Homogeneous Catalysis With Metal Phosphine Complexes

Download as pdf or txt
Download as pdf or txt
You are on page 1of 494
At a glance
Powered by AI
The document provides an overview of a book on homogeneous catalysis using metal phosphine complexes. It discusses topics such as catalysis, phosphine ligands, and transition metal complexes.

The book is about homogeneous catalysis using metal phosphine complexes as catalysts.

The book covers topics such as catalysis mechanisms, ligand effects, transition metal complexes containing phosphine ligands, and applications to reactions such as hydrogenation and hydroformylation.

Homogeneous

Catalysis with Metal


Phosphine Complexes
MODERN INORGANIC CHEMISTRY

Series Editor: John P. Fackler, Jr.


Texas A&M University

METAL INTERACTIONS WITH BORON CLUSTERS


Edited by Russell N. Grimes

HOMOGENEOUS CATALYSIS
WITH METAL PHOSPHINE COMPLEXES
Edited by Louis H. Pignolet

THE JAHN-TELLER EFFECT AND


VIBRONIC INTERACTIONS IN MODERN CHEMISTRY
I. B. Bersuker

A Continuation Order Plan is available for this series. A continuation order will bring
delivery of each new volume immediately upon publication. Volumes are billed only
upon actual shipment. For further information please contact the publisher.
Homogeneous
Catalysis with Metal
Phosphine Complexes

Edited by

Louis H. Pignolet
University of Minnesota
Minneapolis, Minnesota

PLENUM PRESS • NEW YORK AND LONDON


Library of Congress Cataloging in Publication Data
Main entry under title:
Homogeneous catalysis with metal phosphine complexes.
(Modern inorganic chemistry)
Includes bibliographies and index.
1. Catalysis. 2. Phosphine. 3. Complex compounds. I. Pignolet, Louis H., 1943-
II. Series.
QD505.H65 1983 660.2'995 83-17609

ISBN-13: 978-1-4613-3625-9 e-ISBN-13: 978-1-4613-3623-5


DOl: 10.1007/978-1-4613-3623-5

©1983 Plenum Press, New York


Softcover reprint of the hardcover 1st edition 1983
A Division of Plenum Publishing Corporation
233 Spring Street, New York, N.Y. 10013
All rights reserved
No part of this book may be reproduced, stored in a retrieval system, or transmitted
in any form or by any means, electronic, mechanical, photocopying, microfilming,
recording, or otherwise, without written permission from the Publisher
Contributors

Or. Alan Balch • Department of Chemistry, University of California, Davis


California
Or. John M. Brown. Dyson Perrins Laboratory, South Parks Road,
Oxford, England
Or. Penny A. Chaloner • Department of Chemistry, Rutgers University,
Piscataway, New Jersey
Or. Joseph Chaff. Department of Chemistry, School of Chemistry and
Molecular Sciences, The University of Sussex, Brighton, Sussex,
England
Or. Robert H. Crabtree • Sterling Chemistry Laboratory. Yale Univer-
sity, New Haven, Connecticut
Or. Daniel H. Doughty. Exploratory Chemistry, Div. 8315, Sandia
National Laboratories, Livermore, California
Or. Jack W. Faller. Department of Chemistry, Yale University, New
Haven, Connecticut
Or. T. Adrian George • Department of Chemistry, University of
Nebraska-Lincoln, Nebraska
Or. Balint Heil • University of Chemical Industries, Institute of Organic
Chemistry, Hungarian Academy of Sciences, Research Group for
Petrochemistry, Veszprem, Shonherz Zoltan u. 12, Hungary
Or. Norman L. Holy. Department of Chemistry, Western Kentucky
University, Bowling Green, Kentucky
vi CONTRIBUTORS

Dr. James A. Ibers • Department of Chemistry, Northwestern Univer-


sity, Evanston, Illinois
Nancy L. Jones • Department of Chemistry, Northwestern University,
Evanston, I1Iinois
Dr. Laszlo Marko • University of Chemical Industries, Institute of
Organic Chemistry, Hungarian Academy of Sciences, Research Group
for Petrochemistry, Veszprem, Shonherz Zoltan u. 12, Hungary
Dr. Devon W. Meek. Department of Chemistry, Ohio State University,
Columbus, Ohio
Dr. Louis H. Pignolet • Department of Chemistry, University of Min-
nesota, Minneapolis, Minnesota
Dr. Thomas B. Rauchfuss • School of Chemistry, University of I1Iinois,
Urbana-Champaign, I1Iinois
Dr. D. Max Roundhill • Department of Chemistry, Tulane University,
New Orleans, Louisiana
Dr. Alan R. Sanger • Alberta Research Council, 11315-87 Avenue,
Edmonton, Alberta, Canada
Dr. Chad A. Tolman • Central Research and Development Department,
Experimental Station, E. I. du Pont de Nemours and Company, Wil-
mington, Delaware
Dr. Szilard Taras • University of Chemical Industries, Institute of
Organic Chemistry, Hungarian Academy of Sciences, Research Group
for Petrochemistry, Veszprem, Schonherz Zoltan u. 12, Hungary
Preface

The field of transition metal catalysis has experienced incredible growth


during the past decade. The reasons for this are obvious when one considers
the world's energy problems and the need for new and less energy-
demanding syntheses of important chemicals. Heterogeneous catalysis has
played a major industrial role; however, such reactions are generally not
selective and are exceedingly difficult to study. Homogeneous catalysis
suffers from on-site engineering difficulties; however, such reactions usually
provide the desired selectivity. For example, Monsanto's synthesis of
optically-active amino acids employs a chiral homogeneous rhodium
diphosphine catalyst.
Industrial uses of homogeneous catalyst systems are increasing. It is
not by accident that many homogeneous catalysts contain tertiary phosphine
ligands. These ligands possess the correct steric and electronic properties
that are necessary for catalytic reactivity and selectivity. This point will be
emphasized throughout the book. Thus the stage is set for a comprehensive
treatment of the many ways in which phosphine catalyst systems can be
designed, synthesized, and studied.
The book is intended to provide an up-to-date treatise by experts in
the field of homogeneous catalysis with metal phosphine complexes. Reac-
tion mechanisms are emphasized, and the fine tuning of catalyst reactivity
and selectivity is thoroughly discussed. The techniques and methodology
that are required to determine catalytic reaction pathways are comprehen-
sively described, and numerous case studies are presented. The reader will
gain insight into the subject's many facets ranging from the historical
development of the field to the newest techniques in catalyst design and
study.
Contents

Chapter 1
Historical Introduction
Joseph Chatt

1. The Phosphines 1
1.1. The Complexes 3
1.2. Catalysis 8
References 10

Chapter 2
Mechanistic Studies of Catalytic Reactions Using Spectroscopic and
Kinetic Techniques
Chad A. Tolman and Jack W. Faller

1. Introduction . . . . . . . . . . . . . . . . . . 13
1.1. The Nature of Catalyst Systems . . . . . . . 14
1.2. Mechanisms, Rate Laws, and Dominant Species 16
1.3. The Use of Isotopic Techniques . . . . . . . 22
1.4. The Implication s of Product Stereochemistry on Mechan-
isms . . . . . . . . . . . . 34
1.5. Distinguishing Radical Pathways . . . . 43
2. Applications IlIustrating the Methods . . . . 49
2.1. Butene Isomerization by NiL4 and H 2 S0 4 49
2.2. Olefin Hydrogenation with Wilkinson's Catalyst 53
2.3. The Nickel Catalyzed Cyclooligomerization of Butadiene 64
2.4. Rhodium Catalyzed Olefin Hydroformylation 81
References 102
x CONTENTS

Chapter 3
Structurally-Characterized Transition-Metal Phosphine Complexes
of Relevance to Catalytic Reactions
Nancy L. Jones and James A. [bers

1. Introduction . . . . . 111
2. Alkene Hydrogenation 112
3. Hydroformylation 116
4. Olefin Oligomerization 122
5. Metallacycles 126
6. Summary 129
References 132

Chapter 4
Asymmetric Hydrogenation Reactions Using Chiral Diphosphine
Complexes of Rhodium
John M. Brown and Penny A. Chaloner

1. Introduction 137
2. The Variation of Reaction Conditions 139
2.l. Changing the Catalyst 139
2.2. Changing the Substrate 143
2.3. Changing the Solvent 148
2.4. Changing the Pressure 149
2.5. Summary of the Features of Effective Catalysts 149
3. Studies of Reaction Mechanism 149
3.l. Reaction Rates 151
3.2. X-ray Structural Analysis of Catalysts and Reaction Inter-
mediates 151
3.3. Structure and Dynamics in Solution 153
3.4. Conclusions 159
4. The Origin of Stereoselectivity 160
References 163

Chapter 5
Binuclear, Phosphine-Bridged Complexes: Progress and Prospects
Alan L. Balch

1. Introduction . 167
2. Structural Aspects 170
CONTENTS xi

3. Nuclear Magnetic Resonance Spectroscopic Probes 183


4. Reactivity Patterns for Bridged, Binuclear Complexes 188
4.1. Insertions of Small Molecules into Metal-Metal Bonds 189
4.2. Introduction of a Bridging Ligand with Contraction of the
Metal-Metal Separation 193
4.3. Loss of Carbon Monoxide Accompanied by Carbon
Monoxide Bridge Formation 195
4.4. Two-Center, Two-Fragment Oxidative Addition 197
4.5. Two-Center, Three-Fragment Oxidative Addition 200
4.6. The Formation or Stabilization of Other Novel Bridging
Ligands 201
4.7. Capture of a Second Metal Induced by 2(Diphenylphos-
phino )pyridine 204
5. Demonstrated Catalytic Activity of Binuclear Compounds 206
6. Concluding Remarks 208
References 210

Chapter 6
Hydrogenation and Hydroformylation Reactions Using Binuclear
Diphosphine -Bridged Complexes of Rhodium
Alan R. Sanger

1. Introduction . . . . . . . . . . . 216
2. Hydrogenation of Alkenes or Alkynes 217
2.1. Complexes of Ph 2 PCH 2 PPh 2 218
2.2. Complexes of Diphosphines with Longer Chain Lengths 225
3. Hydroformylation of Alkenes 227
4. Concluding Remarks 233
References ........ 235

Chapter 7
Functionalized Tertiary Phosphines and Related Ligands in Organo-
metallic Coordination Chemistry and Catalysis
Thomas B. Rauchfuss

1. Introduction . . 239
2. Ether Phosphines 240
3. Aminophosphines 244
4. Carbonylphosphines 247
5. Alkenylphosphines 251
6. Cyclopentadienylphosphines 253
xii CONTENTS

7. Concluding Remarks 254


References .... 255

Chapter 8
Polydentate Ligands and Their Effects on Catalysis
Devon W. Meek

1. Introduction 257
2. Advantages of Chelating Polydentate Ligands 258
3. Methods for Syntheses of Polydentate Ligands 261
3.1. Grignard and Alkali-Metal Reagents 261
3.2. Other Synthetic Methods 266
4. Polyphosphine Homogeneous Catalysts 274
4.1. Hydrogenation Catalysis with Chelating Triphosphine
Ligands 275
4.2. Hydroformylation Catalysis with Chelating Diphosphine
Ligands 279
5. Solid-Supported Polyphosphine Catalysts 285
5.1. Coordination of Multiple Phosphine Groups from Supported
Monophosphines 285
5.2. A Triphosphine Ligand Attached to Glass Beads 287
5.3. A NP 2 Ligand Attached to Organic Compounds 288
5.4. Chelating Diphosphines Attached to Organic Polymers 288
6. Concluding Remarks 289
References 293

Chapter 9
Cationic Rhodium and Iridium Complexes in Catalysis
Robert H. Crabtree

1. Introduction . . . . . . . . . . . . . . . 297
2. The Effects of Net Ionic Charge . . . . . . 298
2.1. Effects of the Reactivity of Bound Ligands 298
2.2. Effects on the Types of Ligand Bound . . 299
2.3. Effects on Redox Properties . . . . . . 301
2.4. Effects on Solubility and Catalyst Separation 301
2.5. Counter Ion Effects 302
2.6. Mechanistic Effects 302
2.7. Conclusion 303
CONTENTS xiii

3. Synthetic Methods 303


3.1. Halide Abstraction 303
3.2. Protonation 304
3.3. Electrophilic Attack at a Ligand 304
3.4. Redox 304
4. Cationic Rhodium and Iridium Complexes in Catalysis 304
4.1. Hydrogenation of C=C Groups 304
4.2. Hydrogenation of C=O Groups 310
4.3. Alkane Activation 310
4.4. Hydroformylation 313
4.5. Decarbonylation 313
4.6. Water Gas Shift (WGS) 313
4.7. Polymerization 314
5. Concluding Remarks 314
References 314

Chapter 10
Hvdrogenation Reactions of CO and CN Functions Using Rhodium
Complexes
Balint Heil, Laszlo Marko, and Szilard Taros

1. Introduction . . . . . . . . . . . . . . 317
2. Hydrogenation of Aldehydes ..... . 318
3. Hydrogenation of Ketones Not Containing Other Functional
Groups . . . . . . . . . . . . . . . . 319
3.1. Catalytic Activity and Stereoselectivity . . . . . 319
3.2. Enantioselective Hydrogenation of Ketones ..... 324
4. Hydrogenation of Ketones Containing Other Functional Groups 329
4.1. Selective Hydrogenation of Unsaturated Ketones 329
4.2. Hydrogenation of Keto Acids and Keto Esters . . . . . 331
4.3. Hydrogenation of Aminoketones and Other Biologically
Active Derivatives . . . . . . . . . . . . 333
5. Homogeneous Hydrogenation of Carbon-Nitrogen Double
Bonds .......... . 335
6. Abbreviations for Ligand Names 337
References ........ . 338
xiv CONTENTS

Chapter 11
Oecarbonylation Reactions Using Transition Metal Complexes
Daniel H. Doughty and Louis H. Pignolet

1. Introduction . . . . . . . . . . . . . . . . . . . . . 343
2. Discussion of Decarbonylation Mechanism with RhCI(PPh 3 h 347
2.1. Stoichiometric Decarbonylation of Acid Chlorides 347
2.2. Stoichiometric Decarbonylation of Aldehydes 352
2.3. Catalytic Decarbonylation of Acid Chlorides and
Aldehydes .................. 355
3. Catalytic Decarbonylation of Aldehydes with Cationic Diphos-
phine Complexes of Rh(I) . . . . . . . . . . . . . . . . 358
3.1. Discussion of Mechanism with [Rh(P-Pht Complexes 362
4. Catalytic Decarbonylation of Aldehydes Using Cationic Diphos-
phine Complexes of Ir(I) ............. 369
5. Additional Studies with Bis-Chelate Complexes of Rh(I) 371
6. Decarbonylation of Benzoylchloride with [Rh(dppphr 372
References ................... 372

Chapter 12
Homogeneous Catalysis of Oxidation Reactions Using Phosphine
Complexes
D. Max Roundhill

1. Significance of Metal-Catalyzed Oxidation Reactions 377


1.1. Mechanistic Features of Metal-Catalyzed Oxidations 378
2. Transition Metal Phosphine Oxygen Complexes 378
2.1. Synthesis and Structure 378
2.2. Reactions with Electrophiles 380
3. Oxidation of Alkenes 383
4. Oxidation of Other Substrates 385
4.1. Isocyanides 385
4.2. Carbon Monoxide 385
4.3. Aldehydes and Ketones 386
4.4. Cumene 387
4.5. Tertiary Phosphines 387
5. CO-Oxidations 389
5.1. Alkenes and Tertiary Phosphines 390
5.2. Alkenes and Hydrogen 392
5.3. Isocyanides and Carbon Monoxide 392
5.4. Triphenylphosphine and Carbon Monoxide 393
CONTENTS xv

6. Oxygen Atom Transfer from Metal Phosphine Hydroperoxides


and Superoxides . . . . . . . . . . . . . 393
6.1. Metal Hydroperoxides and Alkylperoxides ... . 394
6.2. Metal Peracyls . . . . . . . . . . . . ... . 396
7. Oxygen Atom Transfer from Coordinated Nitrite Ligands 396
7.1. Transfer from Metal Nitrites to Carbon Monoxide and
Triphenylphosphine . . . . . . . . 396
7.2. Transfer from Metal Nitrites to Alkenes 398
References 399

Chapter 13
Catalysis of Nitrogen -Fixing Model Studies
T. Adrian George

1. Introduction . . 405
1.1. Scope and Limitations 406
2. Nitrogen-Fixing Reactions 408
2.1. Titanium, Zirconium, and Hafnium 409
2.2. Vanadium, Niobium, and Tantalum 409
2.3. Chromium, Molybdenum, and Tungsten 413
2.4. Manganese, Technetium, and Rhenium 429
2.5. Iron, Ruthenium, and Osmium 429
2.6. Cobalt, Rhodium, and Iridium . 430
2.7. Nickel, Palladium, and Platinum 430
2.8. Copper . . . . . . . . . . 430
3. Nitriding Reactions . . . . . . . 431
4. Dinitrogen Binding and Reactivity 431
5. Future Prospects 434
6. Concluding Remarks 436
References .... 436

Chapter 14
Polymer-Bound Phosphine Catalysts
Norman L. Holy

1. Introduction and Scope . . . . . . . . . . . . 443


2. Preparation of Polymer-Bound Phosphine Catalysts 444
2.1. Modification of Preformed Polymers 444
2.2. Polymerization of Phosphine Monomers 448
3. Physiochemical Characterization of Catalysts 450
xvi CONTENTS

4. Influence of the Support 454


4.1. Changes in Selectivity 455
4.2. Asymmetric Induction 457
4.3. Matrix Isolation 457
4.4. Coordinative Unsaturation 459
5. Reactions of Ole fins and Dienes 460
5.1. Isomerization 460
5.2. Hydrogenation 461
5.3. Dimerization, Oligomerization, and Polymerization 466
5.4. Addition 470
5.5. Reactions with CO 472
5.6. Metathesis 476
6. Trends 478
6.1. Catalyst Characterization 478
6.2. Reactions 479
References 480

Index 485
1
Historical Introduction
Joseph Chatt

The phosphine metal complexes now finding application in homogeneous


catalysis are those of the tertiary organic phosphines. They were discovered
around the middle of the last century and their ability to combine with
heavy metal salts noted almost immediately, but the application of their
metal complexes to homogeneous catalysis came only after the lapse of
about 100 years.

