thus, under Ar, allyl alcohols were oxidized selectively to aldehydes selectively, and to was not oxidized. Cyclohexanol and adamantanol gave the corresponding ketones. The related system, mentioned in the previous section, catalytically oxidizes cyclohexanol to cyclohexanone160.

1. Catalytic oxidations using ruthenium porphyrins

51

Oxidations of racemic secondary alcohols ArCH(R)OH by system in the presence of HC1 proceed with some kinetic resolution138. The smaller molecules were oxidized with larger selectivity. Thus oxidation of when carried out to ~50% of conversion, yielded ~ 50% methylphenylketone, and the remaining alcohol was enantiomerically enriched (24% ee). Similarly, oxidation of alcohol gave 32% of ketone, with the ee of remaining alcohol being 10%; yielded 25% of ketone, and 25% ee for the remaining alcohol; and yielded 41% of ketone, and 2% ee for residual alcohol138.

3.7

Oxidative dehydrogenation of amines

The oxidation of amines is important biologically7,178,179. In particular, oxidative N-dealkylation (eq. 37) is one specific reaction catalyzed by cytochrome P-450 monooxygenase (Section 1)7,180, and several models

mimicking this reactivity have been reported using iron porphyrins181 and Ru non-porphyrin182 systems. More generally, oxidation of amines catalyzed by transition metal complexes can lead to this dealkylation, formation of amine or dehydrogenation183-185. Ruthenium(II) precursors have been used with a variety of oxidizing agents, including oxygen 186 , t BuOOH 182, 187 , PhIO188, 189, and bromamine-T190 . Reports utilizing high-valent Ru-oxo complexes include the use of (L = N- and P- donors)191 , species (see below); it is this last mentioned type of the system that is the focus of this Section. As first shown in the laboratory here, can oxidatively dehydrogenate primary and secondary amines under stoichiometric and catalytic conditions in benzene at 50°C (with as oxidant)185. For the former, the reaction stoichiometry depends on the number of H-atoms in the of the amine, and for primary amines can be presented primarily by eqs. 38 and 39 (Ru = Ru(TMP)): imines or nitriles are formed with generation of isolable Ru(II)-bis(amine) complexes, and an X-ray structure of was determined185.

52

Maria B. Ezhova and Brian R. James

Pyridine was not oxidized under these conditions185. Reactions of and with can lead to formation of complexes, and corresponding reactions with can lead to complexes of the type which was structurally characterized195; however, the fate of the oxo ligand and reaction stoichiometries in these systems were not determined (see below also)195,196. In the catalytic work185, subsequent to formation of the imine from secondary amines, secondary products such as aldehydes are seen due to imine hydrolysis by the co-product, The maximum turnover after 24 h was 20, observed for the conversion of to and to 18 and 15 turnovers were found for conversion of to Ph(Me)C=NH, and to respectively; after this time, most of the dioxo complex has been syphoned off as the Ru(II)-bis(amine) species, and the catalysis depends on the subsequent conversion of this back to the dioxo species. Figure 26 outlines the essential paths suggested for the catalysis185. An initial 2-electron oxidation of amine to imine by takes place, with formation of Ru(TMP)(O), which was detected; this again presumably disproportionates to and and the latter can then react with to regenerate the dioxo species (Sections 2 and 3.3), or with amine to give species. Imines with an atom can undergo a second dehydrogenation by or possibly Ru(TMP)(O).

Application of the halogenated porphyrin complexes and as catalysts for dehydrogenation of and gave higher rates than the TMP system, but the chloro-systems were visibly less stable (i.e. some bleaching occurred)185.

1. Catalytic oxidations using ruthenium porphyrins

53

More detailed studies on reactions between and amines have revealed, at least by NMR studies (see below), the presence of several Ru-intermediates en route to the final and coproducts (imine or nitrile, and water, eqs. 38, 39)197. UV-Vis studies at ~ M Ru(VI) for the reaction with (R = H, Me, Ph.) in benzene at 20°C show rapid spectral changes with clean isosbestic points for implying the absence of substantial amounts of any intermediates; attempted kinetic studies were thwarted by irreproducibility, although the substituted benzylamines reacted more slowly185,197. The kinetic inconsistencies possibly result from photosensitivity of the or other species (see below) in the light beam197. At the higher concentrations used for NMR, and in the darkness of the probe, kinetic data became reproducible within stoichiometric conditions (20°C under Ar). The stoichiometric reactions (cf. eq. 38) followed the simple rate-law k[Ru] [amine] for loss of Ru(VI), over an amine concentration range of ~ 0.01 - 0.3 M, with k values of 9.74, 1.72 and at 20°C for the R = H, Ph, and Me systems, respectively197, and there was no sign of saturation kinetics as found for the related alcohol systems (see Section 3.6). A kinetic isotope effect measured for the systems is perhaps consistent with formation of an activated complex of the type Ru=O....H---N, via an H-bonding interaction with H-atoms on the N-atom; the suggestion corresponds to that shown in Figure 25 for the alcohol systems where a C-H bond is initially stretched. Cleavage of the N-H bond of the amine versus the C-H bond of the alcohol would account for the generally faster stoichiometric reactions of the species with amines versus alcohols under analogous conditions. In contrast to studies on the alcohol systems, two Ru-intermediates were detected in the NMR studies en route to the bis(amine) product for the slower R = Ph and Me systems (eq.40)197. Intermediate (26) has symmetry and thus two different axial ligands (this being readily diagnosed

by the presence of two signals for the o-Me groups of the mesityl substituents49,60,196,198,199), while intermediate (27) has symmetry and two identical axial ligands. Potential intermediates such as and were ruled out, at least for the system, where the is readily available, as the bis(imine) and (amine)(imine) species with readily identifiable NMR spectra were synthesized independently from The findings are

54

Maria B. Ezhova and Brian R. James

consistent with (27) being the bis(imido) species and indeed a corresponding complex has been isolated and characterized by NMR in the reaction of and (27) was also synthesized from oxidation of by Formation of the bis-imido species implies a net loss of both H-atoms attached to the N-atom of the amine (to give rather than one H-atom from each of the N- and Most relevant to this is the interesting, observed isomerization of the bis(imine) species to the bis(imido) species in benzene at ambient conditions during the oxidation of by which is stoichiometric and forms stepwise the Ru(TMP)(amine)(imine) and then Ru(TMP)(imine)2 species (eq. 41); water and/or Ru species may play a role in such isomerization. Attempts to make species directly

from have not yet been successful. Other evidence for the bis(imido) formulation for (27) includes its reaction with which displaces the imido (or nitrene) moieties to generate the bis(amine) complex and a polymeric nitrene product197. Related nitrene transfer reactions from and (X = H, Cl, I, Me) to pyridine and tertiary phosphines, respectively, have been demonstrated (see below). Substitution of the imido ligand in (porp = TTP, 4-C1-TPP) by to generate has also been reported (see below)200. Identification of (26) is less definitive: most likely it is the oxo-imido species as corresponding species have been characterized (see below). For the system, data for the (26) and (27) intermediates were quite different, and were more consistent with them being the mixed imine/amine species (Me)Ph] and the bis(imine) respectively197. The chemistry of systems differentiated by a Ph vs. Me substituent thus appear to be remarkably different. In some systems with water added initially (prior to its formation as a co-product), significant amounts of were seen197. Mixed Ru(II)-amine/imine complexes of the type have been isolated from the reaction of = the dianion of the chiral 10,15,20-tetrakis[o-(2methoxy-2-phenyl-3,3,3-trifluoropropanolamino)phenyl]porphyrin) with excess esters198,202,203. For example, reaction of the methyl ester of

1. Catalytic oxidations using ruthenium porphyrins

55

L-alanine leads to complexes such as (28) in up to 80% yield Analogous products are formed using the methyl esters of Lvaline (55% yield), L-leucine (44% yield), L-phenylalanine (50% yield), and L-glycine (30% yield)). The complexes were also formed as co-products usually in < 5% yield, although the bis(glycine ester) complex was formed in 30% yield202; a complex was structurally characterized203. The coordinated imine is the expected product from oxidative-dehydrogenation of the reactant amine, and only the E-configuration was seen for the coordinated imines. It seems plausible that a bis(imine) species is formed initially and then one coordinated imine is displaced by the ester amine; the authors did not

mention production of free imine or its potential decomposition products. The amino ester/imino ester species could also be formed from electrochemical oxidation of the Ru(II)-bis (amino ester) complexes202. The reaction of a species with excess L-alanine methyl ester gave a 1:1 mixture of the amino ester/imino ester and bis amino ester complexes202, consistent with disproportionation of the precursor to and (Sections 2 and 3.3). The complexation of the amino esters to Ru(II) did not give any chiral recognition. For example, reaction of with 10 equiv. of rac-valine or -leucine esters yielded a racemic mixture of the corresponding bis (amino ester) complexes (DD : DL : LL = 1 : 2 : 1)203; however, chiral recognition to 52% (for leucine methyl ester) was observed for the complexation of the rac-leucine esters to and the oxidation of the amino esters by the chiral (or perhaps, more precisely, coordination of amino ester to the Ru imino ester complex)

56

Maria B. Ezhova and Brian R. James

was usually selective. Thus, oxidation of rac-valine ester (10 equiv.) yielded the bis(amino ester) complex with 66% ee; ee values for esters of alanine, leucine and phenylalanine were 10, 26, and 0%, respectively203. There are clearly several complications to the simplified scheme proposed for the oxidation of amines (see Fig. 26), although the essential feature of a disproportionation of to and likely remains. The can react with imine as well as amine, and oxo-imido or bis(imido) species have also been identified within certain amine systems, and the concentrations of the dehydrogenated amine species and co-product water, which increase as the stoichiometric or catalytic reactions proceed, will certainly effect the behavior of the system (see Section 3.6.). Reactions of other complexes with excess of primary or secondary amines have been reported195,196,205; with the porp = TPP, TTP, 4Cl-TPP, 3,5-Cl-TPP and 3,4,5-MeO-TPP, and with R = H, cyclohexyl, n-octyl, and n-dodecyl, and (R = Me, Et), the complexes are formed, and X-ray structures of and were solved; however, no information was given on the amine dehydrogenation products that must result from consumption of the oxo ligands. Of interest, Ru(IV)-bis(amido) complexes have been isolated when the secondary amine was used as reactant, and the stability of these complexes was strongly dependent on the substituent on the phenyl groups of TPP: and were isolated presumably via the stoichiometry of eq. 42, although this was not established196; the bis(amido) species can be reoxidized back to the dioxo complex using meta-chloroperbenzoic acid170. Other porphyrinato ligand systems generated a mixture of products196, in contrast with corresponding

chemistry of the complexes that cleanly gave the 206 bis(amido) products . The only X-ray structures of Ru(porp) mixed amido complexes were obtained for Ru(porp)(NHTs)(pz) (porp = TPP207, OEP208), which were isolated during some reactions of with alkene in the presence of pyrazine (see below). Somewhat analogous to eq. 42, reaction of (porp =TTP, 3,4,5-MeO-TPP) with the imine leads to formation of the corresponding methyleneamido complexes of which the trimethoxy derivative was structurally characterized.

1. Catalytic oxidations using ruthenium porphyrins

57

More generally, Ru(porp)-bis(imido) and -oxo(imido) species, products of oxidative-dehydrogenation of bis(amine) complexes, are of considerable interest as metalloporphyrin imido complexes have been postulated in the nitrite reductase cycle209, and can be used as models for metabolism of natural amines201. Further, the imido ligand is isoelectronic with the oxo ligand, and transition metal imido complexes, including metalloporphyrin derivatives, are the key reagents or intermediates in reactions such as oxyamination210, amination211-215 and aziridination201b,207,208,213,214,216 of olefins. Outside of Ru species, metalloporphyrin complexes with imido ligands are well demonstrated and include those of Ti217, Cr218, Mn216, Fe201, and Os68,200,206,219. Well characterized complexes are the ptolylsulfonyl derivatives (porp = TPP, TTP, 4-C1-TPP, 4MeO-TPP, OEP, TMP, prepared by Che’s group via reaction of Ru(porp)(CO)(MeOH) with PhINTs207,208,213; these tosylimido complexes are stable in the solid state for a few days at –15°C, but in at r. t. they decompose after a few hours, and are not very reactive toward water. They readily react with to yield the bis(phosphine) complexes208. Formation of species with other substituents on the N atom and has been proposed in oxidation of and oxidation of primary amine by (see above). has been isolated200, but it quickly hydrolyzes in air to give oxo(imido) and species68,196,200. In comparison, complexes are more stable and have been characterized by 200 X-ray analysis . A mono(imido) complex where X = Me, H, Cl, I) (analogous to has been prepared via reaction with This mono(imido) species can react with tertiary phosphines to yield phosphinimines and complexes (Fig. 28); the imido transfer reaction for the system obeys the rate law = showing standard saturation kinetics behavior199. The electron-deficient arylimide systems were more reactive than the electron-rich analogues [k = 39.5 (X = I), 16.0 (X = H), 13.6 (X = Cl), 25.0°C, in toluene]; and the more reactive the imido complex, the lower its affinity for the phosphine [K = 1.15 (X = I), 13.6 (X = H), 2.39 (X = Cl), . The proposed mechanism involves reversible binding of phosphine to and subsequent rate-limiting irreversible intramolecular imido-group transfer (Figure 28). Unfortunately, no spectroscopic evidence was given for species (29) (or (30)) under conditions where they should be present.

58

Maria B. Ezhova and Brian R. James

Analogous to O-atom transfer from Ru-oxo species, imido transfer can occur from and species. Stoichiometric imido transfer from (31) to alkenes to give aziridines has been demonstrated when porp = TPP, OEP, TTP, 4-C1-TPP, 4MeO-TPP, the reaction being formally analogous to epoxidation using species; for example, when was used in in the presence of pyrazole, aziridines are formed in 22 and 10% yields from cis- and trans-stilbenes, respectively, 75 and 72% yields from cis- and respectively, and in 67-82% yields for p207,208 substituted styrenes . The Ru is recovered as the complex Ru(porp)(NHTs)(pz) (32), the overall stoichiometry being shown in eq. 43207,208. Insertion of the imido group into a C-H bond of alkanes correspondingly gives the tosylamide; e.g. toluene gives the tosyl derivative of benzylamine (eq. 44), in 9% yield208,213,214, while benzyl alcohol gives benzaldehyde in 95% yield (eq. 45)207. The rate-determining step for the aziridination of alkenes by was considered to be formation

of a carboradical intermediate (Fig 29, cf. Fig. 8 for epoxide formation); clean isosbestic points were observed during kinetic studies at 25°C by UVvis spectroscopy, the rate-law was simply in each of and

1. Catalytic oxidations using ruthenium porphyrins

59

styrene, and the small isotopic effects determined at the and at the were considered to be consistent 208 with the nonstereospecificity of such reactions . The net oxidation of benzyl alcohol also obeyed the rate law The insertion of the imido group into alkane C-H bonds occurs selective at the tertiary centers. For example, with adamantane, N-(1adamantyl)tosylamide is formed in 52 and 60% yields from and respectively, under stoichiometric conditions in the presence of pyrazole208. Cyclohexane gave the cyclohexylamide in only 10% yield; ethylbenzene gave (~80%), isopropylbenzene yielded (~75%), while cyclohexene produced 3-(N-tosylamido) cyclohexene (~85%), the insertion occurring at a C-H bond to an carbon. Formation of a carboradical was again considered rate-determining for the alkane reactions, with the relatively large isotope effect being found for tosylamidation of cyclohexane by

(Fig. 30)208.

Application of the more electron deficient complex Ru(TPFPP)CO as a catalyst for such imido transfer reactions with PhINTs as donor, in the absence of pyrazole, led to reasonable yields of the corresponding aziridines and amides214. For example, aziridination of allylbenzene, cyclooctene and

60

Maria B. Ezhova and Brian R. James

1-octene in at 40°C for 2 h (Ru : substrate : PhINTs = 1 : 75 : 150) generated the aziridines with conversions of 32, 43 and 42%, respectively214. The amidation reactions under such conditions (for example, with indane, adamantane and 1,2-dihydronaphthalene) resulted in corresponding amides in high conversions (92-99%) of substrate. Use of a lower temperature, or use of benzene or MeCN, gave poorer results214. Of interest, use a mixture of and (R = Ts, Ns, instead of PhINR itself, for aziridination and amidation of alkenes and alkanes with Mn(TPFPP)Cl as a catalyst, gave excellent substrate conversions, but the corresponding Ru system was not tested214. Enantioselective amidation of saturated C-H bonds has also been reported by Che’s group213. Application of or a mixture of as a catalyst 40°C, Ru : PhINTs : substrate = 1 : 100 : 500) for a range of substrates such as ethylbenzene afforded corresponding amides in yields of 16 to 47% and ee values of 6-48%, all with excess S-configuration213. Attempts to synthesis bis(imido) compounds via oxidative bromination of led to formation of the oxo(imido) species (eq. 46, porp = TPP, TTP, 4-C1-TPP, 3,4,5-MeO-TPP, 3,5-Cl-TPP), that were isolated and characterized68,196,200, where the source of the oxo-ligand was most likely trace amount of water present in the solvent68,196. The oxo(imido) complexes were more stable than the corresponding dioxo species, being air-stable in the solid state, but gradually decomposing in to form the nitrosyl complexes Ru(porp)(NO)(OH). The species react with amines or phosphines to yield the respective bis(amine)- and complexes200, but

the fates of the oxo and imido ligands were not determined. The corresponding and species do not react with and reactivity with is very slow. The nitrido complex is said to be formed via reaction of with aq. in although an attempt to make the nitride complex via oxidation of with mCPBA, or air was not successful220. In contrast, oxidation of complexes (porp = TPP, TTP, 4-C1-TPP) with m-CPBA in a mixture under reflux gave the corresponding species200.

1. Catalytic oxidations using ruthenium porphyrins

4.

61

CONCLUSIONS

Studies on Ru-porphyrin chemistry first appeared in the early 1970s and, as in many other areas of chemical research, the number of publications dealing with the topic has grown exponentially, with much of the work developed from a theme and rationale to mimic the extensive biological chemistry of naturally occurring, enzymatic Fe-porphyrin systems; some of these are involved in oxygenase systems that carry out highly selective oxidation processes of the types which are very attractive from an industrial point of view. As an alternative to using enzymes industrially, efforts remain intense to develop protein-free, “model complexes” that will activate for oxygenation/oxidation processes. Currently the emphasis probably lies in the use of non-porphyrin complexes of Fe and Cu to mimic, for example, methane-monooxygenase221 and oxidase systems44,222, respectively, while the use of Ru, the “slowed down” 2nd-row analogue of Fe, is an obvious choice that has found emphasis in porphyrin-based systems. Certainly, the work from the laboratories here has developed over the last 30 years from the hypothesis that Ru systems would allow for easier detection of intermediates in oxidation pathways and a better understanding of such reactions, this hopefully eventually leading to design of improved catalysts (based on Ru or Fe). The contributions, plus those from many other groups, have led to identification of species with coordinated superoxide and peroxide, and bridging and terminal oxo ligands. Correspondingly, increased understanding of enzyme systems (particularly cytochrome P-450) has contributed to development of the chemistries of Fe- and Ru-porphyrins. The published coordination chemistry of Ru-porphyrins with the metal in oxidation states II to VI is now very extensive, and is discussed in several sections of this Chapter. The ability of Ru(II)-porphyrins to split to generate complexes, and the subsequent transfer of both these coordinated “O-atoms” to organic substrates, both represent truly remarkable and unique chemical processes: to our knowledge, no other metallic center exhibits the reactivity pattern, let alone such “bis(monooxygenase)” activity. There is no doubt that the catalyzed oxygenations of P-, As-, Sb-, and S-based substrates operate by genuine O-atom transfer processes, and involve a key reversible disproportionation of to and Such pathways also pertain in epoxidation of olefins using porphyrin systems (based on sterically hindered porphyrins such as TMP and TDCPP (or OCP)), but activity is slow (a maximum of ~100 turnovers per day) and limited by decomposition chemistry; efforts to heterogenize the systems have been initiated. Use of corresponding with chiral porphyrins has demonstrated their

62

Maria B. Ezhova and Brian R. James

feasibility for asymmetric epoxidation, but ee values of only ~70% have been achieved with small total turnovers ( 100,000 catalytic turnovers, e.g. by as a prototype precatalyst.108

168

T.Funabiki

The functional model oxygenations above mentioned are all for catechol 1,2dioxygenases (CTD) or protocatechuate 3,4-dioxygenases (3,4-PCD). However, there is another group of catechol dioxygenases, chlorocatechol dioxygenases (CCD). These dioxygenases are characteristic of the low substrate specificity and catalyze the intradiol cleavage of not only mono or di-halogenated catechols, but also other various catechols.4 The function and structure of this type of enzymes are not well characterized,58, 59, 65, 66 but development of the functional model oxygenations is a very attractive object from both sides of basic and applied sciences, especially environmental sciences. Funabiki et al. was successful to develop the functional model oxygenation of 3- or 4-chlorocatechol by using TPA as a ligand (eq. 3).109

4. Functional model oxygenations by nonheme iron complexes

169

Formation of chlorocatecholatoiron complexes and the oxygenative decomposition with molecular oxygen were monitored by electronic absorption spectroscopy and by product analysis. Interestingly dehydrochlroination of the initial products was accompanied to give the products without the chloride group. It was found that the addition of 2,6-lutidine as base is effective for promotion of the oxygenation. The cleavage of chlorocatechols by nonheme iron complexes indicated clearly that catechols for the model chemistry are not limited to 3,5-di-tert-butylcatechol and that iron complexes can oxygenate the catechols having electron-withdrawing substituents. From the viewpoint of dehalogenation of halogenated aromatic hydrocarbons, it is important to develop model systems that decompose multi-halogenated catechols, e.g. dichlorocatechol and tetrachlorocatechol. Apart from the model oxygenations, Sorokin et al. developed powerful catalysts for oxidation of tetrachlorocatechol to dichloromaleic acid by using ironphthalocyanine in the presence of or percarbonate as an oxidant.110 The catalyst oxygenates DTBC to give a furanone. A new functional model chemistry on catechol dioxygenases has been developed by Funabiki et al..111 Water-soluble ligands were prepared by sulfonation of tripodal ligands such as TPA and used for catalytic oxygenation of water-soluble catechols such as 4-tert-butylcatechol, 4-chlorocatechol, and protocatechuic acid (eq. 4).

Despite a typical substrate in the enzymatic system, protocatechuic acid was found to be oxygenated by model complexes not in organic solvents but in water. The highly selective and catalytic intradiol oxygenation of 4-tert-butylcatechol indicated that various types of catechols other than 3,5-di-tert-butylcatechol can be used as substrates in the model systems and that oxygenations by nonheme iron systems in water are attractive. The reactivity and selectivity are dependent on the substituent on catechols and of the solution.

170

T.Funabiki

3.1.2 Extradiol Cleavage Oxygenations Structural studies on the catechol dioxygenases that function in the extradiol cleavage oxygenations have progressed remarkably in recent years. However, the functional model studies are far beyond those for the intradiol cleavage. This is because it is hard to isolate the intermediate model complexes, e.g. and even if isolated they are readily converted to the Fe(III) complexes under The first example of the extradiol cleavage by iron complexes was reported by Funabiki et al..86, 87, 112, 113 The extradiol cleavage products were obtained together with the intradiol cleavage products in the system (intra : extra = ca 2 : 1) (eq. 1). In this system, the presence of the species in solution is probable since the oxidation of DTBC to DTBQ as a side reaction is accompanied by the reduction of to Participation of the complex for the extradiol oxygenation was also shown by the different reactivity of from in the in situ system in the presence of 112 1:2 pyridine or water (intra : extra = 1 : 5 Similarly, Lin et al. found that compared with (TACN = 1,4,9-triazacyclononane) pyridine in methanol produced the higher yield of the extradiol cleavage product over the intradiol from pyrocatechol and 3methylcatechol: 1 : 6.7 and 1 : 3 from pyrocatechol.114 The rapid product formation was observed prior to the formation of the complex. The extradiol cleavage of a range of 3- and 4-substituted catechols with electron-donating substituents was observed.115 Different selectivity was observed when monosodium catecholate was used in the absence of pyridine, that may imply a requirement of a proton donor for the extradiol cleavage. Lim et al. synthesized a Fe(II) complex, (BLPA = bis(6-methyl-2-pyridyl)methyl)(2-pyridylmethyl)amine, DTBCH = 3,5-di-tert-butylcatecholate monoanion), that reacts with to give the intradiol (65%) and extradiol (20%) cleavage products.116 The partial direct reaction of the Fe(II) complex with was assumed in addition to the reaction of the Fe(III) complex, which is formed rapidly under but the evidence for the participation of the Fe(II) complex is poor. Jo et al. also synthesized a series of Fe(II)-monoanionic catecholate complexes, 117

The crystal structure of showed that the DBCH ligand binds to the iron center asymmetrically. The complexes react with or NO to afford blue-purple Fe(III)catecholate dianion complexes, that react further with to give a high yield of cleavage products. The products are mainly derived from intradiol

4. Functional model oxygenations by nonheme iron complexes

171

cleavage with a small extent of extradiol cleavage (89 : 3 or 78 : 12). Weiner et al. developed polyoxoanion-supported Fe(II) complex that produces the extradiol cleavage products (ca. 52%) together with DTBQ (40%).118 All-inorganic polyoxoanion as oxidation-resistant ligand forms a Fe(II)-DTBC species that binds with to form 1 : 1 : 1 species (volumetric titration) as shown in Fig. 5. The oxygen species decomposes very slowly to give oxygenated products. No explanation has been given to the very slow (7 days) oxygenation of DTBC, but the system gives an example of catalytic dioxygenase-like system without the need to worry about catalyst lifetime-limiting, ligand-oxidation side reactions.

Instead of the use of complexes, the extradiol oxygenation by 90, 119, 120 complexes has been studied. This mimics the function of the containing enzymes that produce the extradiol oxygenation products from some catechols.60-62 Dei et al. synthesized tri- and tetra-azamacrocyclic ligands and studied the reactivity of the complexes with oxygen, and reported that the complex in acetonitrile yielded 3,5-bis(1,1dimethylethyl)-2H-pyranone (30%) and 4,6-bis(1,1-dimethylethyl)-2H-pyranone (5%) by the extradiol cleavage in addition to DTBQ (65%) (eq. 5).90

Ito et al. used (but not AgOAc) to remove the chloride from the same complex and found that the extradiol cleavage products, two isomeric di-tert-butyl2-pyrones, can be selectively produced (76 : 16).119 It is characteristic that the intradiol cleavage products are formed in only trace amounts. Ogihara et al. used the hydrotris(pyrazolyl)borate ligand to synthesize a five-coordinate

172

T.Funabiki

catecholate complex from which the extradiol oxygenation products (pyrones: 39 + 28%) were obtained together with the intradiol oxygenation product (33%), but not with quinone.120 These results suggested that the presence of the vacant site, i.e. formation of the five-coordinate catecholatoiron species as an intermediate, is the more important for the extradiol cleavage oxygenations rather than the Fe(II) active center. However, the presence of the vacant site is not enough for explanation of the selectivity control by Fe(II) and Fe(III) centers. It is known that the five-coordinate intermediate gives selectively the intradiol cleavage product in the enzymatic system.121 Lim et al. reported formation of both the intradiol (75%) and extradiol (15%) cleavage products from the six-coordinate Fe(III) 122 complex,

Recently, Jo and Que reported the importance of the location of the vacant site for the extradiol oxygenation.123 They compared reactivities of a series of Fe(III) catecholate complexes, containing tridentate ligands that can coordinate to iron in a facial or meridional fashion ( 1,4,7-trimethyl-1,4,7-triazacyclononane, TPY = 2,2’ : 6’,2”terpyridine, BnBPA = N-benzyl N, N-bis(2-pyridylmethyl)arnine).123 The complexes react with in the presence of AgOTf to remove the Cl ligand. The products obtained indicated that the facial tridentate ligand favours the extradiol cleavage rather than the intradiol cleavage. In the proposed mechanism, the facial ligand allows and substrate to occupy the opposite face and form an intermediate that leads to the desired extradiol products (vide infra).

4. Functional model oxygenations by nonheme iron complexes

173

Extradiol cleavage is performed by manganese-containing dioxygenases. Canseshi and Dei synthesized a manganese complex, [Mn(CTH)(DTBC)]Y which structure was analyzed by X-ray diffraction (CTH = (±)l5,7,7,12,14,14-hemxamethyl-1,4,8,11-tetraazacycloetradecane).124 It was found that the two complexes have different physical properties in the solid state, but the spectroscopic properties of their solutions are identical. In highly polar solvents such as DMSO, the spectra of the yellow solutions are consistent with the presence of an chromophore, whereas in weakly polar solvents such as toluene or acetone the spectra of the blue-green solutions are consistent with the presence of a chromophore. No detail has been reported, but interestingly the complex in polar solvents are stable under oxygen, while the complex in weakly polar solvents is quickly oxidized to give DTBQ (60%) and extradiol cleavage products, 3,5- or 4,6-bis(1,2-dimethylethyl)-2H-pyranones (36% + 4%). Funabiki et al. have synthesized a complex but it was converted selectively with molecular oxygen to an intradiol cleavage product.125, 126 3.1.3 Mechanisms of Oxygenations In the mechanism of the intradiol oxygenation, the first step is the activation of catechol by coordination to a Fe(III) center. The chelate coordination of catechols in the enzymatic systems (protocatechuate 3,4-dioxygenase, 3,4-PCD) has been clarified by the X-ray analysis.56, 57, l21, 127-132 In addition, the analysis indicated the binding of catechols to 3,4-PCD forms a five-coordinate complex with the axial Tyr dissociation from two Tyr and two His ligands (Fig. 7).

As described in 3.1.1, various types of catecholatoiron model complexes in the form of the chelate coordination have been isolated and found to give the oxygen insertion products by the reaction with molecular oxygen. Different from the enzyme structure, however, most of them are in the six-coordinate configuration.

174

T.Funabiki

On the other hand, the probability of the monodentate coordination of the catecholate ligand has been proposed in the case of pyrocatechase.133-136 In the model system, Fujii et al.137 and Nishida et al.138, 139 reported supports for the monodentate dianionic catecholate species for the reaction with molecular oxygen based on the EPR spectroscopy and molecular orbital considerations, but no further evidence has been obtained. As described in 3.1.2, five coordinate catecholate complexes are formed in solution from [Fe(L)(DTBC)Cl] (L = TACN, after removing the chloride ligand by silver salt90, 119, l23 or by using a ligand120 in relevance to the extradiol cleavage. Interestingly, these five-coordinate complexes do not give selectively the intradiol cleavage oxygenation. Thus questions remain to be solved whether the five-coordinate intermediate for the protocatechuate 3,4-dioxygenase is specific or the oxygenation mechanisms are different between the enzymatic and model systems. The iron in enzymes retains the high-spin ferric state throughout oxygenation.140-142 However, it has been suggested that activation of the catecholate ligand in the dianionic chelate form to the radical form is important for the reaction with molecular oxygen.87, 143 This was first represented by equilibrium (eq. 6), but later by the radical character (eq. 7).

The radical character of the catecholate ligand has been supported by various spectroscopic data, e.g. UV/VIS, NMR, and XAFS. Characteristic phenolate-toiron(III) LMCT bands at ca. 500 and 900 nm, which were first observed by Funabiki et al.,86, 87, 94 are observed with various functional model catecholatoiron complexes and tried to correlate qualitatively with semiquinonate character of the catecholate ligands and quantitatively with the or DTBQ redox 100, 101, l44 potentials. Recently the LMCT bands were observed both with the catecholatoiron complexes in the solid state and in solution, supporting the radical character of the catecholate ligand as represented by eq. 7.145 It was shown by the absorption edge values in the XANES spectra that the electronic states of iron of the catecholatoiron complexes are between those of and but very different 145 from that of Spin crossover from high- to low-spin states with decreasing

4. Functional model oxygenations by nonheme iron complexes

175

temperature was observed with complex, that may be related to the intense LMCT band throughout the temperature range.145, 146 In the most mechanisms proposed for the intradiol cleavage oxygenations, the direct attack of molecular oxygen upon a carbon atom attached with a phenolate O atom is favored (substrate activation process, A in Fig. 8). The alternative mechanism prefers the direct attack of to the iron center (oxygen activation process, B in Fig. 8).

The A process is advantageous to explain the fact that even hexagonal catecholatoiron complexes such as react with This reason, however, is not applied to the enzymatic system since an open site becomes available for the direct coordination of to the iron center by dissociation of the Tyr ligand. One question arises whether the radical character of the aromatic carbon is strong enough for the direct attack of About this question, Funabiki et al. studied on the catecholatoiron complexes by Extended-Hückel148 and densityfunctional theory149 analyses. Catecholato- and 3,5-di-methylcatecholato(n = 3 or 4) were used as model complexes for convenience. Atomic charge and spin densities (spin, HOMO) of the dimethylcatecholate complex are followings. O1: -0.58 (0.29, 0.14); O2: -0.56 (031, 0.18); Cl: +0.14 (0.043, 0.088); C2: +0.043 (0.059; 0.12); C3: +0.0017 (0.023, 0.024); C4: +0.037 (0.056, 0.075); C5 +0.076 (0.049, 0.082); C6: +0.035 (0.026, 0.0006); Fe: +1.38 (3.91, 0.023). It is apparent that most of the negative and positive charges are located on O and Fe, respectively. Carbons are slightly positively charged. Most of the spin density is also located on

176

T.Funabiki

O1 and O2, and the carbons have very low spin density. Recently, Wheeler et al. have studied on electronic structure of 3,5-di-tert-butylbenzosemiquinone (DTBSQ) by density functional theory analysis.150 Normal population analysis (NPA) atomic charge densities and spin densities in parenthesis (Mulliken net, NPA atomic) are followings. O1: -0.626 (+0.271, +0.272); O2: -0.662 (+0.245, +0.248); C1: +0.368 (+0.031, +0.038); C2: +0.362 (+0.075, +0.076); C3: -0.113 (+0.095, +0.087); C4: 0.257 (+0.116, +0.104); C5: -0.066 (+0.072, +0.075); C6: -0.306 (+0.111, +0.098). Most of the negative charge resides also on O1 and O2 and most of the positive charge resides on C1 and C2. Most of the spin density is also located on O1 and O2. The carbons attached to the tert-butyl groups (C3 and C5) have less spin density than those attached to hydrogens (C4 and C6). In both cases, the spin density on carbons, e.g. C2, is very low and does not link up with the direct attack of the molecular oxygen to C2 in the step of intradiol cleavage. By calculating the total energy change caused by approach to Fe or carbon, Funabiki et al. have proposed the probability of the B process (oxygen activation process) rather than the A process even in the intradiol cleavage.149 In this calculation, the configurational change of the catecholate ligand from chelate to monodentate form is assumed to open a coordination site for As for the reactivity or the semiquinonate ligand with Mialane et al. studied the structure and electronic properties of complex (L = N,N’bis(4-methyl-6-tert-butyl-2-methyl-phenolato)-N,N’-bismethyl-1,2-diaminoethane). 151 In spite of the coordination of the DTBSQ monoanion, the complex was found to be stable under supporting that the semiquinonate carbons do not bind directly with The complex is reduced not to but to that is also stable under Koch et al. also synthesized a complex, This complex was also found to be converted to with superoxide, but stable towards molecular oxygen. Since DTBSQ is proposed to react with as an intermediate in the oxygenative cleavage 153 of by the results indicate that the binding of DTBSQ to stabilizes the DTBSQ moiety against These results support the importance of the form for the C-C bond cleavage, but do not give any or positive support for the direct attack of oxygen to the semiquinonate carbon atom. Recently, experimental supports for the oxygen activation process (B) were reported to explain the extradiol oxygenation of the complex. 123 The process is favored by the complex having facial tridentate ligands rather than tetradentate ligands; in the former and substrate can occupy the opposite face and form an intermediate that leads to the extradiol products (Fig. 9). The A process was proposed for the intradiol oxygenations for the complexes with meridional

4. Functional model oxygenations by nonheme iron complexes

177

tridentate or tetradentate ligands. However, this explanation is not enough for the preferential extradiol oxygenation over the intradiol oxygenation. If the radical spin density on aromatic carbon is similar among the complexes with facial or meridional tridentate ligands and tetradentate ligands, the probability of the oxygen attack to carbon in the B process may be similar. Since the radical character of carbon is different from that of iron in nature, we should be careful to ascribe the selectivity control only to the steric reason. If the ferric center is electron rich enough for binding with molecular oxygen, oxygen may bind first to iron for both intra- and extradiol cleavages, if necessary, with replacing a ligand.

The role of accessibility of to was also discussed on the comparable activity of the ligand system to the fast TPA systems = 2,11154 diaza[3,3](2,6)pyridinophane, Fig. 4). Different from the system, which produces only the intradiol cleavage products, the system produces both intra- and extradiol cleavage products in spite of the tetradentate ligand. The methyl group is thought to be effective to control the species. After the oxygen binding to either iron or aromatic carbon, formation of the peroxide species, has been proposed as an important species, from which one O atom is inserted between the C-C bond. There is little discussion on this process, but explained by the Baeyer-Villiger oxidation process or Criegee rearrangement. None of this type of oxygen adducts has been identified in the iron model systems. Stable rhodium and Iridium porphyrin complexes of the forms, 155 and Ir-O-O-DTBC,156, l57 were isolated, but the very low

178

T.Funabiki

reactivity with is difficult to apply the result to iron systems. In the extradiol oxygenation, there are two possible ways of attachment of oxygen to aromatic carbon, C1 or C6. Funabiki et al. have first proposed the attachment of oxygen to C6 since the attachment at C1 has a possibility of both intra- and extradiol cleavages.87 However, Bugg et al. have preferred a common type of peroxo intermediate (an oxygen adduct at C1) for both intra- and extradiol cleavages.114, l l 5 , 158-163 Factors controlling the different migrations, i.e. acyl migration for intradiol cleavage and alkenyl migration for extradiol oxygenation (Fig. 10), should be clarified to support this mechanism.

Recently, Funabiki et al. have pointed out on the quantum chemical calculation study that the species if formed tends to be stabilized in its form rather than to break the O-O bond toward insertion of one O atom to DTBC.149 The alternative proposal is the attachment of an O atom of between a C-C bond to form an epoxide-like species (Fig. 11). The probability of this process should be discussed in future on the more elaborate calculations.

4. Functional model oxygenations by nonheme iron complexes

179

Little has been studied on the mechanism of the extradiol oxygenations by enzymes and model complexes containing The first step was thought to be the 164 coordination of oxygen to a Fe(II) center, but now the coordination of catechol prior to oxygen is favored,165, 166 Two types of coordination fashion of catechols to have been proposed, i.e. bidentate59, 167-171 and monodentate.172 In the latter case, the loss of only one of the catecholate protons leads to highly asymmetric binding to This leads to charge distribution on the ring, which directs the oxygen attack of the iron-bound dioxygen to a position where extradiol cleavage must occur. Similarly to the intradiol oxygenation, carbon-oxygen bond formation between the semiquinone and superoxide to give an unsaturated lactone intermediate is postulated. Formation of this type of intermediate rather than a dioxetane intermediate was shown by the results of the incorporation by 2,3158 dihydroxyphenylpropionate 1,2-dioxygenase. As for the evidence for a semiquinone intermediate, Spence et al. have observed the cis-trans isomerization of a cyclopropyl radical trap in the oxygenation of cis- and trans-2-(2,3dihydroxyphenyl)cyclopropane-1-carboxylic acid by extradiol oxygenases.159 Formation of 85-95% trans product and 6-15% cis products during the oxygenation was explained by the reversible opening of the cyclopropyl ring of a semiquinone radical intermediate.

3.2 Dioxygenases other than Catechol Dioxygenases Lipoxygenase and dioxygenase are recent targets for model studies on nonheme iron dioxygenases other than catechol dioxygenases. Little progress has been attained in the functional model studies for these oxygenases, but some mechanistic works reported have been reported. As for lipoxygenase, Kim et al. synthesized an air-stable complex as a model for the iron(II) site of lipoxygenase.173, 174 The observed metal-centered transformation of this complex in the reaction with ROOH, i.e. was regarded to parallel the changes observed for lipoxygenase in its reaction with its product hydroperoxide. No application for peroxidation of substrate was studied. Goldsmith et al. studied the model reaction of (PY5 = 2,6bis(bis(2-pyridyl)methoxymethane)pyridine) with linoleic acid and hydrocarbons possessing weak C-H bond, for mimicking the rate-determining step, i.e. the hydrogen atom abstraction from the pentadiene subunit of the substrate by an active ferric hydroxide species to give a ferrous water species and an organic radical.175 It was found that reactivity scales best with the bond dissociation energies of

180

T.Funabiki

substrates, rather than (Scheme 1).

supporting the hydrogen atom abstraction mechanism

acid-dependent dioxygenases were studied by Ha et al.,176 Funabiki and Que et al..179-185 It is proposed that the latter oxygenation involves et al. the binding of to Fe(II), followed by substrate binding, and binding to afford a Fe(III) superoxide species as shown in Scheme 2. Model studies are mostly limited to characterization of complexes, which are inactive for oxygenation. 177, 178

4. Functional model oxygenations by nonheme iron complexes

181

Rieske dioxygenases catalyse enantioselective cis-dihydrokylation of arene and alkene double bonds (eq 8) and utilize a mono-nuclear nonheme iron active site.