1. THE PHOSPHINES

Trimethylphosphine and poly(dimethylphosphane) [P(CH 3 hJn were


the first tertiary phosphines to be discovered. They were discovered in 1847
by Paul Thenard,(l) even before the aliphatic amines. He obtained them
by passing methyl chloride over an impure calcium phosphide at 180-300°C.
Triethylphosphine was prepared similarly from ethyl iodide and an impure
sodium phosphide by Ferdinand Berle in 1855.(2)
These highly reactive substances did not receive the attention they
deserved until after the discovery of the aliphatic amines and their relation
to phosphine excited the interest of Augustus William Hofmann. It was to
him and his co-workers, especially Augustus Cahours (over the period
1857-1871) that we owe the early development of organic phosphorus
chemistry. (3-5) Hofmann was a remarkable man who came from Justus von
Leibig's laboratory at Giesen. He was remarkable not only for his
meticulously accurate researches, as attested by Leibig himself in a footnote

Dr. Joseph Chatt • School of Chemistry and Molecular Sciences, University of Sussex,
Brighton BNl 90J, United Kingdom.
2 JOSEPH CHA IT

to one of his early papers on indigo, but he also contributed substantially


to the establishment of chemistry as a proper subject for academic study.
Coming to London in 1845, he successfully launched, as its Director, the
newly founded Royal College of Chemistry now part of the Imperial College
of Science and Technology in the University of London. He was the fifth
President of the Chemical Society of London, and on his return to Berlin
in 1864 he played an important role in the founding of the "Deutsche
Chemische Gesellschaft zu Berlin" of which he was elected its first Presi-
dent. The difficulties he overcame in preparing the methylphosphines are
well illustrated by his account of the reaction of methyl iodide on sodium
phosphide, his first improved method. "On the application of heat these
substances act on one another with great energy, producing combustible
and detonating compounds, so that the experiment is not without danger.
Often the product of the operation is lost; and if the reaction takes place
without explosion the separation of the constituents of the very complicated
mixture which results can be effected only with the greatest difficuIty.,,(4.5)
He developed more controllable methods, as for example, the reaction of
diethylzinc in ether with phosphorus trichloride and wrote of this: "The
reaction between these two bodies is very violent and readily gives rise to
dangerous explosions if the necessary precautions are neglected." His
precautions were to keep the reaction vessel full of carbon dioxide and
well cooled. Other safer methods, usually starting from phosphine or
phosphonium iodide, were developed later, but none was particularly
convenient. (6)
It says much for Hofmann's and Cahours's skill that they were able
to purify and characterize such highly reactive materials. Certainly to my
knowledge the preparation of trimethyl- and triethylphosphine defeated
some inorganic and organic chemists even in the 1950s. It was this difficulty
of access to the phosphines which delayed the extensive study of tertiary
phosphine complexes for a whole century.
The early development of aromatic phosphine chemistry took place
during the last quarter of the nineteenth century mainly in Michaelis's
school,(7.S) He prepared dichloro(phenyI)phosphine by the passage of the
mixed vapours of benzene and phosphorus trichloride through a red hot
porcelain tube. (7) This was quickly followed by the preparation of
phenylphosphine, dialkylphenylphosphines, and triphenylphosphine. (S-1O)
This latter phosphine was destined some 75 years later to become one of
the most important phosphines in the development of homogeneous cataly-
sis by phosphine metal complexes. Michaelis obtained it by the reaction
of sodium with dichloro(phenyI)phosphine or phosphorus trichloride and
bromo benzene in boiling ether. (S-1O) The development of aromatic phos-
phorus chemistry rapidly overtook that of their intractable aliphatic analogs.
It was well established by the 1920s whereas the aliphatic had to await the
HISTORICAL INTRODUCTION 3

application of Grignard reagents and their reaction with dichloro(phenyl)-


phosphine and phosphorus trichloride, first in the aromatic series(ll) and
then by Hibbert to prepare triethylphosphine in 1906.(12) Hibbert's method
of work-up gave poor yields and was difficult, but it was the best method
available even in the 1930s when the present interest in phosphine complex
chemistry started to stir. The difficulty with triethylphosphine was possibly
the reason that the more easily isolated tripropylphosphine and its higher
homologues were not prepared before 1929.0 3 ,14)

1.1. The Complexes

There are references to metal complexes in the early Hofmann


literature but most of these are to organophosphonium halometallates,
e.g., [PH(CH 3hJ ZnI4' which crystallize from halogen acid solutions on the
addition of the appropriate metal saltY) Nevertheless true coordination
compounds of platinum(II) were isolated almost immediately after the
discovery of the tertiary aliphatic phosphines. (15) This doubtless arose
because Hofmann used the then current technique of amine chemistry to
identify his phosphines through determination of the platinum content of
their sparingly soluble and easily purified hexachloroplatinates(VI). In fact,
he obtained white [PtCh(PMe3hJ and triethylphosphine complexes, but
serious investigation started only in the 1870s(16) when Cahours and Gal(16)
showed that on boiling an aqueous solution of platinum(IV) chloride with
triethylphosphine they obtained the yellow ex -form of [PtCh(PEt 3 hJ and
the white (3- form. They also prepared the salt [Pt(PEt 3 )4JCh. Complexes
of gold(I), copper(I), and palladium(II) were prepared by these and other
workers about the same time, but they aroused little interest beyond the
nature of the isomeric platinum(II) complexes. With the advent of Alfred
Werner and the establishment of the planar configurations of the
platinum(II) diammines, [Pt(NH 3 hCh], the nature of the isomerism was
ellucidated and, by color analogy with the diammines, the yellow ex -isomer
was assigned the cis-configuration and the white (3-isomer, the trans.o7)
The few known analogous platinum(II) complexes, e.g., those of the tertiary
arsines and organic sulphides, had their configurations similarly and incor-
rectly assigned. There the matter rested until the 1930s. In the meantime
other complexes had been prepared, particularly the "mercurichlorides"
[{HgCh(PR 3 )hJ which served as readily purified crystalline derivatives for
identification by melting point of the liquid phosphines.
The present phase in the development of the complex chemistry of
the tertiary organic phosphines started in the early 1930s and was interrup-
ted by the second World War. It is intimately connected with that of the
corresponding arsines and sulphides and less so with the stibines, selenides,
and tellurides-all being easier to prepare than the phosphines.
4 JOSEPH CHA IT

Early in this century, the organic chemistry of arsenic and sulfur


underwent rapid development owing mainly to the discovery that
organoarsenicals have valuable chemotherapeutic properties and, because
of the 1914-1918 war, that some arsenicals and sulphides have sternutatory
or vesciant properties of possible use in warfare. More chemists thus
obtained knowledge of these potentially ligand substances which in the
fullness of time would be used to render transition metal ions soluble in
organic solvents. Naturally the complex chemistry of the more readily
obtainable organic sulphides and organoarsines was developed first, and
the scene was set for new development in the area of organophosphine
complexes by 1930.
It was the application to coordination chemistry of the physicochemical
methods finding their way into organic chemistry during the 1920s and
1930s which raised the need for organic soluble or easily fusible coordina-
tion compounds of high stability. Two notable studies then revived interest
in trialkylphosphine and related complexes, which was to grow in strength
rapidly after the second World War. These studies were attempts by F. G.
Mann(1S) to apply Sugden's parachor(19) to determine whether the coordi-
nate bond, also known as a semipolar double bond, is single or double,
and K. A. Jensen's determination of the dipole moments of complex
compounds. (20)
Mann's study required low-melting compounds for the determination
of their surface tensions in the liquid phase and he chose organo-phosphines,
arsines and -sulphides as ligands with palladium(II) and mercury(II) halides
for his study. He soon showed that the parachor was useless for his
purpose, but discovered that the trialkylarsine (AsR 3 ) complexes, e.g.,
[PdCb(AsR 3 h], lost some of their arsine on fusion in vacuo to give a new
series of halogen bridged complexes, e.g., [Pd 2Cl 4 (AsR 3 h]. Then even the
existence of halogen bridges was contentious. This was resolved by an
X-ray structure paralleled by an extensive study of the chemistry of organo-
phosphine and -arsine complexes in the late 1930s; it increased enormously
the number of known complexes of the tertiary phosphines and arsines. (21,22)
K. A. Jensen's dipole moment study of complex compounds required
complexes soluble in nonpolar solvents, and he also chose organo-sulphide,
-arsine and -phosphine complexes. (20) He established that a high separation
of charge occurs in asymmetric coordination compounds, such as cis-
[PtCb(PR 3 h] (PR 3 = trialkylphosphine) with dipole moments around 11
Debye units, and he showed that all those platinum(II) complexes except
the ammines, hitherto thought to have a trans- configuration, had a cis one
and visa versa. He also demonstrated that tertiary phosphines would stabil-
ize unusual oxidation states by his preparation of [NiBr3{P(C2H5hh]. (23)
The first event of direct relevance to the application of phosphine
complexes to catalysis occurred immediately after the second World War.
HISTORICAL INTRODUCTION 5

This was the appearance of publications by W. Reppe and co-workers in


1948(24.25) followed by Badische Anilin und Soda Fabrik patents.(26) They
showed that various triphenylphosphine complexes of nickel, especially
[Ni(COh(PPhsh](Ph = C 6 H 5), were more effective than other nickel com-
plex catalysts for the polymerization of olefinic and acetylenic substances
and that others, especially [NiBr2(PPh3h], catalyzed the formation of acrylic
acid esters from alcohols (ROH), acetylene, and carbon monoxide:

ROH + co + C2 H 2 --+ CH 2 =CHCOOR

The mechanisms of these catalytic reactions were obscure, but the


demonstration that phosphine complexes had potentially useful catalytic
properties attracted the attention of the petrochemicals industry world-
wide. Thus, the postwar studies of organic soluble transition metal com-
plexes, especially carbonyl, olefin, acetylene, phosphine, and arsine
complexes, occurred mainly in research laboratories financed by industry
or governments and tended to concentrate on the first three of the above
ligands in reaction with molecular hydrogen as showing greater industrial
potential. Phosphines and arsines were involved only incidentally. Despite
Reppe's discovery, the immediate postwar studies of phosphine and arsine
complexes were directly derived from prewar studies and aimed at coordina-
tion chemistry for its own sake. This was true of my work in Imperial
Chemical industries Limited, U.K., and in the study of arsine complexes
by F. P. Dwyer in the University of New South Wales during the 1940s.(27)
This latter work led directly to the establishment of an important school
of arsine complex chemistry in University College, London, under one of
Dwyers's students, R. S. Nyholm. (28) He had an enormous influence on the
development of coordination chemistry, especially of phosphine and arsine
complexes over the two decades from 1950. It was not so much through
his research, important though it was, but more by his infectious enthusiasm,
his ability to produce excellent "popular" reviews, and his interaction with
chemists of all disciplines.
It was during this period from the second World War to around 1965
that the unique character of the phosphines and arsines as ligands was
established. It was also during this period that the fundamentally important
qualities of hydride and ligands attached through carbon to transition metals
were discovered. Here it is not possible to do more than summarize these
developments, t which are elaborated more fully where appropriate in other
chapters.
There was first the ability of phosphine and arsine ligands to stabilize
unusually high oxidation states of the later transition metals, (23) much

t For a bibliography relevant to phosphine chemistry over this period, see Reference 29.
6 JOSEPH CHA IT

exploited by Nyholm.(28) This stabilization is a consequence of their high


ligand field strengths, a factor which renders them compatible with ligands
such as carbon monoxide, hydride ion, saturated organic substances, and
alkyl and aryl groups, all important in homogeneous catalysis. Thus, it was
found that phosphines substituted readily into the binary metal carbonyls
and hydridocarbonyls, that they stabilized as coligands, hydride, alkyl, and
aryl complexes of transition metals, especially of Group VIII. Moreover
phosphines, especially triphenylphosphine, also stabilized low-oxidation
state complexes, e.g., [RhCI(PPh 3 hJ and Pt(PPh 3 )n (n = 3 or 4), a very
important property in preventing the precipitation of the heavier Group
VIII metals under the reducing conditions inherent in such important
reactions as hydrocarbonylation and hydrogenation.
It was also shown that phosphines and arsines as ligands have fairly
high trans -effects, so that although one, two, or rarely three monophosphine
ligands might coordinate strongly to a metal ion or atom, three or more
usually gave easily dissociable complexes thus freeing metal sites for the
activation of reactant molecules. The inability of the metal to hold strongly
a greater number of phosphine or arsine molecules is caused not only by
their high trans-effects but also by their steric bulk. More recently in the
late 1960s it was found that phosphines tend to stabilize dinitrogen com-
plexes which are prepared from molecular nitrogen, often with hydride as
coligand. This has led to attempts to produce phosphine complexes,
especially from molybdenum or tungsten, to mediate catalytically the pro-
duction of nitrogen hydrides or amines directly from molecular nitrogen.
Truly catalytic systems have not yet been discovered but stoichiometric
and cyclic systems with ammonia, hydrazine, or amines as products have
been evolved. (30)
Other qualities of ligands pertinent to the important reactions of
homogeneous catalysis by tertiary phospine complexes were established
also during the period to 1965. These were:
(a) the exceptionally high trans-effects of hydride, alkyl, and aryl
ligands; their high ligand field strengths; and their strong CT-elec-
tron donor abilities as notional anionic ligands, a consequence of
the low electronegativities of hydrogen and carbon as compared
with the more common ligand atoms;
(b) the oxidative addition and related reactions of alkyl and hydrogen
halides; of aromatic and aliphatic hydrocarbon groups; and of the
utmost importance, of dihydrogen by the insertion of electron-rich
metal centers into C-X, H-X, C-H, and H-H bonds, respectively,
e.g.:
HISTORICAL INTRODUCTION 7

(c) the insertion of carbon monoxide into M-C bonds, where C is


the carbon of an alkyl or aryl ligand, to produce acyl or aroyl
complexes;
(d) the insertion of olefins into M-H bonds to produce metal alkyls;
(e) the insertion of ole fins into metal alkyl bonds to produce higher
alkyls and similar insertion of alkynes leading to polymerizations
of various kinds; and
(f) the removal of notionally anionic carbon ligands from the metal
as an uncharged species by addition of hydrogen across the M-C
bond: H2 + M-C --. MH + HC, or by its elimination as an olefin:
M-CHz-CH2R --. M-H + CH 2=CHR.
These reactions provided the basis for the mechanistic understanding
of such industrially important reactions as the hydrocarbonylation of olefins
and the then recently discovered Ziegler-Natta polymerization. They held
out promise of more useful catalyses and led to a flourishing activity in the
area of organotransition metal chemistry world-wide during the 1960s and
1970s.
Another important development during the period to 1965 was the
appearance of certain "key" phosphines as commercial products, especially
triphenylphosphine, tri -n -butylphosphine, dichloro(phenyl)phosphine, and
chlorodi(phenyl)phosphine, the latter two being percursors for the easily
isolated dialkyl(phenyl)- and alkyldi(phenyl)-phosphines. Also during this
time, the development for routine use of new physical methods considerably
eased the investigation of phosphine complexes, first infrared spectroscopy
and its later development to longer wavelengths where metal-halogen
stretching frequencies occur, then the more useful IH nmr, and later 31 p
nmr spectroscopies. These latter particularly allowed quick determinations
of the configurations of phosphine complexes, replacing the time-consuming
measurement of dipole moments, which had hitherto been the only reliable
method, apart from the even more laborious X-ray crystal-structure
determination. Incidently, the dipole moments gave much valuable infor-
mation about the distribution of electric charge in complex compounds. (31)
For the study of the fundamental chemistry of transition metal com-
plexes of phosphines and arsines and their organometallic derivatives, it
was necessary to have stable complexes of known structures and configu-
rations; hence, the studies up to 1945 involved mainly the strongly bonding
trialkylphosphines, and arsine complexes stabilized by chelation as in the
famous 0 -phenylenebis( dimethyl arsine ). For catalysis, it is necessary to
create vacant sites in the coordination shell of the metal where the reactants
can be activated or brought together by coordination for reaction.
Triphenylphosphine and other aromatic phosphines with their greater
steric bulk and more weakly bonding affinity for the metals were ideal
8 JOSEPH CHA IT

for the launching of phosphine complex chemistry into its next phase of
development, that of homogeneous catalysis.

1.2. Catalvsis

The study of homogeneous catalysis by phosphine complexes started


to flourish in the 1950s. It is so recent that here it is necessary only to
indicate the sequence and time-scale of its development. t It was spurred
by K. Ziegler's discovery that the products of the reaction of triethyl-
aluminium with certain complexes of zirconium or titanium were excellent
catalysts for the polymerization of ethene at ordinary temperatures and
pressures. (33) Hitherto pressures of over 1000 atm had been used in the
manufacture of polyethylene. Furthermore, G. Natta immediately extended
Ziegler's process using triethylaluminium and titanium halides to produce
isotactic, i.e., stereoregular, polypropylene, and other stereoregular poly-
mers. (34) These discoveries quickly led to a new and very successful process
for polyethylene and to a new commercially important product in isotactic
polypropylene. It also led to a flurry of research activity into further uses
of transition metal compounds as catalysts, many of which were rendered
homogeneous by the use of triphenylphosphine as a nonreacting ligand to
hold the metal in solution under reducing conditions.
The metals involved were mainly those of Group VIII, and the various
reactions mediated homogeneously by phosphine complexes form the
detailed subjects of the later chapters in this book. The most significant
have involved alkenes and, to a lesser extent, alkynes under reducing
conditions. Of these, hydrogenation has received most study and will be
used to illustrate the general growth of awareness of the unique advantages
of phosphine complexes in catalysis. Other important reactions mediated
advantageously by phosphines as ligands are carbonylation in its various
aspects including hydroformylation, the oligomerization of alkenes, and
their isomerization by double-bond migration.
Despite M. Iguchi's discovery in 1939(35) that certain unstable rhodium
complexes catalyzed the hydrogenation of organic substances, such as
fumaric acid, and Reppe's discovery in the mid 1940s that phosphines
enhanced the activity of some of his nickel-based polymerization and
carbonylation catalysts,(24-26) it was not until 1965 that the next major
breakthrough occurred. This was the discovery of a phosphine-activated
homogeneous hydrogenation catalyst, [RhCI(PPh 3h], known as Wilkinson's
catalyst, which catalyzes the hydrogenation of alkenes at 25°C under one
atmosphere pressure of dihydrogen.(36,37) Its remarkable activity immedi-
ately directed attention to other similar complexes including

t For a general bibliography leading into this phase of development see Reference 32.
HISTORICAL INTRODUCTION 9

[RuHCI(PPh 3hJ, formed by dihydrogen reduction of [RuCh(PPh 3)xJ


(x = 3 or 4), and one of the most active homogeneous catalysts for the
hydrogenation of 1-alkenes. (38)
A very important factor to emerge from the study of these
homogeneous systems was their selectivity as compared with heterogeneous
systems. By the correct choice of conditions, they could be operated to
hydrogenate alkynes in the presence of alkenes, or one alkyne selectively
in the presence of another. The steric bulk of the phosphines usually means
that 1-alkenes are hydrogenated some orders of magnitude faster than
internal alkenes. Dienes can be reduced selectively to monoenes, as for
example by trans- [Pt(SnCI 3)H(PPh 3hJ. (39) The complex [RuCh(PPh3hJ will
even catalyze the reduction by hydrogen of nitro-compounds to amines
without hydrogenating the aromatic ring or reducing other substituents,
including CN.(40)
The selectivity of phosphine complexes in catalysis was further empha-
sized in 1968(41,42) when it was found that the hydrogenation of atropic
acid, CH 2=C(Ph)COOH, catalyzed by [RhCb(P*MePhPrnhJ using (S)-( +)-
methylphenyl-n -propylphosphine, gave asymmetric hydrogenation to
MeC*H(Ph)COOH with a 15% enantiomeric excess. Shortly afterwards,
it was shown that the asymmetric center in the catalyst need not be at
phosphorus but could be advantageously situated in one of the organic
groups of the phosphine. (43) Thus, a rhodium catalyst prepared from
neomenthyldiphenylphosphine gave an enantiomeric excess of 61 % of
(s)-( +)-3-phenylbutanoic acid on hydrogenation of (3 -methylcinnamic acid.
Such asymmetric phosphines have the advantage that they can be obtained
from naturally occurring, optically-active precursors. The best known of
such phosphines is the bidentate ligand (-)-2,3-0-isopropylidene-2,3-
dihydroxy-1,4-bis(diphenyl-phosphino)butane, usually referred to as (-)-
mop and prepared from L( + )-tartaric acid. (44) A rhodium catalyst prepared
from (-)-mop catalyzes the hydrogenation of PhCH=C(NHCOOMe)-
COOH to give (R )-PhCH 2C*H(NHCOOMe)COOH in 72% optical purity,
and recently an enantioselectivity of 90% has been obtained in the reduction
of an appropriate ketone to its alcohol using another chelate diphosphine
rhodium catalyst.(45) Such asymmetrical syntheses are of obvious import-
ance to the pharmaceutical industry.
The great strength of homogeneous catalysis and especially of that
involving phosphine ligands is the opportunity it affords to tailor ligands
so as to enhance the reactivity and selectivity of metal centered catalysts.
Only slight changes in the ligand can effect considerable changes in selec-
tivity. That the reaction occurs in homogeneous solutions also facilitates
the study of the mechanistic steps involved in the catalytic cycle and, by
change of phosphine or metal center, may allow the isolation of stable
substances analogous to metal-containing intermediates of the catalytic
10 JOSEPH CHA IT

cycle. The study of homogeneous catalysis has thus been very useful in
elucidating the important reactions of heterogeneous catalysis, but the
large-scale reactions of the petrochemicals industry are still catalyzed
heterogeneously. Homogeneous catalysis has the great disadvantage that
the catalyst remains in the product and must usually be separated from it.
Ideally, one would prefer a heterogeneous catalyst with the reactivity and
selectivity of the homogeneous. The present thrust in homogeneous
catalysts is not only to increase their reactivity and by tailoring the ligands
to attempt to promote selectivity in the desired reaction, but also to render
them heterogeneous for convenience in use by attaching them to a surface,
hopefully without loss of reactivity or selectivity. (46)

REFERENCES

1. P. Thenard, lahresber. 1847-1848,645-646; Compt. rend. 25, 892-895 (1847).


2. F. B6rle, lahresber. 1855, 590-591; 1. Prakt. Chem. 66, 73-75 (1855); Annalen 97,
334-355 (1856).
3. A. Cahours and A. W. Hofmann, Ann. 104, 1-39 (1957).
4. A. W. Hofmann and A. Cahours, Quart. 1. Chem. Soc. 11, 56-78 (1859).
5. A. W. Hofmann, Phil. Trans. 150,409-448 (1860).
6. - - , Ber. 4, 205-209 (1871).
7. A. Michaelis, Ber. 6, 601-603 (1873); Annalen 181,265-363 (1876).
8. A. Michaelis and L. G1eichman, Ber. 15, 801-804 (1882).
9. A. Michaelis and A. Reese, Ber. 15, 1610 (1882).
10. A. Michaelis and H. V. Soden, Annalen. 229, 295-340 (1885).
11. P. Pfeiffer, Ber. 37, 4620-4623 (1904).
12. H. Hibbest, Ber. 39,160-162 (1906).
13. W. C. Davies and W. J. Jones, 1. Chem. Soc. 1929,33-35.
14. W. C. Davies, P. L. Pearce, and W. J. Jones, 1. Chem. Soc. 1929, 1262-1268.
15. A. W. Hofmann, Annalen 103,357-358 (1857).
16. A. Cahours and H. Gal, Compt. rend. 70, 897-903 (1870); lahresber. 1870, 808-814.
17. A. Werner, Z. Anorg. Chem. 3,265-330 (1893), p. 318.
18. F. G. Mann and D. Purdie, 1. Chem. Soc. 1935,1549-1563.
19. S. Sugden, 1. Chem. Soc. 1924, 1177-1189.
20. K. A. Jensen, Z. Anorg. AUg. Chem. 229,225-251 (1936).
21. J. Chatt and F. G. Mann, 1. Chem. Soc. 1938, 1622-1634 and references therein.
22. F. G. Mann and D. Purdie, 1. Chem. Soc. 1940, 1235-1239 and earlier Parts in the series.
23. K. A. Jensen, Z. Anorg. Allg. Chem. 229.265-281 (1936).
24. See for example, W. Reppe and W. J. Schweckendiek, Annalen 560, 104-116 (1948).
25. W. Reppe cited in J. W. Copenhaver and M. H. Bigelow, Acetylene and Carbon Monoxide
Chemistry (Reinhold, New York, 1949).
26. See for example, W. Reppe and W. Schweckendiek, German Patent 871, 494 (1953);
Chem. Abstr. 52, 10175 (1958).
27. See for example, F. P. Dwyer and D. M. Stewart, 1. Proc. Roy. Soc. N.S. Wales 83, 177
(1949); [Chem. Abstr. 45,7462 (1951)] and references therein.
HISTORICAL INTRODUCTION 11

28. D. P. Craig, "Ronald Sydney Nyholm," Biographical Memories of the Royal Society 18,
445-475 (1972).
29. G. Booth, Adv. Inorg. Chem. Radiochem. 6, 1-69 (1964).
30. J. Chatt, J. R. Dilworth, and R. L. Richards, Chem. Rev. 78, 589-625 (1978).
31. J. Chatt and G. J. Leigh, Angew. Chern. Int. Ed. Engl. 17,400-407 (1978).
32. C. Masters, Homogeneous Transition-Metal Catalysis (Chapman and Hall, London and
New York, 1981).
33. K. Ziegler, E. Holzkamp, H. Breil, and H. Martin, Angew. Chem. 67, 541-547 (1955).
34. G. Natta, I. Polymer Sci. 16, 143-154 (1955).
35. M. Iguchi, I. Chem. Soc. (lpn.) 60, 1787-1792 (1939).
36. J. F. Young, J. A. Osborn, F. H. Jardine, and G. Wilkinson, I. Chem. Soc., Chem. Comm.
1965, 131-132.
37. R. S. Coffey and Imperial Chemical Industries, British Patent no. 1 121 642 (1965).
38. D. Evans, J. A. Osborn, F. H. Jardine, and G. Wilkinson, Nature (London) 208,
1203-1204 (1965).
39. E. N. Frankel, E. A. Emken, H. Itatani, and J. C. Bailar, Jr., I. Org. Chern. 32,1447-1452
(1967).
40. J. F. Knifton, I. Org. Chem. 40, 519-521 (1975); 41,1200-1206 (1976).
41. W. S. Knowles and M. J. Sabacky, I. Chern. Soc., Chern. Commun. 1968,1445-1446.
42. Compare L. Horner, H. Siegel, and H. Blithe, Angew. Chem. Int. Ed. Engl. 7,942 (1968).
43. J. D. Morrison, R. E. Burnett, A. M. Aguiar, C. J. Morrow, and C. Phillips, I. Am.
Chem. Soc. 93, 1301-1303 (1971).
44. H. B. Kagan and T. P. Dang, I. Am. Chem. Soc. 94, 6429-6433 (1972).
45. T. Hayashi, A. Katsumura, M. Konishi, and M. Kumada, Tetrahedron. Lett. 1979,
425-428.
46. See for example, F. R. Hartley and P. N. Vezey, Adv. Organomet. Chem. 15, 189-234
(1977).
2
Mechanistic Studies of
Catalytic Reactions Using
Spectroscopic and Kinetic
Techniques
C. A. Tolman and J. W Faller

1. INTRODUCTION

The rapid growth of the organometallic chemistry of the transition metals


during the last 15-20 years owes much to the development of homogeneous
catalyst systems which are capable of synthesizing organic molecules under
mild conditions and occasionally with remarkable selectivities. Several have
been commercialized and are now used on a large scaleY) A few have
received considerable detailed study-including spectroscopic identification
of the species present in solution under reaction conditions, isolation of
reactive intermediates in some cases, determination of the overall rate law
and measurement of rate and equilibrium constants of several individual
steps, and isotopic labeling studies-so that we have a reasonably clear
picture of how they operate. It is these systems that form the focus of this
chapter. For more general reviews the reader is referred to some recent

Dr. Chad A. Tolman • Central Research and Development Department, Experimental


Station, E. I. du Pont de Nemours and Company, Wilmington, Delaware 19898. Dr. Jack
W. Faller. Department of Chemistry, Yale University, New Haven, Connecticut 06511.

13
14 c. A. TOLMAN AND J. W. FALLER

books, (1-3) which also discuss the special electronic properties of transition
metals which are in part responsible for their catalytic behavior. t

1.1. The Nature of Catalvst Svstems

Before discussing specific examples, we should define what we mean


by the word "catalyst," and point out some of the important general features
of homogeneous catalytic systems. By "catalyst" we mean the transition
metal compound added to a chemical reaction to accelerate its rate or to
change the distribution of products formed.t Generally, many product
molecules are produced per metal atom added. We shall see that catalyst
systems consist of many different metal-containing species and that,
frequently, the compound originally added may change before a substrate
first becomes coordinated; the original compound may not even be recover-
able at the end of the reaction. This fact is recognized by some authors by
introducing terms like "the catalyst precursor," "the real catalyst," or "the
actual catalyst." For example, Collman and Hegedus(2a) define "the actual
catalyst" as "the dominant complex present during the catalytic cycle".
We shall see that the dominant complex present in an active catal1st system
may depend on the substrates used, their concentrations, the concentration
of phosphine ligands, and even the time elapsed after a reaction is initiated.
The dominant complex or the "principal species" present may not even
be involved in the catalytic cycle. If during catalysis we can spectroscopically
observe a complex that is part of the catalytic cycle (one of what we shall
term a "loop species"), it often occurs in observable concentrations because
it precedes the slowest or rate-determining step and is the least active
species in the cycle! This loop species in highest concentration might be
identified as the "principal active species" or the "major active species"
and may in reality be the least active. These difficulties in nomenclature
can produce problems, such as requiring references to "Wilkinson's catalyst
precursor" rather than "Wilkinson's catalyst." The term "catalyst" will
inevitably mean different things to different people and this vagueness leads
us to prefer the simple original empirical definition of "catalyst" given
above.
The catalytic system may also require a cocatalyst. By "cocatalyst"
we mean a compound added to the system to improve the effectiveness of
the catalyst. These are usually nontransition-metal compounds and would

t Masters' book(3) has a particularly nice discussion of the important features of homogeneous
metal catalysts.
is possible to distinguish reactions which are catalytic in the metal from stoichiometric
:j: It
ones if several product molecules are produced per metal atom charged. Some stoichiometric
reactions can be made catalytic, e.g., by reoxidizing the Pd(O) produced in the Wacker
process O .21 by Cu(II), which is in turn regenerated by oxidation with air.
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 15

I
I
t
@
Figure 2.1. A schematic representation of a mechanism of a typical homogeneous catalytic
system. Sl and S2 are substrates, and P product. L represents a phosphorus ligand.

include, for example, Bronsted or Lewis acids or bases, or reducing or


oxidizing agents.
Figure 2.1 shows a schematic representation of a typical homogeneous
catalyst system. Each numbered circle represents a discrete metal complex.
The heart of the system is the loop, consisting in this case of complexes
2-7, which are connected in a cyclic sequence by elementary reaction steps.
It is by repeated cycles around this loop that a single metal atom can
convert many substrate molecules Sl and S2 into product molecules P. In
addition to the loop species in Figure 2.1, there are other species in the
system off the loop (compounds 1 and 8) connected to it by reversible
reactions and another (9) connected by an irreversible reaction (indicated
by the dashed arrow) going from the loop. Note that any compound that
is in the loop or can be connected to a loop species could be charged into
the system as a catalyst. A compound like 9 is commonly referred to as a
"dead catalyst" or a "catalyst deactivation product." Reactions which
irreversibly bleed metal out of the loop are responsible for the finite life
of alI real catalyst systems. A useful measure of catalyst life is the number
of catalytic cycles before deactivation occurs (the turnover number), which
is given by the relative rate of cycling about the loop (product formation)
to catalyst deactivation. A useful measure of instantaneous catalyst activity
is the turnover rate, which is the number of product molecules produced,
per metal atom in the system, per unit time. Depending on the equilibrium
constants connecting complexes 1 through 8 to the loop, only a smaIl
fraction of the metal atoms may be present in the loop at anyone time,
as we shaII see in the subsequent examples.
16 C. A. TOLMAN AND J. W. FALLER

It should be noted that the solid arrows in the loop are drawn in one
direction to indicate the major flow of the reactions; actually, the individual
reaction steps can be, and often are, reversible, so that an equilibrium
mixture of reactants and products can be approached from either direction.
A catalyst system increases the rate of approach to equilibrium but, of
course, cannot affect the equilibrium distribution.
Figure 2.1 shows the formation of only one product. Most catalytic
systems produce more than one product. Each additional product requires
an additional loop in the mechanism, though generally the loops are joined
by reaction pathways or by common catalytic intermediates. As we shall
see, some of these systems can give complex networks of reactions.
In addition to substrates and products which may coordinate and
dissociate, there are usually other ligands, designated L in Figure 2.1, which
also enter or leave the loops but whose role is less apparent. The primary
interest of this chapter is in systems where L is a phosphorus donor ligand.
The behavior of L will generally depend on both its electronic and steric
character. For monodentate ligands, these can be defined in terms of an
electron accepting character v t (based on the A 1 carbonyl stretching
frequency v of Ni(COhL) and a ligand cone angle fJ.(4) Understanding how
to "fine tune" the ligand effects to make catalyst systems more active, more
selective, and longer lasting is a major challenge facing chemists in
homogeneous catalysis. The development of a rational approach to this
fine tuning depends on another major challenge-determining the mechan-
isms of homogeneous catalytic systems.

1.2. Mechanisms, Rate Laws, and Dominant Species


A mechanism is the "road map" of a chemical reaction. It reflects the
sequence of reaction steps and the composition of the intermediates.
Mechanisms are postulated based on spectroscopic studies, rate studies,
product analyses, and isotopic labeling. Certain types of reactions may be
unlikely based on energy considerations (e.g., the Woodward-Hoffman
Rules(5) or the 16 and 18 Electron Rule(6)). There is never enough informa-
tion to prove one mechanism to the exclusion of all others, nor is there
usually enough data to uniquely determine all the rate constants for the
elementary steps of a proposed mechanism. Some possible paths can be
shown to be unimportant. In a typical case, a few rate or equilibrium

t An alternative and sometimes more convenient electronic parameter X is often used, where
X = v(L) - v[P(t-Buh] where vEL] is the frequency in cm- I of the Al carbonyl stretching
mode of Ni(COhL in CH 2 Ch, from Reference 4. Since P(t-Buh gave the lowest frequency
(2056.1 cm -I) observed for a phosphorus ligand, X gives a convenient scale of positive
numbers with values ranging from 0 for P{t-Buh to about 50 cm- I for PF 3 .
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 17

constants can be determined, while others can only be set above or below
certain limits.
A mechanism may be represented by a list of chemical reactions (each
of which is unimolecular or bimolecular), or more graphically by a diagram,
like Figure 2.1, which more clearly shows the flow and cyclic nature of the
reactions.
A rate law describes the dependence of the rate of product formation
(or substrate disappearance) at steady state on the concentrations of
catalyst, cocatalysts, substrates, and ligands. A fact that is not generally
appreciated is that the functional form of the rate law often depends on
the range of experimental variables investigated, as will be shown sub-
sequently.
Several mechanisms may be (and usually are) consistent with a given
rate law, so that other experiments are necessary before we can have any
confidence in a detailed mechanism. In particular, parallel spectroscopic
experiments on the system-preferably on solutions under actual catalytic
conditions-are usually necessary before one interprets the rate law.(7) As
a simple example of a stoichiometric reaction, it is found that the rate of
oxidative addition of CH3I to (C 2 H 4 )Pt(PPh 3h is inhibited in an inverse
first-order manner by the addition of excess C 2 H 4 to the system. (8) Without
spectroscopic experiments one doesn't know whether the inhibition is the
result of association of excess C 2 H 4 to form an inactive (C 2 H 4 hPtL 2
complex, decreasing the concentration of reactive (C 2 H 4 )PtL 2 [reactions
(1) to (3)], or whether C2 H 4 suppresses formation of a reactive PtL 2
complex from an inactive (C 2 H 4 )PtL 2 ((4) and (5)).