Some nonheme monoiron complexes were found to catalyse the cis186-191 dihydroxylation with These catalysts are complexes of tetradentate N4 ligands such as TPA and BPMEN and capable of both epoxidation of olefins and cis-dihydroxylation. Recently Ryu et al. found that produces predominantly the cis-diol product with a diol : epoxide ratio of 3-4 under conditions of limiting as shown in Table 1..190 This is contrast to the preferential epoxide formation with It is reported that complexes forming low-spin Fe-OOH afford diols predominantly, while those forming hi-spin Fe-OOH afford epoxides (vide infra).

4. FUNCTIONAL MODEL SYSTEMS FOR NONHEME IRON MONOOXYGENASES Monooxygenases catalyze the incorporation of one atom of oxygen from molecular oxygen into substrates as shown in eq. 9.

Oxygenations proceed at nonheme mono- and diiron centers in the presence of electron and proton donors that are important for activation of molecular oxygen. The presence of the efficient electron and proton donor systems is characteristic of the enzymatic systems. The most popular nonheme iron monooxygenase is soluble

182

T.Funabiki

methane monooxygenases (sMMO), that bear diiron centers and are able to catalyze monooxygenations of not only methane to methanol, but also other alkanes and alkenes to alcohols, ketones and epoxides.4 The other important types are monoiron enzymes, i.e. pterine-dependent hydroxylases; phenylalanine hydroxylase which catalyzes para-hydroxylation of an aromatic ring of phenylalanine, tyrosine hydroxylase which catalyzes ortho-hydroxylation of an aromatic ring of tyrosine to form a catechol moiety, and triptophan hydroxylase which catalyzes hydroxylation of an aromatic ring of triptophan. Isopenicillin N synthase is also an important monooxygenase, that catalyzes the ring closure reactions of (hydroxycarbonyl)pentanolyl]-L-cysteinyl-D-valine (ACV) without incorporation of oxygen into tiazolidine rings of isopenicillin N. Functional model chemistry for nonheme iron oxygenases has developed remarkably in recent years.5, 192, 193 Not only iron but also other metals were found to catalyze the oxygenase-like oxygenations. Examples shown here will be limited to those of nonheme iron model complexes. Enzymatic monooxygenations require molecular oxygen as an oxygen source, but both of molecular oxygen and activated oxygen, e.g. hydrogen peroxide, can be used in the functional model systems. Continuous challenging works have brought about new iron complexes that can incorporate one oxygen atom to substrates to give alcohols, ketones, epoxides, etc., and various types of ligands were designed for enhancement of activity, selectivity control, and clarification of mechanisms. In the oxygenation mechanisms, formation and structures of active iron-oxygen species are most interested. In the so-called metal-based mechanisms iron-oxo or iron-peroxo species take part in the oxygen transfer to substrates and participation of either the or manifolds has been discussed. In these efforts, and complexes have been isolated and their reactivities with substrates have been studied. It is well assumed that the postulated iron-oxo species and are too unstable to be detected or isolated, but nonspectroscopic evidence for these species came to appear in recent years. On the other hand, participation of free radicals should be considered in most cases of oxygenations. Both of carbon and oxygen centered radicals can take part in the oxygenation processes. Apart from the apparent autoxidation process, it is not easy to differentiate explicitly metal-based mechanisms from free radical mechanisms. Different types of the radical-clock reagents have been developed for detection of radicals of different lifetime, especially in the discussions on the radical-rebound mechanisms. One of important aims of functional model chemistry is to develop efficient oxygenation catalysts that are useful in industries and applied fields. Selective and

4. Functional model oxygenations by nonheme iron complexes

183

efficient oxygenations with the use of molecular oxygen are highly desired. This is far beyond satisfactory, but very much promising.

4.1 Functional Model Oxygenations by Diiron Complexes 4.1.1 Monooxygenation by Diiron Complexes with Peroxides By mimicking soluble methane monooxygenases, monooxygenation of alkanes, alkenes, and arenes by various types of nonheme diiron complexes have been performed with utilization of activated oxygen species, e.g. mchloroperbenzoic acid (m-CPBA), hydrogen peroxide (TBHP), cumene hydroperoxide (CHP), hydrogen peroxide etc.. The reaction proceeds catalytically, but discussions focus on the oxygenation mechanisms, especially on the participation of radicals in the monooxygenation steps and similarity of mechanisms to that of enzymes. Followings are some of the recent examples. Kodera et al. synthesized a rigid complex with 1,2-bis[2-di(pyridyl)methyl-6-pyridyl]ethane that is a dinucleating hexapyridine ligand.194 The complex monooxygenates alkanes (cyclohexane, methylcyclohexane, adamantane) in the presence of m-CPBA with a large turnover frequency and number (TN ).195 Radical–rebound mechanism was suggested based on the reactivity with the additives and kinetic isotope effect. Likewise, the stable complex was activated in the presence of acid chloride RCOCl and DMF to oxygenate hydrocarbons such as cyclohexane.196 Payra et al. used diiron(III) bis(benzimidazole) complex for the quantitative and catalytic epoxidation of styrene in the presence of m-CPBA or NaOCl and N-methylmorpholine N-oxide as an important additives.197 In the absence of oxidant under anaerobic conditions, the reversible oxygen transfer was observed between the complex and triphenylphosphine. A complex with bidentate bipyridine (bipy) ligands 198 was synthesized. The complex possesses Cl ligands that enable the coordination of substrates. This is modification of the model complexes with a variety of tridentate N-based ligands (L) that block the terminal binding sites for substrates.199-202 Using T B H P hydroxylation of ethane (TN: 1.2/3 days), oxygenations of propane (TN: 13/2 days) and cyclohexane (TN: 72/3 days) were performed, resulting in the reactivity sequence of Fish et al. have used TBHP and for oxygenation of cyclohexane, toluene, adamantane, propane, and ethane by another complex, (tmima = [{(1methylimidazol-2-yl)methyl}amine]), that is characteristic for the polyimidazole ligands.203 Formation of intermediate is suggested and the mechanism

184

T.Funabiki

proposed involves the initiation of the reaction by the homolytic reaction with C-H bonds to form a carbon radical, followed by rapid trapping with to form the alkyl hydroperoxide. Decomposition of the peroxide is catalyzed by the initial FeOFe complex to form alcohol, aldehyde or ketone. A complex with an aqua ligand in place of the acetate ligand was synthesized and used for the alkane functionalization reaction.204 Compared with the complex, the complex resulted in the higher turnovers/h (approximately twice) (Table 2).

It was suggested that the loss of the terminal or ligand by TBHP displacement must be rate limiting in the formation of the active Fe oxidant complex. The radical clock experiment using trans-2-phenylmethycyclopropane indicated that the radical rearrangement of the cyclopropylcarbinyl to phenyl-but-3enyl radical is faster than trapping. This conflicts with the no ring-opened product formation by MMO, suggesting the nonradical formation in the MMOcatalyzed C-H functionalization reaction.205 The free radical process in the MMO alkane functionalization is also suggested by the formation of the partially racemized products from optically active (S)- or ethanes.206 Studies on the complex with the aqua ligand were extended to the reaction in aqueous solutions.207, 208 Complexes, (L = TPA or BPIA) and were prepared in situ using and used to the reaction of soluble substrates, i.e., alcohols, at pH 4.2 in the presence of It was shown that the complexes catalyze the homolytic decomposition of TBHP and the generated and radicals initiate homolytic C-H bond abstraction from the alcohol substrate to predominantly provide the aldehyde/ketone product. Reactivity of dinuclear iron complexes with simple bidentate ligands such as 2,2’-bipyridine and 1,10-phenanthroline for oxidation of alkanes, toluene, dimethyl sulfide, trans-stilbene and adamantane was studied in under Ar in the presence of oxidants (TBHP, CHP or ). The effect of the number of the bridging ligand was found remarkable: monobridged >

4. Functional model oxygenations by nonheme iron complexes

185

dibridged >> tribridged (L: bidentate nitrogen ligand, X: potentially labile ligand).209-214 It is suggested that the first step of the reaction is the binding of the oxidant to the iron catalyst probably resulting in the cleavage of the dinuclear unit into monomers as a real active species. The preparation of highly stable, easy to handle, iron complexes with highly labile sites is thought to bring about the advantage of the dinuclear structure over the mononuclear one. As for the cleavage of the O-O bond within the iron-peroxo complex, both of the homolytic (with CHP) and heterolytic (with TBHP or cleavages are proposed. Reactivity of dinuclear iron complexes having tripodal ligands, represented by was investigated in the presence of TBHP under Ar215, 216 with varying the tripodal and the bridging ligands. It was shown that the ratio of (alcohol + ketone)/t-BuOO-adduct) increased as the ligands become more electron donating. Two mechanisms for the decomposition of TBHP have been proposed: a homolytic pathway is initiated by generation of and radicals that form and a heterolytic pathway is initiated by the dissociation of the bridging anion from one iron center to provide a site for the coordinating the alkyl peroxide ion. The latter metal-based mechanism involving an alkylperoxyiron(III) intermediate was supported by the selective alcohol formation 217 and the large kinetic isotope effect value 2-Methyl-1-phenyl-2-propyl hydroperoxide (MPPH) was used as a probe capable of distinguishing between free alkoxyl radical chemistry and radical-free (enzyme mimetic) chemistry.218-220 Miyake et al. studied the reaction of with MPPH and showed that MPPH breaks down by O-O bond homolysis, leading to the formation of the benzyl radical and a high-valent species (Scheme 3).221 The latter oxidizes exogenous substrates such as thioanisole, cyclohexanol, and cyclooctene.

186

T.Funabiki

On the other hand, MacFaul et al. have demonstrated that the hydroxylation of cycloalkanes by involves freely diffusing 222 cycloalkyl radical (Scheme 4). The radical mechanism proposed by MacFaul et al. 220, 222 was also applied to selective oxidation of cyclohexane to cyclohexanol catalyzed by a diiron(II) complex and TBHP (L = 1,4,10,13-tetrakis(2-pyridyl)methyl-1,4,10,13-tetraaz-7,16-dioxacyclo-octadecane). 223

Hydrogen peroxide is no doubt a useful oxygen donor and decomposes to water, but it is usually treated as an aqueous reagent. Thus the effect of water should be considered. Functionalization of hydrocarbons with and using as the oxidant, was studied by Fish et al..224 The results obtained in the oxidation of cyclohexane was consistent with a free-radical chain mechanism in which an initially formed cyclohexyl radical is trapped by oxygen gas to give a cyclohexyl peroxy radical, which abstracts a hydrogen atom to give cyclohexyl peroxide. The selective abstraction of the tertiary hydrogen of adamantane was shown by the high values. Participation of hydroxyl radicals in the oxidation of toluene was also suggested. Similar reactivity was observed by the complex 212, 225 It is also shown that aromatic hydroxylation of the ligand is performed by with a diferric complex prepared in situ (L = N,N’-bis-(2,4,5-trimethoxybenzyl)ethylenediamine N,N’diacetic acid).213, 226 Meckmouche et al. have pointed out that the oxygenation of hydrocarbons utilizing hydrogen peroxide proceeds partly by the metal-based mechanism. 227 Using a diiron(III) complex with a chiral ligand, (pb = 4,5(-)pinene-2,2'-bipyridine), enantio-selective catalytic hydroxylation by hydrogen peroxide was first achieved. White et al. performed an efficient and selective epoxidation of decene and other alkenes with a mononuclear iron complex

4. Functional model oxygenations by nonheme iron complexes

187

(mep : N , N '-dimethyl-N,N'-bis(2-pyridylmethyl)ethane, 1,2-diamine) in the presence of and acetic acid.228 Formation of the carboxylate-bridged diiron(III) as an active species was indicated.

4.1.2 Monooxygenation by Diiron Complexes with Molecular Oxygen Very little has been reported on the monooxygenation with in the presence or absence of reductants. Kitajima et al. reported the monooxygenation with by a dinuclear iron complex in the presence of and Zn powder229 and with Hfacac and Zn powder230 The latter exhibited the greater activity than the former for the dioxygen hydroxylation of alkanes (npentane, cyclohexane, adamantane) and arenes (benzene, toluene, chlorobenzene). Wang et al. have synthesized dinuclear Fe(II) macrocyclic complexes with two dinucleating ligands, [24]RBPyBc and [30]RBBPyBc, containing phenolate pyridine, bipyridine, and amino phenolate groups.231 The complexes react with molecular oxygen and catalyze the monooxygenation of cyclohexane and adamantane in the presence of a two-electron donor, Alcohols are the main product rather than ketones (CyOH/CyO = 1.06 - 1.94). The proposed mechanism involves the formation of a peroxo-bridged dinuclear Fe(III) complex that is converted to a high-valent iron species bridged by an group as a direct intermediate for the reaction with hydrocarbons.

4.2 Monooxygenation by

Diiron Complexes

Dioxygen diiron complexes have been synthesized by mimicking the probable intermediate structures in methane monooxygenase oxygenations. Examples are summarized in the recent review.232-235 One is the diamond core type and the other is the peroxo type. Recently, an unusual species has been suggested in the reaction of with carboxylatebridged diiron(II,II) paddlewheel complexes, but its role in the oxygenation has not been clarified.234, 236

188

T.Funabiki

The type was proposed to correspond to the high-valent species in nonheme diiron enzymes.237, 238 Few examples of this type of complexes, however, are known and characterized by crystallographic analysis.239-241 The reaction between and forms quantitatively that is converted to In addition to the diiron(III) complex, a series of metastable complex with the diamond core was synthesized with using the tetradentate tripodal ligand (TPA and its methylated derivatives).237, 238, 243-245 Reactivity of the TPA complex demonstrated that a species could carry out oxidation reactions of hydrocarbons.246 Cumene is converted to cumyl alcohol and and ethylbenzene to 1-phenylethanol and styrene, but cycloheptane is not oxidized. Two-step mechanism has been proposed to explain the product distribution: formation of alkyl radical in the first and ratedetermining step by the abstraction of a hydrogen atom from the substrate followed by the reaction of the intermediate radical either with the diamond core species or Crystal structure of the high-valent complex with a diamond core was characterized with Resonance Raman evidence for an structure derived from the isomerization of the diamond core was obtained with the complex (Fig. 13).247 Recently, formation of the diamond core species from and TBHP has been shown.248

On the other hand, the reactivity of peroxo complexes toward a variety of substrates was investigated by LeCloux et al..249, 250 The nucleophilic/ basic complex (L = m-xylenediamine bis(Kemp’s triacid imide dinuleating ligand system) reacts with phenols and carboxylic acids to liberate hydrogen peroxide, but not with electrophilic reagents such as olefins or triphenylphosphine, or even with a weak hydrogen donors such as dimethyl benzylamine at -77 °C. At room temperature, the complex reacts with solvents such as THF, toluene, and cyclopentane, to form mixtures of alcohol and ketone products by a radical autoxidation pathway. The complex with

189

4. Functional model oxygenations by nonheme iron complexes less electron-rich pyridine analogue, inactive towards these substrates.

was found

4.3 Functional Model Oxygenations by Iron Species in the Polyoxometalate and Heterogeneous Matrix 4.3.1 Oxygenation by Iron Species in the Homogeneous System In the aim to overcome the catalytic instability towards activated oxygen species (particularly organic ligands), the use of the iron species without organic ligands has been tried for the catalytic system. One of these systems was developed by Zhang et al. who used and p-cyano-N,N-dimethylaniline N-oxide 251 as an oxygen source and by Mizuno et al. who used polyoxometalates, specially di-iron-substituted silicotungstate, (Fig. 14) and as 252-254 an oxygen source.

Homogeneous oxidation reactions were carried out in the presence of 30 % in acetonitrile under Ar. It was shown that the efficiency of utilization to oxygenated products reached to ca. 100% for the oxygenation of cyclohexane. The catalytic system was applied to other alkanes such as n-hexane, n-pentane, and adamantane253 and alkenes such as cyclooctene, 2octene, 1-octene, cyclohexene, styrene and trans-stilbene.254 It is noteworthy that epoxides are selectively (or mainly) formed in the oxygenation of alkenes. Nonradical process is suggested by the fairly stereospecific epoxidation of cis-stilbene. Mizuno et al. found that a water soluble potassium salt of di-iron-substituted silicotungstate, catalyzes the conversion of methanol to methylformate (10.2%) > methanol (0.3%) > formic acid (0.2%) together with (16.6%).255 The work was attempted to demonstrate catalytic oxidation of methane with hydrogen peroxide in water. Interestingly, mono- and tri-iron-substituted silicotungstates is much less active and nonselective for the oxidation. The catalyst

190

T.Funabiki

was used for the oxygenation of cyclohexane and adamantane under in as solvent without addition of reducing reagent or radical initiators.256 Alcohols and ketones were obtained catalytically by the radical-chain path. High turnover numbers were achieved in the selective epoxidation of alkenes with 1 atm in 1,2-dichloroethane/acetonitrile at 356 K. 257 Cyclooctene was converted to cyclooctene oxide (98%) under 82% converstion and 10000 TON (Table 3). Addition of 4-tert-butylcatechol as an alkyl radical inhibitor did not affect the reaction. IR spectra using and indicated the Fe=O band.

4.3.2 Oxygenation by Iron Species in the Heterogeneous System (L = TPA) was applied to the reaction in the The complex was embedded in an amorphous silicate heterogeneous system. surface modified by a combination of hydrophilic poly(ethylene oxide) and hydrophobic poly(propylene oxide). This assembly showed reactivity somewhat higher in comparison to an aqueous micelle system utilizing a surfactant. Functionalization of alkane and alkene in the biphasic catalytic system was performed in the different way by using perfluoroheptane-soluble catalysts.259 Immobilization of iron(III) on to the polymer matrix of cross-linked styrenedivinylbenzene matrix resulted in the epoxidation of cis-cyclooctene and styrene in the presence of TBHP.260 Likewise, oxygenation of cyclohexane was performed by dinuclear iron complexes and in hexagonal mesoporous solid (HPTB = N,N,N',N'-tetrakis(2-benzimidazolylmethy)-2-hydroxy1,3-diaminopropane, HPTP = N,N,N',N'-tetrakis(2-pyridylmethy)-2-hydroxy-l,3261 diaminopropane) in the presence of TBHP or Oxygenations of methane and benzene are thought to proceed by the reaction with that is formed on FeZSM-5 by the loading.262,263 258

4. Functional model oxygenations by nonheme iron complexes

191

4.4 Functional Model Oxygenations by Mono-Iron Species 4.4.1 Oxygenation by Mono-Iron Complex/Activated Oxygen System Since Fenton first used a mixture of and hydrogen peroxide in 1894,264 the combination of the chemicals has been used as the powerful oxidizing reagents. However, the mechanism and the key intermediates in the Fenton chemistry are still under great discussions.265, 266 The chemistry is not so simple as explained by a free hydroxy radical as proposed by Haber and Weiss in 1930s.267, 268 Mechanism of the catalytic decomposition of hydrogen peroxide by nonheme Fe(II) complexes was studied in comparison with that of the heme system involving an oxoiron(IV) or ferryl species (Scheme 5) or that of a Fenton-like nonheme system involving the free hydroxyl radical (Scheme 6).

192

T.Funabiki

The mechanism without formation of the hydroxyl radical was supported by the reactions of with ferrous ions complexed with diethylenetriamineN,N',N",N"-pentaacetate (DTPA),270 with nitriiotriacetate (NTA) or ethylenediamine-N,N'-diacetate (EDDA), 271 and with 8-methyl-1,4-bis(2pyridylmethyl)-l,4,8-triazacycloundecane and 1-methyl-5,9bis(2-pyridylmethyl)-1,5,9-triazacyclododecane Zhang et al. performed the reaction at neutral pH and scavenged a reactive intermediate by 2,2’azinobis(3-ethylbenzothiazoline-6-sulfonate) (ABTS), but not by bromide ion, indicating that the strongly oxidative intermediate is not the hydroxyl radical. Proposed mechanisms for the reactions of and complexes are shown in Scheme 7. This system was applied to hydroxylation of aromatic compounds: Zhang et al.269 and Dunforld283 has summarized the current state of mechanisms in Scheme 7.

In the biomimetic monooxygenations of alkane and alkene, Yamamoto and Kimura reported trans-epoxide formation from olefins with in 284 acetonitrile. Sugimoto and Sawyer performed epoxidation of and monooxygenation of some organic substances by a system. 285, 286 They found further efficient and selective epoxidation of various alkenes and monooxygenation of organic substrates including alkanes by a system in acetonitrile.287-291 The most effective catalyst systems were and or was proposed as the reactive intermediates, which may

4. Functional model oxygenations by nonheme iron complexes

193

be transformed to in the presence of olefins.290 Non free radical process was also proposed for aromatic hydroxylation by Fe11 complexes The oxygenation by this system, i.e. the oxygenated Fenton chemistry, was summarized by Sawyer.293,294 Barton et al. studied the mechanism of alkane functionalization of alkanes (cyclohexane, adamantane, 3-ethylpentane, cyclododecane, etc.) by experiments of the catalytic oxygenation systems using and as iron species and and Geletii and Shilov used and for the oxidation of cyclohexane in the presence of pyridine 301 and compared the ketone/alcohol ratio with the Gif system.302 Barton et al. proposed the intermediate formation of the iron-carbon species and a as the reactive species295-300 and formation of alkyl hydroperoxide as an intermediate. 303, 304 The formation of an intermediate containing an Fe-C bond was further supported by the efficient capture of the bond by iodine and iodide ion to give the corresponding alkyl iodide305 or by PhSeH.306 Using completely regiospecific oxidation of a methylene group of Binor S into a ketone was accomplished.307 Non radical mechanism in this system via an iron-alkyl species was suggested on the bromination of saturated hydrocarbons.308 Using and Barton et al. indicated that the oxygen atoms in the alkyl hydroperoxide do not arise directly from hydrogen peroxide but from formed in situ by the well-known iron(III)catalyzed decomposition of hydrogen peroxide.309 The radical trapping experiment with Tempo (2,2,6.6-tetramethyl-l-piperidinyloxy) also supported for the nonradical nature of the first intermediate in the reaction.309, 310 As an active species, formation of a complex from an isolated complex was 311 proposed. The carboxylate ligands such as picolinate are thought important for the formation of a species as a key intermediate in the hydrocarbon 312 activation chemistry. Important role of the pyridine bases in the Fe(III) complexes for the efficient ketonization within the manifold 313 and of a suitable carboxylic acid as ligand within the manifold was reported.314 In the studies on the role of pyridine, the use of 4-tert-butylpyridine was found to permit the isolation of cyclohexanone and cyclohexanol by simple distillation.315 As an extended application of Gif systems, Barton et al. have developed the carboxylation of saturated hydrocarbons by in 316 pyridine-acetic acid or by These results in Gif Chemistry have been reviewed by Barton. 318, 319 Trapper et al.320 have reinvestigated a typical Gif system to lend support to the proposition of a preponderant, carbon- and oxygen-centered radical pathway rather than the nonradical pathway by Barton and Doller.321 The monomeric

194 complex

T.Funabiki

which was formed by dissolving or in pyridine, produced not only oxo products, but also 2- and 4-adamantylpyridines by the reaction with The formation of adamantylpyridines not only for the tertadamantyl position, but also for the sec-adamantyl sites, especially under Ar is thought to provide direct evidence for the generation of tert- and sec-adamantyl radicals.322 The same mechanistic aspect was obtained by using iron picolinate, It is suggested that the hydroxyl radical plays a key role in Gif oxygenation by (Ar), most likely coupled to substrate-centered alkoxyl radicals under The oxygen-centered radicals perform H-atom abstractions from Gif substrates to generate diffusively free carbon-centered radicals. Liu et al. also studied the hydroxylation of phenol by iron(II) phenanthroline, with Products are catechol > hydroquinone > p-benzoquinone and the efficiency of is 45-90% depending on the reaction condition. It is suggested that the coexistence of and is the key for the hydroxylation by the attack of the hydroxyl radical to phenol. Role of the hydroxyl radical as reactive oxygen species was also suggested in the oxidation of dichloroacetonitrile to phosgene.325 However, the non-radical process in the system was supported by the following results. Kim et al. reported stereospecific alkane hydroxylation with catalyzed by and an intermediate formation of was shown by electrospray ionization mass spectrometry (ESIMS).326 Chen and Que used In the oxidation of cyclohexane, the major product was cyclohexanol (TN: 5.6 within 0.5 h) compared with cyclohexanone (TN: 0.7). The system catalyzed the stereospecific hydroxylation of cis-1,2-dimethylcyclohexane, that supported formation of via (N4Py = N, N-bis(2-pyridylmethyl)-Nbis(2-pyridyl)methylamine) was also used for oxidation of alkanes such as cyclohexane, cyclohexene, adamantane, benzene to alcohols and ketones in up to 31% yield.328 It was proposed that that was detected as a transient purple species, reacts through homolysis of the O-O bond to afford two reactive radical species, and HO·. Two pathways of alkane hydroxylation were suggested in the system329: (A)when the TPA ligand has two or three a generated high-spin directly cleaves the C-H bond, forming the alkyl radicals, that are susceptible to radical epimerization and trapping by (B) when the TPA ligand has substituents, a generated low-spin intermediate derives an species by the heterolytic splitting of the O-O bond. The high-spin system

4. Functional model oxygenations by nonheme iron complexes

195

strongly favors olefin cis-dihydroxylation in which both diol oxygen atoms derive from (Scheme 8).187,330 Asymmetric cis-dihydroxylation of olefins such as cyclooctene, cis-2-heptene and trans-2-octene was first performed by using and 82% ee has been 187 attained in the dihydroxylation of trans-2-octene. The low-spin system carries out highly stereoselective alkane hydroxylation, olefin epoxidation and olefin cis-dihyroxylation. 188, 190, 191, 330 It was shown that steric effects of 6-Methyl substituents on the ligands affect the spin state of the intermediate, which in turn influences the course of the oxidations.

The nonheme iron/peroxide system attracts attention in recent years for the potential use in the catalytic oxidation of environmental pollutants. Oxidative degradation of 2,4,6-trichlorophenol (TCP), a major pollutant produced by paper mills, was efficiently catalyzed by water-soluble iron tetrasulfophthalocyanine with to benzoquinone and ring cleavage products, mainly chloromaleic acid and carbon dioxide.331-334 Hemmert et al. obtained a nonheme iron complex using a new symmetric pentadentate ligand containing four pyridine groups, bis(di-2-pyridylmethyl)amine (BDPMA). The complex was found to catalyze the oxidative degradation of chlorinated phenols (TCP, 2,4dichorophenol (DCP), p-chlorophenol to corresponding benzoquinones with potassium monopersulfate as oxidant in a mixture of acetonitrile-water

196

T.Funabiki

(1/3) at pH 2.335 The turnover rates were 27 (TCP), 22 (DCP) and 23 (pCP) cycles within 30 min. The proposed mechanism involves the iron(IV)-oxo active species, that abstracts one electron and one proton from the phenolic substrate to form a species and a phenolic radical as an intermediate. The formation of the ferryl species, in the initial step was supported by the MD simulation study.336 Horwitz et al. developed nonheme iron complexes with tetraamido macrocyclic ligands (TAML) that are efficient for the dye bleaching reactions with in water from neutral to basic pH. 337 338 Pinacyanol chloride as a reference, azo and quinone dyes were oxidatively degraded to show the bleaching effect. The Fe-TAML catalysts (Fig. 15) were applied to destroy pollutants pentachlorophenol (PCP) and TCPin water, in minutes, under ambient conditions of temperature and pressure.339 Substitution on the benzene ring with hydrophobic tail, produces micelle-forming catalysts. Substitution with the anionic group, increases the negative charge of the catalysts, that is effective to introduce a rate disincentive for the catalyst to attack substrates on negatively charged heterogeneous supports.340 Collins explains characteristics of the TAML activator/peroxide systems as follows: capability of more than 10000 turnovers per hour in certain applicants, weak catalase activity with highly efficient utility of peroxide, water-soluble catalysts that are easy to use and active under both neutral and basic conditions, rapid processes at mild temperatures and under ambient pressure.340 Bleaching of kraft pulp and color in plant effluents were reported as direct or potential examples of environmental applicability.

4. Functional model oxygenations by nonheme iron complexes

197

The selective oxygenation of methane and light alkanes by the Fenton system was performed in the three-phase catalytic membrane reactor, enabling simultaneous reaction and product separation. Frusteri et al. reported that Nafion-based catalytic membranes catalyze the selective oxidation of methane to methanol,341 ethane to ethanol and acetaldehyde,342 and propane to acetone, propionic aldehyde, isopropanol, and n-propanol.343 The partial oxidation proceeds according to a radical mechanism which involves the activation of paraffin on superacid sites and the subsequent reaction of activated paraffins with OH radicals generated according to the Fenton reaction. Bianchi et al. developed a biphasic system (water/acetonitrile) for the selective oxidation of benzene to phenol, with the pyrazinecarboxylic acid derivatives as a ligand to in the presence of trifluoroacetic acid.344 Tert -butyl hydroperoxide (TBHP) is no doubt an important and useful oxidant for monooxygenation of various hydrocarbons. Barton et al. developed oxidation systems which oxidizes cyclohexane and various hydrocarbons by TBHP with Fe(III) salts such as or a cyclohexane soluble iron complex, (TMA: trimethylacetate) in the absence or presence of picolinic acid.345347 A pathway, alkane alkyl hydroperoxide ketone or alcohol, was proposed. Schuchard et al. extended the reaction system to that under 15 bar of oxygen at 70 °C and found the increase in the conversion from 5% to 9% but the decrease in the selectivity.348 Oxidations of cyclohexadienes to aromatic products, anthracene to anthraquinone,349 and alcohols and allylic methylene groups to ketones even in the absence of pyridine/acetic acid350, 351 were reported. Nguyen et al. used a model ligand that mimics the metal-binding domain of the bleomycin for oxidation of cyclohexane to cyclohexanol and cyclohexanone (1 : 1 ratio, 1700% yield on the basis of catalyst concentration) and adamantane and 3-methylpentane. No oxidation of an aromatic C-H bond was observed.352 Low-spin species and were detected by ESR in the reaction of with TBHP and respectively. Formation of a perferryl intermediate from the Fe(III)-peroxo species were thought unlikely. The homolytic cleavage of the O-O bond in to form was thought to take part in the catalytic process by forming cyclohexyl and t-butoxy radicals. Spectroscopic properties and electronic structure, and reactivity for the homolysis of the O-O bond of low-spin (x = 1 or 2) have been studied in detail.353, 354 The role of and manifolds in the oxidation systems were studied by ionic trapping with chloride, azide, and other anions.351, 355, 356 However, since Minisci et al. have established that TBHP chemistry was best interpreted as

198

T.Funabiki

radical,357-360 Barton et al. performed the radical trapping experiments 361-364 and led the conclusion that the results obtained in the system of or can be explained by radical chemistry and oxidation states of iron higher than are not involved.365 Oxygenation of nitroalkanes to the corresponding aldehydes or ketones in relevance to 2-nitropropane dioxygenase was performed by ferroxime(II) dimethylglyoxime)/TBHP in DMF or and both radical species are thought to be involved in the H-abstraction from nitroalkanes. Barton et al. used bis(trimethylsilyl)peroxide for the selective oxidation of alkanes to ketones using and The intermediate formation of which reacts with a methylenic carbon to form bond. By using this oxidant, it was found that unlike the system, no extra ligand was needed if was used.368 A carbon radical from the manifold was proposed to explain the results. Shyu et al. used sodium hypochlorite (NaOCl) for the epoxidation of cisstilbene with iron complexes with salicyladimine ligands, and N,N'-(1,1-dimethylethylene)bis(salicylaldimine), = N,N'-(1,1-dimethylethylene)bis(3-methoxysalicylaldimine). 369 The reaction proceeded readily to produce trans- and cis-epoxides in good yield: 40% yield with cis : trans = 20 : 80, 90% yield with cis : trans = 40 : 60.

4.4.2 Oxygenation by Mono-Iron Complexes/O2 System Monooxygenases catalyze incorporation of one atom of oxygen from molecular oxygen into a substrate accompanied with the formation of water. In this reaction the presence of electron and proton donors is required for activation of molecular oxygen (eq. 9). Thus, selections of electron and proton donors as well as iron complexes are key points for development of functional model Oxygenation systems. There are some results, however, reporting monooxygenation in the absence of electron and proton donors. These systems may not be regarded as functional model systems, but worthy to notify their reactivity. Usually, mechanisms are obscure in these systems. Shue et al. reported ketonization of methylenic carbons such as cyclohexane by (DPAH: 2,6-dicarboxylatopyridine)/ and the dioxygenation of acetylenes, aryl olefins, and catechols as a reaction mimics for dioxygenase.370 The simple system was used for Oxygenation of cyclohexene and methyl linoleate. 371 Catalyst complexes were prepared in situ by mixing or in MeCN with 2,2'-bipyridine, and was found the most reactive. Oxygenation of cyclohexene was performed by complexes in aqueous

4. Functional model oxygenations by nonheme iron complexes

199

solutions.372 is suggested to be reactive with cyclohexene to produce its alcohol in the absence of oxygen. When is present, the oxygen atom transfer proceeds via which is formed from Hirao et. al. developed an efficient system for the catalytic epoxidation reaction of olefins stilbene) with molecular oxygen by the utilization of and the N-heterocyclic podand ligand, BIPA.373, 374 The reaction proceeds at room temperature and in the absence of a co-reductant. The coexistence of 4-ethoxycarbonyl-3-methyl-2-cyclohexen-l-one results in the facile epoxidation. The intermediate was assumed to be involved, but no direct support for the mechanism has been obtained. Iron(III)chloro complexes with salen ligands bearing electronegative substituents were used for the radical chain autoxidation of cyclohexane under 1 atm at 25 °C. About 2/3 of the activity of the best iron porphyrin catalyst was achieved.375 Duprat et al. used Fe(0) (iron rod) for the aromatic hydroxylation in acetic acid.376 Formation of that may be converted to as an active species, was postulated. Most functional model monooxygenations by nonheme iron complexes are performed in the presence of electron and proton donors. Udenfriend used ascorbic acid for monooxygenation with ferrous salts and molecular oxygen in the presence of EDTA as a ligand at neutral pH for aromatic hydroxylation.377, 378 The system was applied to oxygenations of salicylate,67 cyclohexane and cyclohexene,379, 380 naphthalene381 etc. As shown in following equations, ascorbic acid reactivates the catalyst. The active species may or may not involve a free hydroxyl radical. The reaction was improved in the presence of metallic iron which is 382 effective for elimination of the by-product oxalic acid.

Ullrich hydroxylated cyclohexane to cyclohexanol and cyclohexanone (6.9 : 2.8), toluene, acetoanilide, etc., by system in an aqueous media and compared the reactivity with cytochrome P-450.383 This was applied to the hydroxylation of naphthalene. 381 2-Mercaptobenzoic acid may act not only as a sulfur coordinating ligand but a reductant.381 Other thiols were used for hydroxylation of aniline and toluidine in aqueous acetone.384-386 Barton et al. have developed a new procedure for the oxidation of saturated hydrocarbons, e.g. adamantane, cyclohexane, pentane, etc. by using iron powder, carboxylic acids, and hydrogen sulfide in aqueous pyridine under 1 atm and compared the 302, 387 selectivity in the prim-, sec-, tert-C-H bonds. Usefulness of as a reductant

200

T.Funabiki

in the synergistic oxidation of cyclohexane388 and probability of the non-radical oxidation were reported.389 Barton et al. used zinc (Zn) as a reductant. They isolated a trinuclear iron complex, from a reaction system containing iron powder, carboxylic acids, and in aqueous pyridine under 1 atm and found that it oxidized alkanes in the presence of Zn, pyridine, and acetic acid under 1 atm (Gif system). Later, the Gif type oxygenations were performed mostly by using monoiron salts, e.g. Barton et al reported a number of experiments for application of this system to the more complicated compounds such as steroids391-398 and for clarification of mechanism, specially participation of the radical and nonradical processes.295, 296, 399-402 Intermediate formation of alkoxide403 and was proposed based on the effect of on the selectivity, cleavage of methylidene olefins into the ketone or aldehyde, or effect of additives such as thiophenols.300 Recently, Celenligil-Cetin et al. studied the reactivity of ferrous and ferric oxo/peroxo pivalate complexes and claimed that the peroxo species is not directly involved in catalytic Gif-type oxygenation, based on the low reactivity of the peroxo containing complexes in the oxidation of both cisstilbene and adamantane.405 Kitajima et al. used for oxygenation of alkanes and arenes in the presence of hexafluoroacetylacetone (Hhfacac) and Zn powder hydrotris-1-pyrazolylborate).229, 230 was isolated and assumed as an active complex. Balavoine et al. tried to use an electrode in place of zinc in the Gif system for the selective oxidation of saturated hydrocarbons.295, 406-408 Mimoun and Roch used hydrazobenzene (PhNHNHPh) for oxygenation of cyclohexane, cyclohexene, and toluene.409 The most active complex was formed from and carboxylic acid in the presence of hydrazobenzene. Davis et al. used and for the same reaction and proposed the hydroperoxide complex as an active species.410 Sheu et al. used complexes (PA: picolinato) for the monooxygenation of saturated hydrocarbons, especially ketonization of methylenic carbons. 370,411 This system was applied to the hydroxylation of aromatic hydrocarbons as reaction mimic for tyrosine hydroxylase.412 With phenol as a reactant, the dominant product was catechol. Funabiki et al. developed an monooxygenation system using hydroquinones with catecholatoiron(III) complexes in acetonitrile in the presence of pyridine. 413 First the system was applied to hydroxylation of aromatics.413-416 Catalytic activity depends greatly on the substituents on hydroquinones (R-HQ) and increases with the electron-donating property; In the oxygenation of

4. Functional model oxygenations by nonheme iron complexes

201

anisole, selectivity to form either 4-MeOPhOH or PhOH depends greatly on the concentration of pyridine: selective formation of 4-MeOPhOH and PhOH at low and high concentrations of pyridine, respectively. In the oxygenations of p-D-toluene and p-xylene, NIH shifts of p-D and p-Me groups were observed in the fairly high values (55 and 35 %, respectively).415 The oxygenation of the methyl group to form benzyl alcohols and aldehydes was promoted by the high pyridine concentration. The system was used for the hydroxylation of tert-butylphenol to form selectively 4tert-butylcatechol as a model reaction for tyrosine hydroxylase.416 Alkane and alkene were also monooxygenated by the same catalytic system.417,418 Cyclohexane, cyclohexene, and n- and isoalkanes produced highly selectively alcohols at the low pyridine concentration (Table 4).417,418 This is very characteristic and different from other systems that produce ketones selectively, and supports the high-valent ironoxo species as an active species. The formation of ketone was promoted with the increasing pyridine concentration and became dominant in the pyridine solution. The effect of the pyridine concentration is due to the change of iron complexes from a catecholate iron complex to a pyridine iron complex, as supported by the same results with using Not only catalytic activity but also selectivity was greatly dependent on the substituent on hydroquinones. This suggests that coordination of hydroquinones to iron is involved in the oxygenation process416 417,4I8

Takai et al. developed an excellent catalytic system for epoxidation of olefins by tris(l,3-diketonato)iron(III) complexes and aldehydes as a reductant.419 A number of olefinic compounds including styrene analogues and olefinic alcohols were converted to corresponding epoxides in good to quantitative yields with combined use of molecular oxygen and an aldehyde (e.g. 2-ethylbutyraldehyde) at

202

T.Funabiki

room temperature; e.g. styrene was converted to epoxide (89%) and aldehyde (10%) within 9 h in the 100% conversion. Murahashi et al. also used an aldehyde (heptanal) for aerobic oxidation of alkanes and alkylated arenes.420 Iron powder/acetic acid was used as the most effective catalyst, though and were also examined. Oxo-iron species, which may be formed via peracids from acetic acid, was assumed as an active species for hydrogen abstraction from alkanes and epoxidation of olefins. Kesavan et al.421 used nanostructured amorphous iron and alloy like for the oxidation of adamantane under 40 atm and at 28 °C in the presence of isobutyraldehyde and acetic acid. Only alcohols and ketone (l-ol : 2-ol : 2 -one = 17 : 2 : 1) were obtained in the 52 % conversion. Oxo-iron species was also assumed to explain the high selectivity for the tertiary alcohol. Ruiz et al. have used an iron(III)-carbonato complex of orthophenylenebis(oxamato) as a moderately efficient catalyst for the aerobic epoxidation of alkenes with co-oxidation of pivalaldehyde. 422 Characteristic results are the stereodependent epoxidation of cis- and trans-stilbene (conversion rate is trans >> cis) and the rate enhancement by the addition of N-methylimidazole. Iron(IV)acylperoxo or iron(v)-oxo species derived by oxygen-oxygen bond cleavage of the acylperoxo group was assumed as a probable candidate of active species. In this connection, Nam et al. demonstrated the formation of a metal-acylperoxo complex in the auto-oxidation of aldehyde.423

4.4.3 Mono-Iron Oxygen Species In the mechanism of cytochrome P450, it is assumed that the high-valent iron-oxo species is formed from that may derive from the oneelectron reduction of oxy adducts, probably via an It is well assumed that these types of iron-oxygen species are involved in the mechanisms of nonheme iron oxygenases. As an activated bleomycin (BLM), formation of a low spin by the reaction of with or by with was shown kinetically and by ESI-MS.424, 425 As a synthetic high spin species, has been identified by resonance Raman and other spectroscopies.426,427 In recent years, formation of and its deprotonated form with different ligands have appeared. Kim et al. detected species of and by ESI-MS in the co-injection solution of , and TBHP.428 Bernal et al. detected a low-spin species by ESI-MS and ESR in the solution of and (L:N-methyl-N,N',N'tris(2-pyridylmethyl)ethane-l,2-diamine).429 Nguyen et al. detected a low-spin species and by ESR in the reaction of

4. Functional model oxygenations by nonheme iron complexes

203

with TBHP and respectively, in acetonitrile/DMSO.352 Formation of a perferryl intermediate from the Fe(III)-peroxo species were thought unlikely, since the PMA- ligand framework is incapable of stabilizing the high-valent iron center and oxidation of both cyclohexene and norbornene yielded only the allylic oxidation products and the exo-epoxide, respectively. The homolytic cleavage of the O-O bond in to form was thought to take part in the catalytic process by forming cyclohexyl and t-butoxy radicals. deVries et al. used a ligand 2,6-bis[methoxybis(2pyridyl)methyl]pyridine (L) and observed by ESM the transient formation of a from Kim et al. proposed the structure and later for an intermediate detected in the solution of and It was shown by Raman spectroscopy that the O-O bond is significantly weakened as indicated by the lowest 790 Roelfes et al. synthesized pentadentate ligand N4Py with which the purple low-spin was generated by the reaction with and Fe(II) complexes.432 Mialane et al. also confirmed the formation of with L = N,N,N’-tris(2-pyridylmethyl)-N’-methyl-ethane-1,2diamine) by ESI-MS.433 Formation of high-spin blue species from low-spin complexes was reported by Jensen et al..49 They found that purple complex (bztpen: N-benzyl-N,N’,N’-tris(2-pyridylmethyl)-ethane-1,2diamine) can be reversibly deprotonated to give transient blue species showing spectroscopic properties consistent with iron(III)-peroxide complexes, or others. Similarly, Simaan et al. reported the formation of a complex from a purple low-spin complex upon adding a base (L: N-methyl-N,N,N-tris(2-pyridylmethy])ethane-1,2-diamine).50 The complex was characterized by the UV/Vis change from 534 to 740 nm, by the ESR change from g = 7.5 and 5.9 to 9.3 and 4.3, and by ESIMS at m/z = 435. Ho et al. gave a resonance Raman evidence for the interconversion between and 632 and 790 for -1 and 495 and 817 cm for The former has activity for the hydroxylation of cyclohexane, but the latter looses it, supporting the hypothesis that protonation of the peroxo species activates it for participation in the oxygen activation mechanisms by iron enzymes. Similar results for identification of two peroxo complexesas shown in Fig. 16 were reported by Simaan et al..51,52

204

T.Funabiki

As for the formation of high-valent oxo species of nonheme iron complexes, Lange et al. reported about evidence for a nonheme Fe(IV)=O species.434 No direct spectroscopic evidence for the species has been obtained, but the intramolecular hydroxylation of a ligand phenyl moiety was explained by its involvement in the mechanism. This supports the hypothesis that Fe(IV)=O species can be the active species responsible for substrate oxidation in the class of oxygen-activating nonheme iron enzymes. Wada et al. have isolated and characterized species from the reaction of with Reactions of the species with cyclohexene and cyclohexane in the stoichiometric oxidation gave alcohols as the main product with a trace amount of ketones. A higher efficiency was observed in oxidation of to rather than in the case of to indicating that the active species not only acts as a two-electron oxidant but also displays nucleophilicity. It is suggested that the homolytic O-O bond cleavage generates both of and HO, both of which cooperatively perform a C-H bond cleavage step with a subsequent C-O bond formation step in the hydroxylation.