(1)
(2)

RPtL 2 X(C 2 H 4 ) ~ RPtL 2 X + C2H 4 (3)


11K
(C 2 H 4 )PtL 2 <===2 PtL 2 + C2H4 (4)
RX + PtL 2 ~ RPtL 2 X (5)

The rate law in either case is:

-d[Pt(O)]/dt = k[Pt(O)][RX]/(l + K[C 2 H 4 ]) (6)

which reduces to:

-d[Pt(O)]/dt = k[Pt(O)][RX]/K[C 2 H 4 ] (7)


18 C. A. TOLMAN AND J. W. FALLER

provided that (C 2 H 4 hPtL 2 is the major species in the first case or


(C 2 H 4 )PtL 2 is the major one in the second (i.e., K[C 2 H 4 ]» 1).
Spectrophotomeric measurements and IH nmr spectra show that
(C 2 H 4 hPtL 2 does indeed form in solution in the presence of added C2 H 4
and that it provides the intermediate for the relatively rapid exchange of
free and coordinated ethylene which occurs in solution at room tem-
perature. (9)
In a stoichiometric reaction, the rate law tells the difference between
the composition of the major initial species in solution and the composition
of the transition state of the rate-determining step (abbreviated CTSRDS).
Thus, under conditions where (C 2 H 4 hPtL 2 is the major Pt(O) species,

(8)

We know nothing, of course, about the structure of the transition state,


only its composition. Owing to mass balance, this approach is valid even
if rapid equilibria are involved prior to the rate-determinining step;
however, there are situations where deceptively simple rate laws can lead
to difficulty. These possibilities will be discussed after this approach is
illustrated.
In a catalytic system, the rate law distinguishes between the composi-
tion of the dominant species present and the CTSRDSt. With 1-butene
isomerization with NiL4 [L = P(OEth], in the early stages of the reaction
and with small concentrations of added acids, the rate law is

-d[l-BJ/dt = k[NiJ[H+J[l-BJ/[LJ (9)

Since spectroscopic studies show that NiL4 is the major Ni-containing


species under these conditions, (10) the CTSRDS is obtained as in
Equation 10.

(10)

This does not, of course, tell us the order in which H+ and 1-B coordinate
and L dissociates, or whether the H+ is bound to Ni or carbon in the
transition state. Such detailed questions are best answered by examining
the rates and equilibria of individual elementary steps in the reaction, by
breaking the system up and studying its component parts. Thus, one must
address the problem of identifying intermediates on the loop by physical
or chemical methods, as well as suggesting reasonable compositions for all
the species present.
t If there is more than one active species over the range of variables explored, the rate law will
contain more than a simple multiplicative term.
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 19

Spectroscopic studies at low temperature can sometimes permit one


to trap and characterize reactive intermediates that might be unstable at
room temperature; often their structures can be deduced, especially by
nmr. In favorable cases even thermally sensitive complexes can be isolated
and X-ray crystal structures determined, at low temperatures if necessary.
More often, X-ray structures are determined on model systems which are
stable enough to be isolated and handled at room temperature. (See Chapter
2.) In this connection, model compounds containing third-row metals
are often considerably more stable than their congeners in the first or
second rows; however, their catalytic activity may also be substantially
lower.
It is sometimes said facetiously that if a species can be observed directly,
it probably is not a key intermediate in a catalytic reaction. More indirect
methods of inferring the composition and structure of intermediates will
be discussed subsequently (Section 1.3); however, it is appropriate to
consider some guidelines for postulating "reasonable" intermediates in the
catalytic cycle.
One rule of thumb that is often very useful in studying catalytic
mechanisms (or stoichiometric reactions, for that matter) is the 16 and 18
Electron Rule.(6) Simply stated, it says that diamagnetic transition-metal
complexes undergo reactions, including catalytic ones, which proceed by
elementary steps involving only intermediates with 16 or 18 metal valence
electrons. Other pathways are usually unfavorable because of the relatively
high energy of 14- or 20-electron species. Exceptions are known involving
both 14- and odd-electron species; however, 20-electron species are still
extremely rare, and their proposal should be regarded with suspicion unless
the evidence for them is compelling. It has proven possible to isolate
14-electron complexes with very bulky ligands, such as Pd[P(t- BuhJ2 (11)
and RhCI(PCY3h.(12) Furthermore, species having stable 16-electron
configurations, such as Rh I , IrI , PtlI , and Pd lI , are more likely to show
exceptions. There is good kinetic evidence for a RhClL 2 intermediate in
the reaction of H2 with RhCI(PPh 3h (13.14) and for a Bu 2PtPPh3 intermediate
in the decomposition of BU2Pt(PPh3h (15)-though electron counting is
difficult in cases where triphenyl phosphine could donate electrons from
an aromatic ring as it does in Rh(PPh 3h +. (16) Another possibility for
RhCI(PPh 3h. first suggested by Osborn(17) and more recently supported
by the work of deCroon and co-workers,(18) is that the apparent 14-electron
complex is stabilized in benzene solution as a 16-electron 'T/ 2- benzene
solvate. There is growing evidence that many organometallic reactions
actually proceed through paramagnetic odd-electron species,(19,20) and so
fall outside the 16- and 18-Electron Rule. Nevertheless, it still encompasses
a great deal of the chemistry of diamagnetic transition-metal complexes
and often provides a useful starting point for studying mechanisms.
20 r. A. TOLMAN AND J. W. FALLER

The approaches most familiar to organometallic chemists for the infer-


ence of mechanisms from kinetics studies have been discussed by
Pearson. (21-23) It also seems that organometallic chemists interested in
determining mechanisms of homogeneous catalytic reactions can profit a
great deal from examining the approaches by enzymologists and bio-
chemists. Lineweaver and Burke(24) showed that kinetic data can often be
put in convenient linear form for determining rate constants and rate laws
t.
by plotting rate -I against [Sr where [S] is the concentration of substrate
or reactant. King and Aitman l25J described a simplified procedure for
writing rate equations for a cyclic series of reactions-found in both enzyme
and organometallic catalytic systems. Cleland has published extensively on
enzyme kinetics and the relationships between rate laws and mechan-
isms. (26-28) Hopefully, organometallic chemists will begin to apply more of
these methods. Segel's book(29) is recommended for the interested reader.
Examination of the history of mechanisms inferred from gas phase
data emphasizes the need for caution in interpreting kinetics without the
use of ancillary confirmation from other methods. For example, the concept
of CTSRDS assumes a single well-defined rate-determining step. Since the
rate-determining step may vary with changes in concentrations, difficulties
may be encountered, particularly when radical chain processes are involved.
Thus, one may need to answer the question: "What is the mechanism of
the reaction under this set of conditions?;" rather than simply "What is
the mechanism of the reaction?" Furthermore, one assumes he has deter-
mined the rate law completely and may be bamboozled by deceptively
simple rate laws. For example, kinetics studies do not generally allow
detection of solvent involvement. Hence, a k in an equation such as (7)
may actually be

k = k'[solvent]" (11)

and the composition of the transition state in the rate-determining step


determined from the observed rate expression might either include or be
missing several molecules of solvent.
Kinetics studies of gas phase reactions of H2 and 12 dating from the
turn of the century provided the classic textbook example of a bimolecular
reaction until 1967. That "the reaction between hydrogen and iodine is
known to take place at bimolecular collisions involving a single molecule
of each kind,,(21) was a cornerstone for the study of reaction mechanisms
until Sullivan(30) showed that mechanisms involving radical chain processes
were consistent with additional data. Thus the concerted reaction mechan-
ism (12):

(12)
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 21

is no longer the cornerstone of the study of mechanisms that it once was.


This reaction, however, serves to illustrate the difficulty of identifying the
presence of radical chain pathways by conventional kinetics.
One might think that radical reactions would be generally accompanied
by fractional order in the rate law, but often surprisingly simple rate laws
can be found. The decomposition of ethane (13), for instance, gives a rate
law

(13)

that is simply first order in ethane. More detailed studies reviewed by


Laidler(31) show that complex radical paths are involved. Many "first-order"
gas phase processes, such as the isomerization of trans-l,2,-
dideutereoethylene, become second-order processes at low pressure. (32)
This follows from a rate law in which a term in concentration becomes
insignificant at low concentration. Thus, in these cases the CTSRDS
approach might be misleading.
Prior to 1979, the Wacker oxidation of ethylene to acetaldehyde was
generally suggested to proceed via a cis addition of Pd(II) and coordinated
hydroxide based upon kinetic evidence.(33) The original rate law is given
in equation (14):

d[product]/ dt = k [ethylene ][Pd]/[H+][ en (14)

The l/[H+] dependence was originally interpreted in terms of cis attack


(15). The stereochemical studies of Baeckvall, Akermark, and Ljunggren(33)

'. (I
. I
'. ( I
'J -Pd- ----. 'J -Pd- + W
. ~ I (15)
" OH 2 " "OH

showed that trans attack of water, (16), on coordinated ethylene was


consistent not only with the kinetic data, but also with the stereochemical
data.

(16)

Note that this latter mechanism requires nucleophilic attack on a coordin-


ated olefin which is more electrophilic owing to the charge on the complex.
Thus, the rate laws derived from mechanisms incorporating (15) or (16)
differ in the order of solvent, which illustrates the potential pitfalls and
22 C. A. TOLMAN AND J. W. FALLER

difficulties of dealing with solvent in rate expressions. Thus, as suggested


by Atkins,(34) a proof of mechanism in a chemical reaction is less like a
mathematical proof and more like a proof in a court of law. In general, the
greater the number of confirmatory experiments, the better. Nevertheless,
as in the courts, the reputation and rhetoric of the lawyer and the persuasive-
ness of the circumstantial evidence may sway opinions regarding the correct-
ness of a mechanism.

1.3. The Use of Isotopic Techniques

Following the caveats above with regard to possible solvent involve-


ment or unusual rate laws, one can make a reasonable determination of
the CTSRDS and provide reasonable estimates of intermediates in a
catalytic cycle using simple rules, such as the 16- and 18-Electron Rule.
Thus, one can create a diagram, such as Figure 2.1, with reasonable species,
which is consistent with the observed rate law. Unfortunately, this provides
a model, not a proof, of the events involved in the catalysis. The case can
be made stronger by testing the mechanism, or individual steps in it, by
spectroscopic or other more indirect methods.

1.3.1. Isotopic Labeling

Isotopic labeling can often be very helpful and sometimes indispensable


in elucidating the fine details of mechanisms, both from tracing the location
of a particular atom or group throughout the course of the series of steps
from reactant to product, and by utilizing kinetic isotope effects. As the
principles are straightforward and well-known, they will not be detailed
here; however, some recent developments will be reviewed.
The lowering of carbonyl stretching frequencies by the introduction
of 13 CO has provided innumerable systems in which the course of carbon
monoxide substitution, insertion, and migration could be readily followed
by infrared. The advent of Fourier Transform nmr and high field nmr
spectrometers has provided straightforward analysis not only of l3C label-
ing, but of 2 H , and 170, as well as direct detection of many metal nuclei.(35-37)
The use of 170 nmr appears to provide a particularly useful technique for
the study of carbonyls in metal systems with I > t nuclei, such as cobalt,
for which l3C resonances are often broad or unobservable. (38)
The large number of catalytic applications of rhodium suggests that
use of l03Rh nmr spectra should be potentially interesting. Unfortunately,
the sensitivity for this nucleus is quite low and relaxation times are quite
long, which implies that high concentrations and long collection times would
be rquired for the observations. Until recently l03Rh chemical shifts were
only obtained indirectly through spin-decoupling techniques
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 23

(INDOR);(37,39) however, direct detection is practical with modern


instrumentation.(40,41) The concentrations and time required for these
studies make extensive use of l03Rh nmr by conventional Fourier Transform
methods relatively unattractive for catalytic study applications. The
development of a polarization transfer technique (as opposed to an Over-
hauser enhan.cement) known as INEPT,<42) however, promises to allow
detection under reasonable concentration and time conditions. (43) Perhaps
one of the more interesting developments has been the development of
sample systems that allow the determination of nmr spectra under high
pressure, such as 1000 atm of CO IH 2.(44) Even without exceptional equip-
ment, it is practical to collect spectra in heavy-walled tubes (-10 atm), as
in the 31 p nmr studies of rhodium-phosphine hydroformylation
systems. (45,46)
GC-mass spectroscopy also provides a straightforward method of
analyzing isotopic distributions in product mixtures; however, nmr generally
is more useful for evaluating the stereochemistry of isotopic substitution.
This ease of assigning position by nmr presumably accounts for the lack
of general use of radioactive labels by organometallic chemists; however,
this technique is certainly deserving of more attention than it currently
receives. On occasion, microwave spectroscopy can provide an extra-
ordinarily unique characterization of diastereomers, particularly of those
in which stereoisomers involving chiral centers containing deuterium are
involved. (33)
The use of isotopes, as well as the appropriate control experiments
required for valid interpretation, is illustrated well by the reductive elimina-
tion studies of Norton. (47) Reductive elimination is often a key step in
proposed mechanisms and is generally presumed to be intramolecular in
nature; however, crossover experiments have shown that intermolecular
processes can be involved. Thus, the observation that mixed elimination
species and an osmium dimer, OS2(CO)8R2, 10, are found requires an
intermolecular reaction (17, 18).
Os(CO)4H2 + Os(CO)4D2 = H 2, HD, D2 + 10 (17)

The absence of CD 2 H 2 and other control experiments, such as proving


that scrambling of products does not occur under the reaction conditions,
prove that intermolecular reactions are involved.

1.3.2. Spin-State Labeling

When confronted with the problem of labeling a group or an atom, a


synthesis of a new compound using isotopically labeled groups often comes
24 C. A. TOLMAN AND J. W. FALLER

to mind. In relatively rapid reactions (h/2 < 10 sec), it is often possible to


use the effect of spins on the nuclei already present to label a position.
This is particularly useful with phosphorus-containing ligands and with
metals having nuclei with spin one-half.
In essence, within the magnetic field there are two separate types of
phosphorus nuclei, those with spine + !) and those with spin ( - !). Inter-
change of P( +!) with P( -!) can often be detected by nmr. This can
frequently be used to detect rapid intermolecular interchange of phosphine
ligands. Thus, if a group attached to a metal is coupled to a 31 p nucleus
in a phosphine, the resonance for that group is split into a doublet. One-half
of the doublet arises from the P with spin +! and the other from P - !.
Essentially one-half arises from one set of molecules and the other from
another set. Consequently, no matter what intramolecular rearrangements
may be taking place within the species containing P +!, that resonance
should remain distinct from the other with P -!. Once P +! and P - !
exchange, the two resonances average, just as the case of a proton exchange
between water and an alcohol gives a single resonance. Thus, the loss of
coupling to phosphorus indicates intermolecular exchange on the nmr time
scale. The rates can be calculated from line-shape analysis(48.4 9 l, if an
intermediate case is found; however, exchange rates of greater than 1 sec-I
are required. This is an important point, nevertheless, because rates of
over -1 sec -I are required to induce a line-shape change. Thus if the
temperature is lowered and the splitting, or a "static" spectrum, is observed,
it does not mean that exchange has stopped, merely that it has slowed
below a rate of 1 sec-I.
This approach can be useful in distinguishing when a ligand, such as
ethylene, is exchanging rapidly between complexes, because coupling of
the protons of ethylene to both the metal and the phosphine ligands will
be averaged and effectively removed.
Intramolecular rearrangements in metal clusters can also be detected
using the effects of different spin-state interactions. Platinum clusters are
particularly bizarre in that one not only considers the effects of various
195 Pt isotopomers and their relative populations, but also the effects of
nonequivalences resulting from having a coupling not only to the 195 Pt
that is directly attached, but also to a remote 195 pt.(50) The static
Pt 3 (I-L-COh(PR 3 h systems are informative in this regard. The isomeric
distribution for platinum has significant concentrations of several even mass
nuclei with I = 0, but only 195 Pt with a natural abundance of 33.7% has a
significant nonzero spin with I = !. Thus the analysis of spectra of platinum
clusters requires consideration of a system with approximately one-third
of the nuclei with spin! and the remainder with spin zero. The statistical
distribution of 195 Pt among the isotopomers 11 to 14 is shown opposite.
The observed spectra are a superposition of the spectra arising from all
four isotopomers. Portions of the 31 p spectra can be analyzed for 12 by
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 25

~ ~ ~
PRJ

29.63%
PRJ PR 3

44.44%
PRJ PRJ

22.22%
PR 3

3.70%
11 12 13 14
considering 12-Pt( +!) and 12-Pt( -!) separately. The relative ease of
determining 195Pt spectra and the potential use of platinum in catalytically
active systems suggests that the use of nmr in these systems will continue
to increase. The coupling constants of 195Pt to other metal nuclei appear
to be some of the largest known, e.g., lJC95Pt_119Sn) in trans-
[PtCl(SnCh)(PEt3h is 28,954 Hz, (51) and certain aspects of the spectra are
simplified. One should note, however, that with extremely large coupling
constants JIB may be significantly greater than zero, and splittings arising
from couplings between different metals may be large enough relativ~ to
the observed frequencies that the spectra may not be first order. This gives
the unusual feature that 195Pt satellites in 119Sn spectra may not be centered
about the zero-spin platinum resonances.
The use of spin-state labeling requires that interchanges of coupled
nuclei occur on a time scale rapid compared to the relaxation time of the
nuclei. Ordinarily, the rates necessary to average resonances are sufficiently
fast, and this is not a problem; however, there are two situations where
difficulties may be encountered. The coupling between a proton and another
nucleus in two different environments may be of opposite sign; hence, the
averaged coupling constant may be zero if appropriate ratios of the two
environments are populated. This might occur in an intramolecular
exchange which involved a free and bound phosphorus ligand. A situation
can also develop where coupling to a nucleus can be lost owing to rapid
relaxation of that nucleus. This can occur in systems which contain paramag-
netic species. (52) For example, a small fraction of nickel (II) (from a syringe
needle, for instance) can relax 31 p nuclei sufficiently and rapidly to
effectively decouple them. In this case, the rapid relaxation, when the 31 p
coordinates to the nickel, in addition to the rapid exchange between bound
and free phosphine, results in minor shifts of the proton resonances but
complete decoupling of the phosphorus.

1.3.3. Saturation Labeling in nmr

The broadening and eventual coalescence of resonances of nuclei


involved in exchange between sites has provided a convenient tool
for determining exchange rates(48.49l for relatively rapid reactions. The
additional width of the resonances, W, as exchange begins to proceed
26 C. A. TOLMAN AND J. W. FALLER

rapidly enough to be observed as line broadening is related to the rate for


leaving that site by a simple relationship (19). In some cases, different
mechanisms
k=1TW (19)
can be distinguished by the lifetimes of nuclei in various environments,
and since 1966, when line-shape arguments were first used for an
organometallic mechanistic study,(53) many insights into mechanisms have
been provided by line-shape studies. As these methods are now fairly well
known, (48,49) they are mentioned here only to indicate their potential utility
in studying exchange of nuclei which proceed so rapidly that isotopic
labeling would be impractical. For example, a situation may occur where
an isotopic-labeling study of the mechanism of interconversion of two
tautomeric species may be made difficult owing to relatively rapid equilib-
rium, which would scramble the labels before observation was possible.
The rates and at least a knowledge of the permutations of the nuclei would
be forthcoming from an analysis of line shapes as a function of temperature.
Although line-shape analysis is probably the best approach for the
study of rates using nmr, the technique of saturation transfer is more
valuable for investigating mechanisms and the nature of interchanges of
several nuclei.(49) If the rate of exchange is within 10% of the T1 relaxation
rate of a nucleus or faster, i.e., _10- 1 < k < 10 sec-\ the effect can be
observed. In the simple case where there is an exchange between nuclei
in an A site and B site, "saturation" of the nuclei in the A site will be
transferred to the B site when the exchange rate is faster than the relaxation
rate. Thus, irradiating nuclei in the A site not only saturates the A resonance
and decreases its intensity to nearly zero but also decreases the intensity
of the B resonance. The magnitude of the effect on B is a function of the
relative magnitudes of relaxation rate and exchange rate. The exchange
between a methyl group and a methylene hydride (see Equation 20) has
been demonstrated by this technique. (55)
(20)
15 16
Saturation of the methyl group in 15 leads to an observed decrease in the
intensity of only one of the hydride signals in 16. This spin saturation
transfer experiment conclusively demonstrates the elimination of a hydride
from a methyl to produce a methylene and that only one hydride site is
involved to a significant degree. It should be noted that it is impractical to
detect transfer to the other hydride site if the rate is less than a few percent
of that of the primary process.
More complex transfers of sites have been demonstrated by saturation
transfer studies used in the elucidation of sigma-pi rearrangements of allyl
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 27

groups, (56,57) understanding the nature of which is essential for interpreting


asymmetric induction in reactions of allyl metal complexes,
The original saturation transfer methods were developed for field-
swept spectrometers.(5S) With the advent of fixed-field modes of operation,
the use of an external oscillator makes it a reasonably convenient technique.
While it was straightforward to observe an effect, the demonstration that
the relaxation times were comparable in each site or the measurement of
relaxation times in the sites was tedious. (54) With the advent of FT nmr
spectrometers, the potential of effectively labeling sites by spin saturation
has increased owing to the ease of measuring relaxation times. In FT
spectrometers the saturation of a given site is usually accomplished in one
of two ways: "continuous" irradiation or decoupling, or gated irradiation.
The normal mode used for decoupling in an FT instrument is analogous
to using an additional frequency source for the decoupling frequency. (This
frequency may actually be turned on and off rapidly to prevent wreaking
havoc with the detection system.) Although only low power is necessary
to saturate resonances, some decoupling of resonances coupled to the one
being irradiated may occur leading to confusion about intensities. In gated
irradiation, saturation is provided up to the point where a 90° pulse is given
to observe the intensities and then turned off during acquisition of the
data. The saturation observed in the transformed spectrum is that appropri-
ate to the time when the pulse was given, as the loss of saturation is a
z- magnetization property, and the data acquisition measures x, y- magnetiz-
ation produced by the 90° pulse.
Most modern spectrometers provide these options for observing pro-
tons, hence readily allowing saturation transfer experiments, but they have
only recently been used with 13 C (59,60) on modified spectrometers. The
addition of another frequency source, as well as a knowledge of mixers,
wiring diagrams, and the confidence of other users of the instrumentation
are required for these modifications. This often prevents routine use of the
method for a variety of nuclei. This can partially be circumvented by
transfer of Overhauser enhancement instead of saturation; fortunately,
however, Morris and Freeman(611 developed the DANTE pulse sequence
which effectively allows homonuclear irradiation of the nucleus being
observed without the need of additional external frequency sources. More
complex 2-D methods can also provide similar information when many
sites are involved.

1.3.4. Ki netic Isotope Effects

The introduction of isotopes, particularly deuterium, for labels may


also provide the opportunity for further elucidation of the mechanism via
kinetic isotope effects. (62-65) Physical organic chemists, in particular,
28 C. A. TOLMAN AND J. W. FALLER

developed the interpretation of these effects in terms of a mechanism for


a large number of systems, and their approach and vocabulary should prove
useful when applied to catalysis by organometallics.
The differences in zero-point energy cause the dissociation energy for
a C-D bond to be about 2.3 kcal/mol greater than for a C-H bond. (64)
Hence, if no new bonds were formed in the activated complex of a hydrogen
transfer reaction, the activation energy would be 2.3 kcal/mol less for H
than for D. This would translate into an isotope effect of kH/ kD of nearly
fifty at 25°. Many primary isotope effects involving deuterium in organic
systems lie in the range of four to eight. Since new bonds are usually being
formed as old ones are being broken in the activated complex, one must
also consider zero point effects in the activated complex. The simplest
approach leads one to suggest that a bigger isotope effect means more
bond breaking occurring in the transition state. Furthermore, it also suggests
that if one proposes CH-bond breaking in the rate-determining step of a
reaction, strong evidence in its favor would be a signifh:ant isotope effect
on the rate. Unfortunately, this is a naive approach, and the absence of a
large isotope effect may not obviate the proposal, particularly in cases
where metal hydrides are involved. An alternative approach is to describe
the degree of bond breaking or making in the activated complex, as an
early transition state, symmetrical TS, or late TS. (63.65)

C-H-B C-H-B C-H-B


(21)
early TS symmetrical TS late TS

One also might speak of the transition state being more like reactants or
products in attempting to arrive at rationalizations for differences in Ea in
comparing compounds. The important feature here is that a symmetrical
transition state should give the largest isotope effect, because it presents
the largest differences in zero point energy in the TS relative to the ground
state.
In organic systems where a transfer of hydrogen is often between
atoms for which the atom-hydrogen vibrational frequencies are similar,
the concept of a "symmetrical" transition state can be readily visualized.
It would appear that the most "symmetrical" TS is often viewed as one in
which the H-C and H-B bonds in (21) are stretched to the same degree.
More esoteric descriptions involving intersections of hypothetical potential
surfaces are also used; nevertheless, definitions of "symmetrical" in cases
other than those in which C = B are obscure. An empirical approach has
been proposed on the basis that the acidity of the C-H and B-H bonds
should reflect the dissociation energy and, hence, a measure of the potential
surfaces' C-H and B-H bond breaking.(66) The observation that maximum
isotope effects are observed when the B-H and C-H acidities are similar
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 29

gives the approach credibility. Nevertheless, caution is warranted in offering


a "symmetrical" or "unsymmetrical" transition state as a rationalization
of isotope effects, if the bonding in C-H and B-H are substantially different.
Since metal hydride stretching frequencies are low compared to C-H
frequencies (usually 1500-2100 for M-H compared to C-H 2900-
3100 cm- I ), the transfer of hydrogen from metal to a carbon can result in
a situation with major differences in zero-point energy in reactant and
product. More importantly, the likelihood of an "unsymmetrical" TS is
high and the possibility that frequencies are even higher in the activated
complex than in the reactant exists. This is illustrated in the reduction of
a-methylstyrene by HMn(CO)s,(67) Transfer of a hydrogen atom from the
metal hydride (v Mn-H = 1780 cm -1) has been suggested as the rate-deter-
• • (62)
mmmg step .

Although the kinetics in this system are complicated by the possibility


of forming (C 6H s)(CH 3 hCMn(COh initially or by radical cage recombina-
tion and subsequent alpha elimination, the fact remains that at some point
a Mn-H bond is broken and an inverse isotope effect of kH/ kD = 0.4 is
observed.
Inverse isotope effects have been observed for some time with metal
hydride reactions, such as tin hydrides, (68) but their observation in transition
metal reductions may have a considerable impact on the interpretation of
results of catalytic studies. It would appear that a process such as (22) may
be involved in reductions with CO(CO)4H, which also show an inverse
isotope effect (kH/kD = 0.58 for 1,1-diphenylethylene; and 0.43 for 9-
methylidenefluorene).(69,70) There is now evidence for radical involvement
in the cobalt system. (70) Neither the involvement of radicals, nor an unsym-
metrical transition state, however, is required for an inverse isotope effect.
For example, the radicals taken to have arisen from H atom addition in
(22) might have arisen from addition of HMn(CO)s to the double bond
(23) followed by homolytic cleavage (24).

(C 6 H s)C(CH3)=CH 2 + HMn(COls = (C 6 H s)(CH 3hC-Mn(CO)s (23)


(C6HS)(CH3hC-Mn(COls = (C 6H s)(CH 3hC" + (COlsMn° (24)

Distinguishing radical species involved in the catalytic loop instead of being


involved in an unimportant off-loop reaction is difficult but can be accom-
plished via CIDNP experiments(67) (see Section 1.5.3.). An inverse isotope
effect can also arise from the isotope effect on an equilibrium prior to the
rate-determining step.(66) An inverse isotope effect of 0.29 is observed in
30 C. A. TOLMAN AND J. W. FALLER

the reduction of carbonyls with HFe2(CO)8 -,(72) and it has been attributed
to an equilibrium isotope effect on (25), which effectively reduces the
concentration of the reactive deuterated alkyl-Fe2(CO)8 - in the rate-deter-
mining step (26).

DFez(CO)s - + RCH=CHCOR = RCH(D)-CH(COzR)[Fe2(CO)Sr (25)

RCH(D)-CH(COzR)[Fe2(CO)Sr + D+ -. RCHD-CHD(COzR) + Fez(CO)s (26)

Consideration of changes in vibrational frequency suggested a KH/ KD of


approximately 0.5. This requires a rather large equilibrium isotope effect
in (25) to account for the observed kH/ kD of 0.29, as well as any effect
from a subsequent reaction such as (26). Nevertheless, the point to be
made here is that an equilibrium isotope effect on a prior equilibrium can
possibly account for the observation of an inverse kinetic isotope effect.

1.3.5. Equilibrium Isotope Effects

Saunders(73) developed a very useful technique for studying equilibrium


isotope effects on the proton nmr spectra of carbonium ions, which is
known as isotopic perturbation of degeneracy. It is effective in problems
where distinguishing symmetrical from fluxional unsymmetrical compounds
is required. This effect is amplified in 13C spectra(74) and has been used to
distinguish fluxional 1) 1 -cyclopentadienyl metal complexes from 1) S - Cp
complexes. (7S)
The process of a-elimination from methyl groups bound to metals(76)
may be a key step in catalytic alkane activation, and hydride addition to
metal-bound methylene has been suggested as the chain propagation step
in the Fischer-Tropsch reaction. (77-79) The relationship between sym-
metrical, 17, and unsymmetrical bridging methyl groups, 18, and methylene
hydrides, 19 and 20, can be addressed by isotopic perturbation of
degeneracy studies; and other perturbations of nmr spectra by the introduc-
tion of isotopes.

H H, H H. /H
H
H,I/H Hr',I/~b "- C / H
/c,
C /C····':M I M, /M
M/ 'M M II H
M M

17 18 19 20

The presence of an unsymmetrical bridging methyl as in 18 has been clearly


demonstrated in an X-ray structure in which hydrogen atoms were located
in [Fe2(tL-CH3)(tL-CO)(COh(tL-dppm)(1)-CsHs)]PF6.(80) The unsym-
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 31

metrical nature of the C-H-Fe interaction can be implicated by the isotopic


perturbation of degeneracy observed in the proton nmr spectrum of a
partially de ute rated methyl species. This isotopic perturbation method
also led to the earlier suggestion of a similar bridging group in
HOs 3 (CO)10CH 3 • (55)
The single resonance observed in the proton nmr of the iron complex
suggests a low barrier to rotation of the methyl group. Hence, the methyl
resonance of 8 - 2.90 occurs at a chemical shift which is the weighted
average of the two terminal proton environments and the one bridging
environment. The averaged shift may be considered as reflecting the percen-
tage of time a nucleus spends in a given environment. That is, it is given
by equation (27) where the population of protons is p.

(27)

If the two terminal environments are identical, which they often are not,
the equation reduces to (28).
(28)

Hence, for a CH 3 where Pt and Pb are one, the equation simplifies even
further to (29).
(29)

In the mono- and dideutereo species, the zero-point energy differences


will favor H over a 0 in the bridging position owing to the lower vibrational
frequencies expected for a Fe-H-C bridge relative to a terminal C-H.
Therefore, the populations in (27) or (28) will no longer be equal. If K is
the equilibrium constant for a single pairwise exchange between a bridging
and terminal position (30),

K = [188]/[18T] (30)

H H
H... I ,,0 0 ... I,.H,
. . . .c' '-. ....... C' "
M/ ------M M/ ------M

188 18T 18T'

equations (31) and (32) follow from the approximate equation (28).(55)

(CH 20)ave = (0, + Ko, + 0b)/(K + 2) (31)

(CH0 2 )ave = (2Ko, + ob)/(2K + 1) (32)


32 C. A. TOLMAN AND J. W. FALLER

Thus, according to (31) with K < 1 and Ob upfield of 0" an upfield shift in
the proton spectrum of the -CH2D would be expected. The 0 - 3.47 shift
observed for [Fe2(~-CH2D)(~-CO)(COh(~-dppm)(~-C5H5)]PF6 rep-
resents an up field shift of 0.57 ppm. In the deuterium nmr spectrum, the
resonance is observed at 0 -1.88, a down field shift of 1.02 ppm showing
approximately twice the effect required by the statistics (Equation (32)
with K- 1 substituted for K). Equations (31) and (32) neglect an intrinsic
shi!t(73-75) in the proton spectrum arising purely from substitution of D
for H. This effect is relatively small (-0.01 ppm) in proton spectra but
accounts for lack of exact agreement between IH and 2H nmr results and
may become important in evaluating K.
The proton resonance for the CH2D group shows a pronounced tem-
perature dependence. Lowering the temperature 95° changes the shift by
0.25 ppm t08 - 3.72, whereas the CH3resonance shifts only 0.02 ppm. This
temperature dependence is an expected property of the perturbation of an
equilibrium by an isotope effect and should always be verified experi-
mentally before an isotopic perturbation of degeneracy rationalization is
invoked.
A further indication of a bridging methyl group is provided by the
average coupling constants of the protons to 13e. The 13C spectrum shows
a quartet with lI(CH) of 114 Hz. This also represents an average of
lI(CH b), lI(CH t ), and lI(CH t·) as in equation (27) with J's replacing 8's.
A bridging methylene as in H 20S3(COhoCH 2, which contains a type 20
unit, (81-82) has values of lI(CH) of 140 and 143 Hz. (83) The lower s character
in the bonding of methyl would suggest a lower lI(CH) in a symmetrical
methyl bridge, 17, as in AIz(CH3)6(84) or Y2(CH3)z{71 5-C5H5)' (85) The bridg-
ing methyl CH would probably have a lower lI(CH) and the terminal CH
a slightly higher lI(CH) in 18 than observed in 17. Thus, the averaged
lI(CH) in an unsymmetrically bridging methyl, 18, should fall in a range
between 95 Hz (It = 142; Ib = 0) for the extreme methylene hydride 19
form of a weak M- H-C bond and about 125 Hz for the symmetrical form,
17. Solution of analogs of equations (29), (30), and (31) for the couplings
of lI(CH) observed for 121.1(CH3); 118.9(CH2D) and 116.4(CHD 2) for
HOs3(COhoCH3 (55) yield values of lI(CH)t = 150 ± 10 Hz and lI(CH)b =
60 ± 20 HZ.(55) The values for the lI(CH) for the CH2D and CHD 2 also
show the expected variation with temperature.
Perhaps one might now confidently assume that a moderately high-field
three-proton resonance showing an isotopic perturbation of resonance is
indicative of an unsymmetrical bridging methyl, 18. Some caution is advised,
however. The presence of a coupling constant, II (CH), above about 105 Hz,
is a requirement for this situation, as is a variation with temperature of
both lI(CH) and 8 (CH) in the partially deuterated complex. Perhaps the
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 33

most difficult problem to eliminate is that a fluxional methylene hydride


could give rise to many of these observations. The isotopic perturbation of
resonance only indicates the lack of a symmetrical structure, but does not
indicate the nature of the asymmetry. The conversion of methylene hydride
to methyl has been observed to occur with a barrier of about 20 kcal/mol.(75)
Thus a fluxional structure, such as 19 or 20, would show all the properties
of unsymmetrical bridging methyls with the exception of having a slightly
smaller IJ(CH). A rapid equilibrium between a symmetrical methyl and
methylene hydride would be virtually impossible to detect by the nmr
methods discussed above. For this reason, corroborative evidence for
the presence of a single isomer by a technique such as IR should be
available.
Without the potential problem of a different structural form being
isolated in the solid than is present in solution, X-ray and neutron diffraction
studies would be the methods of choice for proving an unsymmetrical
bridging methyl. This problem arises in one of the currently claimed cases
of unsymmetrical bridging methyls, HOs3(COho(CH3), where the equi-
librium with a methylene hydride results in the crystallization of
H 20S 3(COho(CH 2).(74,76) One should note that neutron diffraction experi-
ments also allow a quantitative measurement of the equilibrium isotope
effect. (74)
For simplicity, we have neglected the fluxional process in which the
stronger bond is interchanged between one metal and the other in this
argument. If the chemical shifts of the terminal protons in a static bridging
methyl group, 18, are nonequivalent and there is a reasonably large equi-
librium isotope effect, (30), the rotation of the methyl will not make the
averaged environments for the terminal protons equivalent. Thus, in both
[Fe2(M-CH2D)(M-CO)(COh(M-dppm)(1j-C5H5)]PF6 and HOs 3(COho-
CH 2D the two methyl protons should be nonequivalent if only methyl
rotation giving equilibria between 18B = 18T = 18T' were involved. The
observation of a single signal for these protons implies that the shift
difference is either not resolved or that the methyl protons rapidly switch
from bridging one metal to the other. This nonequivalence could also be
observed in an unsymmetrical methylene hydride system, 19, if H addition
and removal were stereospecific.
The isotopic perturbation method is a powerful one for demonstrating
the unsymmetrical nature of equilibrating systems; however, as indicated
above, the nature of the dissymmetry requires care in interpretation. In
particular, the evaluation of the magnitude of the averaged 13C_1H coupling
and the variation of the shift with temperature are essential. Owing to the
possibility of a methylene-hydride equilibrium, it is also prudent to investi-
gate the possible presence of these tautomers by other physical methods.
34 C. A. TOLMAN AND .J. W. FALLER

1.4. The Implications of Product Stereochemistry for Mechanisms

One of the main potential advantages of homogeneous catalysts is the


ability to control or modify the regioselectivity and stereoselectivity of the
catalysed reactions. It follows, therefore, that the stereochemical con-
sequences attending the key steps of a catalytic cycle may be of prime
importance in understanding and designing catalyst function. In this section
some of the key steps proposed in catalytic cycles will be shown as a vehicle
for illustrating some of the stereochemical methods.
The retention, inversion, or loss of stereochemistry at a chiral center
is one of the prime objectives of a stereochemical study of a reaction. Thus,
displacement reactions, such as a Walden inversion, are classic examples
of such studies. Owing to difficulties, such as beta-eliminations and un-
certainties with regard to absolute configurations, following the
stereochemistry of many key reaction steps was not widely practiced until
the seventies. For example, if a M-C(H)(CH 3 )(C6 H s) was stable and did
not eliminate styrene forming the metal hydride, the absolute configuration
of the metal complex or the stereochemistry might be unknown for the
degradations yielding organics of known configuration.