5.

FROM FUNCTIONAL MODEL TO CATALYSIS

Goals for biomimetic chemistry on oxygenases are clarification of active species, e.g. Fe-OOH or Fe=O, involved in the enzymatic and model oxygenations, development of catalytic systems for the selective oxygenations using as the oxidant, and so on. Efforts for these topics in the last ten years have given rise to various types of nonheme iron complexes that stabilize iron-oxygen species and oxygenate substrates in the di- and mono-oxygenase-like fashion. The high-valent iron-oxo species became popular in the great development of cytochome P450 chemistry. These species was first thought characteristic to heme

4. Functional model oxygenations by nonheme iron complexes

205

enzymes and their models, which involve porphyrin ligands suitable for stabilization of the high-valent state of iron. It is easy to assume and expect formation of the similar iron oxygen species in the nonheme iron systems, but at the same time we realize high barriers for characterization of these species spectroscopically or by isolation. Participation of these species in the oxygenations by nonheme iron complexes was first shown by selectivity of product, e.g. alcohol/ketone ratio or C2/C3 ratio in the alkane or adamantane oxygenations, but it was not necessary for us to wait so long before getting spectroscopic evidence for various types of nonheme iron-oxygen species. Some iron-oxygen species were characterized after isolation. However, we encounter paradoxical facts that stabilized iron-oxygen species exhibit poor reactivity. Many kinds of diiron complexes have been synthesized in relevance to methane monooxygenase. In general, the activities of functional model oxygenations are poor and the complexes are labile to be converted to monoiron species in solutions. The characteristic diiron structure, however, were applied to the diiron-substituted Keggin-type silicotungstate (Fig. 14), that exhibits a high catalytic activity under without addition of reducing reagent. On the other hand, some types of monoiron complexes have been developed to perform enzyme-like oxygenations producing alcohols selectively. In addition, some monoiron complexes are expected to become useful in oxidation of pollutants or bleaching of dyes, regardless of usage of molecular oxygen or activated oxygen such as hydrogen peroxide. Recently, however, reactivity of peroxo adducts of mono- and diiron complexes were compared for the oxidation of sulfides to sulfoxides. The dinuclear catalyst was found to be more reactive and (enantio)selective than its mononuclear counterpart, suggesting that a second metal site affords specific advantages for stereoselective catalysis.436 This result may be helpful for the design of future enantioselective iron catalysts. Nonheme iron oxygenases are characteristic for the abundant types of dioxygenation, compared with heme oxygenases. Progresses in chemistry of catechol dioxygenases are remarkable in these ten years, but little development has been achieved in chemistry of other types of dioxygenases. The selective cleavage of aromatic rings with molecular oxygen is an attractive reaction that converts aromatic to aliphatics. The mechanism of the selectivity control in the oxygen insertion process is an interesting subject in bioinorganic chemistry. It involves chemistry of Fe(III) and Fe(II) for activation of catechols and oxygen. The fruits in that chemistry will be applied to other types of dioxygenases. Recent great progress in the X-ray crystallographic analysis brings about structural information about metalloproteins directly. However, structural model

206

T.Funabiki

investigations are no doubt important, especially for getting information about probable structures of unstable intermediates. Various types of iron-oxygen species isolated or detected contribute greatly to clarify mechanisms of oxygenations. On the other hand, functional model chemistry brings about more dynamic information about the mechanisms and lead to development of catalysis. Functional model investigations reported have been focused mostly on the mechanisms, but achievement of highly selective and highly active reactions with high TON is required for development of catalysts based on functional models. Molecular oxygen is more attractive as an oxygen source than activated species such as hydrogen peroxide. Water is also the more attractive solvent than organic solvents. The efficient catalytic systems of oxygenation using nonheme iron complexes are the challenges for the future.

4. Functional model oxygenations by nonheme iron complexes

207

6. REFERENCES 1. 2. 3. 4. 5. 6.

7.

8. 9. 10. 11. 12. 13.

14. 15. 16. 17. 18. 19. 20. 21.

22.

Holm, R. H.; Solomon, E. I. Chem. Rev. 1996, 96, 2237-3042. Que, L., Jr.; Ho, R. Y. N. Chem. Rev. 1996, 96, 2607-2624. Sono, M.; Roach, M. P.; Coulter, E. D.; Dawson, J. H. Chem. Rev. 1996, 96, 2861-2887. Funabiki, T. Oxygenases and Model Systems; Kluwer Academic Publishers: Dordrecht/Boston/London, 1997; Vol. 19, pp 1-393. Costas, M.; Chen, K.; Que, L., Jr. Coord. Chem. Rev. 2000, 200, 517-544. Watanabe, Y. In Oxygenases and Model Systems, Funabiki, T., Ed.; Kluwer Academic Publisher: Dordrecht/Boston/London, 1997; Vol. 19; pp 223-282. Watanabe, Y.; Fujii, H. In Metal-oxo and Metal-perozo Species in Catalytic Oxidations, Meunier, B., Ed.; Springer Verlag: Berlin, 2000; Vol. 97; pp 6289. Tabushi, I.; Koga, N. J. Am. Chem. Soc. 1979, 101, 6456-6457. Groves, J. T.; Haushalter, R. C.; Nakamura, M.; Nemo, T. E.; Evans, B. J. J. Am. Chem. Soc. 1981, 103, 2884-2886. Groves, J. T. J. Chem. Educ. 1985, 62, 928-931. DiNello, R. K.; Dolphin, D. H. J. Biol. Chem. 1981, 256, 6903-6912. Mandon, D.; Weiss, R.; Jayaraj, K.; Gold, A.; Terner, J.; Bill, E.; Trautwein, A. X. Inorg. Chem. 1992, 31, 4404-4409. Balch, A. L.; Latos-Grazynski, L.; Renner, M. W. J. Am. Chem. Soc. 1985, 107, 2983-2985. Nanthakumar, A.; Goff, H. M. Inorg. Chem. 1991, 30, 4460-4464. Gross, Z.; Nimri, S. Inorg. Chem. 1994, 33, 1731-1732. Fujii, H. J. Am. Chem. Soc. 1993, 115, 4641-4648. Hashimoto, S.; Tastsuno, Y.; Kitagawa, T. J. Am. Chem. Soc. 1987, 109, 8096-8097. Kincaid, J. R.; Schneider, A. J.; Paeng, K.-J. J. Am. Chem. Soc. 1989, 111, 735-737. Hashimoto, S.; Mizutani, Y.; Tatsuno, Y.; Kitagawa, T. J. Am. Chem. Soc. 1991, 113, 6542-6549. Kitagawa, T.; Mizutani, Y. Coord. Chem. Rev. 1994, 135/136, 685-735. Schultz, C. E.; Rutter, R.; Sage, J. T.; Debrunner, P. G.; Hager, L. P. Biochemistry 1984, 23, 4743-4754. Gold, A.; Jayaraj, K.; Doppelt, P.; Weiss, R.; Bill, E.; Ding, X.-Q.; Bominaar, E. L.; Trautwein, A. X.; Winkler, H. New J. Chem. 1989, 13, 169-172.

208

T.Funabiki

23. Esch, J. V.; Roks, M. F. M.; Nolte, R. J. M. J. Am. Chem. Soc. 1986, 108, 6093-6094. 24. Groves, J. T.; Neumann, R. J. Am. Chem. Soc. 1987, 109, 5045-5047. 25. Groves, J. T.; Neumann, R. J. Am. Chem. Soc. 1989, 111, 2900-2909. 26. Groves, J. T.; Solomon, B. U. J. Am. Chem. Soc. 1990, 112, 7796-7797. 27. van der Made, A. W.; Smeets, J. W. H.; Nolte, R. J. M.; Drenth, W. J. Chem. Soc., Chem. Commun. 1983, 1204-1206. 28. Campestrini, S.; Meunier, B. Inorg. Chem. 1992, 31, 1999-2006. 29. Meunier, B. Chem. Rev. 1992, 92, 1411-1456. 30. Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.; Beck, J. S. Nature 1992, 359, 710. 31. Monnier, A.; Schuth, F.; Huo, Q.; Kumer, D.; Margolese, P. I.; Maxwell, R. S.; Stucky, G. D.; Krishnamurthy, M.; Petroff, P.; Firouzi, A.; Janicke, M.; Chmelka, B. F. Science 1995, 267, 865. 32. Yang, H.; Kuperman, A.; Coombs, N.; Mamiche-Afara, S.; Ozin, G. A. Science 1996, 379, 703. 33. Liu, C.-J.; Li, S.-G.; Pang, W.-Q.; Che., C.-M. Chem. Commun. 1997, 65. 34. Liu, C.-J.; Yu, W., -Y.; Li, S.-G.; Che., C.-M. J. Org. Chem. 1998, 63, 7365. 35. Holland, B. T.; Walkup, C.; Stein, A. J, Phys. Chem. B. 1998, 102, 4301. 36. Zhang, L.; Sun, T.; Ying, J. Y. Chem. Commun. 1999, 1103-1104. 37. Jiang, D.-L.; Aida, T. Chem. Commun. 1996, 1523-1524. 38. McCandlish, E.; Miksztal, A. R.; Nappa, M.; Sprenger, A. Q.; Valentine, J. S.; Strong, J. D.; Spiro, T. G. J. Am. Chem. Soc. 1980, 102, 4268-4271. 39. Miksztal, A. R.; Valentine, J. S. Inorg. Chem. 1984, 23, 3548-3552. 40. Burstyn, J. N.; Roe, J. A.; Miksztal, A. R.; Shaevitz, B. A.; Lang, G.; Valentine, J. S. J. Am. Chem. Soc. 1988, 110, 1382-1388. 41. Sisemor, M. F.; Bursyn, J. N.; Valentine, J. S. Angew. Chem.Int. Ed. Engl. 1996, 35, 206. 42. Selke, M.; Sisemore, M. F.; Valentine, J. S. J. Amer. Chem. Soc. 1996, 118, 2008. 43. Sisemore, M. F.; Selke, M.; Bursyn, J. N.; Valentine, J. S. Inorg. Chem. 1997, 36, 979. 44. Selke, M.; Valentine, J. S. J. Amer. Chem. Soc. 1998, 120, 2652-2653. 45. Goto, Y.; Wada, S.; Morishima, I.; Watanabe, Y. J. Inorg. Biochem. 1998, 69, 241-247. 46. Solomon, E. I.; Brunold, T. C.; Davis, M. I.; Kemsley, J. N.; Lee, S.-L.; Lehnert, N.; Neese, F.; Skulan, A. J.; Ynag, Y.-S.; Zhou, J. Chem. Rev. 2000, 100, 235-349.

4. Functional model oxygenations by nonheme iron complexes

209

47. Lubben, M.; Meetsma, A.; Wilkinso, E. C.; Feringa, B.; Que, L., Jr. Angew. Chem. Int. Ed. Engl. 1995, 34, 1512-1514. 48. Ho, R. Y. N.; Roelfes, G.; Hermant, R.; Hage, R.; Feringa, B. L.; Que, L., Jr. Chem, Commun. 1999, 2161-2162. 49. Jensen, K. B.; McKenzie, C. J.; Nielsen, L. P.; Pedersen, J. Z.; Svendsen, H. M. Chem. Commun. 1999, 1313-1314. 50. Simaan, A. J.; Banse, F.; Mialane, P.; Boussac, A.; Un, S.; Kargar-Grisel, T.; Bouchoux, G.; Girerd, J.-J. Eur. J. Inorg. Chem. 1999, 993-996. 51. Simaan, a. J.; Dopner, S.; Banse, F.; Bourcier, S.; Bouchoux, G.; Boussac, A.; Hildebrandt, P.; Girerd, J.-J. Eur. J. Inorg. Chem. 2000, 1627-1633. 52. Simaan, A. J.; Banse, F.; Girerd, J.-J.; Wieghardt, K.; Bill, E. Inorg. Chem 2001, 40, 6358-6540. 53. Hayaishi, O.; Katagiri, M.; Rotherberg, S. J. Am. Chem. Soc 1955, 77, 5450. 54. Ohlendorf, D. H.; Weber, P., C.; Lipscomb, J. D. J. Mol. Biol. 1987, 195, 225227. 55. Ohlendorf, D. H.; Lipscomb, J. D.; Weber, P. C. Nature (London) 1988, 336, 403-405. 56. Ohlendorf, D. H.; Orville, A. M.; Lipscomb, J. D. J. Mol. Biol. 1994, 244, 586608. 57. Vetting, M. W.; D'Argenio, D. A.; Ornston, L. N.; Ohlendorf, D. H. Biochemistry 2000, 39, 7943-7955. 58. Han, S.; Eltis, L. D.; Timmis, K. J.; Muchmore, S. W.; Bolin, J. T. Science 1995, 270, 976-980. 59. Senda, T.; Sugiyama, K.; Narita, H.; Yamamoto, T.; Kimbara, K..; Fukuda, M.; Sato, M.; Yano, K.; Mitsui, Y. J. Mol. Biol. 1996, 255, 735-752. 60. Fujiwara, M.; Golovleva, L. A.; Saeki, Y.; Nozaki, M.; Hayaishi, O. J. Biol. Chem. 1975, 250, 4848-4855. 61. Nakai, C.; Nakazawa, T.; Nozaki, M. Arch. Biochem. Biophys. 1988, 267, 701-713. 62. Nakai, C.; Horiike, K.; Kuramitsu, S.; Hiroyuki, K.; Nozaki, M. J. Biol. Chem. 1990, 265, 660-665. 63. Gibello, A.; Ferrer, E.; Martin, M.; Garrido-Pertierra, A. Biochem. J. 1994, 301, 145-150. 64. Boldt, Y. R.; Whiting, A. K.; Wagner, M. L.; Sadowsky, M. J.; Que, L., Jr.; Wackett, L. P. Biochemistry 1997, 36, 2147-2153. 65. Broderick, J. B.; O'Halloran, T. V. Biochemistry 1991, 30, 7349-7358. 66. Bhat, M. A.; Ishida, T.; Horiike, K.; Vaidyanathan, C. S. Arch. Biochem. Biophys. 1993, 300, 738-746.

210

T.Funabiki

67. Grinstead, R. R. Biochemistry 1964, 3, 1308-1314. 68. Nishinaga, A.; Tojo, T.; Matsuura, T. J. Chem. Soc., Chem. Commun. 1974, 896-897. 69. Moro-oka, Y.; Foote, C. S. J. Am. Chem. Soc. 1976, 98, 1510-1514. 70. White, L. S.; Hellman, E. J.; Que, L. J Org Chem 1982, 47, 3766-3769. 71. Matsuura, T.; Matsushima, H.; Kato, S.; Saito, I. Tetrahedron 1972, 28, 51195129. 72. Matsuura, T. Tetrahedron 1977, 33, 2869-2905. 73. Speier, G.; Tyeklar, Z. J. Chem. Soc., Perkin II 1981, 1176-1179. 74. Sawaki, Y.; Foote, C. S. J. Am. Chem. Soc. 1983, 105, 5035-5040. 75. Farrand, J. C.; Johnson, D. C. J. Org. Chem. 1971, 36, 3606-3612. 76. Pandell, A. J. J. Org. Chem. 1976, 41, 3992-3996. 77. Demmin, T. R.; Rogic, M. M. J. Org. Chem. 1980, 45, 1153-1156. 78. Pandell, A. J. J. Org. Chem 1983, 84, 3908-3912. 79. Ando, W.; Miyazaki, H.; Akasaka, T. J. Chem. Soc., Chem. Commun. 1983, 518-519. 80. Tsuji, J.; Takayanagi, H. J. Am. Chem. Soc. 1974, 96, 7349-7350. 81. Tsuji, J.; Takayanagi, H.; Sakai, I. Tetrahedron Lett. 1975,1245. 82. Tsuji, J.; Takayanagi, H. Chem. Lett. 1980, 65. 83. Rogic, M. M.; Demmin, T. R.; Hammond, W. B. J. Am. Chem. Soc. 1976, 98, 7441. 84. Rogic, M. M.; Demmin, T. R. J. Am. Chem. Soc. 1978, 100, 5472. 85. Funabiki, T.; Sakamoto, H.; Yoshida, S.; Tarama, K. J. Chem. Soc., Chem. Commun. 1979, 754-755. 86. Funabiki, T.; Mizoguchi, A.; Sugimoto, T.; Yoshida, S. Chem. Lett. 1983, 917920. 87. Funabiki, T.; Mizoguchi, A.; Sugimoto, T.; Tada, S.; Tsuji, M.; Sakamoto, H.; 88. 89. 90. 91. 92. 93.

Yoshida, S. J. Am. Chem. Soc. 1986, 108, 2921-2932. Ogo, S.; Yamahara, R.; Funabiki, T.; Masuda, H.; Watanabe, Y. Chem. Lett. 2001, 1062-1063. Cox, D. D.; Benkovic, S. J.; Bloom, L. M.; Bradley, F. C.; Nelson, P. J.; Que, L., Jr.; Wallick, D. E. J. Am. Chem. Soc. 1988, 110, 2026-2032. Dei, A.; Gatteschi, D.; Pardi, L. Inorg. Chem. 1993, 32, 1389-1395. Funabiki, T.; Nagai, Y.; Kojima, H.; Tanaka, T.; Yoshida, S.; Masuda, H. Inorg. Chim. Acta 1998, 275/276, 222-229. Weller, M. G.; Weser, U. J. Am. Chem. Soc. 1982, 104, 3752-3754. Weller, M. G.; Weser, U. Inorg. Chim. Acta 1985, 107, 243-245.

4. Functional model oxygenations by nonheme iron complexes

211

94. Funabiki, T.; Konishi, T.; Kobayashi, S.; Mizoguchi, A.; Takano, M.; Yoshida, S. Chem. Lett. 1987, 719-722. 95. Koch, W. O.; Krüger, H.-J. Angew. Chem. Int. Ed. Engl. 1995, 34, 2671-2674. 96. White, L. S.; Nilsson, P. V.; Pignolet, L. H.; Que, L., Jr. J. Am. Chem. Soc. 1984, 706, 8312-8313. 97. Que, L., Jr.; Kolanczyk, R. C.; White, L. S. J. Am. Chem. Soc. 1987, 709, 5373-5380. 98. Cox, D.D.; Que, L., Jr. J. Am. Chem. Soc. 1988, 110, 8085-8092. 99. Jang, H. G.; Cox, D.D.; Que, L., Jr. J. Am. Chem. Soc. 1991, 113, 9200-9204. 100. Viswanathan, R.; Palaniandavan, M. J. Chem. Soc., Dalian Trans. 1995, 12591266. Viswanathan, R.; Palaniandavan, M.; Balasubramanian, T.; Muthiah, T. P. 101. Inorg. Chem. 1998, 37, 2943-2951. 102. Yamahara, R.; Ogo, S.; Watanabe, Y.; Funabiki, T.; Jitsukawa, K.; Masuda, H.; Einaga, H. Inorg. Chim. Acta 2000, 300-302, 587-596. 103. Mialane, P.; Tchertanov, L.; Banse, F.; Sainton, J.; Girerd, J.-J. Inorg. Chem. 2000, 39, 2440-2444. 104. Duda, M.; Pascaly, M.; Krebs, B. J. Chem. Soc., Chem. Commun. 1997, 835836. 105. Pascaly, M.; Duda, M.; Rompel, A.; Sift, B. H.; Meyer-Klaucke, W.; Krebs, B. Inorg. Chim. Acta 1999, 291, 289-299. 106. Pascaly, M.; Nazikkol, C.; Schweppe, F.; Wiedemann, A.; Zurlinden, K.; Krebs, B. Z. Anorg. Allg. Chem. 2000, 626, 50-55. Mialane, P.; Girerd, J.-J.; Guilhem, J.; Tchertanov, L. Inorg. Chim. Acta 2000, 107. 298, 38-42. 108. Weiner, H.; Finke, R. G. J. Amer. Chem. Soc. 1999, 121, 9831-9842. 109. Funabiki, T.; Yamazaki, T.; Fukui, A.; Tanaka, T.; Yoshida, S. Angew Chem. Int. Ed. Engl. 1998, 37, 513-515. 110. Sorokin, A.; Fraisse, L.; Rabion, A.; Meunier, B. J. Mol. Catal. A, 1997, 117, 103-114. 111. Funabiki, T.; Sugio, D.; Inui, N.; Maeda, M.; Hitomi, Y. Chem. Commun. 2002, 412-413. 112. Funabiki, T.; Yoneda, I.; Ishikawa, M.; Ujiie, M.; Nagai, Y.; Yoshida, S. J. Chem. Soc., Chem. Commun. 1994, 1453-1454. 113. Funabiki, T.; Ishikawa, M.; Nagai, Y.; Yorita, J.; Yoshida, S. J. Chem. Soc., Chem. Commun. 1994, 1951-1952. 114. Lin, G.; Reid, G.; Bugg, T. D. H. Chem. Commun. 2000, 1119-1120. 115. Lin, G.; Reid, G.; Bugg, T. D. H. J. Am. Chem. Soc. 2001, 723, 5030-5039.

212

T.Funabiki

116. Lim, J. H.; Park, T. H.; Lee, H.-J.; Lee, K.-B.; Jang, H. G. Bull. Korean Chem. Soc. 1999, 20, 1428-1432. 117. Jo, D.-H.; Chiou, Y.-M.; Que, L., Jr. Inorg. Chem 2001, 40, 3181-3190. 118. Weiner, H.; Hayashi, Y.; Finke, R. G. Inorg. Chim. Acta 1999, 291,426-437. 119. Ito, M.; Que, L., Jr. Angew. Chem. Int. E. Engl. 1997, 36, 1342-1344. 120. Ogihara, T.; Hikichi, S.; Akita, M.; Moro-oka, Y. Inorg. Chem. 1998, 37, 2614-2615. 121. Orville, A. M.; Lipscomb, J. D.; Ohlendorf, D. H. Biochemistry 1997, 36, 10052-10066. 122. Lim, J. H.; Lee, H. J.; Lee, K.-B.; Jang, H. G. Bull. Korean Chem. Soc. 1997, 18, 1166-1172. 123. Jo, D.-H.; Que, L., Jr. Angew. Chem. Int. Ed. Engl. 2000, 39, 4284-4287. 124. Caneschi, A.; Dei, A. Angew. Chem. Int. Ed. Engl. 1998, 37, 3005-3007. 125. Funabiki, T.; Ito, T.; Matsui, H.; Fukui, A.; Aki, M.; Ogo, S.; Watanabe, Y., The 9th International Conference on Biological Inorganic Chemistry, Minneapolis, USA. 1999, pp 133 126. Funabiki, T.; Ogo, S.; Ito, T.; Matsui, H.; Yamamoto, T.; Tanaka, T.; Hosokawa, Y.; Makihara, N.; Watanabe, Y., 2000 International Chemical Congress of Pacific Basin Societies, Honolulu, Hawaii. 2000, pp INOR 6-1200 127. Earhart, C. A.; Radhakrishnan, R.; Orville, A. M.; Lipscomb J. Mol. Biol. 1994, 236, 374-376. 128. Orville, A. M.; Elango, N.; Ohlendorf, D. H.; Lipscomb, J. D. J. Inorg. Biochem. 1995, 59, 367. 129. Elgren, T. E.; Orville, A. M.; Kelly, K. A.; Lipscomb, J. D.; Ohlendorf, D. H.; Que, L., Jr. Biochemistry 1997, 36, 11504-11513. 130. Orville, A. M.; Elango, N.; Lipscomb, J. D.; Ohlendorf, D. H. Biochemistry 1997, 36, 10039-10051. 131. Orville, A. M.; Lipscomb, J. D. Biochemistry 1997, 36, 14044-14055. 132. Davis, M. I.; Wasinger, E. C.; Westre, T. E.; Zaleski, J. M.; Orville, A. M.; Lipscomb, J. D.; Hedman, B.; Hodgson, K. O.; Solomon, E. I. Inorg. Chem. 1999, 38, 3676-3683. 133. Heistand, R. H., II; Lauffer, R. B.; Fikrig, E.; Que, L., Jr. J. Am. Chem. Soc. 1982, 104, 2789-2796. 134. Lauffer, R. B.; Que, L., Jr. J. Am. Chem. Soc. 1982, 104, 7324-7325. 135. Que, L, Jr. J. Chem. Edu. 1985, 62, 938-943. 136. Que, L., Jr.; Lauffer, R. B.; Lynch, J. B.; Bruce, B. P.; Pyrz, J. W. J. Am. Chem. Soc. 1987,109, 5381-5385.

4. Functional model oxygenations by nonheme iron complexes

213

137. Fujii, S.; Ohya-Nishiguchi, H.; Hirota, N.; Nishinaga, A. Bull. Chem. Soc. Jpn. 1993, 66, 1408-1419. 138. Nishida, Y.; Yamada, K.; Furuhashi, A. Z. Naturforsh. 1990, 45b, 1433-1436. 139. Nishida, Y.; Yoshizawa, K.; Takahashi, S.; Watanabe, I. Z. Naturforsch., C: Biosci. 1992, 47, 209-214. 140. Que, L., Jr.; Lipscomb, J. D.; Zimmermann, R.; Münck, E.; Orme-Johnson, N. R.; Orme-Johnson, W. H. Biochim. Biophys. Acta 1976, 452, 320-334. 141. Whittaker, J. W.; Lipscomb, J. D.; Kent, T. A.; Münck, E. J. Biol. Chem. 1984, 259, 4466-4475. 142. Kent, T. A.; Münck, E.; Pyrz, J. W.; Widom, J.; Que, L., Jr. Inorg. Chem. 1987, 26, 1402-1408. 143. Funabiki, T.; Tada, S.; Yoshioka, T.; Takano, M; Yoshida, S. J. Chem. Soc., Chem. Commun. 1986, 1699-1700. 144. Pascaly, M.; Duda, M.; Schweppe, F.; Zurlinden, K.; Muller, F. K.; Krebs, B. J. Chem. Soc., Dalton Trans. 2001, 828-837. 145. Funabiki, T.; Fukui, A.; Hitomi, Y.; Higuchi, M.; Yamamoto, T.; Tanaka, T.; Tani, F.; Naruta, Y. J. Inorg. Biochem 2002, in press. 146. Simaan, A. J.; Boillot, M.-L.; Riviere, E.; Boussac, A.; Girerd, J.-J. Angew. Chem., Int. Ed. 2000, 39, 196-198. 147. Funabiki, T. In Oxygenates and Model Systems, Funabiki, T., Ed.; Kluwer Academic Publishers: Dordrecht/boston/london, 1997; pp 105-156. 148. Funabiki, T.; Inoue, T.; Kojima, H.; Konishi, T.; Tanaka, T.; Yoshida, S. J. Mol. Catal. 1990, 59, 367-371. 149. Funabiki, T.; Yamazaki, T. J. Mol. Cat. A. 1999, I50, 37-47. 150. Wheeler, D. E.; Rodriguez, J. H.; MacCusker, J. K. J. Phys. Chem. A 1999, 103, 4101-4112.

151. Mialane, P.; Anxolabehere-Mallart, E.; Blondin, G.; Nivorojkine, A.; Gulhem, J.; Tchertanova, L.; Gesario, M.; Ravi, N.; Bominaar, E.; Girerd, J.-J.; Münck, E. Inorg. Chim. Acta 1997, 263, 367-378. 152. Koch, W. O.; Schunemann, V.; Gerdan, M.; Trautwein, A. X.; Krüger, H.-J. Chem. Eur. J. 1998, 4, 1255-1265. 153. Nishinaga, A.; Itahara, T.; Matsuura, T. Bull. Chem. Soc. Jpn. 1974, 47, 18111812. 154. Raffard, N.; Carina, R.; Simaan, A. J.; Sainton, J.; Riviere, E.; Tchertanov, L.; Boucier, S.; Bouchoux, G.; Delroisse, M.; Banse, F.; Girerd, J.-J. Eur. J. Inorg. Chem. 2001, 2249-2254. 155. Bianchini, C.; Frediani, P.; Laschi, F.; Meli, A.; Vizza, F.; Zanello, P. Inorg. Chem. 1990, 29, 3402.

214

T.Funabiki

156. Barbaro, P.; Bianchini, C.; Mealli, C.; Meli, A. J. Am. Chem. Soc. 1991, 113, 3181-3183. 157. Barbaro, P.; Bianchini, C.; Frediani, P.; Meli, A.; Vizza, F. Inorg. Chem. 1992, 31, 1523-1529. 158. Sanvoisin, J.; Langley, G. J.; Bugg, T. D. H. J. Am. Chem. Soc. 1995, 117, 7836-7837. 159. Spence, E. L.; John, L. G.; Bugg, T. D. H. J. Am. Chem. Soc. 1996, 118, 8336-

8343. 160. Winfield, C. J.; Al-Mahrizy, Z.; Gravestock, M.; Bugg, T. D. H. J. Chem. Soc.Perkin Trans. 1 2000, 3277-3289. 161. Bugg, T. D. H.; Gang, L. Chem. Commun. 2001, 941-952. 162. Bugg, T. D. H. Current Opinion in Chemical Biology 2001, 5, 550-555. 163. Eley, K. L.; Crowley, P. J.; Bugg, T. D. H. J. Org. Chem. 2001, 66, 2091-2097. 164. Tai, H. H.; Sih, C. J. J. Biol. Chem. 1970, 245, 5072. 165. Hirata, F.; Nakazawa, A.; Nozaki, M.; Hayaishi, O. J. Biol. Chem. 1971, 246, 5882-5887. 166. Hori, K.; Hashimoto, T.; Nozaki, M. J. Biochem. (Tokyo) 1973, 74, 375-384. 167. Arciero, D. M.; Orville, A. M.; Lipscomb, J. D. J. Biol. Chem. 1985, 260, 14035-14044. 168. Mabrouk, P. A.; Orville, A. M.; Lipscomb, J. D.; Solomon, E. I. J. Am. Chem. Soc. 1991, 113, 4053-4061. 169. Bertini, I.; Briganti, F.; Mangani, S.; Nolting, H. F.; Scozzafava, A. Biochemistry 1994, 33, 10777-10784. 170. Bertini, I.; Briganti, F.; Mangani, S.; Nolting, H. F.; Scozzafava, A. FEBS Lett. 1994, 350, 207-212. 171. Bertini, I.; Briganti, F.; Scozzafava, A. FEBS Lett. 1994, 343, 56-60. 172. Shu, L.; Chiou, Y.-M.; Orville, A. M.; Miller, M. A.; Lipscomb, J. D.; Que, L., Jr. Biochemistry 1995, 34, 6649-6659. 173. Zang, Y.; Elgren, T. E.; Dong, Y.; Que, L., Jr. J. Am. Chem. Soc. 1993, 115, 811-813. 174. Kim, J.; Zang, Y.; Gostas, M.; Harrison, R. G.; Wilkinson, E. C.; Que, L., Jr. J. Biol. Inorg. Chem 2001, 6, 275-284. 175. Goldsmith, C. R.; Jonas, R. T.; Stack, T. D. P. J. Am. Chem. Soc. 2002, 124, 83-96. 176. Ha, E. H.; Ho, R. Y. N.; Kisiel, J. F.; Valentine, J. S. Inorg. Chem. 1995, 34, 2265-2266. 177. Funabiki, T.; Imagawa, H.; Ito, Y., 7th International Symposium on Dioxygen Activation and Homogeneous Catalytic Oxidation, Yoke, UK. 1999, pp P.3

4. Functional model oxygenations by nonheme iron complexes

215

178. Funabiki, T.; Imagawa, H.; Ito, Y., EUROPACAT V, Limeric, Irland. 2001, pp 15-P-07 179. Chiou, Y. M.; Que, L., Jr. J. Am. Chem. Soc. 1992, 114, 7567-7568. 180. Chiou, Y.-M.; Que, L., Jr. Angew. Chem. Int. Ed Engl. 1994, 33, 1886-1888. 181. Chiou, Y.-M.; Que, L., Jr. J. Am. Chem. Soc. 1995, 117, 3999-4013. 182. Chiou, Y.-M.; Que, L., Jr. Inorg. Chem. 1995, 34, 3270-3278. 183. Hegg, E. L.; Ho, R. Y. N.; Que, L., Jr. J. Am. Chem. Soc. 1999, 121, 19721973. 184. Hegg, E. L.; Whiting, A. K.; Saari, R. E.; McCracken, J.; Hausinger, R. P.; Que, L., Jr. Biochemistry 1999, 38, 16714-16726. 185. Ho, R. Y. N.; Mehn, M. P.; Hegg, E. L.; Liu, A.; Ryle, M. J.; Hausinger, R. P.; Que, L., Jr. J. Am. Chem. Soc. 2001, 123, 5022-5029. 186. Chen, K.; Que, L., Jr. Angew. Chem., Int. Ed. 1999, 38, 2227-2229. 187. Costas, M.; Tipton, A. K.; Chen, K.; Du-Hwan, J.; Que, L., Jr. J. Am. Chem. Soc. 2001, 123, 6722-6723. 188. Chen, K.; Costas, M.; Kim, J. H.; Tipton, A. K.; Que, L. Journal of the American Chemical Society 2002, 124, 3026-3035. 189. Chen, K.; Costas, M.; Que, L. J. Chem. Soc., Dalton Trans. 2002, 672-679. 190. Ryu, J. Y.; Kim, J.; Costas, M.; Chen, K.; Nam, W.; Que, L. Chem. Commun. 2002, 1288-1289. 191. Costas, M.; Que, L. Angew. Chem. Int. Ed. Engl. 2002, 41, 2179-+. 192. Funabiki, T. In Oxygenases and Model Systems, Funabiki, T., Ed.; Kluwer Academic Publishers: Dordrecht/boston/london, 1997; Vol. 19; pp 1-18. 193. Kodera, M.; Kano, K.; Funabiki, T. In Oxygenases and Model Systems, Funabiki, T., Ed.; Kluwer Academic Publishers: Dordrecht/boston/london, 1997; pp 283-344. 194. Kodera, M.; Shimakoshi, H.; Nishimura, M.; Okawa, H.; Iijima, S.; Kano, K. Inorg. Chem. 1996, 35, 4967-4973. 195. Kodera, M.; Shimakoshi, H.; Kano, K. Chem. Commun. 1996,1737-1738. 196. Kodera, M.; Taniike, Y.; Itoh, M.; Tanahashi, Y.; Shimakoshi, H.; Kano, K.; Hirota, S.; Iijima, S.; Ohba, M.; Okawa, H. Inorg. Chem 2001, 40, 4821-4822. 197. Payra, P.; Hung, S.-C.; Kwok, W. H.; Johnston, D.; Gallucei, J.; Chan, M. K. Inorg. Chem 2001, 40, 4036-4039. 198. Vincent, J. B.; Huffman, J. C.; Christou, G.; Li, Q.; Nanny, M. A.; Hendrickson, D. N.; Fong, R. H.; Fish, R. H. J. Am. Chem. Soc. 1988, 110, 6898-6900. 199. Wieghardt, K.; Pohl, K.; Gebert, W. Angew. Chem. Int. Ed Engl. 1983, 98, 727-.

216

T.Funabiki

200. Armstrong, W. H.; Spool, A.; Papaethymiou, G. S.; Frankel, R. B.; Lippard, S. J. J. Am. Chem. Soc. 1984, 106, 3653-3667. Toftlund, H.; Murray, K. S.; Zwack, P. R.; Taylor, L. F.; Anderson, O. P. J. 201. Chem. Soc., Chem. Commun. 1986, 191-193. 202. Gomez-Romero, P.; Casan-Pastor, N.; Ben-Hussein, A.; Jameson, G. B. J. Am. Chem. Soc. 1988, 110, 1988-1990. 203. Fish, R. H.; Oberhausen, K. J.; Chen, S.; Richardson, J. F.; Pierce, W.; Buchanan, R. M. Catal. Lett. 1993, 18, 357-365. 204. Buchanan, R. M.; Chen, S.; Richardson, J. F.; Bressan, M.; Forti, L.; Morvillo, A.; Fish, R. H. Inorg. Chem. 1994, 33, 3208-3209. 205. Liu, K. E.; Johnson, C. C.; Newcomb, M.; Lippard J. Am. Chem. Soc. 1993, 115, 939-947. 206. Priestley, N. D.; Floss, H. G.; Froland, W. A.; Lipscomb, J. D.; Williams, P. G.; Morimoto, H. J. Am. Chem. Soc. 1992, 114, 7561-7562. 207. Rabion, A.; Chen, D. S.; Wang, J.; Buchanan, R. M.; Seris, J.-L.; Fish, R. H. J. Am. Chem. Soc. 1995, 117, 12356-12357. 208. Rabion, A.; Buchanan, R. M.; Seris, J.-L.; Fish, R. H. J Mol. Catal. A: Chem 1997, 116, 43-47. 209. Fontecave, M.; Roy, B.; Lambeaux, C. J. Chem. Soc., Chem. Commun. 1991, 939-940. 210. Menage, S.; Vincent, J. M.; Lambeaux, C.; Chottard, G.; Grand, A.; Fontecave, M. Inorg. Chem. 1993, 32, 4766-4773. 211. Vincent, J.-M.; Menage, S.; Lambeaux, C.; Fontecave, M. Tetrahedron Lett. 1994, 35, 6287-6290. 212. Menage, S.; Vincent, J.-M.; Lambeaux, C.; Fontecave, M. J Mol. Catal. A: Chem. 1996, 113, 61-75. 213. Menage, S.; Galey, J.-B.; Hussler, G.; Seite, M.; Fontecave, M. Angew. Chem., Int. Ed. Engl. 1996, 35, 2353-2355. 214. Fontecave, M.; Menage, S.; Duboc-Toia, C.; Vincent, J. M.; Lambeaux, C. Biochem. Soc. Trans. 1997, 25, 65-69. 215. Leising, R. A.; Kim, J.; Perez, M. A.; Que J. Am. Chem. Soc. 1993, 115, 95249530. 216. Leising, R. L.; Norman, R. E.; Que, L., Jr. Inorg. Chem. 1990, 29, 2553-2555. 217. Kim, J.; Harrison, R. G.; Kim, C.; Que, L., Jr. J. Am. Chem. Soc. 1996, 118, 4373-4379. 218. Arends, I. W. C. E.; Ingold, K. U.; Wayner, D. D. M. J. Am. Chem. Soc. 1995, 111, 4710-4711.

4. Functional model oxygenations by nonheme iron complexes

217

219. Sinelgrove, D. W.; MacFaul, P. A.; Ingold, K. U.; Wayner, D. D. M. Tetrahedron Lett. 1996, 37, 823-826. 220. MacFaul, P. A.; Arends, I. W. C. E.; Ingold, K. U.; Wayner, D. D. M. J. Chem. Soc., Dalton Trans. 1997, 135-145. 221. Miyake, H.; Chen, K.; Lange, S. J.; Que, L., Jr. Jnorg. Chem. 2001, 40, 35343538. 222. MacFaul, P. A.; Ingold, K. U.; Wayner, D. D. M.; Que, L., Jr. J. Am. Chem. Soc. 1997, 119, 10594-10598. 223. Vincent, J.-M.; Bearnais-Barbry, S.; Pierre, C.; Verlhac, J.-B. J. Chem. Soc., Dalton Trans. 1999, 1913-1914. 224. Fish, R. H.; Konings, M. S.; Oberhausen, K. J.; Fong, R. H.; Yu, W. M.; Christou, G.; Vincent, J. B.; Coggin, D. K.; Buchanan, R. M. Inorg. Chem. 1991, 30, 3002. 225. Menage, S.; Vincent, J. M.; Lambeaux, C.; Fontecave, M. J. Chem. Soc., Dalton Trans. 1994, 2081-2084. 226. Menage, S.; Galey, J.-B.; Dumats, J.; Hussler, G.; Seite, M.; Luneau, I. G.; Chottard, G.; Fontecave, M. J. Am. Chem. Soc. 1998, 120, 13370-13382. 227. Meckmouche, Y.; Duboc-Toia, C.; Menage, S.; Lambeaux, C.; Marc, F. J. Mol. Cat. A. 2000, 156, 85-89. 228. White, M. C.; Doyle, A. G.; Jacobsen, E. N. J. Am. Chem. Soc. 2001, 123, 7194-7195.

229. Kitajima, N.; Fukui, H.; Morooka, Y. J. Chem. Soc., Chem. Commun. 1988, 485-486. 230. Kitajima, N.; Ito, M.; Fukui, H.; Moro-oka, Y. J. Chem. Soc., Chem. Commun. 1991, 102-104. 231. Wang, Z.; Martell, A. E.; Motekaitis, R. J.; Reibenspies, J. H. Inorg. Chim. Acta 2000, 300-302, 378-383. 232. Suzuki, M.; Furutachi, H.; Okawa, H. Coord. Chem. Rev 2000, 200-202, 105129. 233. Westerheide, L.; Pascaly, M.; Krebs, B. Current Opinion in Chemical Biology 2000, 4, 235-241. 234. Tolman, W. B.; Que, L., Jr. J. Chem. Soc., Dalton Trans. 2002, 653-660. 235. Que, L.; Tolman, W. B. Angew. Chem. Int. Ed. Engl. 2002, 41, 1114-1137. 236. Chavez, F. A.; Ho, R. Y. N.; Pink, M.; Young, V. G., Jr.; Kryatov, S. V.; Rybak-akimova, E. V.; Andres, H.; Münck, E.; Que, L., Jr.,; Tolman, W. B. Angew. Chem., Int. Ed. 2002, 41, 149-152. 237. Que, L., Jr. J. Chem. Soc., Dalton Trans. 1997, 3933-3940. 238. Que, L., Jr.; Dong, Y. Acc. Chem. Res. 1996, 29, 190-196.

218

T.Funabiki

239. Zang, Y.; Dong, Y.; Que, L., Jr.; Kauffmann, K.; Muenck, E. J. Am. Chem. Soc. 1995, 117, 1169-1170. 240. Zheng, H.; Zang, Y.; Dong, Y.; Young, V. G., Jr.; Que, L., Jr. J. Am. Chem. Soc. 1999, 121, 2226-2235. 241. Hsu, H.-F.; Dong, Y.; Shu, L.; Young, V. G., Jr.; Que, L., Jr. J. Am. Chem. Soc. 1999, 121, 5230-5237. 242. Kryatov, S. V.; Rybak-akimova, E. V.; MacMurdo, V. L.; Que, L., Jr. Inorg. Chem. 2001, 40, 2220-2228. 243. Dong, Y.; Fujii, H.; Hendrich, M. P.; Leising, R. A.; Pan, G.; Randall, C. R.; Wilkinson, E. C.; Zang, Y.; Que, L., Jr.; Fox, B. G.; Kauffmann, K.; Münck, E. J. Am. Chem. Soc. 1995, 117, 2778-2792. 244. Dong, Y.; Zang, Y.; Kauffmann, K.; Shu, L.; Wilkinson, E. C.; Münck, E.; Que, L., Jr. J. Am. Chem. Soc. 1997, 119, 12683-12684. 245. Kryatov, S. V.; Ryabak-Akimova, E. V. J. Chem. Soc., Dalton Trans. 1999, 3335-3336. 246. Kim, C.; Dong, Y.; Que, L., Jr. J. Am. Chem. Soc. 1997, 119, 3635-3636. 247. Zheng, H.; Yoo, S. J.; Muenck, E.; Que, L., Jr. J. Am. Chem. Soc. 2000, 122, 3789-3790. 248. Costas, M.; Rohde, l.-U.; Stubna, A.; Ho, R. Y. N.; Quaroni, L.; Münck, E.; Que, L., Jr. J. Am. Chem. Soc. 2001, 123, 12931-12932. 249. LeCloux, D. D.; Barrios, A. M.; Lippard, S. J. Bioorg. Med. Chem. 1999, 7, 763-772. 250. LeCloux, D. D.; Barrios, A. M.; Mizoguchi, T. J.; Lippard, S. J. J. Am. Chem. Soc. 1998, 120, 9001-9014. 251. Zhang, X.; Sasaki, K.; Hill, C. L. J. Amer. Chem. Soc. 1996, 118, 4809-4816. 252. Mizuno, N.; Kiyoto, I.; Nozaki, C.; Misono, M. J. Catal. 1999, 181, 171-174. 253. Mizuno, N.; Nozaki, C.; Kiyoto, I.; Misono, M. J. Am. Chem. Soc. 1998, 120, 9267-9272. 254. Mizuno, N.; Nozaki, C.; Kiyoto, I.; Misono, M. J. Catal. 1999,182, 285-288. 255. Mizuno, N.; Seki, Y.; Nishiyama, Y.; Kiyoto, I.; Misono, M. J. Catal. 1999, 184, 550-552. 256. Nozaki, C.; Misono, M.; Mizuno, N. Chem. Lett. 1998, 1263-1264. 257. Nishiyama, Y.; Nakagawa, Y.; Mizuno, N. Angew Chem. Int. Ed. Engl. 2001, 40,3639-3641. 258. Neimann, K.; Neumann, R.; Rabion, A.; Buchanan, R. M.; Fish, R. H. Inorg. Chem. 1999, 38, 3575-3580. 259. Vincent, J.-M.; Rabio, A.; Yachandra, V. K.; Fish, R. H. Anvew. Chem. Int. Ed. Engl. 1997, 36, 2346-2349.

4. Functional model oxygenations by nonheme iron complexes

219

260. Antony, R.; Tembe, G. L.; Ravindranahan, M.; Ram, R. N. J. Mol. Cat. A. 2001,171,159-168. 261. Knops-Gerrits, P.-P.; Ververckmoes, A.; Schoonheydt. R.; Ichikawa, M.; Jacobs, P. Microporous and Mesopourous Materials 1998, 21, 475-486. 262. Panov, G. I.; Sobolev, V. I.; Dubkov, K. A.; Parmon, V. N.; Ovanesyan, N. S.; Shilov, A. E.; Shteinman, A. A. Reat. Kinet. Catal. Lett. 1997, 61, 251-258. 263. Dubkov, K. A.; I., S. V.; Talsi, E. P.; Rodkin, M. A.; Watkins, N. H.; Shteinman, A. A.; Panov, G. I. J. Mol. Cat. A. 1997, 123,155-161. 264. Fenton, H. J. H. J. Chem. Soc., Chem. Commun. 1894, 65, 899-910. 265. Wardman, P.; Candeias, L. P. Radiation Research 1996, 145, 523-531. 266. Walling, C. Acc. Chem. Res. 1998, 31, 155-157. 267. Haber, F.; Weiss, J. Proc. R. Soc. London 1934, A147, 332-351. 268. Barb, W. G.; Baxendal, J. H.; George, P.; Hargrave, K. R. Trans. Farada Soc. 1951, 47, 462-500. 269. Zhang, X.; Zhang, D.; Busch, D. H.; van Eldik, R. J. Chem. Soc., Dalton Trans. 1999, 2751-2758. 270. Rahhal, S.; Richter, H. W. J. Am. Chem. Soc. 1988, 110, 3126-3133. 271. Rush, J. D.; Koppenol, W. H. J. Am. Chem. Soc. 1988, 110, 4957-4963. 272. Hamilton, G. A.; Friedman, J., P. J. Am. Chem. Soc. 1963, 85, 1008-1009. 273. Hamilton, G. A.; Friedman, J. P.; Campbell, P. M. J. Am. Chem. Soc. 1966, 88, 5266-5268. 274. Hamilton, G. A.; Hanifin, J. W., Jr.; Friedman, J. P. J. Am. Chem. Soc. 1969, 86, 5269-5272. 275. Maissant, J.-M; Bouchoule, C.; Blanchard, M.; Blanchard, M. J. Mol. Catal 1980, 9, 237-240. 276. Hotta, K.; Tamagaki, S.; Suzuki, Y.; Tagaki, W. Chem. Lett. 1981, 789-790. 277. Kubicz, E.; Mlochowski, J.; Skarzewski, J. Pol. J. Chem. 1984, 58, 913-916. 278. Castle, L.; Lindsay Smith, J. R. J. Chem. Soc. Chem. Commun. 1978, 704-705. 279. Castle, L.; Lindsay Smith, J. R. J. Mol. Cat. 1980, 7, 235-243. 280. Dewar, C. A.; Suckling, C. J.; Higiins, R. J. Chem. Research (S) 1979, 336337. 281. Dewar, C. A.; Suckling, C. J.; Higiins, R. J. Chem. Research (M) 1979, 38123830. 282. Brook, A. M.; Castle, L.; Lindsay Smith, J. R.; Higins, R.; Morris, K. P. J. Chem. Soc. Perkin Trans. 1982, 2, 687-692. 283. Dunford, H. B. Coord. Chem. Rev 2002, in press. 284. Yamamoto, T.; Kimura, M. J. Chem. Soc. Chem. Commun. 1977, 948-949. 285. Sugimoto, H.; Sawyer, D. T. J. Am. Chem. Soc. 1984, 106, 4283.

220

T.Funabiki

286. Sugimoto, H.; Sawyer, D. T. J. Am. Chem. Soc. 1985, 107, 3712-3716. 287. Sugimoto, H.; Sawyer, D. T. J. Org. Chem. 1985, 50, 1784-1786. 288. Sugimoto, H.; Spencer, L.; Sawyer, D. T. Proc. Natl. Acad. Sci. U. S. A. 1987, 84, 1731-1733. 289. Sheu, C.; Richert, S. A.; Cofre, P.; Ross, B., Jr.; Sobkowiak, A.; Sawyer, D. T.; Kanofsky, J. R. J. Am. Chem. Soc. 1990, 112, 1936-1942. 290. Tung, H. C.; Kang, C.; Sawyer, D. T. J. Am. Chem. Soc. 1992,114, 3445-3455. 291. Sawyer, D. T.; Kang, C.; Llobet, A.; Redman, C. J. Am. Chem. Soc. 1993, 115, 5817-5818. 292. Hage, J. P.; Llobet, A.; Sawyer, D. T. Bioorg. Med. Chem. 1995, 3, 1383-1388. 293. Sawyer, D. T.; Sobkowiak, A.; Matsushita, T. Acc. Chem. Res. 1996, 29, 409416. 294. Sawyer, D. T. Coord. Chem. Rev 1997, 165, 297-313. 295. Barton, D. H. R.; Halley, F.; Ozbalik, N.; Young, E.; Balavoine, G.; Gref, A.; Boivin, J. New J. Chem. 1989, 13, 177-182. 296. Barton, D. H. R.; Halley, F.; Ozbalik, N.; Schmitt, M.; Young, E.; Balavoine, G. J. Am. Chem. Soc. 1989, 111, 7144-7149. 297. Barton, D. H. R.; Doller, D.; Ozbalik, N.; Balavoine, G.; Gref, A.; Boivin, J. Tetrahedron Lett. 1990, 31, 353-356. 298. About-Jaudet, E.; Barton, D. H. R.; Csuhai, E.; Ozbalik, N. Tetrahedron Lett. 1990, 31, 1657-1660. 299. Barton, D. H. R.; Csuhai, E.; Doller, D.; Ozbalik, N.; Senglet, N. Tetrahedron Lett. 1990, 31, 3097-3100. 300. Barton, D. H. R.; Csuhai, E.; Ozbalik, N. Tetrahedron 1990, 46, 3743-3752. 301. Geletii, Y. V.; Shilov, A. E. Studies in Organic Chemistry 1988, 33, 193-300. 302. Barton, D. H. R.; Gastiger, M. J.; Motherwell, W. B. J. Chem. Soc., Chem. Commun. 1983, 41-43. 303. Barton, D. H. R.; Csuhai, E.; Doller, D.; Balavoine, G. J. Chem. Soc., Chem. Commun. 1990, 1787-1789. 304. Barton, D. H. R.; Doller, D. Stud. Surf. Sci. Catal. 1991, 66,1-10. 305. Barton, D. H. R.; Salgueiro, M. C.; MacKinnon, J. Tetrahedron 1997, 53, 7417-7428. 306. Barton, D. H. R.; Li, T.; Peralez, E. Tetrahedron 1998, 54, 15087-15096. 307. Barton, D. H. R.; Eaton, P. E.; Liu, W. G. Tetrahedron Lett. 1991, 32, 62636264. 308. Barton, D. H. R.; Csuhai, E.; Doller, D. Tetrahedron Lett. 1992, 33, 34133416.

4. Functional model oxygenations by nonheme iron complexes

221

309. Barton, D. H. R.; Beviere, S. D.; Chavasiri, W.; Csuhai, E.; Doller, D.; Liu, W. G. J. Am. Chem. Soc. 1992, 114, 2147-2156. 310. Barton, D. H. R.; Chabot, B. M; Hu, B. Tetrahedron 1996, 52, 10301-10312. 311. Barton, D. H. R.; Beviere, S. D.; Chavasiri, W.; Doller, D.; Liu, W. G.; Reibenspies, J. H. New J. Chem. 1992, 16, 1019-1029. 312. Barton, D. H. R.; Hu, B.; Taylor, D. K.; Rojas Wahl, R. U. Tetrahedron Lett. 1996, 37,1133-1136. 313. Barton, D. H. R.; Hu, B.; Li, T.; MacKinnon, J. Tetrahedron Lett. 1996, 37, 8329-8332. 314. Barton, D. H. R.; Hu, B.; Taylor, D. K.; Wahl, R. U. R. J. Chem. Soc., Perkin Trans. 2 1996, 1031-1041. 315. Barton, D. H. R.; Li, T. Tetrahedron 1998, 54, 1735-1744. 316. Barton, D. H. R.; Beck, A. H.; Delanghe, N. C. Tetrahedron Lett. 1996, 37, 1555-1558. 317. Barton, D. H. R.; Delanghe, N. C. Tetrahedron Lett. 1996, 37, 8137-8140. 318. Barton, D. H. R. J. Mol. Cat. A. 1997, 117, 3-7. 319. Barton, D. H. R. Tetrahedron 1998, 54, 5805-5817. 320. Tapper, A. E.; Long, J. R.; Staples, R. J.; Stravropoulos, P. Angew. Chem. Int. Ed. Engl. 2000, 39, 2343-2346. 321. Barton, D. H. R.; Doller, D. Acc. Chem. Res. 1992, 25, 504-512. 322. Minisci, F.; Vismara, E.; Fontana, F. Heterocycles 1989, 28, 489-519. 323. Kiani, S.; Tapper, A.; Staples, R. J.; Stavropoulos, P. J. Am. Chem. Soc. 2000, 122, 7503-7517. 324. Liu, C.; Ye, X.; Zhan, R.; Wu, Y. J. Mol. Cat. A. 1996, 112, 15-22. 325. Mohamadin, A. M. J. Inorg. Biochem 2001, 84, 97-105. 326. Kim, C.; Chen, K.; Kim, J.; Que, L., Jr.J. Am. Chem. Soc. 1997, 119, 59645965. 327. Chen, K.; Que, L., Jr. Chem. Commun. 1999, 1375-1376. 328. Roelfes, G.; Lubben, M.; Hage, R.; Que, L., Jr.; Feringa, B. L. Chem. Eur. J. 2000, 6, 2152-2159. 329. Chen, K.; Que, L., Jr. J. Am. Chem. Soc. 2001, 123, 6327-6337. 330. Chen, K.; Costas, M.; Que, L., Jr. J. Chem. Soc., Dalton Trans. 2002, 672-679. 331. Sorokin, A.; Meunier, B. Science 1995, 268, 1163-1166. 332. Sorokin, A.; Meunier, B. Chem. Eur. J. 1996, 2, 1308-1317. 333. Sorokin, A.; De Suzzoni-Desard, S.; Poullain, D.; Noel, J. P.; Meunier, B. J. Am. Chem. Soc. 1996, 118, 7410-7411. 334. Hadasch, A.; Sorokin, A.; Rabion, A.; Meunier, B. New J. Chem. 1998, 335. Hemmert, C.; Renz, M.; Meunier, B.J. Mol. Cat. A. 1999, 137, 205-212.

222

T.Funabiki

336. Ensing, B.; Buda, F.; Blochl, P.; Baerends, E. J. Anvew. Chem. Int. Ed. Engl. 2001, 40, 2893-2895. 337. Horwits, C., P.; Fooksman, D. R.; Vuocolo, L. D.; Gordon-Wylie, S. W.; Cox, N. J.; Collins, T. J. J. Am. Chem. Soc. 1998, 120, 4867-4868. 338. Bartos, M. J.; Gordon-Wylie, S. W.; Fox, B. G.; Wright, L. J.; Weintraub, S. T.; Kauffmann, K. E.; Münck, E.; Kostka, K.; Uffelman, E. S.; Richard, C. E. F.; Noon, K. R.; Collins, T. J. Coord. Chem. Rev 1998, 174, 361-390. 339. Gupta, S. S.; Stadler, M.; Noser, C. A.; Ghosh, A.; Seinhoff, B.; Lenoir, D.; Horwits, C., P.; Schramm, K.-W.; Collins, T. J. Science 2002, 296, 326-328. 340. Collins, T. J. Acc. Chem. Res. 2002, Published on Web. 341. Frusteri, F.; Parmaliana, A.; Arena, F.; Giordano, N. J. Chem. Soc., Chem. Commun. 1991, 1332-1334. 342. Frusteri, F.; Arena, F.; Bellitto, S.; Parmaliana, A. Appl. Catai, A 1999, 180, 325-333. 343. Espro, C.; Frusteri, F.; Arena, F.; Parmaliana, A. J. Mol. Cat. A. 2000, 159, 359-364. 344. Bianchi, D.; Bortolo, R.; Tassinari, R.; Ricci, M.; Vignola, R. Angew Chem. Int. Ed. Engl. 2000, 39, 4321-4323. 345. Barton, D. H. R.; Beviere, S. D.; Chavasiri, S.; Doller, D.; Hu, B. Tetrahedron Lett. 1992, 33, 5473-5476. 346. Barton, D. H. R.; Chavasiri, W. Tetrahedron 1994, 50, 19-30. 347. Barton, D. H. R.; Beviere, S. D.; Hill, D. R. Tetrahedron 1994, 50, 2665-2670. 348. Schuchardt, U.; Pereira, R.; Rufo, M. J. Mol. Cat. A. 1998, 135, 257-262. 349. Barton, D. H. R.; Wang, T. L. Tetrahedron 1994, 50, 1011-1032. 350. Barton, D. H. R.; Beviere, S. D.; Chabot, B. M.; Chavasiri, W.; Taylor, D. K. Tetrahedron Lett. 1994, 35, 4681-4684. 351. Barton, D. H. R.; Wang, T.-L. Tetrahedron Lett. 1994, 35, 4307-4310. 352. Nguyen, C.; Guajardo, R. J.; Mascharak, P. K. Inorg. Chem. 1996, 35, 62736281. 353. Lehnert, N.; Ho, R. Y. N.; Que, L.; Solomon, E. I. J. Am. Chem. Soc. 2001, 123, 8271-8290. 354. Lehnert, N.; Ho, R. Y. N.; Que, L., Jr.; Solomon, E. I. J. Am. Chem. Soc. 2001, 123, 12802-12816. 355. Bardin, C.; Barton, D. H. R.; Hu, B.; Rojas-Wahl, R.; Taylor, D. K. Tetrahedron Lett. 1994, 35, 5805-5808. 356. Barton, D. H. R.; Chabot, B. M.; Delanghe, N. C.; Hu, B.; Le Gloahec, V. N.; Wahl, R. U. R. Tetrahedron Lett. 1995, 36, 7007-7010. 357. Minisci, F.; Fontana, F. Tetrahedron Lett. 1994, 35, 1427-1430.

4. Functional model oxygenations by nonheme iron complexes

223

358. Minisci, F.; Fontana, F.; Araneo, S.; Recupero, F. Tetrahedron Lett. 1994, 35, 3759-3762. 359. Minisci, F.; Fontana, F.; Araneo, S.; Recupero, F. J. Chem. Soc., Chem. Commun. 1994, 1823-1824. 360. Minisci, F.; Fontana, F.; Araneo, S.; Recupero, F.; Banfi, S.; Quici, S. J. Am. Chem. Soc. 1995, 117, 226-232. 361. Barton, D. H. R.; Le Gloahec, V. N.; Patin, H. New J. Chem. 1998, 22, 565568. 362. Barton, D. H. R.; Le Gloahec, V. N.; Patin, H.; Launay, F. New J. Chem. 1998, 22, 559-563. 363. Barton, D. H. R.; Le Gloahec, V. N. Tetrahedron Lett. 1998, 39, 4413-4416. 364. Barton, D. H. R.; Le Gloahec, V. N.; Smith, J. Tetrahedron Lett. 1998, 39, 7483-7486. 365. Barton, D. H. R.; Le Gloahec, V. N. Tetrahedron 1998, 54, 15457-15468. 366. Kaizer, J.; Greczi, S.; Speier, G. J. Mol. Catal. A. 2002, 179, 35-39. 367. Barton, D. H. R.; Chabot, B. M. Tetrahedron 1997, 53, 487-510. 368. Barton, D. H. R.; Chabot, B. M. Tetrahedron 1997, 53, 511-520. 369. Shyu, H.-L.; Wei, H.-H.; Lee, G.-H.; Wang, Y. J. Chem. Soc., Dalton Trans. 2000, 915-918. 370. Sheu, C.; Sobkowiak, A.; Jeon, S.; Sawyer, D. T. J. Am. Chem. Soc. 1990, 112, 879-881. 371. Hage, J. P.; Powell, J. A.; Sawyer, D. T. J. Am. Chem. Soc. 1995, 117, 1289712898. 372. Sobkowiak, A.; Narog, D.; Sawyer, D. T. J. Mol. Catal. A: Chem. 2000, 159, 247-256. 373. Hirao, T.; Moriuchi, T.; Ishikawa, T.; Nishimura, K.; Mikami, S.; Ohshiro, Y.; Ikeda, I. J. Mol. Cat.A: Chem. 1996, 113, 117-130. 374. Hirao, T.; Moriuchi, T.; Mikami, S.; Ikeda, I.; Ohshiro, Y. Tetrahedron Letters 1993, 34, 1031-1034. 375. Bottcher, A.; Grinstaff, M. W.; Labinger, J. A.; Gray, H. B. J Mol. Catal. A. 1996, 113, 191-200. 376. Duprat, A. F.; Capdevielle, P.; Maumy, M. J. Chem. Soc. Chem. Commun. 1991, 464-466. 377. Uenfriend, S.; Clark, C. T.; Axelrod, J.; Brodie, B. B. J. Biol. Chem. 1954, 208, 731-739. 378. Brodie, B. B.; Axelrod, J.; Shore, P. A.; Udenfriend, S. J. Biol. Chem. 1954, 208, 741-750. 379. Hamilton, G. A. J. Am. Chem. Soc. 1964, 86, 3391-3392.

224

T.Funabiki

380. Hamilton, G. A.; Workman, R. J.; Woo, L. J. Am. Chem. Soc. 1964, 86, 33903391. 381. Lindsay Smith, J. R.; Shaw, B. A. J.; Foulkes, D. M.; Jeffrey, A. M.; Jerina, D. M. J. Chem. Soc., Perkin II 1977, 1583-1589. 382. Maissant, J.-M.; Bouchoule, C.; Blanchard, M. J. Mol. Catal. 1982, 14, 333340. 383. Ullrich, V. Z. Naturforsch. B. 1969, 24, 699-704. 384. Sakurai, H.; Ogawa, S. Biochhem. Pharm. 1975, 24, 1257-1260. 385. Sakurai, H.; Kito, M. Biochhem. Pharm. 1975, 24, 1647-1650. 386. Sakurai, H.; Kito, M. Chem. Pharm. Bull. 1977, 25, 2330-2335. 387. Barton, D. H. R.; Hay-Motherwell, R. S.; Motherwell, W. B. Tetrahedron Lett. 1983, 24, 1979-1982. 388. Barton, D. H. R.; Li, T.; MacKinnon, J. Chem. Commun. 1997, 557-558. 389. Barton, D. H. R.; Li, T. Chem. Commun. 1998, 821-822. 390. Barton, D. H. R.; Boivin, J.; Ozbalik, N.; Schwartzentruber, K. M. Tetrahedron Lett. 1984, 25, 4219-4222. 391. Barton, D. H. R.; Gokturk, A. K.; Morzycki, J. W.; Motherwell, W. B. J. Chem. Soc., Perkin Trans. 1 1985, 583-585. 392. Barton, D. H. R.; Gokturk, A. K.; Jankowski, K. J. Chem. Soc., Perkin Trans. 1 1985, 2109-2117. 393. Barton, D. H. R.; Boivin, J.; Hill, C. H. J. Chem. Soc., Perkin Trans. 1 1986, 1797-1804. 394. Barton, D. H. R.; Beloeil, J. C.; Billion, A.; Boivin, J.; Lallemand, J. Y.; Lelandais, P.; Mergui, S.; Morellet, N.; Jankowski, K. Nouv. J. Chim. 1986, 10, 439-446. 395. Barton, D. H. R.; Boivin, J.; Crich, D.; Hill, C. H. J. Chem. Soc., Perkin Trans. 1 1986, 1811-1814. 396. Barton, D. H. R.; Beloeil, J. C.; Billion, A.; Boivin, J.; Lallemand, J. Y.; Mergui, S. Helv. Chim. Acta 1987, 70, 273-280. 397. Barton, D. H. R.; Beloeil, J. C.; Billion, A.; Boivin, J.; Lallemand, J. Y.; Lelandais, P.; Mergui, S. Helv. Chim. Acta 1987, 70, 2187-2200. 398. Barton, D. H. R.; Boivin, J.; Lelandais, P. J. Chem. Soc., Perkin Trans. 1 1989, 463-468. 399. Barton, D. H. R.; Boivin, J.; Ozbalik, N.; Schwartzentruber, K. M.; Jankowski, K. Tetrahedron Lett. 1985, 26, 447-450. 400. Barton, D. H. R.; Boivin, J.; Gastiger, M.; Morzycki, J.; Hay-Motherwell, R. S.; Motherwell, W. B.; Ozbalik, N.; Schwartzentruber, K. M. J. Chem. Soc., Perkin Trans. 1 1986, 947-955.

4. Functional model oxygenations by nonheme iron complexes

225

401. Barton, D. H. R.; Boivin, J.; Motherwell, W. B.; Ozbalik, N.; Schwartzentruber, K. M.; Jankowski, K. Nouv. J. Chim. 1986, 10, 387-398. 402. Balavoine, G.; Barton, D. H. R.; Boivin, J.; Lecoupanec, P.; Lelandais, P. New J.Chem. 1989, 13, 691-700. 403. Barton, D. H. R.; Csuhai, E.; Ozbalik, N. Tetrahedron Lett. 1990, 31, 28172820. 404. Barton, D. H. R.; Lee, K. W.; Mehl, W.; Ozbalik, N.; Zhang, L. Tetrahedron 1990, 46, 3753-3768. 405. Celenligil-Cetin, R.; Staples, R. J.; Stavropoulos, P. Inorg. Chem. 2000, 39, 5838-5846. 406. Balavoine, G.; Barton, D. H. R.; Boivin, J.; Gref, A.; Ozbalik, N.; Riviere, H. Tetrahedron Lett. 1986, 27, 2849-2852. 407. Balavoine, G.; Barton, D. H. R.; Boivin, J.; Gref, A.; Ozbalik, N.; Riviere, H. J. Chem. Soc., Chem. Commun. 1986, 1727-1729. 408. Balavoine, G.; Barton, D. H. R.; Boivin, J.; Gref, A.; Le Coupanec, P.; Ozbalik, N.; Pestana, J. A. X.; Riviere, H. Tetrahedron 1988, 44, 1091-1106. 409. Mimoun, H.; Seree de Roch, I. Tetrahedron 1975, 31, 777-784. 410. Davis, R.; Durrant, J. L. A.; Khan, M. A. Polyhedron 1988, 7, 425-438. 411. Sheu, C.; Sawyer, D. T. J. Am. Chem. Soc. 1990, 112, 8212-8214. 412. Hage, J. P.; Sawyer, D. T. J. Am. Chem. Soc. 1995, 117, 5617-5621. 413. Funabiki, T.; Tsujimoto, M.; Ozawa, S.; Yoshida, S. Chem. Lett. 1989, 12671268. 414. Funabiki, T.; Toyoda, T.; Ishida, H.; Tsujimoto, M.; Ozawa, S.; Yoshida, S. J. Mol. Catal. 1990, 61, 235-246. 415. Funabiki, T.; Toyoda, T.; Yoshida, S. Chem. Lett. 1992, 1279-1282. 416. Funabiki, T.; Yokomizo, T.; Suzuki, S.; Yoshida, S. J. Chem. Soc., Chem. Commun. 1997, 151-152. 417. Funabiki, T.; Ishida, H.; Yoshida, S. Chem. Lett. 1991, 1819-1822. 418. Funabiki, T.; Kashiba, K.; Toyoda, T.; Yoshida, S. Chem. Lett. 1992, 23032306. 419. Takai, T.; Hata, E.; Yamada, T.; Mukaiyama, T. Bull. Chem. Soc. Jpn. 1991, 64,2513-2518. 420. Murahashi, S.; Oda, Y.; Naota, T. J. Am. Chem. Soc. 1992, 114, 7913-7914. 421. Kesavan, V.; Sivanand, P. S.; Chandrasekaran, S.; Koltypin, Y.; Gedanken, A. Angew. Chem. Int. Ed. Engl. 1999, 38, 3521-3523. 422. Ruiz, R.; Triannidis, M.; Aukauloo, A.; Journaux, Y.; Fernandez, I.; Pedro, J. R.; Cervera, B.; Castro, I.; Munoz, M. C. Chem. Commun. 1997, 2283-2284.

226

T.Funabiki

423. Nam, W.; Kim, H. J.; Kim, S. H.; Ho, R. Y. N.; Valentine, J. S. Inorg. Chem. 1996, 35, 1045-1049. 424. Sam, J. W.; Tang, X.-J.; Peisach, J. J. Am. Chem. Soc. 1994, 116, 5250-5256. 425. Stubbe, J.; W., K. J.; Wu, W.; Vanderwall, D. E. Acc. Chem. Res. 1996, 29, 322-330. 426. Ahmad, S.; McCallum, J. D.; Shiemke, A. K.; Appleman, E. H.; Loehr, T. M.; Sanders-Loeher, J. Inorg. Chem 1988, 27, 2230-2233. 427. Neese, F.; Solomon, E. I. J. Am. Chem. Soc. 1998, 120, 12829-12848. 428. Kim, J.; Larka, E.; Wilkinson, E. C.; Que, L. Angew. Chem., Int. Ed. Engl. 1995, 34, 2048-2051. 429. Bernal, I.; Jensen, M. I.; Hense, K. B.; MacKenzie, C. J.; Toftlund, H.; Tuchangues, J.-P. J. Chem. Soc., Dalton Trans. 1995, 3667-3675. 430. de Vries, M. E.; La Crois, R. M.; Roelfes, G.; Kooijman, H.; Spek, A.; Hage, R.; Feringa, B. L. Chem. Commun. 1997, 1549-1950. 431. Ho, R. Y. N.; Roelfes, G.; Feringa, B. L.; Que, L. J. Am. Chem. Soc. 1999, 121, 264-265. 432. Roelfes, G.; Lubben, M.; Chen, K.; Ho, R. Y. N.; Meetsma, A.; Genseberger, S.; Hermant, R. M.; Hage, R.; Mandal, S. K.; Young, V. G., Jr.; Zang, Y.; Kooijman, H.; Spek, A. L.; Que, L.; Feringa, B. L. Inorg. Chem. 1999, 38, 1929-1936. 433. Mialane, P.; Nivorojkine, A.; Pratviel, G.; Azema, L.; Slany, M.; Godde, F.; Simaan, A.; Banse, F.; Kargar-Grisel, T.; Bouchoux, G.; Sainton, J.; Horner, O.; Guilhem, J.; Tchertanova, L.; Meunier, B.; Girerd, J.-J. Inorg. Chem. 1999, 38, 1085-1092. 434. Lange, S. J.; Miyake, H.; Que, L., Jr. J. Am. Chem. Soc. 1999, 121, 6330-6331. 435. Wada, A.; Ogo, S.; Nagatomo, S.; Kitagawa, T.; Watanabe, Y. Inorg. Chem 2002, 41, 616-618. 436. Mekmouche, Y.; Hummel, H.; Ho, R. Y. N.; Que, L.; Schunemann, V.; Thomas, F.; Trautwein, A. X.; Lebrun, C.; Gorgy, K.; Lepretre, J. C.; Collomb, M. N.; Deronzier, A.; Fontecave, M.; Menage, S. Chemistry-a European Journal 2002, 8, 1196-1204.

Chapter 5 Catalysts for selective aerobic oxidation under ambient conditions Thioether sulfoxidation catalyzed by gold complexes Eric Boring, Yurii V. Geletii and Craig L. Hill Department of Chemistry, Emory University, Atlanta, Georgia 30322

Abstract: Diversity-based methods for catalyst discovery coupled with the knowledge of lead systems for the catalysis of organic oxidation reactions has led to the development of new species that actually catalyze rapid and selective (non-radical-chain), reductant-free, oxidation under ambient conditions (room temperature and 1.0 atmosphere of air). The first process of focus is selective sulfoxidation of thioethers (organic sulfides). The principal work reviewed here involves homogeneous catalysis, but highly reactive heterogeneous formulations have already been identified. The stoichiometry is that characteristic of dioxygenase enzymes: (sulfoxide). Oxidative dehydrogenation, a less desirable net process, is not seen. Studies have primarily been conducted with 2-chloroethyl ethyl sulfide (CEES), which is both notoriously unreactive and a useful simulant for mustard. Extensive kinetics and product studies have identified the active catalyst, at least in acetonitrile solution, to be (1), and the rate limiting step to be reaction of 1 with another molecule of the thioether substrate. Reoxidation of the resulting Au(I) to Au(III) by is a fast subsequent step. The solvent kinetic isotope effect rate of sulfoxidation when Cl is replaced by Br, and multiparameter fitting of the kinetic data establish that the mechanism of the rate-limiting step itself involves a bimolecular attack of CEES on a Au(III)-bound halide and it does not involve Isotope labeling studies with indicate that and not or is the source of oxygen in the sulfoxide product. Interestingly, is consumed and subsequently regenerated in the mechanism. Despite the impressive (unique) reactivity attributes above, these recently developed catalytic systems have some limitations that include an induction period and inhibition by sulfoxide product. However, these two difficulties are eliminated in other solvents or in nontoxic developmentally attractive perfluoropolyether (PFPE) media. Another potential problem, is catalyst inactivation by precipitation of the Au as colloidal Au(0), but this can be largely avoided by use of appropriate reaction conditions. Finally, these Aucatalyzed aerobic sulfoxidation reactions can be co-catalyzed by some d-block ions. Cu(II) is particularly effective in this context resulting in substantial increases in reaction rate at low

227 L.I. Simándi (ed.), Advances in Catalytic Activation of Dioxygen by Metal Complexes, 227-264. © 2003 Kluwer Academic Publishers. Printed in the Netherlands.

228

E. Boring, Yu. V. Geletii and C. L. Hill

Cu(II) concentrations. Co-catalysis by the d-block ions also results in elimination of the induction period in some cases.

Key words: Amino acids, Catalytic oxidation under ambient conditions, 2-chloroethyl ethyl sulfide, Co-catalysis by copper, Decontamination, Effect of ligands, Gold complexes, Mustard gas, Perfluorinated solvents, Solvent effect, Sulfoxide, Thioether

1. INTRODUCTION The selective reductant-free catalytic oxidation of organic compounds by has been a major goal in both fundamental and applied chemistry for years. Nearly all catalysts for oxidations, whether they be biological or abiological (synthetic/industrial) contain redox-active metal centers.1-7 “Selective” generally means that the radical-chain mechanisms that dominate the ubiquitous oxidations of nearly all organic and many inorganic materials (autoxidations) aren’t operable. While there are a handful of radical-chain oxidations that proceed with reasonably high selectivity, it is the nature of the substrate itself in these cases that ensures high selectivity and not the chemistry (control of the elementary processes in the mechanism). One of these rare examples is the oxidation of p-xylene to p-terephthalic acid, one of the largest scale commercial homogeneous catalytic processes globally. Despite proceeding by a homolytic mechanism (radical chains and many radical intermediates), the pterephthalic acid product is produced in a very high selectivity >99.9% primarily because its unusual structure makes it quite stable under the reaction conditions.8 The DuPont process for aerobic oxidation of cyclohexane (adipic acid manufacture) is more indicative. While this also remains a large scale process given its low cost, it proceeds with poor selectivity.9 One of the several reasons for the intrinsically low selectivity (and consequent generically low desirability) of conventional radical-chain oxidations, is that many oxygen-based and/or radical intermediates in such processes are capable of reacting simultaneously with the substrate. Unfortunately the metal centers in such systems, both catalytic and structural, typically exacerbate the problem by generating additional nonselective but kinetically potent reactive intermediates via both redox and Furthermore, the mechanistic electrophilic non-redox processes.10 complexity in conventional radical-chain oxidations, including the

5. Catalysts for aerobic oxidation under ambient conditions

229

ensemble of freely diffusing time-dependent and concentration-dependent reactive intermediates, makes these oxidations difficult to control.9 Many of these limitations are avoided if a reducing agent is used. Indeed, a multitude of biological and abiological/industrial oxidation processes use a reductant because the resulting peroxo and other intermediate-redox-state forms of oxygen generally exhibit greater reactivity. More significantly however, these reactive oxygen species are frequently more selective and their chemistry more controllable, with or without the involvement of metal centers, than the chemistry of the active oxygen and oxygenated intermediate species operable during reductant-free oxidations. However, if a reducing agent is required, its cost must be factored into the overall economics, and this renders many an otherwise potentially attractive process unworkable practically. Given these points, it is no wonder that there has been and is now considerable interest in developing molecules and materials that catalyze the selective (non-radical chain) reductant-free oxidation of many classes of organic substrates using only There are just a handful of homogeneous catalysts for such processes, and despite the fact all of these have one or more significant (rate, selectivity and/or stability) limitations, each of these studies has garnered much attention.1-7 Furthermore, none of these systems functions effectively under ambient conditions (1 atm of air o and ~22 C). This is unfortunate because a major intellectual impetus as well as programmatic or developmental driver of such catalytic chemistry is the realization of materials (coatings, fabrics, cosmetics, others) that catalyze the degradation of the ubiquitous toxic agents in our environment without chemical or physical assistance (without the requirement for heat, light, water, solvents, activators, etc.).11-13 Such materials could economically function in a host of locations (the home, the workplace, the car, etc.) for a host of applications and thus benefit mankind. The toxic agents of relevance include sulfur compounds (thioethers, thiols and with mustard (formula: being one deleterious thioether of much current national and international concern because it is a widely prevalent chemical warfare agent.14-20 Several nitrogen compounds, including pyridine, nicotine, trimethyl amine, aldehydes, halogenated compounds and other volatile organic compounds (VOCs) also constitute everyday threats to human health.11-13,21 This chapter summarizes quite recent research on a new type of system based on the coinage metals that does in fact catalyze selective reducingagent-free oxidation of an important target substrate, 2-chloroethyl ethyl sulfide (henceforth CEES for convenience), a very effective simulant for mustard.22,23 The review addresses the genesis of this new type of system and its systematic experimental elaboration including a quite thorough analysis of mechanism. In this context, the selectivity (virtually quantitative

230

E. Boring, Yu. V. Geletii and C. L. Hill

for the desired minimally-toxic sulfoxide product, CEESO), the stoichiometry and the evidence for a non-radical chain mechanism are presented and discussed. This work represents a foray into “functionally smart” materials, materials that not only sense and adapt to ambient conditions but also execute important functions when appropriate. In this instance, the function is catalyzing the facile oxidative degradation of toxic sulfur compounds if and whenever they are present.