1.4.1. NMR Determinations of Erythro and Threo (CHDh Fragments

Snyder(86-88) originated the study of dideutereo-2-phenylethyl groups


for determining the stereochemistry of displacement reactions. White-
sides(89) initiated the use of these systems in the study of organometallic
systems through examination of the reaction of erythro-3,3-dimethylbutyl-
1,2-d2 -p-bromobenzenesulfonate, 21, with [(CsHs)Fe(COhr (33) and the
subsequent carbonyl insertion reaction promoted by triphenylphosphine
(34).
R R
D--6-- H (Cs H s )Fe(CO)2 - ~ H--6--D
(33)
HAfAD HAf'D
O-S02C6 H4 Br Fe(COh(C 5 H 5 )
erythro-21 threo-22
R R
H--6--D
H~D (34)
22 + PPh 3 = HAf'D HAf'D
C(O) C(O)
/ /
[(R)-Fe(C 5 H 5 )(PPh 3 )(CO)] [(S)- Fe( C5 H 5 )(PPh 3 )( CO)]

(R, S,R)-23 (S, S, R)-23

Unfortunately, the difficulty of the syntheses and the yields of the erythro
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 35

sulfonates dissuaded many potential users. Although the yields were even-
tually improved,(90) the synthetic procedures were still lengthy. Alternate
routes involving hydroboration(91) and hydroalumination(92) might also be
moderately attractive; however, the most straightforward synthesis of rela-
tively pure 21-(R = t-Bu) appears to be via sequential Cp2Zr reactions.(93)
The most efficient synthesis of threo -22-(R = Ph) is via Rh-catalyzed hydro-
genation of commercially available (3- methoxystyrene. (94) Although the
brosylate, 21, contains an effective leaving group for strong nucleophiles,
such as [CsHsFe(COhr or pyridinebis(dimethylglyoximato)-cobalt(I), the
triflate is more generally useful with less nucleophilic organometallic
anions. (9S)
Fortunately, the stereochemistry of the (CHDh moiety can be deter-
mined readily by nmr experiments. The trans configuration of the protons
in the erythro isomers implies a larger 3J (HH)vic coupling than in the gauche
configuration found in the threo isomers. Hence, the coupling in erythro- 22-
(R = t-Bu) is 13.1 Hz, whereas that in threo-22 is 4.5 HZ.(89) Since a chiral
iron center is formed in the phosphine addition-carbonyl-insertion reaction
(34), the methylene protons are diastereotopic(96) and thus (R,S,R)- 23 and
(S,S,R)- 23 are different, which requires careful interpretation of the ob-
served spectra. Nevertheless, this method has become one of the most
straightforward ways of examining changes in chirality in reactions of
organometallics.
In equation (34) the ligand-promoted carbonyl-insertion reaction
occurs with ~90% retention of configuration. Treatment of 22-(R = t-Bu)
with bromine in pentane, chloroform, or carbon disulfide gives reaction
(35) with >90% inversion of configuration.(90)
R R
H--&D + Br2 _ D--&H
(35)
~D H~D
Fe(CO),(CsHs) Br
threo-22- rR = t-Bu) erythro-24

It is tempting to extrapolate these results to all alkyl metal systems,


and although carbonyl insertions generally appear to take place with reten-
tion in the alkyl group, mirabile dictu, halogenations occur via several paths
in other organometallics. Thus, bromination of the R = t- Bu analog of 22
with Cp2Zr takes place with retention.(93) Since the mechanism for the iron
system appears to occur via oxidation of the complex followed by backside
nucleophilic attack of halide ion, (94,97) (36), the highly oxidized Zr(IV)
reaction proceeds via a different route owing to the inaccessability of the
oxidative route. Halogenations of cis-[threo -Ph~ CHDh-Mn(CO)4PEt3]
proceed with retentions of from 17% to 77% depending upon the con-
ditions.(98) The more bizarre occurrence, however, is that the bromination
36 C. A. TOLMAN AND J. W. FALLER

(35) of threo-22-(R = Ph) yields threo_24.(94,97) Hence, exactly opposite


conclusions are drawn for the same metal systems with the only difference
being (CHDhC(CH3 h vs. (CHDhC6 H s. Baird has used the (CHDhPh
.
funchona I'Ity extensive
. Iy for stereoch ' I stud'les'
emlCa (9497-100) an d attn'b utes
this difference to the potential for formation of the bridged phenonium
ion. (94,97) A key indicator of the phenonium ion or radical involvement is
the observation that a-l3C-Iabeled alkyl metal yields an equal mixture of
a- and {3-labeled product,

(36)

A similar conclusion also resulted from a study of the dideuteriomethylene


complex, ('1/ 5 -C sH s)Fe(COh-(CD 2 )(CH 2 )Ph. (101) Essentially, the bridged
species retains the "cis" orientation of the de uteri a in this intermediate,
25, in equation (37), which accounts for retention in threo-22.
H D
~ /
~c.
threo-22-(R = Ph) + 12 - . I: ':::C 6 HS -. threo-24 (37)

~\,
H D

25

Iodide attack trans to the bridging phenyl at either carbon yields the threo
isomer, but will yield the (RS) or (SR) isomer depending on which carbon
is attacked. The basic assumption used in the interpretation of the erythro
and threo isomer interconversions is that the configuration of the carbon
a to the R group remains constant. Hence, if inversion occurs at both
centers, it would not be detected. Thus, an observation interpreted as
"retention" with the Ph(CHDh-system should be viewed with suspicion;
nevertheless, racemization or inversion would be interpreted correctly.
Studies with optically-active (CHD)Ph corroborate the observed inversion
with bromination found with -(CHDht-Bu;(97) hence, care must be taken
with interpretations using the (CHDhPh system, and it would appear that
the butyl system provides a more reliable indication of processes at a
primary carbon,
Since the stereochemistry of the (CHDh group can be determined so
conveniently directly from the nmr, it will remain a popular technique, and
its utility is not limited to studies of previously prepared X- (CHDhR
compounds. Further evidence for the trans attack of water on coordinated
olefin in the Wacker reaction (see Equation 16 in Section 1.2) was provided
by the isolation of a {3- alkoxyethylpalladium complex from the reaction of
a cis-dideuterioethylene complex (38) with MeOH.
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 37

(38)

threo-25

The vicinal coupling constants for the (CH 2h analog of 25 showed vicinal
coupling constants of 4.8 and 12.3 Hz. The 3J(HH)vic of 4.8 Hz in the
deuterated product proved that it was threo- 25 and the attack of methoxide
was trans to the metal. (102)

1.4.2. nmr Determinations of Other Diastereomers

The use of threo and erythro alkyl isomers can be extended and in
some cases the preparative aspects improved by the use of other groups
on the a-carbon besides Hand D. For example, (RR, SS) and (RS, SR)-
C6H s(CHF)(CHD)Br have been used in studies of oxidative addition.(93)
Although a secondary carbon is introduced, (RS, SR)- Ph(CHF)-
(CHBrC0 2Et) has also been used successfully. Both of these systems,
however, have the potential of interpretational problems arising if apparent
"retention" is observed. Cis- and trans- substituted cyclohexyl gr.oups pro-
vide an opportunity to utilize well-known coupling patterns for the study
of stereochemistry.(91.103) Thus, the oxidative addition of cyclohexyl
bromides to [(py )(dmg hCo(I)r proceeds with inversiony03,37)
Br

Br~
~Br+CO(dmghPY --+ ~
~CO(dmg)2py
(39)

Care is required here, however, as the 3J (HH)vic coupling constants to


metal-substituted CH may not be well established; furthermore, C-M bond
lengths are sufficiently long that an axial metal substituent is not unexpected.
The inversion or retention of stereochemistry in pseudo tetrahedral
complexes of metals, such as (7] S- CsHs)Fe(Me)(CO)(PPh3), 26, can be
informative with regard to the effects of the reaction at the metal center.
The absolute configuration at the metal center can be designated by Rand
S following rules which take polyhapto ligands as pseudo atoms correspond-
ing to the mass (actually the atomic number) of the atoms bound to the
meta.I (104 .lOS) Hence, ( 7] S - C
sHs) woul d correspond to a mass of 60 or an
atomic number of 30. This yields priorities of (7]s -CsHs) > PPh 3 > CO >
CH 3. A (7] s -1-CHr3-C6Hs-CsHs)Fe moiety is also chiral and can be
designated by Cahn-Ingold-Prelog nomenclature, (lOS) as well. Thus,
(RR -SS) and (RS-SR) diastereomers of (7] s-l-CHr3-C6HS-
38 C. A. TOLMAN AND J. W. FALLER

H3C~C6H5

12 I (40)
---+ Fe.
1/ l "CO
PPh 3

(RS)-27 (RR)-28

C sH s)Fe(Me)(CO)(PPh 3 ), 27, exhibit different nmr properties and can


be separated by crystallizationyo 6 J The reaction with iodine, (40), can then
be studied by nmr assuming that the chirality of the CpFe group is retained.
The nmr results indicate that the reaction proceeds with only partial
selectivity. Unfortunately, it is impractical to distinguish an (RS-SR) pair
from a (RR-SS) pair by nmr, and X-ray crystallography is required to
establish the stereochemistry. Two features are clear, however, without
the X-ray results: a) reaction did not occur with pure retention or inversion;
and b) partial epimerization of 27 occurred during the reaction. This led
to the proposal that initial attack of the iodine resulted in a 5-coordinate
pseudo-square pyramidal intermediate, 30, as shown in Equation 41.

(RSJ-27 + 12 --. (41)

30

Since these 5 -coordinate complexes can readily rearrange, (107) the reverse
reaction of 30 ~ 27 provides a route for the epimerization at the iron center.

1.4.3. Enantiomeric Ratio Determination by Polarimetry

The use of an optically active cyclopropyl bromide with [(py)-


(dmg hCo(l) rgave a complex for which optical rotations could not be
measured owing to the high-extinction coefficientsY oS ) Thus, the possibility
of racemization in the formation of the complex could not be investigated.
More often, however, one is thwarted by the uncertainty of the relationship
between chiroptical properties and absolute configuration. One of the
earliest organometallic stereochemical studies(109) showed that decarbony-
lation of a manganese complex occurred with retention of the "sign" of
rotation (42):

( + J-(PhCH 2 )(CH)(CH 3 )COMn(CO)s --. (+ HPhCH 2 )(CH)(CH 3 )Mn(CO)s (42)


MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 39

Unfortunately, this only proved that racemization had not occurred, as


there had not been a way to relate sign of rotation to configuration.
In a stereochemical study where one organic product of known absolute
configuration is converted into an organic product of known absolute
configuration, the interpretation is straightforward. Hence, in the decar-
bonylation of ( - )-(R)-2-methyl-2-phenylbutanal with Rh(PPh 3 hCI, (43),
the 94% optical purity and relative configuration of the product were
readily determined from measurements of optical rotationY 10)

<;:H3
CH 3 CH z ....... : /CHO
C (43 )
I
C6 HS
(- )-(R)-31 ( + )-(S)-32

Note that the decarbonylation proceeds with retention, even though the
chirality descriptor, R, is reversed, as is the sign of the rotation. This
reversal of the descriptor is a result of changing relative assignment priorities
of the groups attached to the chiral carbon and sometimes leads to confusion
in descriptions of stereochemical experiments.
The optical rotations, particularly in the uv, of chiral centers of
R (CHD)R' are sufficiently large that it is practical to use polarimetry to
measure optical purity.(111) Generally, however, a sequence of steps to
produce a known configuration organic product from a known configuration
substrate is required, since the chiroptical properties of the metal complexes
are not known. For example, the sequence of predominate inversion in
oxidative addition and retention in carbonyl insertion has been investigated
using Ph(CHD)Cl, Equation 44 (R = D)Y12)

Ph Ph PEt 3 Ph 0 PEt 3
/ co " C-C-Pd-X
II I
x-c " ,C-Pd-X
I ------+
\'R R'/ I Rj I
H H PEt 3 H PEt 3

33 34 35

1 MeOH (44)

Ph 0
".C-C-OMe
II
R'/
H
36

Although the situation is improving, there are still a relatively small number
of metal compounds for which the absolute configurations have been firmly
40 C. A. TOLMAN AND J. W. FALLER

established. Although there has been some success with relating the
configuration of olefins bound to platinum (13 ) and allyls bound to pal-
ladium(114) to chiroptical properties with quadrant rules, it has proven
extremely risky to attempt correlations with chiral metal centers. Absolute
configuration studies of (7]s_C sH s)Mo(S7.11S) and (7]s-C sHs)Fe 016 ) com-
pounds have suggested possible correlations, but generally only very minor
changes can be tolerated before significant changes in CD or ORD spectra
make interpretations unreliable. For example, a change in conformation
of an allyl group from (R)-[(7]s-C sH s)Mo(NO)(CO)endo-(allyl)t to (R)-
[(7]s -CsHs)Mo(NO)(CO)exo-(allyl)t can reverse the signs of some CD
absorptions.(S7) Thus, it appears that until a signifiant number of absolute
configurations have been determined for' metal centers, most examinations
of metal-center stereochemical changes in reactions will require X-ray
structure determinations. Physical methods providing a measure of
diastereomeric purity will be useful, but establishing relative configurations
may be difficult.

1.4.4. Binding of Chiral Substrates-Chiral Shift Reagents

In attempting to elucidate aspects of reaction mechanisms based on


inversion or retention of configuration, the problem of identifying enan-
tiomeric composition or optical purity is encountered. Polarimetric methods
are subject to uncertainties of chemical purity, variations with solvent
composition, and questionable literature values for the standard rotation
of a 100.0% pure enantiomer. One may utilize the approach of making a
derivative with an optically-active reagent and detecting the ratio of
diastereomers by nmr or gas chromatography.(17) One of the most con-
venient and reliable techniques is based on the chiral lanthanide shift
reagents introduced by Whitesides. (18 ) Their use has become standard
practice for the determination of enantiomeric ratio and enantiomeric
excess. When these reagents (usually derived from optically pure camphor)
interact with donor atoms in an enantiomeric pair, the diastereomeric
interactions yield separate nmr signals for each enantiomer which can then
be directly integrated. The use of these reagents in the determination of
optical purity has been recently reviewed. (119) The most effective reagents
for binding to weak donors often contain fluorinated alkyl groups to increase
the coordination efficiency; hence, some of the initial complexes studied
were based on trifluoromethylhydroxymethylen-d -camphor, such as
Eu(facamh. The europium complex prepared from di[d -campholyl]-
methanato ligand, Eu(dcmh, however, appears to generally give the largest
diastereomerically induced separations of resonances.(119) These reagents,
which are commercially available, provide the most straightforward
method of determining the enantiomeric excess in products containing
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 41

donor groups, such as esters, ketones, and alcohols. Bridging carbonyls in


organometallic compounds, (120) as well as cyanides and isocyanates, (121)
also bind to lanthanide shift reagents and consequently should allow optical
purity determinations via this method.
The "enantiomer shift differences," ~~8, are predominately induced
by the differing equilibrium constants between the enantiomers and the
paramagnetic shift reagent (45), (46).
KR
(R)-substrate + Eu(dcmh ~ [(R)-substrate]-[Eu(dcm)]3 (45)

KS
(S)-substrate + Eu(dcmh ~ [(S)-substrate]-[Eu(dcm)]3 (46)

Since europium-substrate complex itself would have very large shifts and
the observed shifts are averages of bound and free substrate, the greatest
shifts would usually be observed for the substrate with the largest K. When
considering catalytic cycles involving chiral intermediates, it may be necessary
to consider differences arising from diastereomeric interactions, such as
equations 45 and 46. While the chiral discrimination in the binding is the
major source of the shift, the intrinsic shifts in the diastereomeric complexes
are not equal. Thus, it is possible to even use the reagents for optical purity
determinations in cases where KR and Ks are effectively equal, as with
Ph(CHD)OH.