2. DISCOVERY OF CATALYTIC OXIDATION SYSTEM By combining heteropolyanions (polyoxometalates; POMs for convenience) and selected cations in acetonitrile, more than 150 combinations were assayed for their catalytic activity towards selective CEES oxidation to CEESO by dioxygen under ambient (room temperature and atmospheric pressure) conditions. The main criteria in choosing POMs were their ability to undergo reversible redox transformations and to catalyze homogeneous oxidations either by peroxides or other terminal oxidants. The cations chosen included redox-active transition metal ions or cations conventionally used as the counterions in POMs. In control experiments the chloride, nitrate or perchlorate salts of the same transition metal ions were also examined. The list of these catalytic systems and some selected results were recently published.22 Most of the screened combinations (catalysts) showed little if any catalytic activity. Only two catalysts exhibited considerable activity, both of which included The catalysts contained 1 equivalent of or and five equivalents of A mixing of with these POMs resulted in a formation of a white precipitate, which elemental analysis showed to be NaCl. It was hypothesized that removal of a chloride ligand from produced an active catalyst. To explore this idea Ag(I) a stronger halide abstractor in acetonitrile than Na(I), was used to remove a chloride anion from the Au center. The rate of CEES oxidation by this POM-free system based on Au and Ag was significant, and by varying the ratios of and the rate was increased several-fold. Interestingly, the Au(III)/Ag(I) system appeared to be inactive when was replaced with Subsequently, by varying the ratios of different Ag(I) salts it was established that the rate was very sensitive to the Au(III): ratio. A detailed study revealed that the most active catalyst was formed when and were combined in a 1 : 1 : 1 ratio, suggesting a formation of the complex. This catalytic system is significantly faster than the two most reactive catalysts in the literature for

5. Catalysts for aerobic oxidation under ambient conditions

231

the

homogeneous oxidation of thioethers, namely and The Ru-based catalyst oxidizes tetrahydrothiophene (THT) to tetrahydrothiophene oxide (THTO) at elevated temperature and ~8 atm of with a rate of only 3.4 turnovers (at 110 °C). The latter catalyst is more active, producing 17.6 turnovers of THTO after 30 min, but the conditions are at elevated temperature and pressure (60 °C and 14 bar of Both these systems are completely inactive under ambient conditions. In contrast, our first unoptimized system was producing nearly comparable turnovers to the Cebased system (35.4 equivalents of CEESO per Au(III) after 4 hr) under ambient conditions. It’s worth mentioning that CEES is significantly harder to oxidize than THT, both thermodynamically and kinetically, since it is less reducing and less nucleophilic.16 The selectivity (CEESO formed based on consumed CEES) was also noteworthy.16 CEESO was the only product formed in the reaction, and further oxidation to did not occur. The simultaneous measurement of dioxygen and CEES consumption, as well as CEESO formation revealed the following reaction stoichiometry:

Given the unprecedented reactivity and selectivity of the discovered system (mixture of Au(III) and Ag(I) salts), we chose it to investigate the reaction kinetics in detail. The principal system of focus comprises 1, 0.75, and 1.25 equivalents of the and precursors respectively. The general features of this catalytic system, the complex rate law and other kinetic features of this reaction, and the reaction mechanism are reported below.

3. STOICHIOMETRIC Au(III) REDUCTION BY THIOETHERS Stoichiometric thioether oxidation by Au(III) complexes has been extensively studied in the literature. The major features of this reaction are the following: (1) is not reactive towards thioethers under ambient conditions; (2) halide and thioether ligands exchange in rapid pre-equilibria; (3) the rate-limiting redox step involves thioether reduction of Au(III) forming Au(I) and sulfoxide; (4) the formation constant for the Au(III) complex with 1 thioether ligand > the formation constant for the Au(III) complex with 2 thioether ligands (5) Au(III) complexes with 3 thioether ligands are not known; and (6) the lability of the transient Au(III) complexes renders them nonisolable.26-30 Our Au(III)-based catalytic

232

E. Boring, Yu. V. Geletii and C. L. Hill

system for oxidation shares all the above features but has additional components and therefore is more mechanistically complex.

4.

IN SITU CATALYST PREPARATION

As is the case with both Ag salt precursors, and by themselves are completely inactive as catalysts. Moreover, the addition of a soluble Ag(I) salt (e.g. or to does not result in precipitation of AgCl, as expected via eq 2:

However, the subsequent addition of CEES to this Au(III)/Ag(I) solution leads to the immediate disappearance of the yellow chromophore of Au(III), and solid AgCl forms (confirmed by elemental analysis after isolation). The stoichiometry for this catalyst preparation reaction was determined to be eq 3:

If solid AgCl is filtered off and the solvent is removed, the remaining solid contains Au, CEES and/or CEESO moieties but no Ag. Attempted recrystallization of this material was unsuccessful, and Au(III) was reduced to a catalytically inactive Au(0) colloid. It is observed that once Au(0) forms in the system, it is not reoxidized, leading to an irreversible inactivation of the catalyst. However, use of the correct ratios of precursors in the presence of or even ambient air, produces soluble and stable Aubased catalysts, some of which retain activity for at least one week. Most Au(I) complexes, with the exception of disproportionate in and several disproportionation equilibria have been measured.31-34 Although disproportionation is less favourable kinetically in solution, it is possible that the catalyst inactivates by hydrolysis of the Au(I) intermediate followed by its reduction to form Au(0).35 In acetonitrile, catalytic thioether oxidation may exhibit a short induction phase which is followed by the main reaction (Figure 1). This induction period depends on the concentrations of CEES, Au, and At higher concentrations the induction period is shorter (Table 1 shows the dependence of the induction period). During the induction period the system remains colorless, neither CEES nor are consumed, and no CEESO is formed. At the end of the induction period, the solution rapidly becomes yellow. The electronic absorption spectrum of this yellow species is fully

5. Catalysts for aerobic oxidation under ambient conditions

233

consistent with a conventional square-planar Au(III) complexes.36 The concentration of Au(III) remains constant after the induction period, i.e. during the main reaction.

5. REACTION STOICHIOMETRY Once the catalyst is formed in situ (eq 1) and reoxidized by dioxygen during the induction period to the active Au(III) form, the main catalytic reaction then proceeds. The stoichiometry of the reaction, eq 1, was established by quantifying consumption, CEES consumption and CEESO formation. In general, other stoichiometries for thioether oxidations are also possible including the more common "oxidative dehydrogenation" process (eq 4 for substrate = These reactions are more favourable for

234

easily dehydrogenated hydrocarbons.

E. Boring, Yu. V. Geletii and C. L. Hill

substrates

including

alcohols

and

aliphatic

However, substrates with oxidizable lone pairs (such as thioethers, RSR) are usually oxidized through “oxygen atom transfer” mechanism, eq 5:

Peroxides, high-valent transition metal oxo species, or peroxo radicals are the typical “oxygen donor” reactive species, “O”, in sulfoxidation reactions.38-40 A conversion of to peroxo species or other forms of reduced oxygen requires a consumption of a 2-electron sacrificial reducing agent, The most frequently used reductants are NADH, NADPH (particularly in enzyme-catalyzed oxidation processes) or ascorbate and thiols. An exemplary “monooxygenase” stoichiometry of sulfoxidation by dioxygen is eq 6:

The "dioxygenase" stoichiometry, eq 1, is optimal because all of the oxidizing capacity of dioxygen is used and both oxygen atoms are accounted for in the desired product. For both stoichiometries, eq 1 and 6, the question arises whether the oxygen atom in the sulfoxide originates from dioxygen or from water, which is always present in our system in small amounts. The experiments with unambiguously showed that our Au-based catalytic oxidation always resulted in the formation of 100% labeled (the mechanism and details are discussed below). Determination of the stoichiometry also involves the product selectivity since sulfoxides can undergo further oxidation to the corresponding sulfones. This is particularly important in our work since mustard sulfoxide is much less toxic than sulfone and the sulfoxide is not a vesicant while both the sulfide and sulfone are.41 In our case, the selectivity for CEESO is exceptionally high and the sulfone, has never been detected. It is worth mentioning that a high selectivity is commonly observed at low conversions for most oxidation reactions. In acetonitrile, total CEES conversion was impossible to achieve due to the inhibition of the reaction by the CEESO product. However, the same catalytic system with trifluoroethanol as a solvent (see Section 15) can completely oxidized CEES, while the selectivity for CEESO formation was close to 100%.

5. Catalysts for aerobic oxidation under ambient conditions

235

6. EMPIRICAL REACTION RATE LAW Typical kinetic curves of CEES consumption and CEESO formation are presented in Figure 1. The dotted, dashed and solid lines in Figure 1 arise from fitting of experimental data to the proposed mechanism and are discussed below. While it may not be readily apparent, the kinetics of the main reaction after the induction period does not obey any simple kinetic law. The reaction gradually slows down as a result of a mild inhibition by CEESO product. The dotted line is a simple exponential assuming no such inhibition takes place. The primary experimentally determined rate parameter was +d[CEESO]/dt, but -d[CEES]/dt and were also evaluated in many cases. These values are indicated henceforth as "rate". The reaction rate was determined as a function of the concentrations of CEES, Au(III), Ag(I), together in a constant 1:2:1 mole ratio, DMSO as a model for the CEESO product and

Figures 2-4 and Table 2 give the rate dependencies of the main reaction, eq 1, on the concentrations of CEES, total Au(III), and DMSO. The maximum rate is achieved at Au : : ratio 1 : 2 : 1 , suggesting that the dominant transition state complex contains one and two groups. Further, it was established that replacement of by or generated species with little or no catalytic activity. In contrast, replacement of by formed a more reactive complex, confirming a significant and specific role for (or in the active catalyst (Figure 2). The rate increases with the concentration of the thioether

236

E. Boring, Yu. V. Geletii and C. L. Hill

with saturation observed at high CEES , suggesting a replacement of a and/or ligand by CEES to form an inactive Au(III)-complex.

The presence of both and appear to be critical for catalytic activity. Since the activity of the catalysts depends on the ratio of Au : : this ratio was kept constant at 1 : 2 : 1 in the determination of the dependence. The rate increases quadratically with with the reaction rate order 1 < n < 2. The dependence (see Figure 7 in our recent publication22) is inversely hyperbolic, which is consistent with once again competing with and/or CEES for open coordination sites on the catalytically active Au center. This, in turn, decreases the concentration of the catalytically active Au complexes. The rate of the main reaction is independent of concentration (Table 1), implying that Au(III) reduction and not Au(I) reoxidation is rate limiting.

5. Catalysts for aerobic oxidation under ambient conditions

237

7. RATE LIMITING STEP A zero order reaction rate with respect to dioxygen concentration suggests that the rate limiting step is Au(III) reduction by CEES, not reoxidation of Au(I) by dioxygen. For futher proof of the rate limiting step, we determined whether Au(I) and/or Au(III) was the dominant redox state of Au during the main reaction. This was assessed by measuring the absorbance for Au(III) after the induction period as function of CEES and concentrations, while all other reactants and conditions were kept constant. The electronic absorption spectrum (350 - 450 nm) of Au-complexes during the main reaction is very similar to that of conventional isolable square-planar Au(III) complexes for is colorless After the induction period, the observed spectra and their intensity remained unchanged and were independent of both reactants ([CEES] from 0.05–0.37 M and from 50100%) within experimental error. Thus, the aggregate Au species during catalytic CEESO formation was almost entirely in the form of Au(III). Based on the two definitive results described above, it became obvious that the reduction of Au(III) species and not the oxidation of Au(I) species is rate limiting in these aerobic oxidation reactions.

238

E. Boring, Yu. V. Geletii and C. L. Hill

It is worth mentioning that thioethers are catalytically oxidized to sulfoxides in nitromethane-aqueous nitric acid in the presence of with the rate-limiting step being the reoxidation of Au(I) to Au(III).28 This involves the use of an excess of relative to Au; however, in our systems and not is the terminal oxidant. This follows from the fact that only one equivalent of is used in our reactions, but 200 equivalents of CEESO product per equivalent of can be obtained.

8. PROPOSED REACTION MECHANISM As mentioned above, Figure 2 indicates that the most catalytically active species contains in a 1 : 2 : 1 ratio. Dimeric or oligomeric d8 square planar Au(III) complexes are very unlikely,43 therefore a monomeric Au(III) complex with 2 ligands and 1 ligand has the capacity to bind one more ligand strongly, most likely a CEES ligand. Five-coordinate Au(III) complexes are very rare. The only 5-coordinate Au(III) complexes with appreciable stability contain chelating ligands such as bromodicyano(l,10-phenanthroline)gold(III) isolated from dimethylformamide.44 If there is an interaction between a positive Au(III) center and a negative axial ligand counterion, this ligand would likely be because associates very weakly with Au(III) (and not at all with Au(I)) and probably can be ruled out in our case. A proposed mechanism that is compatible with the stoichiometric thioether-Au(III) reactions in the literature and all our data is given in Scheme 1. Literature data on the stoichiometric Au(III) reduction by thioethers provide evidence for the rapid exchange of all ligands on Au(III) prior to thioether oxidation.26,27,45-47 In our mechanism these pre-equilibria are summarized by eqs 7 and 8 in Scheme 1. Mixed Au(III) complexes are formed with and CEES: a complex with one CEES ligand (1), with formation constant and a complex with two CEES ligands (2), with formation constant These ligands drive eq 7 to the right, while and CEESO (or DMSO, data in Table 2), drive it to the left. Additionally, or other solvent molecules may also bind to Au(III) or shift both equilbria, eqs 7 and 8. Au(III) complexes with one thioether ligand, analogous to 1, are considerably more abundant in solution than complexes with two thioether ligands, such as 2 Au(III) complex with three thioether ligands are not considered because there is no data indicating these species form in solutions

5. Catalysts for aerobic oxidation under ambient conditions

239

As discussed above, all the available information (section 7) is consistent with the rate limiting redox step being the simultaneous oxidation of thioether and reduction of Au(III). A mechanism involving the rate-limiting formation of the required Au(III) complex prior to undergoing redox chemistry is not supported by any of the literature studies. The kinetics and equilibrium investigations establish that in the oxidation of and by and of by the slowest step is the bimolecular reaction of thioether with Au(III) complexes containing one or two thioether ligands. Both these systems are stoichiometric; they do not undergo catalytic turnover. These systems differ from ours as they do not contain and are used in polar, protic solvent systems like 95:5 or 100% water.

240

E. Boring, Yu. V. Geletii and C. L. Hill

In our studies polar, aprotic solvents such as have primarily been used (see also sections 15 and 16). Nevertheless, the Au(III)-thioether coordination and the redox chemistry in the two systems are otherwise very similar. The two reactions depicted in eq 9 have the rate constants for 1 and for 2. Since 1 is the most reactive complex, eq 9 is written in terms of 1 and its redox products, and The subsequent steps, eqs 10 and 11, are fast and consequently kinetically inaccessible processes. Of the several possible mechanisms for the rate-determining redox step, eq 9, but one is more consistent with all available data (see the next section for details) and involves bimolecular attack of thioether on a coordinated chloride ligand of Au(III). This mechanism has been proposed for some stoichiometric thioether oxidations by Au(III).45,46,48 None of the data on our system contradicts this mechanism. It is also consistent with the increase in rate when is replaced by under otherwise identical conditions (see also section 12). Inner sphere ligand transfer redox processes, including reduction of Au(III), are known to proceed faster for compared with ligands.46, 49 For example, is reduced by sulfite and hydrogen sulfite ca. 10 times faster than Interestingly, this kinetic preference is in the opposite direction from the reaction enthalpy since the metal centers with the have higher potentials and are stronger oxidants than their Since is large and then The equilibrium expressions from eqs 78, Scheme 1, and reaction mass balance expressions afford eqs 12 and 13.

Eqs 13 -15 can be used to obtain the expressions for [1] and [2], in eqs 16 and 17.

5. Catalysts for aerobic oxidation under ambient conditions

241

Combining eqs 11, 16 and 17 gives eqs 18 and 19 for the reaction rate.

Eq 19 has been used to fit the experimental data, namely the dependencies of rate versus [CEES] and The best fit is a solid curve shown in Figure 4 with and The errors were estimated at the 95% confidence limit. Since the complex eq 19 simplifies to eq 20.

It is evident without curve fitting that and complex 2 shows a low catalytic activity towards CEES oxidation. At high [CEES] nearly all Au is present as Au(III), and eq 19 then simplifies to eq 21 (details are in Boring, et al.)22:

Figure 4 shows saturation kinetics at [CEES] > 0.4M, which clearly suggests that Thus, the CEES substrate slows the reaction by shifting eq 8 to the right in favor of the less reactive 2 by substituting CEES for and/or from the more reactive 1. Eq 20 also demonstrates how sensitive the overall reaction rate is to a change in and While the rate is linearly proportional to the dependence on is more complex and varies with and [CEES] as illustrated in Figures 3 and 4 by dashed and dotted lines computed at different values of The effect of solvents and ligand substitution on and is addressed below. The dependence of the rate on Figure 3, was also fitted to eq 20, the solid curve is with and which are in good agreement with those obtained from the fit of the rate vs. [CEES] data and Surprisingly, our value for is very similar to that for complexation of to in 5% aqueous methanol: However, in the latter system is ~ 12.27 While these two systems are very similar, they have different ligands (no and more in the latter one) and solvent.

242

E. Boring, Yu. V. Geletii and C. L. Hill

After the slow Au(III) reduction by thioether, the process involves conversion of the oxidized sulfur center to the sulfoxide. The first intermediate of this conversion is a chlorosulfonium ion (eq 9, Scheme 1). These species are known to hydrolyze rapidly to yield sulfoxides.50 The study with isotope labeled is entirely consistent with this mechanism (see below). As mentioned above, the reaction slows down with time due to inhibition by CEESO products. CEESO replaces or in the active complex 1 and gives an inactive complex analogous to 2. This inhibition has been modeled by DMSO, a molecule structurally and electronically similar to the CEESO product, and is addressed below (section 13). A full evaluation and nonlinear fitting of the kinetics of self-inhibition by CEESO is discussed in a recent paper.23 The result of this fitting is shown as a solid line in Figure 1, and the anticipated CEES consumption (an exponential decay with if there were no inhibition by CEESO is given as a dotted line (see below).

9. MECHANISMS RULED OUT Several other possible mechanisms for the rate-limiting redox step are discussed below and ruled out. The first one assumes that a slow formation of the required Au(III) complex (for example, 1) is the rate limiting step, and redox transformation is fast. However, this mechanism is not supported by the literature studies on stoichiometric Au(III) reduction by thioethers, nor by the kinetic data obtained in this research. Clearly the reduction of Au(III) is the slow step, but this in turn may proceed through different possible mechanisms. One mechanism for Au(III) reduction by thioethers involves the intramolecular reductive elimination of 1 and 2 to form a thioxonium salt (eq 9 in Scheme 1) and a Au(I) complex, eqs 22 and 23.

5. Catalysts for aerobic oxidation under ambient conditions

243

This mechanism has been proposed for some thioether oxidations by Au(III)26,45,47 and the conversion of 2 to and CEESO can be assisted by eq 24.26 However, this mechanism is also very unlikely in our catalytic system; analysis of the reaction kinetics unambiguously rules it out. If eqs 22 and 23 were rate limiting, then the overall rate would be eq 25.

Combining eqs 16, 17 and 25 gives eqs 26 and 27 for the reaction rate.

This rate law is incompatable with our experimental kinetic data. If then the reaction rate should increase quadratically with which is in agreement with Figure 3, while the reaction rate should decrease with increasing [CEES], which is inconsistent with our data (Figure 4). then the rate should increase with both and [CEES] finally reaching saturation with an observed reaction rate order < 1. This is in disagreement with the data presented in Figure 3. Thus, only one of the rate dependences, either that for [CEES] or that for can be explained by this mechanism, but not both simultaneously. Incorporation of eq 24 in the above analysis leads to the same expression as eq 27, where is replaced by Thus, the addition of eq 24 into the reaction mechanism does not affect the theoretical reaction rate law, eq 27. Additional evidence against the unimolecular collapse of 2, assisted by a water molecule, being the rate limiting step (eq 24) derives from a nonexistent kinetic isotopic effect when was used instead of The kinetics data are in accord with functioning as an inhibitor competing with or CEES for a coordination site in the Au(III) complexes and thus driving eq 7, Scheme 1, to the left. In principal, a radical-chain thioether oxidation initiated by Au(III) complexes may also take place. Such a mechanism has been proposed by Riley et al.25 for aerobic thioether oxidation catalyzed by Ce salts. This sulfoxidation reaction proceeds in at elevated temperatures (> 70 °C). The proposed mechanism includes the reduction of Ce(IV) by thioether to form the radical cation which is efficiently trapped by to give The latter radical, being a strong oxidant, reoxidizes Ce(III) to

244

E. Boring, Yu. V. Geletii and C. L. Hill

Ce(IV) and forms another reactive intermediate which in turn reacts with to produce 2 molecules of sulfoxide product. According to this mechanism, the oxygen atom in the sulfoxide should be incorporated from not Since in our system 100% oxygen-18 incorporation from was observed, such a radical-chain mechanism can be definitively ruled out. Additional evidence against this mechanism derives from a competitive oxidation of different thioethers (see section 18). Au(III)-based catalysts appear to be highly discriminatory with respect to thioethers when a combination of thioethers are oxidized together. The high-energy intermediate, proposed to be generated in the Ce-based systems, would likely not able to discriminate highly between different thioethers.25 Finally, another possible mechanism for aerobic sulfoxidation is proposed by Riley and co-workers for the reaction catalyzed by This chemistry proceeds at rather high temperatures (>100 °C) and requires the use of alcohols as solvents. The reactive species in this system is which is produced as an intermediate in the oxidation of Ru(II) to Ru(IV) by Thus, the Ru(IV) formed is reduced back to Ru(II) by alcohol, a solvent molecule. The labelling studies in this system show that no oxygen-18 incorporation from occurs, in direct and total opposition to our Au(III)-based catalytic system (100% incorporation from Additional evidence against such a mechanism derives from the stoichiometry itself (eq 1 in our case versus eq 6 in Ru-based system) and selectivity (~100% yield of CEESO at high CEES conversion, but further sulfoxide oxidation to sulfone in the Ru system studies.

10. ORIGIN OF OXYGEN IN SULFOXIDE PRODUCT; ROLE OF IN SULFOXIDATION The simplest procedure for sulfoxide synthesis is thioether oxidation by and acids are efficient catalysts for this reaction.51 In the presence of protons is a potential intermediate in Au(I) reoxidation by eq 28:

HC1 and are formed during the catalytic process, eq 9, and one of the reactants, is acidic. Thus CEES could be, at least partly, oxidized by formed during this process. To assess the role of and to address the formation of a thioxonium salt intermediate (by establishing the incorporation of oxygen-18 into CEESO from isotope labeling studies were done by replacing regular with In addition, CEES oxidation by in the presence of our Au(III)-catalysts was also assessed.

5. Catalysts for aerobic oxidation under ambient conditions

245

Oxidation of CEES to CEESO by alone occurs very slowly (if at all), but the reaction is efficiently catalyzed by p-toluenesulfonic acid. In the presence of this strong acid or our Au(III)-based catalyst, the reaction proceeds upon mixing. Thus, if is formed as an intermediate in our system (eq 11), it should immediately react with CEES. In this case one half of the total CEESO formed in a catalytic reaction could derive from reaction of with CEES. However, oxygen-18 labeling experiments show that the CEESO produced in the early part of the reaction (up to 10 turnovers) is and The somewhat lower in the CEESO than in the used derives in part from water present in the initial This Au catalyst precursor compound is hygroscopic and very hard to dry. The percentage of in the CEESO also decreases with an increase of CEES conversion. For example, the incorporation into CEESO product was ~60 atom % after 60 turnovers. The dependence of incorporation percentage on the extent of CEES conversion is consistent with a consumption of labeled water during catalytic turnovers. The water concentration remains unchanged in the overall reaction: one water molecule is consumed in eq 9 but it is immediately regenerated in eq 11 resulting in a dilution of pool with Thus, CEESO is produced in the reaction involving exclusively but not or If alone were involved, the ratio would not exceed 50%.

11. REOXIDATION OF Au(I) BY DIOXYGEN. CATALYST PREPARATION FROM Au(I) COMPLEX The isotope labeling and kinetic studies of CEES sulfoxidation by reveals that reoxidation of Au(I) by in the presence of protons does not proceed via eq 28, and is not formed as an intermediate. As indicated above, this re-oxidation reaction is not a rate limiting process, therefore its mechanism can not be assessed kinetically. This reaction is also not precedented in the literature. Au(I) forms linear, trigonal planar or tetrahedral complexes, with the linear geometry being the most common.43 The complexes and are colorless (for lignad, easily prepared by reduction of in ethanol and isolable with bulky ligands 52, and stable in acetonitrile.33,34,42 The complex is of considerable commercial importance, since it is used in the extraction of gold from its ores and in gold electroplating applications. It is polymeric with a linear Complexes of Au(I) with thioethers of formula are usually prepared by reduction of in the presence of appropriate dialkylsulfide, eq 29.29,43,54

246

E. Boring, Yu. V. Geletii and C. L. Hill

Alternatively, the complexes can be prepared by direct addition of the thioether to In solutions with an excess thioether, the thioether ligand is in rapid exchange with In (1:1) the complex, is formed after the addition of 2 eq of thioether to 1 eq of This suggests that thioethers bind strongly to Au(I). The equilibrium constants and have been determined for the displacement of in by and other ligands33,42, and log is 10.6 for the latter one. Because Au(I) is a 2-electron reducing agent while is a 4-electron oxidizing agent, the reaction between them requires either formation (in protic media) or the formation of other intermediates. So far no experimental data on this re-oxidation reaction are available, therefore only a highly speculative mechanism could be proposed, Scheme 2:

5. Catalysts for aerobic oxidation under ambient conditions

247

Based on structural properties of Au(I) complexes, the formation of a binuclear complex 4 bridged by two Cl-ligands with a linear moiety is reasonable, eq 30, Scheme 2. Dioxygen binding to this complex promoted by the presence of two thioether ligands results in the formation of a mixed Au(III)/Au(I) complex 5, with a square-planar peroxo-Au(III) and linear Au(I) moieties. Mixedvalent Au-complexes are precedented in the literature. The vibrational spectra of some dialkyl sulphide complexes of gold (III) and gold (I) halides have been reported.56 A mixed-valence complex of gold with dimethyl sulfoxide has been reported.57,58 The unimolecular rearrangement of 5 results in the formation of a complex 6, eq 33. Transfer of terminal oxygen atom from the peroxo-group to the sulfur atom of the thioether, eq 32, can be definitively ruled out since it would result in only 50% incorporation of oxygen-18 from labeled water during the catalytic reaction. A heterolytic cleavage of O-O bond in 6, assisted by protons, would result in the formation of the catalytically active complex 1, eq 34, without intermediate formation. Scheme 2, while highly speculative, is compatible with the structural properties of Au(I) and Au(III) complexes in the literature and our current experimental data. Scheme 2 suggests that an active catalyst can be prepared starting from a Au(I) complex and proper amounts of nitrate and chloride salts in the presence of a strong acid. We attempted to prepare such a catalyst by mixing p-toluenesulfonic acid, and CEES in acetonitrile under oxygen and have found that the rate of catalysis increases with increasing acid concentration. For example, the reaction occurs very slowly if at all when no acid is present, but reactivity is clearly visible when one equivalent of p-toluenesulfonic acid acid is present. When three equivalents are used, the reactivity is comparable to the system.

12. EFFECT OF LIGANDS ON REACTIVITY As described above and in recently published mechanistic studies22 a ratio of catalyst components of 1 produced the most reactive catalyst, indicating that is necessary for high catalytic activity. To assess this necessity, was systematically replaced with other ligands and Surprisingly, the stoichiometric substitution of with produced a catalyst with ~2fold higher activity than in the case of (Figure 2). However, this new system was not thoroughly investigated because is more toxic than and would be less desirable for many applications. Other ligands led to

248

E. Boring, Yu. V. Geletii and C. L. Hill

completely inactive catalytic systems. Thus, or are definitely necessary for high catalytic activity. Above it was noted that the replacement of with increases the initial reaction rate. This increase in rate is consistent with the involvement of a ligand transfer reaction in the rate limiting step, eq 9 in Scheme 1, in the proposed mechanism (for detailed discussion see section 8 and 9). However, there is another factor, which considerably affects the overall reaction rate. In acetonitrile at [CEES] > 0.1 M a considerable part of total Au(III) is in the form of the inactive complex 2, which results in a saturation of the initial rate with increasing CEES concentration (Figure 4). Eq 20 perfectly describes this complex experimental dependence (solid line on Fig 3). Because a ligand binds more strongly than to Au(III)58, (eq 8, Scheme 1) is lower for the Br-Au complex than the Cl-Au complex. Nitrite is a softer ligand than nitrate, and subsequently it also binds stronger to Au(III) (which prefers to complex with soft ligands) than nitrate.59 Several gold (III) nitrite complexes have been identified60-62, but none have been isolated. Thus, both and evidently drive eq 8 to the left thereby decreasing Changing the value of has a dramatic effect on the reaction rate according to eq 20. The magnitude of this effect also depends on CEES and concentrations and is demonstrated by dashed and dotted lines in Figures 3 and 4. These theoretical dependencies are computed using different values for It is clearly seen that a decrease in the equilibrium constant, results in a considerable increase in the reaction rate. Thus, higher activity for and is at least partly explained by a decrease of for these ligands. A more thorough investigation shows that this ligand replacement also reduced the induction period and decreased the inhibition of the overall reaction by CEESO product (Figures 2). As a consequence, a significantly higher conversion of CEES to CEESO can be achieved in the system with a ligand. Inhibition of the reaction by CEESO product arises from formation of the inactive complex 2’ with the sulfur ligands, analogous to 2, but with one of the CEES ligands replaced by a CEESO22 ( see Scheme 3 and section 13 below). The analogous complex with two sulfoxide ligands, 2”, is also inactive. A replacement of bromide or nitrite with CEESO in a complex similar to 1 can be described by equilibrium constants and which are analogous to Because a replacement of bromide or nitrite by CEESO is less favourable is lower for bromide or nitrite than for chloride and nitrate), product inhibition is less pronounced. An induction period in CEES oxidation is likely to be the result of slow catalyst formation during the reoxidation of Au(I) by dioxygen at the beginning of the reaction.22 However, since reoxidation is not a rate limiting step during the main process, a kinetic evaluation of the ligand effect on the rate of Au(I) reoxidation is not possible.

5. Catalysts for aerobic oxidation under ambient conditions

249

13. PRODUCT INHIBITION (DMSO EFFECT) As noted above, the reaction slows down as product CEESO accumulates (Figure 1). Inhibition by CEESO product has been proven by using DMSO as a model and also by nonlinear least squares fitting of kinetics of CEES consumption.22 However, a more thorough evaluation of the DMSO dependence reveals that at low DMSO concentrations (< 0.03 M), the initial reaction rate increases to a maximum and then subsequently decreases with increasing [DMSO]. A maximum rate is observed at ~10 mM DMSO (Table 2). A more complex model compared to the one previously described22 to interpret this dual effect of DMSO (initial acceleration followed by inhibition) is given in Scheme 3.

Equilibria for the reaction of DMSO with Au(III) are similar to those of CEES with Au(III). DMSO forms complexes 7, 2’ and 2”. Complex 7 is an

250

E. Boring, Yu. V. Geletii and C. L. Hill

analogue of 1, while 2’, 2” and 3’ are analogues of 2 and 3, respectively. Complexes 3 and 3’ are exchanging CEES and CEESO ligands in fast equilibria.

Equilibria and mass balance expressions lead to eqs 39-40 for the concentrations of 1 and 723:

where Eqs 39 and 40 give eq 41 for the reaction rate law:

Nonlinear least squares fitting of initial rate versus [DMSO] dependencies at different initial CEES concentrations (some data are given in Table 2) is in agreement with the experimental data (Figure 4 in Boring et al.)23. The results of this fitting show that is roughly three orders of magnitude higher than In summary, two DMSO properties explain its dual effect on the reaction rate: its stronger binding to Au(III) compared with thioethers and the higher reactivity of Au(III)-sulfoxide complex towards thioether oxidation than the analogous Au(III)-thioether complex DMSO is a stronger electron withdrawing ligand compared to CEES. Therefore, electron density on the Au center in complex 7 is lower resulting in a higher potential than for 1, consequently Indeed, a study of the square planar complexes (where L are different thioethers and their sulfoxides), which is isoelectronic to Au(III), reveals that the potentials of the sulfoxide complexes are 200-250 mV higher than those the corresponding thioether complexes.63 Crystal structure studies of and provide additional evidence. The structures indicate that a covalent bond between DMSO and Pt(II) is stronger than between DMS and Pt(II). Steric hindrance may be another factor in the preferential binding of DMSO relative to CEES (CEES is more bulky). Because sulfoxides have a stronger affinity to Au(III), they form inactive 2’ and 2” complexes at lower concentrations and thus the inhibition is observed when [DMSO] < [CEES]. CEESO, the product of CEES oxidation, forms a similar sulfoxide complexes 2 ’” ( eq 42).

5. Catalysts for aerobic oxidation under ambient conditions

251

At a higher conversion of CEES, more inactive Au(III)-complexes are formed. As a result the reaction slows down faster than expected accounting only for CEES consumption. In other words, the reaction is self-inhibited by the reaction product, CEESO. This self-inhibition is clear in Figure 1, a line with (+) symbols represent the kinetics of CEES consumption assuming that no selfinhibition takes place (exponential kinetics). The extent of self-inhibition is evidently dependent on the binding properties between sulfoxides and Au(III), represented by the equilibrium constant (eq 42). A stronger binding of the sulfoxide to Au(III) (higher results in a stronger self-inhibition. Dashed and dotted lines in Figure 1 represent the theoretical curves describing CEES consumption using different values of The best fit is obtained when (a solid line in Figure 1). The formation of inactive complexes is accompanied by a complexation of a nitrate anion and a large Au(III) cation, 2’,2” or 2’” . The formation of these charged species should be more favourable in a polar solvent such as This explains why self-inhibition is more pronounced in than in (for additional discussion see section 15). Au(III)-Br complexes are larger than similar Au(III)-Cl complexes because the ionic radius of in crystals is higher than that of 1.96A and 1.81, respectively. The enthalpy of solvation also decreases in the order implying that the formation of the analogous bromide complexes, 2’ and 2” is less favourable. Therefore, as mentioned in section 12, product inhibition is less severe when a Cl-ligand is replaced with Br.

14. CO-CATALYSIS BY TRANSITION METAL IONS Since the reduction of a Au(III) complex by CEES is the rate limiting step, it was speculated that transition metal ions in high oxidation states might catalyze the reduction of Au(III). Both Fe(III) and Cu(II) were screened for their activity in acetonitrile with and ligands. Both these metals eliminate induction period (Figure 5), but only Cu(II) gave a rate enhancement (i.e. co-catalyst activity), particularly in the presence of (see Figure 2 in recent article).23 For example, the initial oxidation rate at ~3.0 M CEES is approximately 2 times faster for than for while is 8 times faster than It is important to note that addition of Cu(II) salts to the system eliminates the inhibition effect of the sulfoxide product, allowing for almost the complete oxidation of CEES to CEESO. Co-catalysis by a copper salt is even more

252

E. Boring, Yu. V. Geletii and C. L. Hill

pronounced in heterogeneous systems if perfluorinated polyethers (PFPE oils) are used as a solvent (see section 16). This co-catalysis by Cu(II) can be attributed to the formation of Au(III)-HalCu(II) complexes, where Hal is or The oxidation of CEES by this Au(III)/Cu(II) is more facile due to the additional withdrawal of electron density from Au(III) by Cu(II) (a more electropositive Au center results in an increase in the overall reaction rate).

Au(III)-Hal-Cu(II) moieties are precedented in the literature. For example, a copper(II) chloroaurate(III) complex with a Au(III)-Hal-Cu(II) unit is formed by the neutralization of with and has been characterized (GMELIN registry number 177659). Additionally, a withdrawal of electron density from Au(III) by Cu(II) makes the Au(III)ionic bond stronger, resulting in a less favourable ligand substitution of nitrate by CEESO, and consequently less product inhibition.

15. SOLVENT EFFECTS Acetonitrile is a polar organic solvent and therefore many inorganic salts and organic compounds are reasonably soluble in it. Additionally it is extremely robust towards oxidation and is aprotic. Therefore most of the research reviewed here involved the use of acetonitrile as the solvent. However, it is rather toxic (40 ppm is the practical exposure limit (PEL), http://www.msdsonline.com) limiting its use in personal care products including topical skin protectants (TSPs) and other skin creams. Another

5. Catalysts for aerobic oxidation under ambient conditions

253

disadvantage of acetonitrile is that when it is used as a solvent, an induction period can occur during the catalytic reaction. For these reasons we looked at the possibility of using other solvents. Figure 6 shows two of the four solvents (trifluoroethanol and acetonitrile) in which catalysis of CEES oxidation to CEESO occurs. Three of the four solvents, trifluoroethanol, nitromethane and 1,2-dichloroethane, do not exhibit a measurable induction period. The highest rate is observed in trifluoroethanol. It is also noteworthy to mention that a complete conversion of CEES to CEESO can be achieved in trifluoroethanol. When ethanol or acetone is used, the Au(III) is immediately reduced to colloidal Au(0), which cannot be reoxidized by In tert-butanol, Au(III) is reduced to Au(I) but reoxidation of Au(I) to Au(III) is not observed. In pyridine, a mixing of and CEES does not result in precipitation of AgCl indicating that the active catalyst is never formed. In perfluoropolyether (PFPE) solvents at least one component is insoluble, so these reactions are not entirely homogeneous. These systems are classified as heterogeneous and are discussed below.

Binding of solvent molecules to Au(III), eq 7 in Scheme 1, as well as a shift of the equilibrium between the cation 2 and neutral 1 (eq 8) should be considered. For example, acetonitrile binding to both Au(III) and Au(I) complexes is well precedented35,69,70, and the solvolytic equilibrium, eq 43, has been studied in (95:5, v/v) by Canovese et al.71

254

E. Boring, Yu. V. Geletii and C. L. Hill

The value of was found to be implying that 26% of the total Au(III) (at 5 mM) is in the form In contrast, trifluoroethanol, nitromethane, and 1,2-dichloroethane would coordinate weakly if at all with the Au complexes. The equilibrium constants are likely to be very sensitive to the nature of the solvent. For example, the equilibrium constant for the displacement of heterocyclic amines (for example pyridine, py) by from Au(III), eq 44, changes from 25 in to 0.085 in

The equilibrium (eq 8 in Scheme 1) between a neutral and positively charged complex, 1 and 2 respectively, depends on the solvent polarity. The more polar the solvent, the more eq 8 is driven to the right, which increases and thereby decreases the concentration of the active complex 1 and the overall reaction rate (see discussion on the effect of in section 12). However, the solvent effect on the reaction rate is more complex. The overall rate is controlled not only by the concentration of the reactive complex 1, but also by the rate of the reduction of 1 by CEES, eq 9 in Scheme 1. In this rate limiting step two charged species, chlorosulfonium and nitrate ions, are formed. Their charge separation should be more favourable in polar solvents. Thus, predicting the effect of the solvent on the overall rate could be difficult a priori because and have dependencies in opposite directions. The highest oxidation rate is observed in trifluoroethanol, a less polar solvent than acetonitrile. This higher rate is attributable to a shift of eq 8 to the left This is also consistent with the weak product inhibition observed in trifluoroethanol. Formation of the inactive 2’” cation and nitrate anion, eq 42, is not as favourable in lower polarity solvents resulting in less self-inhibition. For that reason, CEES oxidation to CEESO proceeds to almost completion in trifluoroethanol, but significant inhibition occurs in acetonitrile (Figure 6). It has also been observed that an induction period is not significant, if it exists at all, in solvents other than acetonitrile. The length of this induction period depends on dioxygen concentration and therefore is likely to depend on the rate of Au(I) oxidation by dioxygen.22 Since this reaction is not rate limiting in the catalytic process it is impossible to assess a solvent’s effect on the rate of Au(I) reoxidation.

5. Catalysts for aerobic oxidation under ambient conditions

255

16. HETEROGENEOUS SYSTEMS Since it was found that the reaction proceeds faster, with no induction period and is less self-inhibited in trifluoroethanol, the next logical step was to replace trifluoroethanol with high molecular weight perfluorinated polyethers (PFPEs). The PFPEs are the principal components of TSPs and are effectively nontoxic. The systems appears to be active for selective aerobic sulfoxidation of CEES in the PFPE oil Galden D® 02 , (17 turnovers per hour, see Table 1 in a recent article)23 and highly active in Fomblin MF-300®,a PFPE surfactant with terminal carboxylic acid functional groups (1170 turnovers per hour, Figure 7).