1.4.5. Binding of Prochiral Substrates-Asymmetric Induction

In the studies of asymmetric hydrogenation mechanisms using rhodium


complexes of chiral bis-phosphine complexes (see Chapter 4), there are
parallels to the binding of substrates by the europium complexes containing
chiral-diketone ligands. There are some important features in the process
which selects the binding of substrates and their impact on mechanistic
paths that need be addressed. Bosnich and his co_workers,(122-123) have
considered the design of chiral his -phosphines and have presented evidence
that the major source of the discriminatory interaction was the chiral array
of phenyl groups. On this basis, as well as on the basis of studies to identify
important intermediates, one would presumably be well on the way to
rationally designing a catalyst to produce high-optical yields. The import-
ance of allowing the substrate to bind with the proper stereochemistry is
reminiscent of the importance attached to "lock-and-key" models of en-
zymes. Thus, on observing an intermediate in a catalytic cycle, one might
proceed to design a system for which the substrate of the appropriate
chirality would be bound. The advantages and potential pitfalls of this
approach are well-illustrated by studies of asymmetric hydrogenation.
42 C. A. TOLMAN AND J. W. FALLER

The rhodium-catalyzed homogeneous hydrogenation of ole fins using


(PPhhRhCl has been studied extensively (see Section 2.2), and similar
mechanistic pathways might be anticipated in the asymmetric hydrogena-
tion reactions. It should be noted that the most effective asymmetric
hydrogenation catalysts(123-125) are based on the cationic bisphosphine
hydrogenation catalysts developed by Schrock and Osborn. (126-127) Further-
more, the chelating phosphine, chiraphos, gives some of the highest optical
yields. The chiral process has been discussed by Bosnich. (122)
Aspects of mechanism relating to the order of attack of olefin and
hydrogen have received considerable attention in consideration of the
reduction with (PPh 3hRhCI and cationic triphenylphosphine
analogs. (128-130)

(PPh 3 hRh + + olefin ~ (PPh 3 hRh(olefin( ~ (PPh 3 hRhH 2 (olefin( (48)

The cycle is completed, and the product is obtained by hydride addition


from a cis-hydride followed by reductive elimination. The "hydride" route
(47) is now generally accepted.(13l) The evidence for the
(ROHh(PPh3hRhH~ cation as an intermediate was obtained by direct nmr
observation of this cis_dihydride(126-127) after hydrogenation of the precur-
sor in alcohol. The "unsaturate" route (48), however, was implicated with
chelating phosphines by the observation that only a [(diphos)Rh(ROHht
species is observed upon hydrogenation of [(norbor-
nadiene)Rh(diphos)t.o 32 ) Thus, the change of two PPh 3 to a diphos ligand
completely changes the principal pathway. This illustrates the care that
must be exercised in extending the results from one system to another.
The addition of prochiral substrates, such as esters of a-acetamidocin-
namic acids, to a solution of the solvate of a chiral diphosphine rhodium
complex allowed the observation of at least one of the two diastereomeric
. 49 .(133-138)
compIexes, E quatlOn
H C0 2 Et
Rh[(S,S)chiraphos](solventh + "" C=C /
Ph / "" NH(CO)Me
tf
37

Me" Ph Ph
1/ Ph H O M e
Me .....'\~P" ~,- C/
_____ Rh--.. I (49)
P"""'- N
Ph/ "Ph C0 2 Et

(S, S)-38 (R, R)-38


MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 43

Presuming that the minor stereoisomer in solution is not the least


soluble, (138) the crystal structure(136) indicates that the major isomer in
solution is (S, S)-38, where (S, S) refers to the configuration at the 1- and
2-carbons of the bound olefin. The rate-determining step is the addition
of hydrogen to give 39, and the equilibrium shown in equation (49) occurs
rapidly.(138) After hydrogen addition, cis-hydride transfer from the metal
to the olefin gives a metal alkyl, 40, followed by reductive elimination
yielding the product, Equations 50-51.
/NH(CO)Me
olefin H
37 ~ (5,5)-38 -4 (5,5)-39 -4 (R)-40 -4 PhCH2C~H (50)
Kss
'COOEt
(5)-41

,COOEt
olefin H .,-
37 ~ (R, R)-38 ~ (R, R)-39 -4 (5)-40 -4 PhCH 2C-H (51)
KRR ",

'NH(CO)Me

(R)-41

The predominant product (>95% ee) of this reaction, however, is the ester
of N-acetyl-(R)-phenylalanine, (R)-41. This requires that the
Rh(chiraphos) (olefin) intermediate, (R,R)-38, which is in lowest con-
centration, yields most of the product. This illustrates again that the dominant
species observed is not necessarily the most reactive, nor is it necessarily
a loop species in the most important catalytic cycle.
The important consideration, therefore, is the relative rate of addition
of H2 to the Rh (chiraphos) (olefin) intermediates, (S, S)-38 and (R, R)-
38.(138) Hence, as discussed in more detail by Bosnich,(122) if all prior
equilibria (KRR and Kss) are fast, the only relative energies of any con-
sequence are those of the transition states between 38 and 39. Under these
circumstances the only diastereomeric discriminatory of importance are
those in these transition states, and not those of the prior equilibria. Thus,
in general, when attempting to design catalysts with greater enantioselec-
tivity or stereoselectivity, it is essential that relative transition-state energies
be considered, either instead of, or in addition to, those of ground states.

1.5. Distinguishing Radical Pathways

Although one might hope that the key steps in catalytic systems, such
as oxidative addition or reductive elimination, would proceed via a uni-
versally valid mechanism, it is clear that many of these steps may occur by
several paths. Several paths may even occur within the same system and
minor perturbations may alter the primary route. The success of the 16-
and 18-Electron Rule in rationalizing many aspects of organometallic
chemistry and catalytic systems has tended to lead to nearly exclusive
44 C. A. TOLMAN AND J. W. FALLER

consideration of reaction mechanisms involving only even numbers of


electrons. There are often enough alternate mechanistic paths conceivable,
using the 16- and 18-Electron Rule, that the inclusion of additional odd-
electron paths produces an unwieldy number of candidates for the mechan-
ism. It is now clear, however, that one-electron processes are increasingly
being encountered and cannot be neglected. By the mid-seventies,
Kochi(139) had investigated radical participation in nontransition metal
systems extensively, but the studies of oxidative addition mechanisms by
Osborn(140-141) and Lappert(142-1,,3) pointed out the importance of radical
paths in organometallic systems which had been presumed to occur by
conventional two-electron SN2 paths.
Although one might expect that an oxidative addition is the most likely
candidate for a radical path, even the displacement of carbonyl by phosphine
may proceed by a radical path. For example, HRe(CO)s does not readily
undergo phosphine substitution under scrupulously clean conditions, but
does react rapidly under conditions in which the Re(CO)s radical is
formedY44)
Stable organometallic radicals are rare, owing to their proclivity for
dimerization, and most that are known owe their stability to steric effects
which prevent dimerization. Thus, V(CO)6 and derivatives of (1'/3-allyl)-
Fe(COh(14S) have been known for nearly twenty years. Sublimation of
some organometallic dimers with relatively weak bonds onto a liquid
nitrogen-cooled probe has provided a method of preparing radicals, such
as CO(CO)4(146) and (1'/s-CsH s)Cr(COh(147) for ESR study. The major path
for carbonyl substitution in some metal-metal bonded dimers is believed
to involve breaking of the dimer bond, substitution on the radical, and
reformation of the dimer. (148)
Relatively stable organometallic radicals, such as [(1'/ s-CsHs)Fe-
(diphos)CH 3t and [Mo(COh(diphosht, can be prepared by chemical
oxidation,(149.1S0) and the paramagnetic hydride FeH(diphosh is available
through reduction of the [FeHCI(diphosht. The largest number, however,
are available through electrochemical techniques. (lS1) Electron rich systems
are oxidized relatively easily and E 1/2 measurements provide a quantitative
measure of the facility of oxidation.
The importance of radical pathways can usually be correlated with the
ease of electron transfer.
Generally, ESR spectra of organometallic radicals are unexceptional
and are similar to those of organic radicals with g - 2. For example,
trans-[Mo(COh(dmpehr has a g = 2.053 and a coupling of ap = 25 Gis
resolved. (1S2) A value of g = 2.085 and poorly resolved fine structures are
found for HFe(diphos);ys3) The relative sharpness of the resonances is
important, particularly with iron compounds, because inadvertent oxidation
could give rise to a low spin d S complex. Since these d S complexes have
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 45

one unpaired electron and a g - 2, they might be confused with an


organometallic radical. These species, however, often have rapid relaxation
times at room temperature so that very broad resonances are observed. (1S3)
Variations in line width with temperature are observed for [Fez(CO)s]'
with g = 2.0385 and a width -3 G at -80° and -20 Gat 25 o .(1S4.1SS) The
spectrum of [Fe3(COh2]', g = 2.0016, is sharp at -80°C and clearly shows
- 3 G couplings to equivalent l3C nuclei.

1.5.1. Direct Observation of Radicals

The possibility of a radical path in a reaction is often suggested by an


unexpected racemization or an unusual reactivity order (see Section 1.5.4).
In an effort to prove that the mechanism involves radical participation, it
might appear that direct observation of the radicals by ESR would be the
method of choice. Generally, however, one wishes to demonstrate not only
the presence of radicals, but that the major path of the reaction proceeds
via a radical route. The extreme sensitivity of ESR may make detection
of radicals relatively straightforward, but making quantitative estimates to
prove that a radical is a loop species is another matter. Nevertheless, direct
ESR detection of an intermediate is a quite useful approach.
Treatment of cis-[Mo(COh(dmpeh] with Ph 3 CCl in the cavity of an
ESR spectrometer yielded a spectrum of trans- [Mo( CO)z(dmpe hJ", which
was identified by its characteristic values of g and apY SZ ) Metal-centered
radicals have also been observed in oxidative additions to
[Mo(Nzh (diphos h]( 1S6) and (7J S- CsHs)Fe(diphos )MgBr. (1S7) Organic
radical fragments have been detected in the reaction of isopropyl iodide
with [(7Js-CsHs)Fe(COhr in THF.(15S) In this case the characteristic septet
of doublets of the (CH3 hHC" radical was observed in a flow cell in which
the reactants were mixed at room temperature.
The failure to observe radicals by ESR does not, however, allow them
to be discounted from the mechanism. The flow experiments with i-PrI
and [(7Js-CsH5)Fe(COhr would suggest that [(7Js-CsHs)Fe(COhJ" should
have been detected. Presumably the very short lifetime of the radical
precludes its observation. In some situations, one might have an
organometallic anion acting solely as a one-electron donor and then
dimerizing rapidly enough that it is not observedys9) Unfortunately, owing
to the difficulties in quantifying the experiments, the observation of ESR
signals does not prove that radicals are involved on the main reaction path.

1.5.2. Radical Spin Traps

An alternative to the direct observation of the rather elusive radicals


themselves is to provide a substrate with which they will react and produce
46 C. A. TOLMAN AND J. W. FALLER

a radical that is stable indefinitely. This spin-trapping procedure has been


used primarily for organic radicals,(160-161) but has been extended to metal-
based organometallics by Lappert. (142) The technique usually involves a
nitroso compound, which reacts with the transient radical to form a nitroxide
(see Equation 52). The resultant nitroxide can then be characterized by

o
I
R-N=O+R' ...... N° (52)
R/ ""R'
the spin splittings of nuclei two or three bonds removed from the nitroxide
nitrogen. The most commonly used trap is 2-methyl-2-nitrosopropaneY62)
The spectra are simplified with this trap as the t-butyl protons in the
resulting trapped radicals do not couple to the electron. Furthermore, it
exists in solution primarily as the monomer, whereas many nitroso com-
pounds are unreactive dimers.(163) The major drawback is that it tends to
decompose by loss of NO forming (t-BuhNO at a rate sufficient to interfere
in the spectra of the trapped radical. To overcome this problem, other
traps such as nitrosodurene were developed. They often exist in solution
predominantly as the dimer, but sufficient monomer is usually present for
most applications(164); complications may result, however, if the equilibrium
is not established rapidly.
It is important to establish by appropriate control experiments that
the trapped radicals were produced exclusively by the reaction under
consideration. Lappert(143) has discussed these precautions, such as avoiding
the effects of exposure to light and making certain that the reagents
themselves do not react with the spin trap. The importance of these controls
is demonstrated by the study of the oxidative addition of tosyl halides to
PtO species, where it was found that tosyl iodide + R NO gave a clean ESR
. (143)
signal for [(MeC 6 H 4 S0 2 )(R)NO].

1.5.3. CIDNP

Chemically induced dynamic nuclear polarization (CIDNP) arises in


some radical pair processes and is characterized by intensity inversion and
enhancement of some of the resonances in the nmr spectra of radical
combination products during a reaction. (65 ) This provides definitive
evidence for the involvement of radicals, but as with ESR, it is difficult to
quantitatively estimate the fraction of the total reaction proceeding via the
radical path. The conditions under which a CIDNP effect will be observed
are rather limited, and it has only been observed in a few cases with
organometallics.(166) In particular, situations in which a metal-based radical
can be implicated are extremely rare.(67,167) Nevertheless, with quantitative
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 47

measurements of enhancement factors, convincing evidence can be pro-


vided for a reaction proceeding predominantly by a radical path.
Two cases of particular relevance are the CIDNP effects observed in
the oxidative addition of isopropyl iodide to Pt(PEt 3 h (167) and the reduction
of a-methylstyrene with HMn(CO)s:(67)
(CH 3 hCHI + Pt(PEt 3 h ---. propane + propene + 2,3-dimethylbutane
+ PtHI(PEt 3 h + PtI 2(PEt 3 h (53)

PhC(CH 3 )=CH 2 + 2HMn(CO)5 ---. PhCH(CH 3 h + Mn2(COho (54)

The i-PrI reaction produces the expected products of an isopropyl radical


reaction, and CIDNP enhancements and multiplet effects were observed in
the propene resonances and isopropyl iodide resonances. In the HMn(CO)s
reduction, enhancements were observed in the Mn-H resonance and in
the methyl resonance of the product.
As reaction (54) was studied quantitatively and is simpler, its results
will be used to illustrate the method. CIDNP effects can be interpreted in
terms of a radical pair mechanism. At some stage a geminate radical pair,
42, is formed and the competition among electronic relaxation times, the

PhC(Me)=CH 2 + HMn(COls ~ Ph(MehCC", ·Mn(COls (55)


42
42 cagee,cape. Ph(Me)zC"+·Mn(CO)5 (56)
back reaction to starting materials, and the escape of the radicals from the
solvent cage around the pair produces the unusual polarizations of the
protons observed in the reactants and products. Measurement of the relaxa-
tion times and the rates of decay of the intensities of the cIDNP-enhanced
resonances provides a measure of the absolute intensity enhancements.
The magnitude of the enhancements, - 300, supports the view that most
of the product is being formed by the radical path. The calculation of the
CIDNP effect involves the dephasing rate of the radical pair, which is related
to the relative g values of the radicals. Thus, one can obtain some estimate
of the value of g of one of the radicals, if the other is known. This provides
an additional tool for estimating the involvement of a metal-based radical.

1.5.4. Indirect Methods

Although loss of stereochemistry in a reaction might provide an indica-


tion that radical reactions are occurring, retention or inversion does not
necessarily preclude radical reactions. There are numerous examples in
. l'Iterature (168-170) were
t h e orgamc h i"
stereose . 0 bserved 10
ectlvlty IS . rad'lca1
reactions. For example, homolytic decomposition necessarily generates two
48 C. A. TOLMAN AND J. W. FALLER

radicals in close proximity. These fragments may diffuse away or react


within the solvent cage. If reactions take place before the radicals diffuse
away from each other, high stereoselectivity can be observed. The lifetime
of a radical pair is estimated to be -10- 11 s, and after escape from the
cage racemization usually occurs.
Inference of a radical path can also be made on the basis of structure
and reactivity arguments. For example, two-electron SN2 processes gen-
erally show a maximum rate in halide displacements from alkyls of Me >
Et > i- Pr > t - Bu. The reverse reactivity order is generally characteristic
of radicals. R I > RBr > R CI reactivity is also expected for radical reactions.
Initiation of polymerization of styrene or acrylonitrile is also an indicator
of the presence of radicals.
In addition to a radical pair process, a radical chain process may be
involved in some oxidative additions. (140,171)

pta + RO -+ Pt-R (57)

Pt-R + RX -+ X-Pt-R + RO (58)

In such cases, a chain initiation step is required; hence, if the reaction can
be initiated with light, AIBN, or benzoyl peroxide, a radical chain process
is indicated. Since the chain involves the continual presence of a radical
after initiation, inhibition of the reaction by a radical trap, such as galvanoxyl
or duroquinone, also provides evidence for a radical chain mechanism.
Because some radical traps react with organometallics, however, lack of
inhibition may be the result of deactivation of the trap.
An excellent measure of radical involvement for certain types of studies
relies upon the rearrangement rates of organic radicals that might be formed
under the reaction conditions.o 72 ) Thus, if a free 5-hexenyl radical is formed
in the oxidative addition of 5-hexenyl bromide, the product will contain a
cyclopentylmethyl group. The rearrangement of the radical occurs at a rate
of 105 s-t, so that if the radical has a lifetime of more than about 10-4 s,
it will rearrange. This technique for the demonstration of radical paths has
been used in a number of cases. (154.155) A similar approach using cyclo-
propylcarbinyl halides(156,173) (krearrangement - 108 S-l) suggests that
approximately 30% of the reaction follows a radical path with iodide.