However, both these systems are heterogeneous because is only partially soluble in Fomblin MF-300® and completely insoluble in Galden D-02® (perfluorinated oil). In Fomblin MF-300® the catalyst partially dissolves when CEES is added because the Au(III) precursor, is soluble. With the co-catalyst activity of Cu(II) in acetonitrile in mind, the sulfate salts of Fe(III), Cu(II), Mn(II), V(IV), Ti(IV), Co(II), and Ni(II) were also evaluated in Fomblin MF-300®. A synergistic effect is observed when all of these redox active metals are added to the

256

E. Boring, Yu. V. Geletii and C. L. Hill

system (see Table 2 in a recent paper).23 The most active co-catalyst is Cu(II). When 2 equivalents of is combined with 1 equivalent of the catalyst is 3.8 and 6.5 times more effective after 10 minutes of reaction time than when one of the components is omitted and respectively). It is noteworthy that is catalytically inactive without a feature that is present in fluorinated media as well as in acetonitrile. Also, the recently reported data (Figure 6 in a recent paper)23 indicate that inhibition by product sulfoxide is less pronounced in the system. Importantly, product selectivity in the heterogeneous systems is the same as in acetonitrile, eq 123, namely that no sulfoxide overoxidation to sulfone is observed in these systems within the limits of instrumental detection. This is very important for mustard gas (HD) oxidative detoxification, because the sulfoxide, “HD(O)”, is significantly less toxic than the sulfone, Since these systems are heterogeneous, their full kinetic evaluation is not possible. However, they are very similar to homogeneous ones because the same product (CEESO) and selectivity (~100% CEESO) are observed; both halide and are required for activity; and Cu(II) is a highly active cocatalyst. These similarities suggest that the key features of the mechanisms in the two types of media are also very similar,

17. EFFECT OF AMINO ACIDS A key goal of this work is to develop a catalytic system that could be incorporated into TSPs for the oxidative detoxification of mustard gas. Cornified layers of skin (epidermal) cells contain different amino acids which could bind to active Au(III) catalytic complexes and thus could reduce or eliminate their activity. Therefore, the effect of amino acids containing such functions as alkyl, amide, amine, carboxylate, imidazole, indole, alcohol, phenol, disulfide, thioether, and guanidino side chain groups on the catalytic activity for aerobic CEES oxidation were evaluated in heterogeneous systems using Fomblin MF-300® as a solvent. These Au(III)based catalysts remain active in the presence of most amino acids. Only a few amino acids exhibit moderate inhibition of the reaction. The inhibitory effect is as follows: tryptophan (indole) (most inhibiting) > methionine (thioether) > tyrosine (phenol) > leucine (alkane) > histidine (imidazole) > arginine (guanidine) > asparagine (amide) > serine (alcohol), aspartate (carboxylate) > cystine (disulfide) (least inhibiting).23 Thus, if Au(III) centers in the suspended Au(III)-based catalysts in a deployed TSPs have direct molecular contact with the amino acids in the skin, then the epidermal

5. Catalysts for aerobic oxidation under ambient conditions

257

polypeptides would probably have little impact on the catalytic aerobic oxidative decontamination (sulfoxidation) of HD.

18. OXIDATION OF THIOETHERS OTHER THAN CEES The majority of this comprehensive study has been concentrated on the chemistry of CEES oxidation, since it closely resembles one toxic target, mustard gas. However, the Au-based catalysts reported here are able to oxidize a variety of thioethers, as well as disulfides. In fact, of the six organosulfur compounds tested, CEES under optimized catalytic conditions is actually oxidized at a much lower rate and shows more significant inhibition than the other substrates. For comparison, when THT was evaluated, the optimized Au-based catalyst under ambient conditions yields ~27 turnovers in 30 min, while Riley’s Ce-based system yields 17.6 turnovers at elevated temperature and pressure over the same time period.25 Table 3 gives the substrates that were evaluated at similar concentrations, the induction period associated with each substrate, as well as the initial rates of oxidation. The thioether oxidations were found to be completely selective in that only sulfoxide was formed, while the products arising from dimethyl disulfide oxidation have not yet been characterized (cleavage of S-S bond is suspected). The thioethers thus far studied can be divided into two groups. The first group are those that exhibit no visible induction period (Figure 8), and the second group are those that do exhibit a visible induction period (Figure 9). Thioethers with higher initial oxidation rates, tend to have shorter induction periods.

An interesting feature of our Au-based catalysts is the high substrate selectivity during competitive thioether oxidation (Figure 10). When equal amounts of two thioethers (0.36 M total thioether concentration), THT and CEES, are added simultaneously to the system, the initial rates of oxidation of the two substrates are and

258

E. Boring, Yu. V. Geletii and C. L. Hill

respectively. Under the same conditions, the rates of oxidation for the same two substrates tested separately are and Thus while oxidation of THT in the presence of CEES is roughly twice slower than when THT is tested alone, the oxidation rate of CEES in the presence of THT is 52 times slower than when CEES is tested alone.

Thus while oxidation of THT in the presence of CEES is roughly twice slower than when THT is tested alone, the oxidation rate of CEES in the presence of THT is 52 times slower than when CEES is tested alone.

5. Catalysts for aerobic oxidation under ambient conditions

259

Importantly, the ratio of initial rates is 180 in competitive oxidation, while it is only 7 when THT and CEES are oxidized separately. An additional point to be made is that in the competitive oxidation, CEES starts to be oxidized with a much higher rate only when almost all (97%) THT is consumed. This rate is similar to that of CEES oxidation when it is tested alone, indicating that in the mixed substrate system the catalyst has extremely high substrate selectivity and that THTO has very little if any inhibitive effect on the oxidation of CEES. Thus, this study shows that thioethers other than CEES can also be oxidized in this system with efficiencies even higher than that for CEES. Competitive oxidation reveals that this catalyst discriminates between different thioether substrates with considerable efficiency.

19. EXPERIMENTAL DETAILS General procedures. Quantification of thioethers and sulfoxides was performed on an HP5890 Gas Chromatograph equipped with a FID detector and a 5% phenyl methyl silicone capillary column. As an internal standard for GC analyses, 1,3-dichlorobenzene was used. Identification of products was performed using a HP 5890 GC with a 5% phenyl methyl silicone capillary column and a 5971A Mass Selective Detector. Quantification of Au(III) concentration was performed using a HP 8452A Diode Array Spectrophotometer. The oxygen concentration was varied using a Series 810 Mass Trak flow-meter with dried argon as the other gas.

260

E. Boring, Yu. V. Geletii and C. L. Hill

General procedure for sample preparation in acetonitrile. Once the discovery had been made that was highly active for thioether oxidation, stock solutions of each component were prepared in 20-mL vials using anhydrous and exposure to light was prevented by wrapping each vial with aluminum foil. can replace as a source that acts as a poor halide abstractor. All reagents were dried using activated molecular sieves after they had been dissolved in anhydrous Appropriate amounts of each stock solution were added via syringe to a 20-mL glass vial fitted with a PTFE septum that was first purged with The atmospheric pressure was adjusted to 1.0 atm and the reactions were carried out at In all cases, thioethers or sulfoxides were the last component to be added. The same procedure as above was used except that the total volume was adjusted using a solvent other than acetonitrile (stock solutions of each components were prepared in acetonitrile). Determining the stoichiometry of sulfoxide with respect to The stoichiometry of sulfoxide with respect to was determined by measuring consumption using a volumetric method and sulfoxide formation by GC. Evaluation of the reaction kinetics. The reaction kinetics were evaluated and curves fit using the Solver subprogram in Microsoft Excel. Rate laws were determined by varying one component of the system while keeping all others constant. In order to determine the effect on reaction rate of the product sulfoxide, DMSO was used as the model for CEESO. Determining the source of oxygen in product sulfoxide. The source of oxygen of the product sulfoxide was determined by using and monitoring the mass abundances by GC-MS. In a separate experime,t concentrated was added after the addition of thioether. Substitution of and with other anions. In separate experiments, was replaced with equimolar amounts of TMAOH. TBA and TMA are abbreviations for tetra-n-butylammonium and tetramethylammonium cations, respectively. In a separate experiment was replaced with an equimolar amount of Measuring product selectivity using DMSO. Using similar reaction conditions as above, DMSO was added after the addition of CEES and the quantity of DMSO and were monitored by GC. Determining the effect of DMSO (a product sulfoxide model) on the rate of catalysis. Varying amounts of DMSO were added to the normal reaction components to assess the effect on induction period, rate of catalysis, and self-inhibition. Assessment of Cu(II) and Fe(III) on rate of CEES oxidation by Using similar reaction conditions as above in two separate experiments, two equivalents of and were added before the addition of CEES.

5. Catalysts for aerobic oxidation under ambient conditions

261

CEES oxidation in PFPE media. Using as the Au source, as the source, and transition metal sulfates as the transition metal source, reactivity towards CEES oxidation was assessed in two types of PFPE media: Galden D-02®, a perfluorinated polyether oil, and Fomblin MF-300®, a perfluourinated polyether surfactant. was synthesized using a literature procedure.36 TEA is an abbreviation for tetran-ethylammonium cation. Effect of amino acids on the rate of CEES oxidation. Using the same procedure as in the previous section, the inhibitory effect of various amino acids were evaluated in Fomblin MF-300®, a PFPE surfactant. Competitive oxidation between CEES and THT using Using similar reaction conditions as above, two different thioethers, GEES and THT, were added simultaneously in equimolar amounts and the reactivity was compared to that of CEES alone.

20. CONCLUSIONS Diversity-based methods combined with mechanistic information on currently successful catalytic systems for oxidations have led to the discovery of catalysts that addresses two of the holy grails in catalysis and oxidation: complexes that catalyze selective (non-radical-chain) oxidation by without the requirement of a sacrificial reductant, and complexes that catalyze rapid reactions with the ambient environment (room temperature and 1.0 atmosphere of air). The reaction studied here is thioether (organic sulfide) sulfoxidation via the “dioxygenase stoichiometry”: (sulfoxide), and the principal catalyst in homogeneous acetonitrile is (1). The selective sulfoxidation of thioethers is of interest in decontamination (mustard destruction) and organic synthesis. Extensive kinetics, product, spectroscopic studies have established all the elementary processes in the mechanism and the detailed aspects of the rate-limiting step which involves reaction of 1 with a molecule of the thioether substrate. The limitations exhibited by this initial system in acetonitrile solution, namely the presence of a significant induction period and inhibition by the sulfoxide product, are eliminated by the use of other solvents or developmentally attractive and nontoxic perfluoropolyether (PFPE) media. These aerobic selective catalytic sulfoxidation reactions are also co-catalyzed by some d-block ions. All the mechanistic and energetic information established in this initial research reviewed here augurs well for further logical development of this new environmentally attractive (green) catalytic chemistry.

262

E. Boring, Yu. V. Geletii and C. L. Hill

References 1. Groves, J. T.; Quinn, R. J. Am. Chem. Soc. 1985,107, 5790-5792. 2. Hill, C. L.; Khenkin, A. M; Weeks, M. S. ACS Symp. Ser. 1993,523, 67-80. 3. Neumann, R.; Dahan, M. Nature 1997,388, 353-355. 4. Hill, C. L.; Weinstock, I. A. Nature (London) 1997,388, 332-333. 5. Neumann, R.; Dahan, M. J. Am. Chem. Soc. 1998,120, 11969-11976. 6. Hill, C. L. Nature 1999,401, 436-437. 7. Brink, G.-J. ten; Arends, I. W. C. E.; Sheldon, R. A. Science 2000,287, 1636-1639. 8. Parshall, G. W.; Ittel, S. D. Homogeneous Catalysis. The Applications and Chemistry of Catalysis by Soluble Transition Metal Complexes; 2nd ed.; Wiley-Interscience: New York, 1992. 9. Tolman, C. A.; Druliner, J. D.; Nappa, M. J.; Herron, N. Alkane Oxidation Studies in Du Pont's Central research Department; Hill, C. L., Ed.; Wiley: New York, 1989, pp Chapter 10. 10. Sheldon, R. A.; Kochi, J. K. Metal-Catalyzed Oxidations of Organic Compounds; Academic Press: New York, 1981. 11. Xu, L.; Boring, E.; Hill, C. J. Catal. 2000,195, 394-405. 12. Fukumoto, K.; Onoda, S.; Sugiura, M.; Horii, M.; Hayashi, H. Material for Removing Offensive Odor; Kabushiki Kaisha Toyota Chuo,Kenkyusho, Aichi-ken, Japan: United States, 1997. 13. Dimotakis, E. D.; Cal, M. P.; Economy, J. Environ. Sci. Technol. 1995,29, 1876-1880. 14. Menger, F. M.; Elrington, A. R. J. Am. Chem. Soc. 1990,112, 8201-8203. 15. Yang, Y.; Szafraniec, L. L.; Beaudry, W. T.; Davis, F. A. J. Org. Chem. 1990,55, 36643666. 16. Gall, R. D.; Faraj, M.; Hill, C. L. Inorg. Chem. 1994,33, 5015-21. 17. Hill, C. L.; Gall, R. D. J. Mol. Catal. A: Chem. 1996,114, 103-111. 18. Johnson, R. P.; Hill, C. L. J. Appl. Toxicol. 1999,19, S71-S75. 19. Koper, O.; Lucas, E.; Klabunde, K. J. J. Appl. Toxicol. 1999,19, S59-S70. 20. Reviews of decontamination: a) Yang, Y.-C.; Baker, J. A.; Ward, J. R. Chem. Rev. 1992, 92, 1729-1743, b) Organic Indoor Air Pollutants: Occurrence - Measurement Evaluation. Tunga Salthammer, (Ed.), Wiley-VCH: Weinheim, 1999; c) Cheremisinoff, P. N.; Abraham, J. Encyclopedia of Environmental Control Technology: Work Area Hazards; Gulf Publishing Co.: Houston, 1995; Vol. 8, p.51, d) Harrison, R. M. In Air Pollution and Health, R. E. Hester and R. M. Harrison, (Eds.), Royal Society of Chemistry: Cambridge, 1998; Vol 10, pp. 101-126. 21. Ashley, D. L.; Bonin, M. A.; Cardinali, F. L.; McCraw, J. M.; Wooten, J. V. Environ. Health Perspectives 1996,104, 871-877. 22. Boring, E. A.; Geletii, Yu. V.; Hill, C. L. J. Am. Chem. Soc. 2001,123, 1625-1635. 23. Boring, E. A.; Gueletii, Yu. V.; Hill, C. L. J. Mol. Catal. A, 2001,176, 49-63. 24. Riley, D. P. Inorg. Chem. 1983,22, 1965-1967. 25. Riley, D. P.; Smith, M. R.; Correa, P. E. J. Am. Chem. Soc. 1988,110, 177-180. 26. Ericson, A.; Elding, L. I.; Elmroth, S. K. C. J. Chem. Soc., Dalton Trans. 1997,7, 11591164. 27. Annibale, G.; Canovese, L.; Cattalini, L.; Natile, G. J. Chem. Soc., Dalton Trans. 1980, 1017-1021. 28. Gasparrini, F.; Giovannoli, M.; Misiti, D.; Natile, G.; Palmieri, G. Tetrahedron 1983,39, 3181-3184. 29. De Filippo, D.; Devillinova, F.; Preti, C. Inorg. Chim. Acta 1971,5, 103-108. 30. Natile, G.; Bordignon, E.; Cattalini, L. Inorg. Chem. 1976,15, 246-248.

5. Catalysts for aerobic oxidation under ambient conditions

263

31. Schmid, G. M.; Curley-Fiorino, M. E. Encylopedia of Electrochemistry of the Elements; M. Dekker: New York, 1975; Vol. 4. 32. Skibsted, L. H.; Bjerrum, J. Acta Chem. Scand. A 1977,31, 155-156. 33. Goolsby, D.; Sawyer, D. T. Anal. Chem. 1968,40, 1978. 34. Fenske, G. P.; Mason, W. R. Inorg. Chem. 1974,13, 1783-1786. 35. Kissner, R.; Welti, G.; Geier, G. J. Chem. Soc., Dalton Trans. 1997,10, 1773-1777. 36. Mason, W. R.; Cray, H. B. Inorg. Chem. 1968,7, 55-58. 37. Chambers, R. C.; Hill, C. L. J. Am. Chem. Soc. 1990,112, 8427-33. 38. Mata, E. G. Phosphorus, Sulfur Silicon Relat. Elem. 1996,117, 231-286. 39. Mashkina, A. V. Catal. Rev. 1990,32, 105-161. 40. Capozzi, G.; Drabowicz, J.; Kielbasinski, P.; Menichetti, S.; Mikolajczyk, M.; Nativi, C.; Schank, K.; Schott, N.; Zoller, U. The Syntheses of Sulphones, Sulphoxides and Cyclic Sulphides: Updates from the Chemistry of Functional Groups; John Wiley & Sons: Chichester, 1994. 41. Marrs, T. C.; Maynard, R. L.; Sidell, F. R. Chemical Warfare Agents: Toxicology and Treatment; Wiley & Sons: Chichester, New York, 1996, pp 141. 42. Roulet, R.; Lan, N. O.; Mason, W. R.; Fenske, J., G. P. Helv. Chim. Acta 1973,56, 24052418. 43. Puddephatt, R. J. Comprehensive Coordination Chemistry: The Synthesis, Reactions, Properties, and Applications of Coordination Compounds; 1 ed.; Wilkinson, G., Gillard, R. D. and McCleverty, J. A., Ed.; Pergamon Press: Oxford, 1987; Vol. 5. 44. Marangoni, G.; Pitteri, B.; Bertolasi, Y.; Gilli, G.; Ferretti, V. J. Chem. Soc., Dalton Trans. 1986, 1941. 45. Elding, L. I.; Skibsted, L. H. Inorg. Chem. 1986,25, 4084-4087. 46. Berglund, J.; Elding, L. I. Inorg. Chem. 1995,34, 513-519. 47. Elmroth, S.; Elding, L. I. Inorg. Chem. 1996,35, 2337-2342. 48. Elding, L. I.; Olsson, L. F. Inorg. Chem. 1982,21, 779-784. 49. Kochi, J. K. Oxidation-Reduction Reactions of Free Radicals and Metal Complexes; Kochi, J. K., Ed.; Wiley: New York, 1973; Vol. 1, pp 591-685. 50. Smith, S. G.; Winstein, S. Tetrahedron 1958,3, 317. 51. Drabowicz, J.; Kielbasinski, P.; Mikolajczyk, M. Synthesis of sulphoxides; Patai, S. and Rappoport, Z., Ed.; John Wiley & Sons: Chichester, 1994, pp 529-648. 52. Braunstein, P.; Clark, R. J. H. J. Chem. Soc., Dalton Trans. 1973, 1845. 53. Puddephatt, R. J. The Chemistry of Gold; Elsevier Scientific: New York, 1978. 54. Uson, R.; Laguna, A.; Vicente, J. J. Organomet. Chem. 1977,131, 471-475. 55. Dash, K. C.; Schmidbaur, H. Chem. Ber. 1973,106, 1221-1225. 56. Allen, E. A.; Wilkinson, W. Spectrochim. Acta 1972,28A, 2257-2262. 57. Schoenfelner, B. A.; Potts, R. A. J. Inorg. Nucl. Chem. 1981,43, 1051-1053. 58. Elding, L. I.; Groening, A.-B. Acta Chem. Scand. 1978,32, 867-877. 59. Douglas, B. E. Concepts and Models of Inorganic Chemistry; 3rd ed. ed.; Wiley: New York, 1994. 60. Cattalini, L.; Tobe, M. L. Inorg. Chem. 1966,5, 1145-1150. 61. Cattalini, L.; Orio, A.; Tobe, M. L. J. Am. Chem. Soc. 1967,89, 3130-3134. 62. Cattalini, L; Orio, A.; Tobe, M. L. Inorg. Chem. 1967,6, 75-78. 63. Davies, J. A.; Hasselkus, C. S.; Scimar, C. N.; Sood, A.; Uma, V. J. Chem. Soc., Dalton Trans. 1985,1985, 209-211. 64. Kapoor, P.; Lövqvist, K.; Oskarsson, Å. J. Mol. Struct. 1998,470, 39-47. 65. Horn, G. W.; Kumar, R.; Maverick, A. W.; Fronczek, F. R.; Watkins, S. F. Acta Cryst. C 1990,C46, 135-136. 66. Melanson, R.; Rochon, F. D. Can. J. Chem. 1975,53, 2371-2374.

264

E. Boring, Yu. V. Geletii and C. L. Hill

67. For data in water, see Atkins, P. W. Physical Chemistry; W. H. Freeman: New York, 1990. 68. Mylius, F. Z Anorg. Chem. 1911,70, 203-231. 69. Bravo, O.; Iwamoto, R. T. Inorg. Chim. Acta 1969,3, 663-666. 70. Tsvelodub, L. D.; Malkova, V. I. Sib. Khim. Zh. 1991,3, 72-77. 71. Canovese, L.; Cattalini, L.; Tomaselli, M.; Tobe, M. J. Chem, Soc., Dalton Trans. 1991,1991,307-314. 72. Cattalini, L.; Ricevuto, V.; Orio, A.; Tobe, M. L. Inorg. Chem. 1968,7, 51-55.

Chapter 6 Catalytic oxidations using cobalt(II) complexes

László I. Simándi Chemical Research Center, Institute of Chemistry, Hungarian Academy of Sciences, H-1525 Budapest, P.O. Box 17, Hungary

Abstract: The reactivity of cobalt(II) complexes toward dioxygen has long been recognized. Synthetic oxygen carriers reversibly form mononuclear superoxo and dinuclear complexes. The bound can be removed by pumping or flushing with an inert gas and this cycle can be repeated many times over. However, there is always some loss of reversibility, leading to cobalt- or ligand-centered oxidation. Dioxygen complexes are generally regarded as the source of activity in cobalt-catalyzed oxidations. The term dioxygen activation is used to describe the oxidation of added oxidizable substances via interaction with intermediate dioxygen complexes or their conversion products. Observations of this behavior have prompted extensive research into the vast area of homogeneous catalytic oxidation using cobalt complexes. In this review the progress made in the study of various cobalt-based catalyst systems in the last decade is surveyed. Catalytic oxidations by different classes of cobalt(II) complexes with salen, porphyrin, phthalocyanin, dioxime, amine, pyridine, cyclidene, peptide and carboxylato ligands are discussed with specific reference to the products formed and the underlying reaction mechanisms. For each catalyst type the oxidation of various substrates is reviewed, including substituted phenols, lignin phenolics, catechols, anilines, thiols, alcohols, diols, and alkenes. Oxygen insertions, NO oxidation and oxidations via alkylperoxo complexes are treated. Obviously due to the paramagnetic nature of cobalt(II) complexes, free-radical mechanisms are predominant in cobalt-catalyzed oxidations, permitting insight into the nature of reaction intermediates by the ESR technique. Key words: Catalytic oxidation, homogeneous catalysis, dioxygen activation, dioxygen complexes, biomimetic oxidation, functional metalloenzyme models, oxidation mechanisms, oxidative dehydrogenation, oxygen insertion, alkene epoxidation, catecholase reaction

265 L.I. Simándi (ed.), Advances in Catalytic Activation of Dioxygen by Metal Complexes, 265-328. © 2003 Kluwer Academic Publishers. Printed in the Netherlands.

266

L.I. Simándi

1. INTRODUCTION Various cobalt (II) complexes, usually referred to as synthetic dioxygen carriers, are known to interact reversibly with dioxygen under ambient conditions, affording predominantly dioxygen complexes of different types, which have been extensively reviewed 1-6. The ranges of O-O bond lengths in the most common structural types are shown in Table 16.

In its triplet ground state the reactivity of dioxygen, having two unpaired electrons in its degenerate antibonding orbitals, is controlled by the rules of spin conservation. The kinetic barrier imposed by these conditions is sufficiently large to ensure survival of all living organisms despite their thermodynamic instability toward oxidation in an atmosphere containing dioxygen. The kinetic inertness of dioxygen can be overcome by coordination to a metal center, which eliminates the restrictions of spin conservation. The interaction of with a transition metal ion surrounded by suitable ligand(s), i.e. dioxygen complex formation, can be broadly regarded as of two major types, viz., (i) reversible binding (oxygenation), and (ii) activation of dioxygen. Reversibly bound can be removed from the complex by pumping or flushing with an inert gas ( Ar), and this cycle can be repeated many times over. However, there is always a gradual loss of reversibility, as demonstrated for cobalt(II)-based synthetic dioxygen carriers. This is due to irreversible oxidation of the central metal ion and/or the ligand(s) surrounding the metal. In the presence of a suitable external substrate, its catalytic oxidation may take place, leading to a catalytic cycle, in which the activated dioxygen species is the key intermediate. Substrate oxidation and catalyst regeneration should occur in the successive cycles. The catalytic activity usually decreases with time and is ultimately lost due to metal centered and/or ligand based irreversible oxidations.

6. Catalytic Oxidations using Cobalt(II) Complexes

267

2. COBALT DIOXYGEN COMPLEXES The majority of cobalt dioxygen complexes can be classified as of the mononuclear superoxo or dinuclear type7. 1-7 For other types the reader is referred to the review literature . Dioxygen complexes are often formed reversibly and can be characterized by equilibrium constants. In catalytic systems usually enters the catalytic cycle via equilibria of the types (1) and (2):

These reactions can be regarded as autoxidation of although both steps (1) and (2) are substitution (complex formation) reactions. True autoxidation with loss of an electron by occurs on protonation of the dimer, leading to the formation of as in the case of

The cobalt-bound (activated) dioxygen exhibits higher reactivity toward certain substrates than does free In kinetic studies the above equilibria may be treated as rapid pre-equilibria, which maintain a near constant concentration of the active intermediate(s). The rate constants for binding are usually very large and kinetic studies require the stopped-flow technique, in which a solution of the cobalt complex prepared under an inert gas is rapidly mixed with an solution. The reaction is then monitored spectrophotometrically. The problem of inert storage was eliminated in the case of cobaloxime(II), by preparing the cobalt complex in the mixing chamber from and cobalt perchlorate9. This was made possible by the fact that the formation of is much faster than its subsequent reaction with Recently, a new method has been reported for studying rapid biological reactions involving dioxygen10. It was used to investigate the reduction of dioxygen to water by cytochrome c oxidase. Photolysis of a synthetic caged dioxygen carrier, produces dioxygen in situ on a nanosecond or faster time scale. This avoids complications due to the fate of photodissociated CO in a conventional CO flow-flash experiment. The kinetics of dioxygen binding to the cobalt(II) complexes of

268

L.I. Simándi

1,4,7,10-tetraazacyclododecane and 1,4,8,11-tetraazacyclotridecane has recently been studied11. The general features of oxygenation equilibria and autoxidation involving cobalt-based lacunar cyclidene dioxygen carriers have been extensively studied by Busch and coworkers with special reference to dioxygen affinities and their correlation with structural features 3,4,12,13. Lacunar systems have considerable activities for reversible dioxygen binding in their cavities under ambient conditions.

The autoxidation of dioxygen carriers may involve both metal-centered oxidation to and ligand oxidation, leading to ligand destruction. Typical ligand oxidations involve oxidative dehydrogenation of to imine bonds as in the case of cobalt (II) complexes of Pydien14, and to carbonyl group oxidations in macrocyclic chelates15,16:

6. Catalytic Oxidations using Cobalt(II) Complexes

269

Recent progress in bimetallic dioxygen complexes containing compartmental ligands has been reviewed.17

3. OXIDATIONS CATALYZED BY Co(salen) COMPLEXES The cobalt(II) complexes of salen type ligands have long been used as catalysts for the oxidation of substituted phenols. The subject has been extensively reviewed22,23. In this chapter emphasis is placed on recent advances.

3.1 Oxidation of substituted phenols 3.1.1 2,6-di-tert-butylphenol

The properties of oxygen adducts of complexes (Figure 3) have been studied by cyclic voltammetry, IR, electronic and ESR spectroscopy. The complexes exhibit oxygen-binding ability in the presence of an axial ligand (py), and end-on superoxocobalt type dioxygen adducts are formed18. Electron-donating substituents gave higher concentrations of the superoxo complex. The Meand derivatives give superior catalytic activity in the oxygenation of 2,6-di-tert-butylphenol to the corresponding quinone.

The oxidation of 2,3,6-tri-tert-butylphenol and 2,4,6-tri-tert-butylaniline

270

L.I. Simándi

with molecular oxygen and tert-butylhydroperoxide was investigated using biomimetic Mn-, Fe- and Co-complexes as catalysts19. The catalytic activity and product distribution were determined and compared with those observed in the reactions of the well-known Co(salen) complex. Supercritical is gaining importance as a reaction medium. The oxidation of substituted phenols by O2 has been studied in supercritical using [{N,N’-bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediaminato(2- )}cobalt(II)] as catalyst20 Both oxidase and oxygenase activities are observed with 2,6-di-tert-butylphenol as substrate at 70°C and 207 bar total pressure, at an O2:phenol:catalyst ratio of 1500:20:1, and a methylimidazole to catalyst ratio of 1.28. Total conversion to 2,6-di-tert-butyl-1,4-benzoquinone and 3,5,3’,5’-tetra-tert-butyl-4,4'-diphenoquinone took place in 21 h.

3.2 Oxidation of 3,5-di-tert-butylcatechol Binuclear Co(II) complexes derived from 2,6-diformyl-4-methylphenol and various aromatic monoamines have been prepared and investigated.21 The Co(II) complexes have the composition where L represents the organic ligand. The complexes are active catalysts in the oxidation of 3,5-di-tert-butylcatechol.

3.3 Oxidation of lignin phenolics Earlier work on the Co(salen)-catalyzed of p-substituted phenols to p-benzoquinones22,23 has been extended to include substrates that serve as models for lignin phenolic subunits. Lignin is a renewable source of carbon and its oxidation to p-benzoquinone derivatives would allow conversion to useful intermediates. Certain p-substituted phenolics can be oxidized to p-benzoquinones with dioxygen using the Co(salen) complexes A and B in Figure 4 as catalysts24.

6. Catalytic Oxidations using Cobalt(II) Complexes

271

These 5-coordinate cobalt complexes form mononuclear superoxo and bridged dinuclear complexes, which are reactive toward phenolic substrates, thereby initiating catalytic oxidation cycles23:

Typical reaction conditions are shown in Figure 5 for syringyl alcohol as substrate. In the presence of the catalysts in Figure 4, 2,6-dimethoxybenzoquinone is formed in 71 and 88% yield, respectively.

The p-substituted phenolic substrates and the yields of their oxidation products are listed in Figure 6. The proposed mechanism of oxidation is shown in Figure 7 25,26. The superoxo complex abstracts the phenolic hydrogen, affording a phenoxy radical, which is trapped by a second complex or dioxygen, giving intermediate (I) . When phenol oxidation is carried out stoichiometrically with type systems, structural analogs of I have been isolated and characterized27. Elimination of formaldehyde from I produces quinone and, when a catalytically active Co-hydroxy species28. The complex formed in the H-abstraction step breaks down to regenerate the starting catalyst26. The catalytic effect depend on the ease of removal of the phenolic hydrogen, the rate decreasing in the following order of p-substituents: MeO (123), Me (28) CN (1). The propenoidic phenols, E-methyl ferulate (1), E-4-hydroxycinnamic acid methyl ester (4) and E-3-chloro-4-hydroxycinnamic acid methyl ester (7), can be catalytically oxidized with dioxygen in the presence of [Co(salen)] (Figure 8)29. The yields depend on the solvent and the phenyl substituents. EPR and electronic spectra suggest the involvement of a coordinated o-benzosemiquinone type radical as the active intermediate. Oxidative processes are involved in both the polymerization of phenylpropendioic phenols to lignin and the degradation of lignin in the environment30,31. Model systems for the oxidative degradation of phenolic phenylpropenoids provide mechanistic information about the biological cycle of lignin32. Co(salen) was found to catalyze the oxidation of the lignin

272

L.I. Simándi

6. Catalytic Oxidations using Cobalt(II) Complexes

273

model E-ferulic acid to vanillic aldehyde, vanillic acid methyl ester and homovanillic aldehyde (Figure 9) at 1 bar of and

These transformations correspond to double bond cleavage with O-atom insertion and aromatic hydroxylation. ESR studies suggest an organometallic radical intermediate in chloroform and methanol and a superoxocobalt(III) species in pyridine.

274

L.I. Simándi

3.4 Nitrogen monoxide reacts with nitrogen monoxide to give the nitrosyl complex where is a dinucleating macrocycle, having a salen- and a saldien-like metal binding site (Figure 10). The Co occupies the salen and the Pb the saldien site (Figure 11)34.

6. Catalytic Oxidations using Cobalt(II) Complexes

275

The NO coordinates to the Co at the axial site trans to the bridging dmf oxygen, providing a six-coordinate geometry at the metal.

The nitrosyl complex is oxidized with molecular oxygen to the nitro complex

3.5 Oxidative dehydrogenation In -saturated dmf solution, the Co(II) complex of the tetrahydrosalen ligand undergoes slow oxidative dehydrogenation at both C-N bonds and the cobalt(II) salen complexes (CoL) are formed35 (Figure 12). The overall process is autocatalytic and involves reduction of to water, but is probably an intermediate.

276

L.I. Simándi

3.6 Oxidation of quercetin Quercetinase is a dioxygenase, catalyzing the insertion of into quercetin and 3-hydroxyflavones, leading to oxidative cleavage of the heterocyclic ring, affording the corresponding depsides and carbon monoxide (Figure 13)36. Model studies of oxygenations have been reported using copper(II) complexes37. Co(salen) exhibits a remarkable activity in the cleavage of model substrate (Figure 14)38. Cyclic voltammetry of the substrate anion-catalyst binary intermediate complex indicates that it is the substrate anion involved in an ion-pair complex with the cationic that inserts dioxygen, yielding the product depside39 HL2.

3.7 Mercaptoethanol Co(salen) immobilized on silica, layered double hydroxides and NaX zeolite act as photocatalysts for the oxidation of 2-mercaptoethanol and sodium thiosulfate40.

3.8 Alkenes and alcohols In the presence of cobalt(II) Schiff-base complexes some ketoesters and

6. Catalytic Oxidations using Cobalt(II) Complexes

277

aldehydes promote the oxidation of alkenes and alcohols with The added carbonyl compounds assist in the formation of dioxygen complexes and act as reducing agents during oxygen atom transfer to the organic substrates. The cobalt(II) Schiff-base complexes shown in Figure 15 act as catalysts for the oxidation of a wide range of organic substrates (e.g. alkenes, alcohols, benzylic compounds and aliphatic hydrocarbons) with dioxygen in the presence of aliphatic aldehydes, ketones or ketoesters.

EPR studies in acetonitrile at room temperature have shown that aliphatic carbonyl compounds promote the formation of superoxocobalt(III) complexes, which are responsible for the catalytic effect53. According to the general procedure for catalytic oxidation of alkenes, benzylic compounds and alcohols, the mixture of 5 mmol substrate, 5 mol% of cobalt(II) complex, 10 mmol were stirred under 1 atm for 20-35 hours at room temperature or 50-60°C. The observed types of reactions include alkene epoxidation, allylic and benzylic oxidation, and alkane hydroxylation. The ligands on the cobalt atom control the chemoselectivity of oxidation. The proposed reaction mechanism is shown in Figure 16. The catalytic activity is due to the superoxocobalt(III) complex which binds a ester or aldehyde (left and right end of the horizontal row, respectively). Intramolecular H-atom transfer affords an enolato-cobalt complex (left) or a coordinated acyl radical (right). Both of these release an

278

L.I. Simándi

oxocobalt(IV) species, which is regarded as the active intermediate responsible for the oxidation products formed via O-atom transfer. These are shown above and under the horizontal row. Oxometal species have been variously invoked as intermediates in Otransfer reactions, especially in connection with single oxygen atom donors, such as iodosylbenzene, persulfate, N-oxides and peroxy compounds54. Oxocobalt(IV) was proposed as the active intermediate in the cobaloxime(II)-catalyzed oxidation of hydrazobenzene with

3.9 Alkene epoxidation In the presence of methyl, tert-butyl, or (-)-menthyl esters of 2oxocyclopentanecarboxylic acids, the cobalt(II)-salen type complexes and Jacobsen-type manganese(III) complexes shown in Figure 17 are active catalysts for alkene epoxidation with dioxygen56. In these epoxidations (Figure 18), alkyl-1-hydroxy-2-oxocyclopentanecarboxylates and 1-alkyl-2-oxo-hexanedicarboxylic acids are formed as cooxidation products.

6. Catalytic Oxidations using Cobalt(II) Complexes

279

The (-)-menthyl/cobalt system is selective for epoxide formation but the products are racemic in line with radical epoxidation in solution rather than

280

L.I. Simándi

at the cobalt complex. The experimental results are consistent with chain reactions involving the free radicals shown in Figure 19. The Jacobsen-type manganese complex gives lower yields of epoxides (40-60%), but for 2,2-dimethylchromene and styrene the epoxides are optically active (12-60% ee).

3.10 Primary amines Bis[3 -(salicylideneimino)propyl] methylaminecobalt(II) (CoSMDPT) 57 catalyzes the oxidation of primary amines with dioxygen . Benzylamine, 4methyl-benzylamine and 4-methoxybenzylamine gave aldehyde intermediates, which were converted to Schiff bases with the starting amines. The oxidation of 1-phenyl-ethylamine gave 2,4,6-triphenyl-3-aza-3,5-heptadiene via disproportionation of the intermediate Schiff base.

4. OXIDATIONS CATALYZED BY COBALOXIMES The cobaloxime(II) derivatives where H2dmg is dimethylglyoxime, and L is py and catalyze the oxidative dehydrogenations (Figure 20) and oxygen insertions of some organic substrates at room temperature and atmospheric dioxygen pressure58-61.

where Q – p-benzoquinone, AB – azobenzene. In the following discussion and stand for cobaloxime(II) and cobaloxime(III), i.e. and respectively, and L is py, or In the reaction mechanisms proposed for cobaloxime-catalyzed oxidations, the superoxocobaloxime(III), formed directly from cobaloxime(II) and is the key intermediate. It is also the source of several

6. Catalytic Oxidations using Cobalt(II) Complexes

281

other active intermediates reactive toward different substrates. Thus, formed from is the source o f hydroxocobaloxime(III), and of oxocobaloxime(IV), The complex is capable of both H-atom and electron abstraction, both possibilities being feasible routes for oxidative dehydrogenations. Oxygen insertion reactions were in some cases interpreted in terms of transient oxocobaloxime(IV). The formation routes and reactions of these active oxygen species are summarized below. Dioxygen complex formation

Oxidative dehydrogenation

where

is

or

and

is a free radical.

Electron and proton transfer

Oxygen insertion

where S is

, RNC, or PhNO.

4.1 Oxidation of o-phenylenediamine The cobaloxime(II) derivatives and where is dimethylglyoxime, are catalysts for the oxidation of ophenylenediamine (OPD) by atmospheric at ambient temperature63. In acetone, methyl ethyl ketone or cyclohexanone as solvents, cyclization to 2,2-disubstituted dihydrobenzimidazoles (DHB) is followed by

282

L.I. Simándi

dehydrogenation to 2,2-disubstituted 2H-benzimidazoles (2HB) (Figure 21). Acetaldehyde affords 2-methylbenzimidazole in a similar way62 (Figure 22).

4.2 Oxidation of 2-aminophenol Cobaloxime(II) derivatives where and catalyze the oxidative dehydrogenation of 2-aminophenol (ap) at room temperature and 1 bar in The reaction product is 2-aminophenoxazine-3-one (apx) formed in quantitative yield according to the stoichiometric equation of Figure 23. This system can be regarded as a functional model of Phenoxazinone synthase, which is involved in the biosynthesis of Actinomycin D (AD), a

6. Catalytic Oxidations using Cobalt(II) Complexes

283

naturally occurring antineoplastic agent, from the 2-aminophenol derivative A (Figure 24).

The kinetics of oxidation of 2-aminophenol (ap) in the presence of [Co has been interpreted64 in terms of a mechanism consisting of the steps shown in Figure 25. Upon dissolution of the catalyst precursor in MeOH one of the axial ligands dissociates, generating the active pentacoordinate catalyst, denoted by The key intermediate is the superoxocobaloxime derivative which abstracts an H-atom from the ap substrate via an H-bonded species X. The aminophenoxyl radical produced is further oxidized to -benzoquinone monoimine (bqmi), which is an intermediate on the path to apx formation. EPR studies provide evidence for formation of the free radical as intermediate.65

4.3

Oxidation of 3,5-di-tert-butylcatechol

The oxidation of 3,5-di-tert-butylcatechol by to the corresponding 1,2-benzoquinone (DTBQ) is catalyzed by the cobaloxime

284

L.I. Simándi

in methanol66 and benzene.67 The overall stoichiometry is given by the equation:

An intermediate catecholatocobaloxime (III) has been isolated from the reacting mixture. Its molecular structure determined by x-ray diffraction66 revealed an axially bonded unidentate catecholato ligand, which is formed via free-radical coupling between the cobaloxime(II) and the semiquinone anion radical The latter has been detected by EPR spectroscopy during the reaction together with its cobaloxime(III)-bonded derivative which indicates a free-radical mechanism for catechol oxidation. The mechanism proposed for the catalytic oxidation on the basis of the detectable intermediates is shown in Figure 26. In this scheme stands for also referred to as cobaloxime(II). The double catalytic cycle is joined together by and -activation takes place via formation of superoxocobaloxime(III), followed by Hatom abstraction from affording During the catalytic reaction, and catechol oxidation are taking place according to the lower cycle, while the cobaloxime species in the upper cycle are present at steady state or equilibrium concentrations.

6. Catalytic Oxidations using Cobalt(II) Complexes

285

A detailed kinetic study68 of the cobaloxime(II) catalyzed oxidation of DBCatH2 has led to the kinetic equation (4).

which is consistent with the reaction mechanism:

The kinetic behavior requires the formation of intermediate X (Figure 27), which decomposes in the rate-determining H-atom abstraction step. In the steady state following the fast initial phase the overall stoichiometry requires that be reduced to rather than therefore, the hydroperoxocobaloxime formed in step (7) should undergo disproportionation, regenerating one half of the absorbed and producing the hydroxocobaloxime (step 10). Also, for a sustained catalytic cycle to occur in the steady state, a route to the product DTBQ and a path regenerating the catalyst are necessary. These requirements of a catalytic cycle are fulfilled by addition of electron transfer steps (11) and (12) to the mechanism. Formation of the cobaloxime derivative exhibiting the 8-line ESR signal observed during the catalytic oxidation can be explained by reaction (12). It is also involved in equilibrium (13), which has been demonstrated by reacting cobaloxime(II) with DTBQ under when the same ESR spectrum was obtained. A remarkable feature of mechanism (5) - (13) is that the dioxygen activation steps (5) and (6) are involved in both the rapid initial phase and the steady state. They control the rate of dioxygen uptake by the reacting solution, which, however, differs very strongly in these two phases. This

286

L.I. Simándi

implies that as the system reaches the steady state very shortly after mixing the reactants, the concentration of drops to a fraction of the starting value, so that the consumption rate falls from the stopped-flow level9 to values susceptible to monitoring by gas volumetry.

6. Catalytic Oxidations using Cobalt(II) Complexes

287

4.4 Oxidative cleavage of a stilbene derivative In the presence of the cobaloxime(II) catalyst at room temperature and 1 atm or air, undergoes oxidative dehydrogenation to the corresponding stilbenequinone (StQ), and parallel oxidative cleavage at the C=C double bond to afford 2,6-di-tert-butyl-4-hydroxybenzaldehyde (Ald, Figure 28)69. The relative contribution of these two stoichiometries depends on the degree of conversion.

In the steady state the relative rate of formation of StQ and Ald is approximately constant throughout a given run, indicating that they are formed in competitive reactions, via a common free-radical intermediate (Figure 29) detected by EPR spectroscopy during the reaction.

288

L.I. Simándi

The kinetics of the catalytic oxidation has been studied by a volumetric technique at constant pressure. The observed kinetic behavior was consistent with the reaction mechanism (14) - (20), where represent the moiety, respectively.

The rate law corresponding to the proposed mechanism can be derived by assuming that steps (14) and (15) are reversible and step (16) is ratedetermining. Disproportionation (17) can be regarded as fast9, consequently, the concentration of is negligible. As the kinetics were determined from initial rates and the formation of Ald is delayed relative to that of StQ, steps (19) and (20) can be disregarded and the concentration of CoOOR- is also negligible. These considerations lead to rate equation (21), which is consistent with the kinetic behavior.

6. Catalytic Oxidations using Cobalt(II) Complexes

289

Formation of the aldehyde product requires a reaction between and in which the latter is attacked in its mesomeric form with the unpaired electron at one of the olefinic carbon atoms. The alkylperoxocobaloxime(III) formed will decompose to the aldehyde probably via a dioxetane intermediate. The required protons are derived from the water present in ordinary benzene. The proposed reaction mechanism is depicted in Figure 30

4.5 Oxygen insertions 4.5.1 Oxygen insertion to terminal P, C and N atoms

Cobaloxime(II) catalyzes formal O-atom insertion into triphenylphosphine, alkyl isocyanides and nitrosobenzene 58,59:

290

L.I. Simándi

Similarly, O-atom insertion was also observed in the case of benzimidazole, which was converted to bis(N-oxide) in the presence of cobaloxime(II) (cf. Figure 21)63. Double oxygen insertion takes place in the oxidative bond cleavage of (cf. Figure 28)69: the olefinic bond is split, affording two substituted benzaldehydes. 4.6.2 Insertion of

into alkylcobaloximes

Upon photoirradiation of alkylcobaloximes under dioxygen, the insertion of into the Co-C bond is observed70,71. The (alkyldioxy)cobaloxime formed is then degraded into a carbonyl compound or an alcohol by reduction or heating (Figure 31)72-75.

Irradiation of alkylcobaloximes 1a-h under oxygen in chloroform afforded a carbonyl compound and an alcohol as main organic products (Figure 32)76. Chlorocobaloxime was isolated after the reaction in all cases. The reaction proceeds through a stable intermediate (alkyldioxy)cobaloxime. The alkyl ligand controls the selectivity to the oxygenated organic products.

6. Catalytic Oxidations using Cobalt(II) Complexes

291

The reactions of molecular oxygen with photoproduced cobaloximes(II) from the cobalt(III) complexes and (where is dimethylglyoxime and py is pyridine) have been investigated by flash photolysis methods77. In the first step mononuclear superoxocobalt(III) species are formed, which then react with photoproduced alkylcobaloxime(III) to yield Cobalt(II) complexes of 9,10-phenanthrenedionedioxime and benzoquinonedioxime were prepared and investigated by EPR in the solid state78. Dioxygen adducts are presumably formed as intermediates, affording ultimately the tris-(oximato)cobalt(III) chelates. A detailed kinetic study of dioxygen insertion into 27 organocobaloximes derived from dimethyl-, dicyclohexyl- and diphenylglyoxime has been performed under thermal and photochemical conditions79. The rate of insertion on the nature of the equatorial glyoxime, the axial ligand, the organic group and the solvent.

292

L.I. Simándi

5. OXIDATIONS CATALYZED BY COBALT(II) PORPHYRINS Metalloporphyrins are versatile oxidation catalysts. The field is dominated by iron porphyrins, which serve as synthetic models for heme type oxygenases. They have been extensively studied as witnessed by the number of review articles and books on the subject. It is only possible here to cite a few items of the vast selection80-88. Cobalt(II) porphyrins bind dioxygen reversibly and thereby mediate porphyrin degradation89,90, which can be regarded as a catalytic oxidation. In poorly coordinating solvents (dichloromethane and THF) cobalt(II) octaethylporphyrin is converted to cobalt(III) octaethyloxaporphyrin dichloride, a cobalt verdoheme analog, containing an oxidized porphyrin ring. of a type complex are formed91,92 (Figure 33).

In an extensive series of studies Lyons and coworkers have shown that halogenated metalloporphyrins are exceptionally active catalysts for the selective reaction of alkanes with molecular oxygen93-100. The greater the degree of halogenation of the ring, the greater is the catalytic activity of the metal complex. Complexes of iron are generally more active than those of cobalt, manganese, or chromium101. The product profile of isobutane oxidation is characteristic of radical reactions but also sensitive to the nature of the metal center. The selectivity to tert-butanol is about 90% or better. Among cobalt complexes catalyzing the oxidation of alkanes (isobutane or propane), is the most active101, surpassing Co(acac)2 and Co(BPI)(OAc), or (BPI = bispyridyliminoisoindoline)102,103.

6. Catalytic Oxidations using Cobalt(II) Complexes

293

Besides being very active catalysts for alkane oxidation by the iron perhaloporphyrins are also most active in the decomposition of alkyl hydroperoxides. The nature of products formed depends on the structure of the aliphatic substrate and can be rationalized by a catalytic pathway very efficiently generating alkyl and alkoxy radicals at low temperatures. Obviously, similar considerations apply also to the mechanism of oxidation catalysis by (Figure 34). Cobalt(II)-porphyrins a-d (Figure 35) are versatile catalysts promoting the oxidation of the organic substrates listed in (Figure 36) by a combination of molecular oxygen and 2-methylpropanal under ambient conditions104. Typically 10 mmol of hydrocarbon and 20 mmol of the aldehyde are stirred in an autoclave in 15 mL acetonitrile for 12-15 hours under at room temperature. Although not stated explicitly by the authors, acyl free radicals are obviously the key intermediates, converted by to acylperoxy radicals responsible for the oxidation.

294

L.I. Simándi

Alkenes, allylic or benzylic substrates and alcohols are also converted to the corresponding oxidized products at ambient conditions (Figure 37)105.

6. Catalytic Oxidations using Cobalt(II) Complexes

295

The approach of fluorous biphase systems106 utilizes the low miscibilities of fluorocarbons with most organic solvents, which allow easy separation of a catalyst soluble in the fluorocarbon phase from the product soluble in the organic phase106,107. The fluorocarbon-soluble cobalt complex of the tetraarylporphyrin ligand in Figure 38 catalyzes the epoxidation of alkenes

296

L.I. Simándi

under fluorous biphase conditions in the presence of 2-methylpropanal (Table 2). Advantages of this approach include higher substrate to catalyst ratios (1000:1), facile product separation and recycling of the expensive fluorous solvent.

and its fluorinated analog catalyze the oxidation of ethane, propane and cyclohexane by (100 psi, 85°C, 20-24 hours) to the corresponding alcohols in the presence of CO, in a mixture of trifluoroacetic acid and water. Further oxidation to aldehydes and acids also takes place108. It is proposed that CO converts the known superoxo species to an oxocobalt(IV) complex, which is capable of abstracting an H-atom from the alkanes (eq. 23). This possibility has been suggested earlier for cobaloxime(II)-catalyzed oxidations60,61.

6. Catalytic Oxidations using Cobalt(II) Complexes

297

Remarkably, primary C-H bonds are more reactive than the weaker secondary C-H bonds or C-H bonds in to an alcohol functionality. Under illumination by a 450 W high pressure mercury lamp at 30°C and 1 atm benzaldehyde and p-chlorobenzaldehyde undergo oxidation by molecular oxygen in the presence of Co(II) tetraphenylporphyrin, Co(II)[meso-tetra (benzoyloxy-phenyl)porphyrin] (CoTBCOPP) and Co(II)[meso-tetra(benzenesulfonyloxy-phenyl)]porphyrin (CoTBSOPP)109. The activity of the photocatalyst decreases in that order. An induction period in uptake was observed and accelerates the oxidation. The proposed reaction mechanism involves a free-radical chain process intiated by a superoxocobalt(III) porphyrin species, which abstracts an H-atom from the aldehyde, generating an radical. The latter participates in the following propagation and termination steps: Initiation

Propagation

Termination

298

L.I. Simándi

Step (26) differs from what the authors have proposed. The effect of light is assigned to generation of free radicals, which act as initiators and shorten the induction period. The oxidation of alkylaryl sulfides with dioxygen/2,2-dimethypropanal is catalyzed by Co(II)-tetraarylporphyrin and a perfluoroalkyl-substituted Co(II)-phthalocyanine in a fluorous organic biphasic system110. Sulfoxide was generally obtained as the main product, together with variable quantities of sulfone (0-100%), depending on the nature of the substrate. The reaction probably proceeds through a free-radical oxidative process initiated by the Co(II)-porphyrin. The catalysts used are decomposed under the free-radical reaction conditions.

6. OXIDATION WITH COBALT(II) PHTHALOCYANINES The interaction of cobalt(II) phthalocyanine with ammonia and dioxygen has been studied by ESR spectroscopy. Both 1:1 and 1:2 adducts with are formed. The phthalocyanine ring is oxidized to a cation radical when the binunlear adduct is formed111. Cobalt(II) phthalocyanines are efficient oxidation catalysts112. They have been used for the oxidation of ascorbic acid113, cysteine114, mercaptoethanol115, hydrazine116, hydroxylamine117 and sulfite118. In pyridine solution the nitroso derivative PcCoNO can be converted under an atmosphere (50 atm) to the nitrito derivative which 119,120 oxidizes triphenylphosphine or 1-octene by O-atom transfer . The tetra-tert-butylphthalocyanine complex of cobalt(II) catalyzes the oxidation of styrene to 1-phenylethanol with at room temperature in the presence of (Figure 39). The analogous reactions with the Mn(III) and Fe(III) complexes are inhibited by the freeradical scavenger TEMPO, indicating a free-radical mechnism. The lack of inhibition of the cobalt(II)-catalyzed oxygenation points to the involvement of a intermediate. The zinc complex is catalytically inactive. In the presence of water-soluble cobalt phthalocyaninetetra(sodium sulfonate) (CoPcTS), 3,4-dimethoxybenzyl alcohol (a lignin model) is catalytically oxidized by dioxygen to 3,4-dimethoxy-benzaldehyde123-127. Typical reaction conditions are 1 atm and 85°C, 12 hours at pH 11, yield 100%.The product yield decreases with decreasing pH; it is 84% at pH 10 and 18% at pH 8. A cationic latex particles had little effect on the yield or rate of oxidations. PcTS complexes of Fe(II), Cu(II) and Ni(II) gave less than 1% yield under comparable conditions.

6. Catalytic Oxidations using Cobalt(II) Complexes

299

The kinetics and mechanism of the oxidation of 2-aminophenol (ap) to 2-aminophenoxazin-3-one (apx) under ambient conditions, catalyzed by the recently synthesized tetrakis(3,5-di-tert-butyl-4-hydroxyphenyl)-dodecachlorophthalocyaninatocobalt(II), (Figure 41) have been studied by spectrophotometry128,129. The rate of ap formation is first-order in and obeys Michaelis-Menten type kinetics with respect to [ap]. The suggested mechanism involves rate-determining inner-sphere electron transfer from coordinated ap to coordinated in the superoxo complex. Cobalt(II)-phthalocyanine [Co(II)pc] supported on active carbon exhibits catalytic activity in the oxidation of sulfide ions to elemental sulfur by dioxygen in aqueous solution131. Hydrophobization of the carbon surface facilitates activation of dioxygen by adsorption. Sulfide ion activation occurs mainly on the supported Co(II)Pc. Water soluble cobalt(II) 2,9,16,23-tetrasulfophthalocyanine, zinc(II)2,9,16,23-tetrasulfo-phthalocyanine, zinc(II)tetracarboxyphthalocyanine, and non-metallic sulfophthalocyanine complexes are catalysts and photocatalysts for the oxidation of sulfide, sulfite and thiosulfate ions by dioxygen132. Typical conditions are 293 K, 1 atm and pH 9.24 in aqueous solution.The cobalt phthalocyanine complexes show high catalytic activity only in the

300

L.I. Simándi

oxidation of sulfide, but oxidation is incomplete and not enhanced by irradiation with visible light.

Zinc phthalocyanine complexes show good catalytic activity only upon simultaneous irradiation with visible light132. Interaction between dioxygen and the long-lived triplet state of these complexes produces reactive singlet dioxygen, which also interacts with compounds containing sulfur in various oxidation states. Immobilized cobalt(II) and zinc(II)phthalocyanine anchored on silica or intercalated in the galleries and cavities of layered double hydroxides (hydrotalcite) and NaX zeolite catalyze or photocatalyze the oxidation of 2mercaptoethanol and sodium thiosulfate133. The activity of the immobilized catalysts for oxidation and photooxidation is lower than that of the complexes in homogeneous phase. This is due to the hindered diffusion of dioxygen and sulfur-containing compounds to the active catalyst sites. The six-coordinate low-spin cobalt(III) complexes and (pc = phthalocyaninate, py = pyridine, dce = 1,2-dichloroethane and thf = tetrahydrofuran) oxidize terminal olefins to the corresponding methyl ketones134. Dioxygen activation and oxygen-atom transfer reactions are possible in terms of the redox couple135. Particulate and solubilized cobalt(II) phthalocyanines (PCs) exhibit catalytic activity in the oxidative decomposition of erythrosine with hydrogen peroxide136.

6. Catalytic Oxidations using Cobalt(II) Complexes

301

7. OXIDATIONS CATALYZED BY COBALT(II) AMINE COMPLEXES 7.1 Catalytic oxidation of substituted anilines In the presence of cobalt(II) salts, o-phenylenediamine (OPD) undergoes facile and selective catalytic dehydrogenation by dioxygen137,138. The organic products are 2,3-diaminophenazine in methanol, and 2Hbenzimidazoles in acetone and other aldehyde or ketone type solvents (Figure 42).

Kinetic studies have shown that the catalytic activity is due to the species, which is reactive toward dioxygen139. Under oxidative conditions, the square-planar o-benzosemiquinonediimine (s-BQDI) complex and the square-pyramidal are formed; they have been characterized by X-ray diffraction.140 The complexes (X = As, Sb, P) have been isolated from the system and analyzed by X/ray diffraction141. Triphenylphosphine added to the system is oxidized to indicating that there is an active superoxo or intermediate in the 142 reacting mixture . Solutions of 2-aminothiophenol (HAT) in methanol or THF, containing cobalt(II) perchlorate rapidly absorb dioxygen at room temperature and atmospheric pressure143, producing 2,2’-diaminodiphenyl disulfide (DADS) in 90% yield (Figure 43). The rate of the catalytic reaction as a function of the 2-aminothiophenol concentration shows a maximum at a cobalt(II) to HAT ratio of 1:2, indicating that a complex of the composition is the major catalytic species. The proposed reaction mechanism for HAT oxidation is shown in Figure 44.

302

L.I. Simándi

Under similar conditions 2-aminophenol (ap) is catalytically converted to the mixture of products shown in Figure 45143,144. The catalytic activity can be ascribed to the complex, which binds dioxygen to form the superoxo species An intramolecular redox reaction then produces the o-benzoquinone monoimine (bqmi) intermediate, which undergoes oxidative dimerization to 2-aminophenoxazine-3-one (apx) and 2,2’-dihydroxyazobenzene (dhab).

6. Catalytic Oxidations using Cobalt(II) Complexes

303

N-benzylidene-2-hydroxyaniline (Bha) and its related derivatives (a-f, Figure 47) can be catalytically oxidized by in the presence of the hydroxo-bridged dicobalt complex (Figure 46) in DMF at 90°C145.

The oxidation products were the corresponding 2-substituted benzoxazoles (Box), formed in yields higher than 90% (Figure 47). Hydrolysis of the starting compound with water produced 2-aminophenol in a side reaction. It was oxidized to 2-aminophenoxazine-3-one. Galvinoxyl radical did not affect the rate of oxidation, therefore no free radical reactions seem to be involved.

304

L.I. Simándi

Kinetic studies reveal first order dependence on both the catalyst concentration and dioxygen pressure. The dependence of the initial rate on the substrate concentration shows a saturation type behavior, indicating reversible initial coordination of the substrate to (A), with subsequent coordination of to the catalyst, producing intermediate X, which is oxidized in the subsequent rate-determining step (Figure 48).

The proposed structure of intermediate X is shown in Figure 49.

6. Catalytic Oxidations using Cobalt(II) Complexes

305

In an atmosphere of dioxygen, the catalytic oxidation of 1,2diaminobenzene takes place in the presence of the dinuclear cobalt complex of the 24-member macrocyclic ligand OBISDIEN 146. In water at 25°C the only oxidation product is 2,3-diaminophenazine, which is formed via a complex mechanism involving o-benzoquinonediimine and a complex as intermediates. The mechanism shown in Figure 50 has been proposed to interpret the observed kinetic behavior.

306

7.2

L.I. Simándi

Oxidation of miscellaneous substrates

The derivatives of the dicobalt(III) OBISDIEN complex, formed under O2 atmosphere may also bind a bridging substrate, such as oxalic147,148 or ketomalonic149 acid and subsequently undergo intramolecular redox reaction with substrate oxidation. Other substrates studied in the OBISDIEN-dicobalt-dioxygen system included phosphonoformic acid, malonic acid, ethylenediamine, glycine, catechols and others150. The size-fit relationship is important for the formation of the dioxygen complex containing the bridging substrate and for the redox reaction to occur. The substrate becomes oxidized only after it forms the reactive intermediate complex.

6. Catalytic Oxidations using Cobalt(II) Complexes

307

Phosphite ion coordinated to dinuclear complex is oxidized to phosphate via apparent O-atom insertion into the P-H bond, which resembles a hydroxylation reaction151. The proposed mechanism (Figure 51) of phosphite to phosphate oxidation within the complex involves the intermediate in Figure 52.

8. OXIDATIONS CATALYZED BY COBALT(II) PYRIDINE COMPLEXES 8.1

epoxidation

In the presence of t-BuOOH and or as catalyst, undergoes oxidation by to pinene oxide (PO), trans-verbenol (Vol) and verbenone (Vone) (Figure 53)152-154. Typically 8-12 mol% t-BuOOH and 0.15 mol% catalyst are used in at 60-100°C. In reaction times of 24 hours, a maximum of 60% verbenone is obtained. Initial Co(II) to Co(III) oxidation of the catalyst by t-BuOOH affords butoxy and butylperoxy radicals. The former abstracts an allylic hydrogen from αpinene, producing an allyl radical, which reacts with yielding transverbenol and verbenone. The butylperoxy radical adds to the double bond of The resulting alkylperoxy radical decomposes to pinene oxide.

308

L.I. Simándi

An efficient catalytic system for oxygenation of phenols to p- or oquinones with molecular oxygen was achieved by utilization of a catalyst consisting of and the multidentate N-heterocyclic podand ligand, N,N'-bis-2-(2-pyridyl)ethyl)-2,6-pyridinedicarboxamide, 2-BPEPA (Figure 54)155.

In acetonitrile the bipyridinecobalt(II) complex activates dioxygen via the reversible formation of a complex156, which dehydrogenates (oxidizes) N-methylanilines, benzyl alcohol, and aldehydes (with subsequent autoxidation). In the absence of substrate, the complex reduces dioxygen via residual water in the solvent to generate HOOH.

9. COBALT-FENTON SYSTEMS According to recent reports157, Fenton chemistry involves the formation of hydroperoxide (ROOH) adducts (A) of reduced transition metal ions [iron(II), copper(I), and cobalt(II)] via nucleophilic addition, e.g.

where B = py or These reactive intermediates (A) bind dioxygen to form adducts of the type (B), which react selectively with methylenic carbon centers of hydrocarbons to form ketones:

and with arylolefins to form dioxygenation products:

6. Catalytic Oxidations using Cobalt(II) Complexes

This phenomenon has been termed oxygenated Fenton chemistry. It has also been observed with in 4:1 MeCN/py, which effects ketonization of the methylenic centers of cyclohexane, cyclohexene and ethylbenzene158-160.

10.

CATALYZED OXIDATIONS

10.1 Under Mukaiyama’s conditions161, aromatic (benzoin, 4,4’-dimethylbenzoin and 4,4’-dimethoxybenzoin [anisoin]) are readily oxidized by dioxygen at room temperature in the presence of excess aldehyde or aldoacetal with catalytic amounts of or under homogeneous conditions162. The stoichiometric equation is shown in Figure 55.

Typical reaction conditions are: 1.2 mmol anisoin, 0.020 mmol 4 mL 1,2-dichloroethane, 3.6 mmol iso-butyraldehyde (0.9 mmol added initially, the remainder added dropwise over 1 hour at room temperature). Yield of diketone 94%.

10.2 Substituted phenols The oxidation of substituted phenols has both biological and synthetic relevance163-166. catalyzes the oxidation of substituted phenols Pac by in the presence of 3-methylbutanal at 40°C in 1,2-dichloroethane (Figure 56). No oxidation takes place in the absence of a catalyst. The major oxidation product of 2,6-dimethylphenol is the corresponding diphenoquinone DQa, whereas 2,6-di-t-butylphenol affords comparable amounts of benzoquinone BQb and DQb. and act similarly.

309

310

L.I. Simándi

3-Methylbutanal is the source of the corresponding hydroperoxide, which is the active oxidant in the reaction.

10.3 Pinanediols In the presence of a 3-fold excess of sacrificial 2-methylpropanal, effects the aerobic oxidative cleavage (Figure 57) of (+)- and (-)pinanediols (PDa,b) to the enantiomerically pure (+)- and (-)-cis-pinonic acids (PAa,b)167. In 1,2 dichloroethane at room temperature, 88% yield was observed in a reaction time of 6 hours.

A variety of secondary and benzylic alcohols can be oxidized by in 168,169 the presence of 2-methylpropanal, a sacrificial aldehyde both in the presence and absence of a metal catalyst in a homogeneous phase. The results are compared in Figure 58. is the most efficient catalyst as compared with Cu, Ni, Pd, Fe, Mn acac complexes. Depending on the substrate, yield enhancements by a factor of 1.5 - 24 are observed for

6. Catalytic Oxidations using Cobalt(II) Complexes

311

identical reaction times.

Similar yields were obtained with a heterogenized cobalt catalyst prepared by copolymerization of [2-(acetoacetoxy)ethyl methacrylate(1-)] with N,N-dimethylacrylamide and N,N’-methylenebis(acrylamide)168.

11. OXIDATIONS VIA ALKYLPEROXOCOBALT COMPLEXES The butylperoxocobalt(III) complex of the pentadentate ligand N,N-bis[2-(2-pyridyl)ethyl]pyridine-2,6-dicarboxamide Figure 59) oxidizes alkanes upon thermal decomposition170. The complex can be prepared by reacting with in dichloromethane.

312

L.I. Simándi

When cyclohexane is used as the substrate, cyclohexanol, cyclohexanone and cyclohexyl chloride are the products. In single turnover oxidation of cyclohexane at the optimum temperature of 80°C, a maximum yield of 59% of the oxidized products is obtained. The mechanism of cyclohexane oxidation involves homolytic scission of the 0-0 bond exclusively. The radicals generated abstract an Hatom from cyclohexane to afford cyclohexyl radicals, which in turn react with dioxygen and produce cyclohexanol (CyOH) and cyclohexanone (CyO) presumably via the Russell-type termination reaction171:

The oxidation of cyclohexane can be either stoichiometric or catalytic. In the presence of excess TBHP, higher yields of oxidized products are obtained, indicating multiple turnovers. Twelve additional Co(III)-alkylperoxo complexes have been synthesized, including those of the ligand N,N-bis[2-(1-pyrazolyl)-ethyl]-pyridine2,6-dicarboxamide (Figure 60), with various primary, secondary, and tertiary R groups172. When the various complexes are warmed (60-80°C) in dichloromethane in the presence of cyclohexane(CyH), the formation of cyclohexanol (CyOH ) and cyclohexanone (CyO) is observed in good yields. Homolysis of the O-O bond in the complexes generating radicals is responsible for the alkane oxidation. Since species are converted into complexes at the end of a single turnover in stoichiometric oxidations, catalytic systems can be generated by the addition of excess ROOH to the reaction mixtures. Catalytic oxidations proceed at considerable rates at moderate temperatures

6. Catalytic Oxidations using Cobalt(II) Complexes

313

and involve species as a key intermediate. Turnover numbers in excess of 100 and ca. 10% conversion of CyH to CyOH and CyO are achieved in 4 h in most catalytic oxidations.

12. OXIDATIONS WITH Co-CYCLIDENE COMPLEXES Busch and coworkers173-175 have synthesized a series of remarkable vaulted and lacunar cobalt(II) cyclidene complexes, which reversibly bind dioxygen and undergo autoxidation to cobalt(III) species. In some cases they exhibit catalytic properties in the oxidation of phenol derivatives. The totally synthetic superstructured cobalt(II) cyclidene complexes (Figure 61) CoA (vaulted), CoB (unbridged) and CoC (lacunar), function as both oxidase and oxygenase models in oxygenation of substituted phenols176. The vaulted complex CoA catalyzes the oxidation of 2,6-di-tert-butylphenol in acetonitrile solution. Typically, at 25°C and 1 atm a conversion of 37% is obtained in 24 hours, the products being 5% 2,6-di-tert-butyl-1,4-benzoquinone (DTBQ) and 22% 3,5,3’,5’-tetra-tertbutyl-4,4’-diphenoquinone (TTBDQ ). Irreversible loss of the catalyst occurs due to autoxidation. The product distribution is determined by the competing reactions shown in Figure 62. Previous studies on the catalytic oxidation of phenols led to the conclusion that the phenoxyl free radical required for interpretation of the observed product pattern and kinetic behavior is formed via H-atom abstraction by the omnipresent superoxocobalt species. In this work an alternative mechanism is proposed, involving electron transfer between the same pair of reactants, followed by the loss of a proton, as described in Figure 63. Electron transfer is pictured to occur through the delocalized system of the phenol, which is supported by the results of molecular mechanics studies173.

314

L.I. Simándi

The detailed reaction mechanism proposed is shown in Figure 64. The main mechanistic features of the catalytic oxidation/oxygenation of 2,6-di-tert-butylphenol are analogous to those described earlier177. The phenoxy radical generated by one-electron transfer can be attacked by either the superoxo complex (oxygenase action) or another phenoxy radical to form either a adduct or TTBDQ, respectively. The adduct releases benzoquinone and yields the species; the latter abstracts an electron from another phenol molecule, releasing water and regenerating the catalyst.

6. Catalytic Oxidations using Cobalt(II) Complexes

315

316

L.I. Simándi

13. OXIDATIONS WITH COBALT PEPTIDE COMPLEXES Recently, there has been increasing interest in metallopeptides as nucleic acid modification reagents178 that can act as affinity cleavage appendages to nucleic acid binding domains179,180. The development of site-selective DNA strand scission agents may involve the substrate oxidation reactions of metallo-Gly-Gly-His and its derivatives181. Photirradiated mixtures of Co(II) and under ambient totally converted form I DNA to form II182. An oxygenated Co(III)-peptide complex is presumably formed upon photoirradiation, which is capable of inducing DNA strand scission by OH radical generation via His-Co coordination in close proximity to DNA.

6. Catalytic Oxidations using Cobalt(II) Complexes

317

The stereoisomerism and equilibrium properties of dioxygen carrying cobalt(II) complexes of histamine and its derivatives have recently been reinvestigated183. The Co(II) - glycyl-L-histidine-imidazole Co(II)-L-histidylglycineimidazole and the corresponding L-histidine and glycylglycine systems have been studied184. Potentiometric, gas-volumetric and spectroscopic (UV/VIS, near IR, ESR) measurements indicated the presence of two types of ternary species: with parent complexes containing a deprotonated amide group (as for glycylglycine and similar dipeptides) and with parent histidinelike active complexes. The overall stability constants of the ternary complexes have been determined. An increase in reversibility of dioxygen uptake was found in both cases relative to the systems without imidazole. The oxygen uptake by Co(II) complexes with a group of diastereoisomeric dipeptides, consisting of alanine and leucine in various chiral forms, has been studied in aqueous solution185.

15. CARBOXYLATOCOBALT COMPLEXES AND SALTS Cobalt(II) ion has been found to catalyze the dioxygen driven oxidation of N-(phosphonomethyl)iminodiacetic acid (PMIDA ) to N-(phosphonomethyl)glycine (PMG ) in aqueous solution186. Additional products are and formic acid. This homogeneous catalytic conversion is novel and represents, in effect, an oxidative dealkylation of one carboxymethyl moiety, yielding the Nsubstituted glycine. The reaction is selective to the desired product PMG when carried out at the natural pH of the free acid substrate (approximately 1-2) and when carried out at substrate loadings less than 5% by weight. PMG is the active agent in the herbicide Roundup187.

Kinetic studies on dilute systems have been carried out. The reaction is first order in substrate and The oxygen pressure dependence exhibits saturation kinetics, while the selectivity increases as the oxygen pressure increases. The rate is also inversely proportional to The proposed mechanism consists of the reaction sequence (a) – (d), involving complexation of Co(II) and generation of the active oxidant via oxidation by Its oxidative dealkylation to the product PMG is shown in Figure 65.

318

L.I. Simándi

In the presence of diacetyl, the acetylation of adamantane (Ad) is catalyzed by in acetic acid at 60°C under 1 atm of (Figure 66). After 2 hours of reaction, the products are 1-acetyladamantane (AAD) (47%) 1,3-diacetyl-adamantane (20%) and 3-acetyladamantan-1-ol (6%)188.

6. Catalytic Oxidations using Cobalt(II) Complexes

No reaction takes place in the absence of with Co(II) or in the presence of with Co(III), which points to a cobalt-dioxygen complex as key intermediate. It reacts with diacetyl to produce acetyl free radicals, which in turn are converted to acetylperoxyl radicals via reaction with The latter are capable of abstracting an H-atom from adamantane, which constitutes the hydrocarbon activating step (Figure 67).

Cobalt salts have been extensively used as catalysts for the oxidation of monoterpenes with for the flavor and fragrance industry189, 190,191. The addition of NaBr to increases the conversions of limonene (L), (aP) and (bP) in autoxidations, by dioxygen (Figure 68)192,193.

319

320

L.I. Simándi

The best selectivities to various allylic oxygenated products (11 – 24 %) were achieved at a ratio of 1. Substitution of Co(II) by Mn(II) up to 30% had no effect on the conversion, but shifted the selectivity significantly towards alcohols and acetates. This was ascribed to a decrease in the overall redox potential, which suppressed further oxidation and side reactions of the primary products. Cobalt(II) acetate is slightly soluble in boiling nonane and decane and promotes autoxidation to ketones as major products It is proposed that the reaction takes place via a cobalt(III)-peroxy species which reacts with the alkanes to form cobalt(II)-alkylhydroperoxide complexes. These complexes decompose to form preferentially ketones, with concomitant regeneration of cobalt(II) acetate194.

16. MISCELLANEOUS COBALT CATALYSTS N-hydroxyphthalimide, a radical catalyst, has been reported to promote the oxidation of various organic substrates such as diols, alkylbenzenes, cycloalkanes and adamantanes by dioxygen under mild conditions195-199. The oxidation takes place both in the absence and presence of or Isobutane is converted to t-butyl alcohol with high selectivity at 100°C and an air pressure of 10 atm in benzonitrile over 8 hours200. Acetone and tbutyl hydroperoxide are formed in smaller amounts, the latter being an intermediate of the reaction as indicated by the maximum observed in its concentration. The catalyst used is a combination of N-hydroxyphthalimide (NHPI ) and or may also be used as the active cobalt species, the former being the most active. NHPI is a source of phthalimidoxyl free radicals (PINO), produced by H-atom abstraction from NHPI by the superoxocobalt(III) species or the complex PINO in turn abstracts an H-atom from isobutane to produce an isobutyl radical, which is trapped by dioxygen to afford the tbutylhydroperoxy radical (A) and ultimately t-butyl hydroperoxide. The

6. Catalytic Oxidations using Cobalt(II) Complexes

321

latter is readily decomposed by metal ions to the t-butoxy radical, which is then cleaved to acetone (see Figure 69).

In addition to isobutane, 2-methylbutane, 3-methylpentane and 2,3dimethylbutane also undergo oxidation in the presence of the system200. The exposure of solid [Tp”Co(CO)] (Tp” = hydrotris(3-isopropyl-5methylpyrazolyl)-borate) to excess gas afforded the dioxygen complex A close analog of this paramagnetic complex (Tp’ = hydrotris(3-tert-butyl-5-methylpyrazolyl)-borate)202 decomposes in solution to yield the doubly bridged and more reactive transient intermediate which subsequently abstracts H-atoms from the ligand (Figure 70)203.

322

L.I. Simándi

The chemistry of cobalt-dioxygen complexes of hydrotris(pyrazolyl)borate ligands (Tp(R)) has been reviewed204 with special reference to the formation of low-valent metal-peroxo and high-valent metal-oxo species, such as Co(II)-superoxo, alkylperoxo and dinuclear complexes. In the hydrotris(3,5-diisopropyl-1-pyrazolyl)borate ligand system oxygenation of the proximal isopropyl substituents on is mediated by the and the alkylperoxocobalt species.

17. CONCLUSIONS The cobalt(II) species most widely used as catalysts in oxidations by are complexes with salen, porphyrin, phthalocyanine, acac, dimethylglyoximato, amine, pyridine, cyclidene and carboxylato ligands. Low-spin cobalt(II) complexes are inherently reactive toward dioxygen due to their d7 electron configuration. Many of the binding reactions are very fast, therefore, few rate constants have been determined and special techniques such as stopped-flow or flash photolysis had to be employed in kinetic work. The superoxocobalt(III) species formed are in most cases the active intermediates in catalytic reactions, which occur readily with substituted phenol or aniline derivatives, producing quinone type dehydrogenation products. The key step in catalytic cycles is often H-atom abstraction or electron transfer to the superoxo complex from the substrate, which is converted to a free radical, possibly reacting further with cobalt(III)

6. Catalytic Oxidations using Cobalt(II) Complexes

323

derivatives formed via cobalt-centered oxidation. This is a pathway for the regeneration of cobalt(II) for further catalytic cycles. Alternatively, the intermediate phenoxyl radical may be attacked by the superoxo complex at the para-position, leading to oxygen insertion products. A procedure widely applied for effecting O-atom insertions (olefin epoxidation, hydrocarbon hydroxylation or ketonization) with cobalt(II) catalyts is the addition of sacrificial 2-methylpropanal. It is converted via Hatom abstraction to an acyl radical, which upon reaction with produces an acylperoxyl radical. H-atom abstraction by the latter leads to a hydroperoxide, which is capable of effecting O-atom insertions via freeradical chain reactions. The sacrificial aldehyde is lost via oxidation to an acid or an ester.

18. REFERENCES Jones, R.D.; Summerville, D.A.; Basolo, F. Chem. Rev., 1979, 79, 139. Niederhoffer, E.C.; Timmons, J.H.; Martell, A.E. Chem. Rev., 1984, 84, 137. Busch, D.H. in Oxygen Complexes and Oxygen Activation by Transition Metals. Martell, A.E., Sawyer, D.T., Eds.; Plenum Publishing Corporation: New York, 1988. 4. Busch, D.H.; Alcock, N.W. Chem. Rev., 1994, 34, 585. Bianchini, C.; Zoellner, R.W. In Advances in Inorganic Chemistry, Sykes, A.G.; Ed.; 5. Vol. 44, Academic Press: San Diego, 1997; p. 263. Simándi, L.I. Catalytic Activation of Dioxygen by Metal Complexes, Kluwer Academic 6. Publishers: Dordrecht, 1992, Chapter 1. Vaska, L. Acc. Chem. Res., 1976, 9, 175. 7. Eaton, D.R.; O’Reilly, A. Inorg. Chem., 1987, 26, 4185. 8. Simándi, L.I.; Savage, C.R,; Schelly, Z.A.; Németh, S. Inorg. Chem., 1982, 21, 2765. 9. 10. Van Eps, N; Szundi, I.; Einarsdottir, O. Biochemistry, 2000, 39, 14576. 11. Cabani, S.; Ceccanti, N.; Pardini, R.; Tiné, M.R. Polyhedron, 1999, 18, 3295. 12. Warburton, P.R.; Busch, D.H. Dynamics of Iron(II) and Cobalt(II) Dioxygen Carriers. In Perspectives of Bioinorganic Chemistry, Hay, R.W., Dilworth, J.R., Nolan, K.B., Eds.; JAI Press Ltd.: London, 1993, Vol. 2, pp. 1-79. 13. Busch, D.H.; Stephenson, N.A. Inclusion Compounds Volume 5: Inorganic and Physical Aspects of Inclusion, Atwood, J., Davies, E., MacNicol, D. Eds.; Oxford University Press: Oxford, 1991; pp.276-310. 14. Martell, A.E. in Oxygen Complexes and Oxygen Activation by Transition Metals; Martell, A.E., Sawyer, D.T., Eds.; Plenum Publishing Corporation: New York, 1988. 15. Durham, B.; Anderson, T.J.; Switzer, J.A.; Endicott, J.F.; Glick, M.D. Inorg. Chem., 1977, 16, 271. 16. Weiss, M.C.; Goedken, V.L. J. Am. Chem. Soc., 1976, 98, 389. 17. Suzuki, M.; Furutachi, H; Okawa, H. Coord. Chem. Rev., 2000, 200, 105. 18. Araki, K.; Kuboki, T.; Otohata, M.; Kishimoto, N.; Yamada, M.; Shiraishi, S. J. Chem. Soc., Dalton Trans., 1993, 3647. 19. Knaudt, J.; Forster, S.; Bartsch, U.; Rieker, A.; Jager, E.G. Z. Naturforsch., Sect. B-A, J. Chem. Sci., 2000, 55, 86. 20. Musie, G.T.; Wei, M., Subramaniam, B.; Busch, D.H. Inorg. Chem., 2001, 40, 3336.

1. 2. 3.

324

L.I. Simándi

21. Srinivas, B.; Zacharias, P.S. Transit. Metal Chem., 1991, 16, 521. 22. Nishinaga, A. in Oxygenases and Model Systems, Funabiki, T., Ed.; Kluwer Academic Publishers: Dordrecht, Boston, London, 1997, p. 157-194.

23. Simándi, L.I. Catalytic Activation of Dioxygen by Metal Complexes, Kluwer Academic 24. 25. 26. 27. 28. 29. 30. 31. 32. 33.

34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56.

Publishers: Dordrecht 1992, Chapter 5. Bozell, J.J.; Hames, B.R.; Dimmel, D.R., J. Org. Chem., 1995, 60, 2398. Nishinaga, A.; Tomita, H. J. Mol. Catal., 1980, 7, 179. Zombek, A.; Drago, R.S.; Corden, B.B.; Gaul, J.H. J. Am Chem. Soc., 1981, 103, 7580. Nishinaga, A.; Tomita, H.; Nishizawa, K.; Matsuura, T.; Ooi, S.; Hirotsu, K. J. Chem. Soc., Dalton Trans., 1981, 1504 . Nishinaga, A.; Tomita, H.; Matsuura, T. Tetrahedron Lett., 1979, 2893. Bolzacchini, E.; Chiavetto, L.B.; Canevali, C.; Morazzoni, F.; Orlandi, M.; Rindone, B. J. Mol. Catal. A-Chemical, 1996, 112, 347. Gold, M.H.; Wariishi, H.; Akileshwaran, L.; Mino, Y.; Loher, T.M. Lignin Enzymic and Microbial Degradation, INRA Publications, Paris, 1987. Higuchi T., Biosynthesis and Biodegradation of Wood Components, Academic Press, New York, 1985. Basolo, F.; Hoffmann, B.M.; Ibers, J.A. Acc. Chem. Res., 1975, 8, 384. Bassoli, B.; Brambilla, A.; Bolzacchini, E.; Chioccara, F.; Morazzoni, F.; Orlandi, M.; Rindone, B. Green Chemistry, Designing Chemistry for the Environment, ACS Symposium Series 626, (P.T. Anastas, T.C. Williamson, Eds), American Chemical Society, Washington DC, 1996. Furutachi H.; Okawa, H. Bull. Chem. Soc. Japan, 1998, 71, 671. Böttcher, A.; Elias, H.; Jager, E.G.; Langfelderova, H.; Mazur, M.; Müller, L.; Paulus Oka, T.; Simpson, F.J.; Krishnamurty, H.G. Can. J. Microbiol., 1972, 18, 493. Balogh-Hergovich E.,; Speier, G. J. Mol. Catal., 1992, 71, 1. Nishinaga, A.; Tojo, T.; Matsuura, T. J. Chem. Soc., Chem. Commun., 1974, 896. Nishinaga, A.; Kuwashige, T.; Tsutsui, T.; Mashino, T.; Maruyama, K. J. Chem. Soc., Dalton Trans., 1994, 32, 805. Iliev, V.; Ileva, A.; Bilyarska, L. J. Mol. Catal. A-Chemical, 1997, 126, 99. Bhatia, B.; Punniyamurthy, T.; Iqbal,. J. J.Org. Chem., 1993, 58, 5518. Bhatia, S.; Punniyamurthy, T.; Bhatia, B.; Iqbal,. J. Tetrahedron, 1993, 49, 6101. Punniyamurthy, T.; Bhatia, B.; Iqbal,. J. Tetrahedron Lett., 1993, 34, 4657. Punniyamurthy, T.; Iqbal,. J. Tetrahedron Lett., 1994, 35, 4003. Punniyamurthy, T.; Iqbal,. J. Tetrahedron Lett., 1994, 35, 4007. Kalra, S.J.S.; Punniyamurthy, T.; Iqbal, J. Tetrahedron Lett., 1994, 35, 4847. Punniyamurthy, T.; Bhatia, B.; Iqbal, J. J.Org. Chem.,1994, 59, 850. Punniyamurthy, T.; Kalra, S.J.S.; Iqbal, J. Tetrahedron Lett, 1995, 36, 8497. Reddy, M.M.; Punniyamurthy, T.; Iqbal, J. Tetrahedron Lett., 1995, 36, 159. Punniyamurthy, T.; Reddy, M.M.; Kalra, S.J.S.; Iqbal,. J. J. Pure Appl. Chem., 1996, 619. Maikap, G.C.; Guhathakurta, D.; Iqbal, J. Syn. Lett.,1995, 189. Iqbal, J.; Bhatia, S.; Reddy, M.M. Synth. Commun., 1993, 23, 2285. Punniyamurthy, T.; Bhatia, B.; Reddy, M.M.; Maikap, G.C.; Iqbal, J. Tetrahedron, 1997, 53, 7649. Simándi, L.I. Int. Rev. Phys. Chem., 1989, 8, 21. Simándi, L.I.; Fülep-Poszmik, A.; Németh, S. J. Mol. Catal., 1988, 48, 266. Rhodes, B.; Rowling, S.; Tidswell, P.; Woodward, S.; Brown, S.M. J. Mol. Catal.AChemical, 1997, 116, 375.

6. Catalytic Oxidations using Cobalt(II) Complexes 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94.

325

Ganeshpure, P.A; Sudalai, A.; Satish, S. Proc. Indian Acad. Sci.-Chem. Sci., 1991, 103, 741. Németh, S.; Szeverényi, Z.; Simándi, L.I. Inorg. Chim, Acta Lett., 1980, 44, L107. Németh, S.; Simándi, L.I. Inorg. Chim. Acta Lett., 1982, 64, L21. Németh, S.; Fülep-Poszmik, A.; Simándi, L.I. Acta Chim. Acad. Sci. Hung., 1982, 110, 461. Németh, S.; Fülep-Poszmik, A.; Simándi, L.I. J. Mol. Catal., 1988, 48, 265. Németh, S.; Simándi, L.I. J. Mol. Catal, 1982, 14, 87. Németh, S.; Simándi, L.I. J. Mol. Catal., 1982, 14, 241. Simándi, L.I.; Bama, T.M.; Németh, S. J. Chem. Soc. Dalton, 1996, 473. Simándi, L.I.; Barna, T.M.; Korecz, L,; Rockenbauer, A. Tetrahedron Lett., 1993, 34, 717. Simándi, L.I.; Barna, T.M.; Argay, Gy.; Simándi, T. L. Inorg. Chem., 1995, 34, 6337. Simándi, T. L.; Simándi, L.I. React. Kinet. Catal. Lett., 1998, 65, 301. Simándi, L.I.; Simándi, T. L. J. Chem. Soc., Dalton Trans., 1998, 3275. Simándi, L.I.; Simándi, T.L.: J. Mol. Catal. A: Chemical, 1997, 117, 299. Giannotti, C.; Gaudemer, A.; Fontaine, C. Tetrahedron Lett., 1970, 3209. Merienne, C.; Giannotti, C.; Gaudemer, A. J. Organomet. Chem., 1973, 54, 281. Bied-Charreton, C.; Gaudemer, A. J. Am. Chem. Soc., 1976, 98, 3997. Jensen, F.R.; Kiskis, R.C. J. Organomet. Chem., 1973, 49, C46. Giannotti, C.; Fontaine, C. J. Organomet. Chem., 1973, 52, C41. Gupta, B.D.; Roy, S. Inorg. Chim. Acta, 1985, 108, 261. Kijima, M.; Yamashita, H.; Sato, T. J. Organomet. Chem., 1994, 474,177. Selvaraj, S.; Natarajan, P. Indian J. Chem., Sect. A, 1995, 34, 253. Venkatalakshmi, N.; Rajasekharan, M. V. Proc. Indian Acad. Sci.-Chem.Sci., 1995, 107, 327. Gupta, B.D., Kanth, V.V.; Singh, V. J. Organomet. Chem., 1998, 570, 1. Metalloporphyrins in Catalytic Oxidations, Sheldon, R.A., Ed. ; Marcel Dekker: New York, 1994. Metalloporphyrin Catalyzed Oxidations, Kluwer: Dordrecht, 1994. Grinstaff, M.W.; Hill, M.G.; Labinger, J.A.; Gray, H.B. Science, 1994, 264. Murahashi, S.I.; Naoya, T.; Komiya, N. Tetrahedron Lett., 1995, 36, 8059. Tabushi, I. Coord. Chem. Rev., 1988, 86,1. Watanabe, Y. in Oxygenases and Model Systems, T. Funabiki, Ed.; Kluwer: Dordrecht,1997; p. 223. Shimada, H.; Sliger, S.G.; Yeom, H.; Ishimura, Y. in Oxygenases and Model Systems, T. Funabiki, Ed.; Kluwer: Dordrecht, 1997; p. 195. Mansuy, D.; Battioni, P. In Bioinorganic Catalysis, J. Reedijk, E. Bouwman, Eds.; Edition, Marcel Dekker: New York, Basel, 1999, p. 323. Sheldon, R.A.; Kochi, J.K. Metal-catalyzed Oxidations of Organic Compounds, Academic Press: New York, London, Sydney, San Francisco, 1981. Walker, F.A. J. Am. Chem. Soc., 1970, 92, 4235. Basolo, F.; Hoffman, B.M.; Ibers, J.A. Acc. Chem. Res., 1975, 8, 384. Balch, A.L.; Mazzanti, M.; Olmstead, M.M. J. Chem. Soc., Chem. Commun., 1994, 269. Balch, A.L.; Mazzanti, M.; St. Claire, T.N.; Olmstead, M.M. Inorg. Chem., 1995, 34, 2194. Ellis, Jr., P.E.; Lyons, J.E. J. Chem. Soc. Chem. Commun., 1989,1188. Ellis, Jr., P.E.; Lyons, J.E. J. Chem. Soc. Chem. Commun., 1989,1190.

326 95. 96. 97. 98. 99.

100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. 131.

L.I. Simándi Ellis, Jr., P.E.; Lyons, J.E. J. Chem. Soc. Chem. Commun., 1989, 1316. Ellis, Jr., P.E.; Lyons, J.E. Catal. Lett., 1989, 3, 389. Lyons, J.E.; Ellis, Jr., P.E. Catal. Lett., 1991, 8, 45. Ellis, Jr., P.E.; Lyons, J.E. Coord Chem. Rev., 1990, 105, 18. Lyons, J.E.; Ellis, Jr., P.E.; Durante, V.A. In Studies in Surface Science and Catalysis, R. Graselli, Ed.; Elsevier: New York, 1991, Vol. 67, p. 99. Lyons, J.E.; Ellis, P.E. In Metalloporphyrins in Catalytic Oxidations, R.A. Sheldon, Ed.; Dekker: New York, 1994, p. 291. Lyons, J.E.; Ellis. P.E.; Myers, H.K. J. Catal, 1995, 155, 59. Tolman, C.A.; Druliner, J.P.; Krusic, P.J.; Nappa, M.J.; Seidel, W.C.; Williams, I.D.; Ittel, S.D. J. Mol. Catal., 1988, 48, 129. Sussine, L.; Brazi, E.; Robine, A.; Mimoun, H.; Fischer, J.; Weiss, R. J. Am. Chem. Soc.,1985, 107, 3534. Mandal, A.K.; Khanna; V.; Iqbal, J. Tetrahedron Lett., 1996, 37, 3769. Mandal, A.K., Iqbal, J. Tetrahedron, 1997, 53, 7641. Horváth, I.T.; Rábai, J. Science, 1994, 266, 72. Pozzi, G.; Montanari, F.; Quici, S. Chem. Commun., 1997, 69 Tang, H; Shen, C.Y.; Lin, M.R.; Sen, Y. Inorg. Chim. Acta, 2000, 300, 1109. Chen, H.; An, T.C.; Fang, Y.J.; Zhu, K. J. Mol. Catal. A-Chemical, 1999, 147, 165. Colonna, S.; Gaggero, N.; Montanari, F.; Pozzi, G.; Quici, S. Eur. J. Org. Chem., 2001, 1, 181. Zwart, J.; van Wolput, J.H.M.C. J. Mol. Catal., 1979, 5, 51. Manassen, J. J. Catal. Rev. Sci. Eng., 1974, 9, 223. Ogata, Y.; Marumo, K.; Kwan, T. Chem. Pharm. Bull., 1969, 17, 1194. Dolansky, J.; Wagnerová, D.M.; Veprek-Siska, J. Coll. Czech. Chem. Commun., 1976, 41, 2326. Maas, Th.A.M.M.; Kuijer, M.; Zwart, J. J. Chem. Soc., Chem. Commun., 1976, 86. Wagnerová, D.M.; Schwertnerová, E.; Veprek-Siska, J. Coll. Czech. Chem. Commun., 1973, 38, 756. Wagnerová, D.M.; Schwertnerová, E.; Veprek-Siska, J. Coll. Czech. Chem. Commun., 1974, 39, 3036. Wagnerová, D.M.; Coll. Czech. Chem. Commun., 1987, 52,1223. Ercolani, C.; Pennesi, G. Inorg. Chim. Acta, 1985, 101, L41. Ercolani, C.; Paoletti, A.M.; Pennesi, G.; Rossi, G. J.Chem. Soc., Dalton Trans., 1991, 1317. Sugimori, T.; Horike, S.; Tsumura, S.; Handa, M.; Kasuga, K. Inorg. Chim. Acta, 1998, 283. 275. Kasuga, K.; Tsuboi, K.; Handa, M.; Sugimori, T.; Sogabe, K. Inorg. Chem. Commun., 1999, 2, 507. Zhu, W.M.; Ford, W.T. J. Mol. Catal., 1993, 78, 367. TurkH.; Ford, W.T. J. Org. Chem., 1991, 56, 1253. Srinivasan, S.; Ford, W.T. J. Mol. Catal., 1991, 64, 291. Hassanein, M.; Ford, W.T. J. Org. Chem., 1989, 54, 3106. Turk, H.; Ford, W.T. J. Org. Chem., 1988, 53, 460. Szeverényi, Z.;. Milaeva, E.R; Simandi, L.I. J. Mol. Catal., 1991, 67, 251. Szeverényi, Z.; Milaeva, E.R.; Simándi, L.I. in Dioxygen Activation and Homogeneous Catalytic Oxidation (L.I. Simándi, Ed.), Elsevier, Amsterdam, 1991, p. 171. Milaeva, E.R.; Szeverényi, Z.; Simándi, L.I. Inorg. Chim. Acta, 1990, 167, 139. Vassileva, M.; Andreev, A.; Schulz-Ekloff, G.;. Wohrle, D. React. Kinet. Catal. Lett., 1993, 50, 139.

6. Catalytic Oxidations using Cobalt(II) Complexes

327

132. Iliev, V.; Ileva, A. J. Mol. Catal. A-Chemical, 1995, 103, 147. 133. Iliev, V.; Ileva, A.; Bilyarska, L. J. Mol. Catal. A-Chemical, 1997, 126, 99. 134. Ercolani, C.; Paoletti, A.M.; Pennesi, G.; Rossi, G. J. Chem. Soc. Dalton Trans., 1991, 1317. 135. Ercolani, C.; Pennesi, G. Inorg. Chim. Acta, 1985, 101, L41. 136. Watanabe, T.; Archer, R.D. J. Mol. Catal., 1994, 93, 253. 137. Simándi, L.I.; Németh, S. J. Mol. Catal., 1982, 14, 241. 138. Simándi, L.I.; Barna, T.; Szeverényi, Z.; Németh, S. Pure Appl. Chem., 1992, 64,1511. 139. Simándi, L.I.; Németh, S. Inorg. Chem., 1983, 22, 3151. 140. Peng, S.-M.; Chen,C.; Liaw, D.S.; Wang, Y. Inorg. Chim. Acta, 1985, 101, L31. 141. Simándi, L.I.; Németh, S.; Argay, Gy.; Kálmán, A. Chim. Acta, 1989, 166, 31. 142. Simándi, L.I.; Németh, S. J. Mol. Catal., 1984, 22, 341. 143. Simándi, L.I.; Németh, S.; Rumelis, N. J. Mol. Catal., 1987, 42, 357. 144. Simándi, L.I.; Barna, T.; Szeverényi, Z.; Németh, S. Pure Appl. Chem., 1992, 64, 1511. 145. Chang, S.Y.; Cheng, Y.H.; Uang, B.J.; Cheng, C.P. J. Chem. Soc., Dalton Trans., 1999, 2769. 146. Rosso, N.D.; Szpoganicz, B.; Martell, A.E. Inorg. Chim. Acta, 1999, 287, 193. 147. Martell, A.E.; Motekaitis, R.J. J. Chem. Soc., Chem. Commun., 1988, 915. 148. Martell, A.E.; Motekaitis, R.J. J. Am. Chem. Soc., 1988, 110, 8059. 149. Motekaitis, R.J.; Martell, A.E. Inorg. Chem., 1991, 30, 694. 150. Motekaitis, R.J.; Martell, A.E. Inorg. Chem., 1992, 31, 5534. 151. Motekaitis, R.J.; Martell, A.E. Inorg. Chem., 1994, 33, 1032. 152. Lajunen, M.K.; Koskinen, A.M.P. Tetrahedron Lett., 1994, 35, 4461. 153. Lajunen, M.K.; Maunula, T.; Koskinen, A.M.P. Tetrahedron, 2000, 56, 8522. 154. Lajunen, M.K. J. Mol. Catal. A: Chemical, 2001, 169, 33. 155. Moriuchi, T.; Hirao, T.; Ishikawa, T.; Ohshiro, Y.; Ikeda, I. J. Mol. Catal. A-Chemical, 1995, 95, LI. 156. Sobkowiak, A.; Sawyer, D.T. J. Am. Chem. Soc., 1991, 113, 9520. 157. Sawyer, D.T. Coord. Chem. Rev., 1993, 165, 297. 158. Sawyer D.T.; Kang, C.; Llobet, A.; Redman, C.,J. Am. Chem. Soc., 1993, 115, 5817. 159. Kang, C.; Redman, C.; Cepak, V.; Sawyer, D.T. Bioorg. Med. Chem., 1993, 1, 125. 160. Tung, H.C.; Kang, C.; Sawyer, D.T. J. Am. Chem. Soc., 1993. 114, 3445. 161. Mukaiyama, T. in The Activation of Dioxygen and Homogeneous Catalytic Oxidation (Barton, D.H.R.; Martell, A.E.; Sawyer, D.T., Eds), Plenum Press, New York, 1993. 162. DellAnna, M.M.; Mastrorilli, P.; Nobile, C.F.; Lopez, L. J. Mol. Catal. A-Chemical, 1996, 111, 33. 163. Mastrorilli, P.; Muscio, F.; Suranna, G.P.; Nobile, C.F.; Latronico, M. J. Mol. Catal. AChemical, 2001, 165, 81. 164. Solomon, E.I.; Sundaram, U.M.; Machonkin, T.E. Chem. Rev., 1966, 96, 2563. 165. Nasir, M.S.; Cohen, B.I.; Karlin, K.D. J. Am. Chem. Soc., 1992, 114, 2482. 166. Araki, K.; Kuboki, T.; Otohata, M.; Kishimoto, N.; Yamada, M.; Shiraishi, S. J. Chem. Soc. Dalton Trans., 1993, 3647. 167. Mastrorilli, P.; Suranna, G.P.; Nobile, C.F.; Farinola, G.; Lopez, L. J. Mol. Catal. AChemical, 2000, 156, 279. 168. Giannandrea, R.; Mastorilli, P.; Nobile, C.F.; Suranna, G.P. J. Mol. Catal., 1994, 99, 27. 169. Cicco, S.R.; Latronico, M.; Mastrorilli, P.; Suranna, G.P.; Nobile, C.F., J. Mol. Catal AChemical, 2001, 165, 135. 170. Chavez, F.A.; Nguyen, C.V.; Olmstead, M.M.; Mascharak, P.K. Inorg. Chem., 1996, 35, 6282. 171. Russell, G.A. J. Am. Chem. Soc., 1957, 79, 3871.

328

L.I. Simándi

172. Chavez, F.A.; Rowland, J.M.; Olmstead, M.M.; Mascharak, P.K. J. Am. Chem. Soc., 1998, 120, 9015 173. Busch, D.H.; Alcock, N.W. Chem. Rev., 1994, 94, 585. 174. Busch, D.H.; Jackson, P.J.; Kojima, M; Chmielewski, P.; Matsumoto, N.; Stevens, J.C.; Wei, W.; Nosco, D.; Herron, N.; Ze, N.; Warburton, P.R.; Masarwa, M.; Stephenson, N.A.; Christophy, G., Alcock, N.W. Inorg. Chem., 1994, 33, 910. 175. Masarwa, M.; Warburton, P.R.; Evans, W.E.; Busch, D.H. Inorg. Chem., 1993, 32. 3826. 176. Deng, Z., Busch, D.H. Inorg. Chem., 1995, 34, 6380. 177. Yamada, M.; Araki, K.; Shiraishi, S. J. Chem. Soc., Perkin Trans., 1990, 2687. 178. Long, E.C.; Eason, P.D.; Liang, Q.Met. lonsBiol. Sys. 1996, 33, 427. 179. Dervan, P.B. Methods Enzymol. 1991, 208, 497. 180. Harford, C.; Narindrasorasak, S.; Sarkar, B. Biochemistry, 1996, 35, 4271. 181. Cuenoud, B.; Tarasow, T.M.; Shepartz, A. Tetrahedron Lett., 1992, 3, 895. 182. Ananias, D.C.; Long, E.C. Inorg. Chem., 1997, 36, 2469. 183. Gajda, T.; Henry, B.; Delpuech, J.-J. Inorg. Chem., 1997, 36, 1850. 184. Kufelnicki, A.; Swiatek, M. Polish J. Chem., 1995, 69, 206. 185. Kufelnicki, A; Swiatek, M. Polish J. Chem., 1999, 73, 579. 186. Riley, D.P.; Fields, D.L.; Rivers, W. J. Am. Chem. Soc., 1991, 113, 3371. 187. Hershman, A.; Bauer, D.J. U.S. Patent 4,264,776. 188. Kishi, A.; Kato, S.; Sakaguchi, S.; Ishii, Y. Chem. Commun., 1999, 1421. 189. Encinar, J.M; Beltran, F.J.; Frades, J.M. Chem. Eng. Technol., 1993, 16, 68. 190. El Firdoussi, L.; Benharref, A.; Allaud, S.; Karim, A.; Castanet, Y.; Mortreux, A.; Petit, F. J.Mol. Catal., 1992, 72, L1. 191. De Carvalho, M.-E.; Meunier, B. N.J. Chem., 1986, 10, 223. 192. deFatima, M.; Gomes, T.; Antunes, O.A.C. J. Mol. Catal. A-Chem., 1997, 121, 145. 193. Partenheimer, W. Catal. Today, 1995, 23, 69. 194. Goosen, A; Morgan, D.H. South African J. Chem., 1994, 47, 59. 195. Yoshino, Y.; Hayashi, Y.; Iwahama, T.; Sakaguchi, S.; Ishii, Y. J. Org. Chem., 1997, 62, 6810. 196. Ishii, Y.; Kato, S.; Sakaguchi, S. Tetrahedron Lett., 1996, 37, 4993. 197. Ishii, Y.; Iwahama, T.; Sakaguchi, S.; Nakayama, K.; Nishiyama, Y. J. Org. Chem., 1996, 61, 4520. 198. Ishii, Y.; Nakayama, K.; Takeno, M.; Sakaguchi, S.; Iwahama, T.; Nishiyama, Y. J. Org. Chem., 1995, 60, 3934. 199. Iwahama, T.; Sakaguchi, S.; Nishiyama, Y.; Ishii, Y. Tetrahedron Lett.,1995, 36, 6923. 200. Sakaguchi, S.; Kato, S.; Iwahama, T.; Ishii, Y., Bull. Chem. Soc. Japan, 1998, 71, 1237. 201. Reinaud, O.M.; Theopold, K.H. J. Am. Chem. Soc., 1994, 116, 6979. 202. Egan, J.W., Jr.; Haggerty, A.L.; Rheingold, A.L.; Sendlinger, S.C.; Theopold, K.H. J. Am. Chem. Soc., 1990, 112, 2445. 203. Reinaud, O.M.; Yap, G.P.A.; Rheingold, A.L.; Theopold, K.H. Angew. Chem. Int. Ed. Engl.,1995, 34, 2051. 204. Hikichi, S.; Akita, M.; Moro-Oka, Y. Coord. Chem. Rev., 2000, 198, 61.

Acknowledgement. The author’s work described in this review was supported by the Hungarian Science Fund (OTKA Grant No. T029036).

329

Subject Index 3,5-di-tert-butyl-5-(formyl)-2-furanone

1

163 184

1,10-phenanthroline

2 2,2’-bipyridine 2,3,6-tri-tert-butylphenol

184

269 oxidation of 2,3-dihydroxyphenylpropionate 1,2179 dioxygenase 195 2,4,6-trichlorophenol 2,4,6-tri-tert-butylaniline 269 oxidation of 2,4-di-tert-butylmuconic acid anhydride

3,5-di-tert-butylcatechol oxidation of 3-ethylpentane 3-methylcatechol 3-methylpentane

3 3,3',5,5'-tetra-tert-butyl-4,4'dihydroxystilbene 287, 290 168,173 3,4-PCD 3,5-di-tert-burylcatechol 163 3,5-di-tert-butyl-l,2-benzoquinone 163 163 3,5-di-tert-butyl-2-pyrone 3,5-di-tert-butyl-5-(carboxymethyl)-2163, 167 furanone

193 170 197

4 4,6-bis(1,1 -dimethylethyl)-2H-pyranone

171 4-chlorocatechol 4-tert-butylcatechol

168,169 169,201

5

163 2,6-di-tert-butylphenol 269, 270, 313, 314 oxidation of 2-aminophenol oxidation of 282, 283, 284,299, 302, 303 2-aminothiophenol 301, 302 oxidation of 2-chloroethyl ethyl sulfide.. 227,228,229 199 2-mercaptobenzoic acid 276, 300 2-mercaptoethanol 2-methyl-1-phenyl-2-propyl 185 hydroperoxide 2-methylpropanal 293, 294, 310, 311, 323

270, 283, 286

181

6 195 179, 188

A ABTS acetonitrile

192

227, 230, 245, 247, 251, 252, 253, 260, 261 activation of dioxygen 266, 299 acyl radical 277, 323 adamantane 183, 184, 189, 190, 193, 197, 199, 200, 202, 205 acetylation of 318, 319 alcohols 182, 194, 197, 201, 204, 205 oxidation of 265, 276, 294, 296, 310, 311, 320 198, 201 aldehydes alkanes 160, 182, 187, 189, 192, 197, 198, 200, 202 alkene epoxidation 265, 277, 278

330 160, 182, 189, 190, 202, 265, 77, 295, 296 184, 193, 197 alkyl hydroperoxide alkyl isocyanides 289 oxidation of 158 alkylperoxides 289 alkylperoxocobaloxime(III) 185 alkylperoxyiron(III) 2, 60, 61 amidation 84, 87, 102, 103 amine oxidase 256, 261 amino acid 166 aminopyridine ligands 202 amorphous iron 199 aniline 197 anthracene 82 ascorbate oxidase 199 ascorbic acid 160, 167, 179 asymmetric 267 autoxidation of 2, 59, 60 aziridination

alkenes

B Baeyer-Villiger oxidation BDPMA BIPA bipyridinecobalt(II)

177 195 167, 184, 199 308

167 166 166 bleomycin BLPA BnBPA BPG BPIA BPMCN BPMEN BQPA Busch, D.H bztpen

197, 202 170, 172 172, 174

165 167, 184 188, 195 181, 194

166 268, 313, 323, 328

203

C C=C cleavage

2

values 186 carbon radical 184, 198 100, 118, 121 Casella, L catalases 158, 159 catalyst inactivation 227 catechol 1,2-dioxygenase 162 catechol 1,2-dioxygenases 162, 168 160, 162, 168, catechol dioxygenases 169, 170, 179, 205 Catechol dioxygenases 161, 162 catechol oxidases.81, 82, 83, 84, 100, 123

CCD

168

227, 229, 235, 240, 245, 247, 250, 252, 259, 261 CEESO 230, 233, 238, 242, 248, 250, 253, 260 chiral porphyrins 1, 2, 29, 30, 62 162, 168 chlorocatechol dioxygenases chloroperbenzoic acid 183 chlorosulfonium ion 242 CHP 183, 184 cis-1,2-dimethylcyclohexane 194 cis-cyclooctene 190 cis-dihydroxylation 195 292, 309, 310, 311, 320 280, 282 Co(III)-alkylperoxo oxidation via 312 Co(salen) 269, 270, 271, 273, 276 Co(salen) complexes 270 267, 278, 284, 287, cobaloxime(II) 290, 296 cobalt(II) complexes 265, 267, 269, 317,

CEES

322 cobalt(II) cychdene complexes 313,314 cobalt(II) OBISDIEN 305, 306 cobalt(II) octaethylporphyrin 292 298, 300 cobalt(II) phthalocyanine cobalt(II) porphyrins 292 co-catalysis 252 co-catalyst 251, 255, 256 colloidal Au(0) 227, 253

331 244, 259 competitive oxidation 159 compound I 228, 251, 252 copper 104, 123 aliphatic hydroxylation 98, 99 aromatic hydroxylation

87, 91, 94, 96, 100, 105 84, 102, 123 cofactor biogenesis 79 dioxygen activation 79, 84, 93, 94, 95, 100, 105, 108, 123 79, 114, 116 DNA cleavage 93, 108 hydroperoxo 80, 123 industrial processes 79, 100 model compounds 79, 82, 91, 98, 104 monooxygenase 99, 104 N-dealkylation 81 nitrite reductases ortho-oxygenation of phenols 79, 83 108, 123 oxidase models 80, 82, 97, 102 oxidative coupling 81 oxygenases 87, 88, 90, 93, 94, 97 peroxo phenanthroline DNA oxidation 112, 123 87 phenoxide bridged 79, 110 phenoxyl radical 82 proteins (table) 87, 88, 90 superoxo 82 copper methane monooxygenase Criegee rearrangement 177 168 CTD 173 CTH dioxygen adducts

318 catalysis by 2 Cu-containing proteins 183 cumene hydroperoxide 188 cumyl alcohol 186 cycloalkyl radical 188 cycloheptane 183, 189, 190, 197, 198, cyclohexane 199, 200, 203, 204 185, 186, 193, 197, 199 cyclohexanol

189, 194,198 186 cyclohexyl peroxide 185, 189, 190, 195 cyclooctene 179 cyclopropyl radical 158 cytochrome c oxidase cytocnrome c oxidases 86, 87, 123 cytocnrome P-450 2, 6, 7, 8, 52 ,62, 159

cyclohexene

D 195 DCP dehydrochlroination 169 dehydrogenation of phenols 2 175 density-functional theory 276 depsides 247 dialkyl sulphide 187, 188 diamond core dichlorocatechol 169 257 dimethyl disulfide dimethyl sulfide 184 247 dimethyl sulfoxide dinuclear iron complexes 184, 185, 190 dioxoruthenium(VI) 2, 29 dioxygenase 227, 234, 261 161 dioxygenase-model oxygenations dioxygenases 2 DNA cleavage 114, 116 dopamine 79, 82, 104 198 DPAH DTBC 164, 165, 167, 169, 170, 174, 178 163, 164, 166, 167, 174, 176 DTBSQ 173, 176, 177

E EDDA 192 effect of solvent 241 E-methyl ferulate 271 epoxidation l, 2, 9, 21, 26, 32, 41, 5 9 , 6 2 , 160, 183, 186, 189, 192, 195, 198, 201 epoxides 182, 189, 198, 201 ESI-MS 194, 202

332 ethylbenzene 188 Extended-Hückel 175 extradiol 160, 161, 163, 170, 173, 176, 179

F facial ligand Fe(IV)=O

172 159,204 176 176 161, 194, 204 161, 182, 185, 193, 196, 204 Fenton 191, 193 161, 182, 194, 200, 201 193 FeZSM-5 190 fluorous biphase systems 295

G galactose oxidase Gif system Glaser process gold complexes Gorun, S.M.

2,11, 82, 85, 108 193,200 80 227 94,120

H 204 198 162, 169, 184, 189, 193, 195,204 Haber-Weiss reaction 2,4 halogenated metalloporphyrins 292 halogenated porphyrins 2 187, 200 HDA 164 HDP 165 heme protein models 159 heme-copper oxidases 86 hemerythrin 158 hemocyanin 81, 82, 84,123,158 hemoglobin 2,6,65,158

high-valent histidine HPTB HPTP hydrazobenzene hydrocarbons oxidation of hydrogen peroxide

161 162,166 190 190 200, 278, 280

277, 308 158, 182, 183, 186, 191, 193, 205 hydrogen sulfide 199 hydroquinones 200 hydrotris(pyrazolyl)borate ligand 171 hydroxy radical 191 hydroxylation 1, 2, 21, 29, 33, 62, 161, 182, 186, 192, 199, 200, 204

I immobilized Co(II) phthalocyanine 300 induction period 227, 232, 237, 248, 251, 253, 257, 261 inhibition 227, 234, 242, 248, 251, 254, 261 intradiol 160, 161, 166, 168, 170, 171, 172, 178, 179 iron(III)-peroxo 159 iron-oxo 161, 182, 201, 202, 204 iron-peroxo 182, 185 isopenicillin N synthase 160, 182

K Kemp’s triacid imide 188 ketones 182, 187, 194, 197, 201, 204 ketonization 161, 193, 198, 200 kinetic isotope effect 183, 185 kinetics of dioxygen binding 267 Kitajima,N. 89, 94, 119, 120 Klinman,J.P 84, 102, 118 Knowles,P.F 84, 118 Kodera,M. 93, 107, 120, 121

333

L

N

laccase 82 lacunar cobalt(II) cyclidene complex. 313 Lewis acidity 164, 166 lignin phenolic subunits 270 lignin phenolics oxidation of 270 lipoxygenase 160, 179 LMCT 164, 165, 167, 174 177

167, 176 81 197 292, 325, 326

Lou Gehrig’s disease low-spin Lyons, IE.

M macrocylic cobalt(II) complexes 268 Masuda, H. 93, 120 m-CPBA 183 172, 174 membrane reactor 197 mep 187 meridional 172, 177 metal-based mechanisms 182 metalloenzymes 158 metalloporphyrins 2, 8, 9 methane 160, 161, 182, 187, 97, 205 methane monooxygenase 160, 161, 187, 205 methanol 170, 182, 189, 197 methyl linoleate 198 methylcyclohexane 183 MMO 184 molecular oxygen 158, 159, 163, 169, 173, 181, 187, 201, 205 monooxygenases. 2, 3, 160, 181, 182, 198 monoterpenes oxidation of 319 MPPH 185 mustard gas 2, 256, 257 myoglobin 2, 6, 12, 65, 158

N4Py 194, 203 NaOCl 183, 198 N-benzylidene-2-hydroxyaniline oxidation of 303 NIH shifts 201 nitrate 230, 247, 248, 251, 252, 254 nitrene complexes 2 nitrene transfer 2, 55 nitrilotriacetate 164, 192 nitrite 248 nitrogen monoxide oxidation of 274 nitrosobenzene oxidation of 289 nitrous oxide 2, 34 N-methylmorpholine N-oxide 183 nonlinear fitting 242 NTA 164, 192

O O-atom donors 1, 2, 9, 11, 29, 63 O-atom insertion .273, 289, 290, 307, 323 o-benzosemiquinone-diimine 301 olefins oxidation of 2, 9, 14, 21, 24, 26, 30, 34, 38, 58, 62 O-O-bond lengths 266 o-phenylenediamine oxidation of 281, 301 organocobaloximes 291 organosulfur compounds 257 ortho-hydroxylation 182 O-transfer 278 oxidases 2, 3, 11 oxidations 158, 197 oxidative dehydrogenation of amines 2 of alcohols 2 oxidative dehydrogenations 280, 281 oxidative detoxification 256 oxocobalt(IV) 278, 296

334 oxo-ferryl 159 oxygenases158, 159, 160, 161, 163, 179, 182, 202, 204, 205 oxygenated Fenton chemistry 309 oxygenations 158, 163, 168, 170, 172, 175, 179, 182, 187, 199, 200, 206

P PA

167, 192, 198, 200

pb 186 pCP 195 PCP 196 PDA 165 pentadentate ligand 195, 203 peptidylglycine monooxygenase 79, 82, 104 peracids 158, 159, 202 perfluorinated polyethers 251, 255 perhaloporphyrins cobalt(II) 293 peroxidases 158, 159 peroxo-iron(III) 160 phenol oxidation 79, 83 phenylalanine hydroxylase 160, 182 phosphines oxidation of 1, 2, 14, 16, 55, 58, 61 phosphite oxidation 307 pivalaldehyde 202 PMAH 197 pollutant 195 pollutants 162, 195, 205 polyoxoanion 167, 171 porphyrins 1, 2, 8, 9,10,14, 21, 27, 28, 37, 62 primary amines oxidation of 280 propane 183, 197 propenoidic phenols oxidation of 271, 273 protocatechuate 3,4-dioxygenasel62, 173, 174 protocatechuic acid 169

p-substituted phenols oxidation of 270, 272, 273 pterin-dependent hydroxylases 160 p-xylene 201 pyridine 163, 165, 170, 183, 187, 189, 197, 199, 200, 203 pyrocatechase pyrocatechol

2, 4, 161, 174 161, 170

Q quantum chemical calculation quantum chemical methods quercetin oxidation of quercetin model

178 161 276 276

R radical autoxidation 188 radical cation 243 radical character 174, 175, 177 radical-chain mechanism 228, 244 radical-clock reagents 182 radical-rebound mechanism 183 rate constants for binding 267 rebound mechanism 2 reductive activation of 2 reductive elimination 242 Reedijk, J. 89, 117, 119 Reglier, M. 106 reoxidation 236, 238, 244, 245, 248, 253, 254 reversible binding 266 Rieske dioxygenases 181 ruthenium imine/amine complexes of 2 ruthenium amido complexes 2 ruthenium porphyrins 1, 2 ruthenium(IV) disproportionation of 2

335

S salicylate 199 saturated hydrocarbons oxidation of 1, 2, 19, 28, 41, 63 shape selective 160 singlet state 158 sMMO 160, 182 solvent kinetic isotope effect 227 spin density 175, 177 Stack, T.D.P. 92, 96, 109, 120, 121, 122 steroid epoxidation 2 structural model complexes 158 Structural type 266 substituted phenols oxidation of265, 269, 270, 309, 310, 313 sulfone 234,244,256 sulfoxidation 227, 234, 255, 257, 261 sulfoxide 227, 230, 234, 242, 244, 247, 256, 257, 261 superoxide dismutase 82 superoxide dismutases 158 superoxide ion 158 superoxocobaloxime(III) 280, 284 supported catalysts 2 synergistic effect 255 synthetic dioxygen carriers 266 syringyl alcohol oxidation of 271

T TACN TAML TBHP

170, 171, 172, 174 196 183, 184, 188, 190, 197, 202 184, 185, 198 184, 185 TCP 195, 196 Tempo 193 tert-butylphenol 201 tetera-chlorocatechol 169 tetraamido macrocyclic ligands 196 tetraazamacrocyclic ligand 167

tetradentate ligands 168, 176 tetrahydrosalen dehydrogenation of 275 tetrahydrothiophene 231 thioanisole 185 thioether 227, 235, 238, 242, 243, 244, 246, 249, 250, 256, 259, 261 thioethers oxidation of 1, 2, 14, 17, 19 TMA 197 tmima 183, 186 Tolman, W.B. 91, 99, 119, 120, 121 toluene 173, 183, 184, 186, 199, 201 toluidine 199 topical skin protectant 252 TPA 165, 167, 168, 170, 175, 177, 188, 190, 194, 197, 202 171, 174 TPY 172, 174 trans-2-octene 195 trans-2-phenylmethycyclopropane 184 trans-stilbene 184, 189, 202 tridentate ligand 163, 172 trifluoroethanol 234, 253, 254, 255 triplet state 158 tripodal ligands 164, 165, 167, 169, 185 trispicMeen 166 tryptophan dioxygenase 2, 4 tryptophan hydroxylase 160 tyrosinases79, 81, 82, 83, 86, 98, 100, 102, 104, 123 tyrosine 160, 162, 166, 182, 200, 201 tyrosine hydroxylase 160, 182, 200, 201

W Wacker process water-soluble Water-soluble ligands Wieghardt, K

80 169, 195 169 110, 121

X XANES

174

336

Z zinc

200

acid 160, 161, 180 acid-dependent dioxygenases 160 188

diferric complex diiron(H)

203 203 186 186

cation

159

Catalysis by Metal Complexes Series Editors: R. Ugo, University of Milan, Milan, Italy B.R. James, University of British Colombia, Vancouver, Canada

1.

F.J. McQuillin: Homogeneous Hydrogenation in Organic Chemistry. 1976 ISBN 90-277-0646-8

2.

P.M. Henry: Palladium Catalyzed Oxidation of Hydrocarbons. 1980 ISBN 90-277-0986-6

3.

R.A. Sheldon: Chemicals from Synthesis Gas. Catalytic Reactions of CO and ISBN 90-277-1489-4 1983

4.

W. Keim (ed.): Catalysis in

5.

A.E. Shilov: Activation of Saturated Hydrocarbons by Transition Metal Complexes. 1984 ISBN 90-277-1628-5

6.

F.R. Hartley: Supported Metal Complexes. A New Generation of Catalysts. 1985 ISBN 90-277-1855-5

7.

Y. Iwasawa (ed.): Tailored Metal Catalysts. 1986

8.

R.S. Dickson: Homogeneous Catalysis with Compounds of Rhodium and Indium. ISBN 90-277-1880-6 1985

9.

G. Strukul (ed.): Catalytic Oxidations with Hydrogen Peroxide as Oxidant. 1993 ISBN 0-7923-1771-8

10.

A. Mortreux and F. Petit (eds.): Industrial Applications of Homogeneous Catalaysis. 1988 ISBN 90-2772-2520-9

11.

N. Farrell: Transition Metal Complexes as Drugs and Chemotherapeutic Agents. ISBN 90-2772-2828-3 1989

12.

A.F. Noels, M. Graziani and A.J. Hubert (eds.): Metal Promoted Selectivity in Organic Synthesis. 1991 ISBN 0-7923-1184-1

13.

L.I. Simándi (ed.): Catalytic Activation of Dioxygen by Metal Complexes. 1992 ISBN 0-7923-1896-X

14.

K. Kalyanasundaram and M. Grätzel (eds.): Photosensitization and Photocatalysis Using Inorganic and Organometalic Compounds. 1993 ISBN 0-7923-2261-4

15.

P.A. Chaloner, M.A. Esteruelas, F. Joó and L.A. Oro: Homogeneous Hydrogenation. 1994 ISBN 0-7923-2474-9

16.

G. Braca (ed.): Oxygenates by Homologation or CO Hydrogenation with Metal Complexes. 1994 ISBN 0-7923-2628-8

17.

F. Montanari and L. Casella (eds.): Metalloporphyrins Catalyzed Oxidations. 1994 ISBN 0-7923-2657-1

Chemistry. 1983

ISBN 90-277-1527-0

ISBN 90-277-1866-0

18.

P.W.N.M. van Leeuwen, K. Morokuma and J.H. van Lenthe (eds.): Theoretical Aspects of Homogeneous Catalisis. Applications of Ab Initio Molecular Orbital Theory. 1995 ISBN 0-7923-3107-9

19.

T. Funabiki (ed.): Oxygenases and Model Systems. 1997

20.

S. Cenini and F. Ragaini: Catalytic Reductive Carbonylation of Organic Nitro Compounds. 1997 ISBN 0-7923-4307-7

21.

A.E. Shilov and G.P. Shul’pin: Activation and Catalytic Reactions of Saturated Hydrocarbons in the Presence of Metal Complexes. 2000 ISBN 0-7923-6101-6

22.

P.W.N.M. van Leeuwen and C. Claver (eds.): Rhodium Catalyzed Hydroformylation. 2000 ISBN 0-7923-6551-8

23.

F. Joó: Aqueous Organometallic Catalysis. 2001

24.

R.A. Sánchez-Delgado: Organometallic Modeling of the Hydrodesulfurization and ISBN 1 -4020-0535-0 Hydrodenitrogenation Reactions. 2002

25.

F. Maseras and A. Lledós (eds.): Computational Modeling of Homogeneous Catalysis. 2002 ISBN 1-4020-0933-X

ISBN 0-7923-4240-2

ISBN 1-4020-0195-9

KLUWER ACADEMIC PUBLISHERS – BOSTON / DORDRECHT / LONDON * Volume 1 is previously published under the Series Title: Homogeneous Catalysis in Organic and iNorganic Chemistry