(59)
+
30%
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 49

2. APPLICA TlONS ILLUSTRA TlNG THE METHODS

2.1. Butene Isomerization bV NiL4 and H2 S04

This system(10) nicely illustrates many of the principles outlined in the


first section. Most of the work was done using Ni[P(OEthJ4 or Ni[P(OMehJ4
as catalysts in methanol solvent. In the presence of a strong acid co catalyst,
such as H 2S0 4 or HCI (but not without cocatalyst), butenes can be isom-
erized within minutes at room temperature to an equilibrium mixture of
the linear isomers. This is probably the most active olefin isomerization
catalyst system known.
Using 1-butene (the least stable isomer, abbreviated 1-B) as the starting
olefin, the rate law (at the beginning of the reaction) is given by
Equation 60,

-d[l-B]/dt = kK[NiJo[H+][l-B]/(l + K[H+])[L] (60)

where [NiJo is the initial concentration of NiL 4, and k = 0.7


-1
S and
K = 50 M- 1 t at 25° for L = P(OEth.
A plot of k obs - 1 against [H+r 1 will give a straight line since

kOb,-1 = [L](1 + l/K[H+])/k[Ni]o (61)

The intercept on the ordinate as [H+r 1 -+ 0 ([H+J -+ (0) gives the maximum
rate attainable for fixed [LJ and [NiJo-when all the nickel is converted to
HNiL 4+.
Spectroscopic studies (vis/uv and 1H and 31 p nmr) of solutions under
catalytic olefin isomerization conditions show NiL4 and HNiL4 +, with more
of the latter as the acid concentration is increased. Gradually the catalyst
dies as these species are converted to Ni(II), in a reaction which is inverse
first order in [L]. In the absence of olefin, H2 is evolved, while in the
presence of butene substrate, butane is formed. The system gives about
300 catalytic cycles at 25° and about 3000 at 0°, indicating that catalyst
degradation has a higher activation energy than olefin isomerization.
In an experiment at room temperature using D 2S0 4 in CH 3 0D, 39%
of the original 1-butene was isomerized in 15 s. There was some d 1 -1-
butene (0.5%), but more than 99% of the 2-butene products were undeuter-
ated. The ratio of isomerization to deuteration is about 170. In spite of
this, we are confident that the reaction involves olefin insertion into a nickel
hydride to form a nickel alkyl intermediate, followed by de-insertion of

t We have chosen to use M to represent moll-I and mM to represent mmoll- I in hopes


that this usage will be more familiar to readers.
50 C. A. TOLMAN AND J. W. FALLER

NVE
18

18

16

18

16

Figure 2.2. The mechanism of butene isomerization by Ni[P(OEth]4 and H 2 S0 4 .

isomerized olefin. Additional supporting evidence is provided by the quanti-


tative analysis by microwave spectroscopy of various d I-propylene isomers
formed on exposure of do-propylene to solutions of Ni[P(OEthJ4 and DCl
in CH 3 OD. (74 ) The analysis indicates that insertion into the NiD bond
occurs with relative rates of 4: 1 to form DCH 2 CH(CH 3 )Ni and
CH 3 CHDCH 2 Ni intermediates. The presence of deuterium on C 2 in the
product propylene cannot be accounted for by a 1T'- allyl hydride mechanism.
On long exposure of I-butene to a similar solution, one obtains a
thermodynamic mixture of the isomers (about 6% I-butene, 25% cis-2-
butene, and 69% trans- 2- butene) along with statistical scrambling of the
deuterium originally present in solution with all of the hydrogens of the
C4 H s YO)
The mechanism shown in Figure 2.2 satisfactorily accounts for the
observations and is consistent with the 16- and 18-Electron Rule.(6) The
number of metal valence electrons (NVE) is written to the right of each
species.
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 51

Ignoring the back reaction of product 2-B (letting k-6 --+ 0) and apply-
ing the method of Cleland, (27) we obtain the rate law given in Equation
60, where K = K 1 and

(62)

where the rate constants are defined in Figure 2.2. Equation 62, clearly
shows the impossibility of determining each of the individual rate constants
from the overall rate law. We also cannot tell which of steps 3, 4, 5, or 6
is rate determining. In principle, a choice could be made from the deuterium
isotope effect. The more rapid isomerization in CH 30D suggests that olefin
insertion is rate determining (breaking an Ni-H bond and forming a stronger
C-H bond), but the situation is complicated by the more rapid degradation
of L which also occurs in CH 30D.
This is 'l system where none of the loop species is present in sufficient
concentration to be spectroscopically detectable. No alkyl is observed even
when C 2H 4 is added to the HNiL4 + solutions. With butadiene, however,
the additional stability provided by coordination of the double bond makes
the trihapto (7T-crotyl)NiL 3+ stable enough to be isolated. (175)
Spectroscopic and kinetic studies give considerable insight into the
sequence and rates of the individual steps of the isomerization mechanism.
Ni[P(OEthJ4 itself does not dissociate to the NiL3 complex to a detectable
extent, even at 70°. Using the very sensitive vis/uv method, Kd is estimated
to be less than 10- 10 M- 1 at 70°,(176) and must be considerably less than
this at 25°. The 31 p NMR singlet for Ni[P(OEthJ4 indicates that the complex
has the expected tetrahedral structure. Adding P(OEth gives a separate
sharp signal, showing that ligand exchange-in the absence of added acid-
is slow on the NMR time scale. Though there appear to be no X-ray
structures of NiL4 or NiL3 complexes, Pt(PF3)4(177) and Pt(PPh 3h(178) have
been done and shown to be tetrahedral and trigonal, respectively.
The half-life for dissociation of the first ligand from Ni[P(OEthJ4,
measured by capturing the NiL3 intermediate by isonitrile(179) or CO,(180)
is several hours at 25°-far too slow to allow ligand dissociation to be the
first step. Studies in the reaction of Ni[P(OEthJ4 with 0Iefins(180) show that
these reactions proceed at the same slow rates, and that highly activated
olefins (such as maleic anhydride) must be used before any olefin complexes
are detectable. The reaction of H+ with NiL 4, on the other hand, is so fast
that stop-flow techniques must be used. For reaction (63) at 25° using
HCI0 4, k1 = 1550 M- 1 s-t, k-1 = 45 S-1.(181) The calculated KJ of 35 M- 1
is in satisfactory agreement with the value of 48 ± 14 M-t, determined

(63)
52 C. A. TOLMAN AND J. W. FALLER

with room temperature equilibrium measurements using H 2S0 4, and the


K of 50 M- 1 in the rate equation 60 for butene isomerization.
IH and 31 p NMR spectra of HNiL4 + at all accessible temperatures
show that all of the P nuclei are equivalent on the NMR time scale, in
spite of the fact that HNiL4 + probably has a trigonal bipyramidal structure
in its lowest energy form. This rapid exchange of axial and equatorial
ligands, quite common for 5-coordinate complexes(182) must be part of the
reason for the catalytic activity of nickel, since it readily provides a mutually
cis orientation of the hydride and olefin in the HNiL 3(Olt complexes
suggested by the [Lr 1 dependence of the rate law.
While hydrido-olefin complexes are not observed in this system, a
related platinum complex trans-HPt(PEt3h(C2H4t is quite stableY83) In
this case, at least part of the stability is attributable to the mutually trans
orientation of Hand C2H 4; a cis orientation appears to be required for
insertion.
HNiL3 + is not detectable in solutions of HNiL4 +, and an upper limit
of 4 x 10- 5 M can be set on K 2 .(181)

(64)

The value of k2 can be determined to be 0.015 s -1 at 25° by trapping the


HNiL3 + with butadiene to form 1T-crotyl complexes. (175) Thus, the lower
limit on L2 is about 400M- 1 S-l. Comparison of k2 with kl shows that
protonating the NiL4 accelerates L dissociation by a factor of more than 102.
Figure 2.2 is oversimplified in that it does not show the formation of
cis-2-butene, the reaction of H+ with HNiL3 + to liberate H 2, Ni(II), and
L (the dominant path for nickel oxidation in the absence of substrate), nor
the acid catalyzed decomposition of P(OEth to HPO(OEt)z(181) which also
occurs.
While the kinetics of olefin isomerization have only been examined
with L = P(OEth and P(OMeh, the equilibrium constants for protonation
of NiL4 complexes in Table 1 indicate that K decreases rapidly as /I
increases, suggesting that NiL4 complexes with strongly electron withdraw-
ing ligands are likely to be poor catalysts. Ni[Ph 2PCH 2CH 2PPh 2J2 is also
expected to be a poor catalyst because of the difficulty in obtaining a
coordination site for olefin binding. It is interesting in this respect that
HNi[Ph 2PCH2CH 2PPh 2h + reacts only very slowly with butadiene (tl/2 =
8 h at 50° vs.less than 0.1 h for HNi[P(OEthJ4 + at 25°). HCo[P(OEthJ4
does not react with butadiene at allY84)
The HNiL4 + complexes are effective catalysts for isomerization of a
variety of olefins. 1,4-pentadiene is isomerized to the 1,3-isomer, which
then reacts rapidly and irreversibly with the nickel hydride to form 1,3-
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 53

Table 1. Equilibrium Constants for Protonation of NiL4


Complexes

Ligand K(M-')a v(cm-')b

PhzPCHzCHzPPhz 410.0 2066.7


PPh(OEth 107.0 2074.2
P(OEth 33.0 2076.3
P(OMeh 35.0 2079.5
P(OCHzCHzClh 1.2 2084.0
P(OCH zCCI 3 h <0.1 2091.7

" Measured in most cases at 0" in methanol using H 2 S0 4 , from C. A. Tolman, Inorg.
Chern.H,3128(1972).
b Carbonyl stretching frequencies used to characterize the electronic character of the

ligands. See reference 4.

dimethyl-1T-allyl nickel complexesY 12 ) Functionally substituted olefins,


such as methylpenteneoates, can be readily isomerized at 25° to an equi-
librium mixture of the isomers. (180)

2.2. Olefin Hvdrogenation with Wilkinson's Catalvst

2.2.1. Background and Rate Laws

This fascinating catalyst system, which has been studied in considerable


detail, (185) is based on RhC1(PPh 3 h and was apparently discovered indepen-
dently in Wilkinson's laboratory by J. A. Osborn(17) and at leI by R. S.
Cofiey;(186) nevertheless, Wilkinson's name has stuck. Ole fins can be
smoothly hydrogenated in minutes at room temperature with millimolar
Rh, and an H2 pressure of 1 atm. Hydrogenation rates increase with both
olefin concentration and H2 pressure as shown by plots of reciprocal rates
of cyc10hexene hydrogenation against 1/(cyc1ohexene) and 1/PHz in Figures
2.3A and 2.3B. The dependence of rate on catalyst concentration is com-
plex, and in the absence of added L is approximately half-order in
[Rh]. (12,187) The rate becomes first-order in [Rh] and nearly inverse first-
order in [L] at moderate (-0.01 M) concentrations of added PPh 3 • Several
rate laws advanced by various workers have been summarized by
Halpern(188) (Table 2), who reports that his Equation 65 (for cyc1ohexene)
is consistent with rates over wide ranges of concentrations, though he has
not published the experimental measurements of catalytic hydrogenations
under steady-state conditions.

-d[S]/dt = [Rh]tot{a + b[L]/[H 2 ] + c[L]/[S]}-1 (65)


54 c. A. TOLMAN AND J. W. FALLER

- I
0.5 1.0 1.5 2.0
1 I (cyene) mol- 1 l

15
B
I')
I
a
0
X
VI
10
b
-0
E

.,c
.,
~..,
- 5
..,
......

1.0 2.0 3.0 4.0 5.0


1/PH2 cm- 1 X 10 2

Figure 2.3. Reciprocal plots of the rate of hydrogenation of cycIohexene (abbreviated cyene)
by Wilkinson's catalyst in benzene at 25°: A, plotted against reciprocal olefin concentration
at O.625(a), 1.25(b) and 1.875(c) x 10- 3 M RhCl(PPh 3 h; B, plotted against reciprocal hydro-
gen pressure at 2.5(a) and 1.25(b) x 10- 3 M RhCI(PPh 3 h, from Reference 17.

Spectroscopic studies under catalytic olefin hydrogenation conditions


without added L show the presence of RhCIL 3 (43), H 2 RhCIL 3 (44),
(RhCIL 2 h (45), and H 2 (RhCIL 2 h (46). (In the case of ethylene
(C 2 H 4)RhCIL 2 (47) is also observed.)(14) The amount of the hydrides
increases with H2 pressure, while the dimeric species are increased by
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 55

Table 2. Rate Laws· for RhCI(PPh 3J3 Catalyzed Olefin Hydrogenation

No. Rate law Reference

Rate = [Rh]tot{a/[H 2 ][S] + b/[H 2 ] + c/[S]}-l 17


2 [Rh],ot{(a + b[L])/[H 2 ][S] + c/[H 2 ] + d/[S]}-l 193
3 [Rh]tot{a + (b + c[L])/[H 2 ][S] + (d + e[L])/[H 2 ] + (f + g[L])/[S]}-l 187
4 [Rh]tot{a + b[L]/[H 2 ] + c[L]/[S]}-l b 188
5 [Rh]tot{(a[S] + b[L])[L]/(d[S] + e[L])[S] + (e[S] + f[L])/[H 2 ]}-t c

a Taken from reference 118. Here [L] and [S] are the concentrations of PPh, and substrate olefin.
e For cyclohexene.
C For styrene, based on unpublished results of J. Halpern and T. Okamoto.

temperature and suppressed by added L. Structures 43 to 47 are established


by 31 p , I H, and 13 C NMR spectra of the L = PPh 3 and P(p-tolylh complexes
in solution(14) and by the X-ray crystal structures of RhCI(PPh 3 h and
(C 2 F 4)RhCI(PPh 3 h in the solid state.o 89 ) RhCI(PPh 3 h can crystallize in
two crystalline modifications with different orientations of the crowded
phenyl rings-a red a-form with a tetrahedral distortion in which the r.m.s.
deviation of the coordinated atoms from the mean coordination plane is
0.43 A, and an orange (3-form with an r.m.s. deviation of 0.28 AY90) A

L
L-I-H
I
cr//
Rh/"j
1'H
I
L

43 44

L
L-Cl-[-H
j "'i"1
Rh Rh
/ " /" I'
L--Cl--,H
L

45 46

47
56 C. A. TOLMAN AND J. W. FALLER

related, less crowded complex with a chelating tridentate ligand,


RhCl[Ph 2P(CH 2hPPh(CH 2hPPh 2], shows an r.m.s. deviation of only
0.08 A, and Rh-P bond lengths which are systematically shorter by 0.02 A
than those in RhCl(PPh 3 hY91,192)
The dimer itself is quite a good hydrogenation catalyst, though it is
not very soluble in the case of L = PPh 3 , and is subject to poisoning by
small amounts of adventitious O 2, Addition of small amounts of free L
have an accelerating affect on the hydrogenation rate, while larger amounts
inhibit with a [Lr 1 concentration dependence as shown in Figure 2.4 for
L = P(p-tolylh.
Isotopic labeling studies using D2 and l,4-dimethyicyciohexene show
that although the d 2-cyciohexane is the major product (about 90%)
significant amounts of do, db d 3 , and d 4 products can also be detected,
indicating stepwise addition and reversible formation of (alkyl)RhH inter-
mediates,(193) which are also indicated by the occurrence of olefin double-
bond migration under hydrogenation conditionsY94) Kinetic isotope effects
(kH/ k D ) in cyciohexene hydrogenation ranging from 0.90 to 1.17 have
been reported, but Siegel and Ohrt(195) have pointed out that the value
observed depends on the particular conditions of the experiment.

1.5
c
'E
]"
~ 1.0
~
Q.
;:)

N
:I: 0.5

2 3 4 5 6 7 8 9 10 ~O
RATIO ADDEO L: Rh

Figure 2.4. Rate of hydrogen uptake at ambient temperature and 1 atm by 100 ml 1 M
cyclohexene in toluene containing 1.3 x 10- 3 M Rh added as [RhCl[P(p-tolylhhlz and
various amounts of added P(p-tolylh, from Reference 14.
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 57

~-_2

Figure 2.5. The mechanism of cyC\ohexene (S) hydrogenation by Wilkinson's catalyst. L is


PPh 3 or P(p-tolyllJ.

2.2.2. The Mechanism

Considerable insight into the mechanism shown in Figure 2.5 has been
achieved by studying individual steps in the system; rate and equilibrium
constants measured so far are given in Table 3. Halpern and Wong(13) used
stop-flow techniques to measure the rapid reaction of H2 with RhCI(PPh 3h
to form H 2RhCIL 3-a reaction which is inhibited by L but becomes
independent of [H 2] at low [L], when k1 becomes rate determining. The
value of k1 = 0.718- 1 implies a half-life of L dissociation from RhCIL 3 of
1 s. RhCIL 2 is never present in significant quantities, but reacts so rapidly
with H2 that below ligand concentrations of about 0.15 M it provides the
major pathway for H2 absorption; at higher [L] the direct reaction of H2
with RhCIL 3 dominates(14) and k2 can be determined. While RhCI(PPh 3h
cannot be observed spectroscopically, the 14-electron complex
RhCI(PCY3h with bulkier ligands has been isolated.(12)
Recently de Croon and co-workers(18) explored the rate of cyclohexene
hydrogenation by RhCI(PPh 3h in benzene over a wide range of substrate
concentrations (0.2 to 8.2 M) and found that the rate increases with
increasing substrate concentration up to about 3 M then decreases by a
factor of 2 on increasing the cyclohexane concentration to 8.2 M! Their
explanation is that coordination of the benzene is important in stabilizing
RhCI(PPh 3h as RhCI(PPh3h(T/2-benzene). Very high substrate concentra-
tions decrease the concentration of available benzene. Thus RhCIL 2 in
Figure 2.5 should perhaps be written as RhCIL 2B. Note that the kinetic
measurements of Halpern and Wong were made in benzene, and that they
58 C. A. TOLMAN AND J. W. FALLER

Table 3. Values of Individual Rate and Equilibrium Constants Determined"


in Wilkinson's Hvdrogenation System

Reaction k; L; K; Combinations Reference

0.71 S-I <10- 5 M 197


2 4.8M- I S-I 9 X 10 3 M- I b 13, 14
(12.5 M- I S -I)b (40 X 10 3 M-I)b

3 500s- 1 «1 Me 14
4 >7 x 104 M- I S-I k 4 / L I = 0.9 13
5 K5K 3 = 3.4 X 10-4 d 197
6 0.20s- le L6/ k 7 0 . 1t 197
7 0"
8 <6 X 10- 6 M K~K8 = 3.3 X 10-4 Mh 14
(2.5 X 10-4 M)
9 5.4M- I S-I 14
(12.5 M- I S-I)b (5.5 X 10 3 M-I)b
10 KIKJO« 1;0.4forC2 H4 14
(<< 1)

(1.7) for C 2 H 4

a At 25°, with L = PPh 3 (or P(p-tolyl)3 in parentheses) in benzene, with S = cyclohexene, unless listed
otherwise. Reaction numbers refer to Figure 2.5, with forward reactions in the directions indicated by
arrows.
b Taken from reference 14 but converted units assuming an H2 solubility in benzene or toluene at 25° of
2.0 X 10- 3 M/atm.
, Rapid and reversible by nmr. See Figure 2.6.
d Both steps 5 and 3 were assumed to be rapid and reversible.
e This could be the slow step in the inner loop. in accord with the deuterium isotope effect.

f We estimate this value in order to account for the non-d 2 products when D2 is used .
• Step 7 is irreversible.
h K~ /Ks
2RhClL 3 ~ [RhClL 2 l2 + 2L.

would have no way to exclude this possibility, based on their data. One
way to check deCroon's hypothesis would be to use mesitylene or durene
as the solvent, rather than benzene. The methyl groups should block T/ 2
coordination, severely retarding the rate of H2 uptake.
The rapid and reversible nature of step 3 in Figure 2.5 is shown by
31 p NMR studies. Figure 2.6 shows spectra before and after addition of
H2 to RhCI(PPh 3 h. At the high [RhJ used, the dimer and free PPh 3 were
not observed. The absence of P-P coupling and Rh-P coupling for the
upfield P nucleus in the spectrum (B) at 30° after adding H2 shows that
the unique phosphine (the one trans to H in structure 44) undergoes rapid
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 59

30-

T
Figure 2.6. 31 p {lH} nmr spectra of 0.14 M RhCl(PPh 3 h in CH 2 Cl 2 at the temperatures
shown: A, before adding hydrogen; B, after adding hydrogen; and C, after sweeping nitrogen
through the solution, the arrow indicates the position of free PPh 3 • From Reference 196.
60 C. A. TOLMAN AND J. W. FALLER

tntermolecular exchange at this temperatureY96) A value for k3 at 25° of


500 M- 1 S-l can be estimated by line-shape analysis. A kinetic study of
the reaction of cyclohexene with preformed H 2RhCI(PPh 3 h serves to
support steps 3, 5, and 6.0 97 ) The value of k6 (and therefore also of Ks in
Reference 13) is questionable, however, since it depends on an extrapola-
tion of a plot of k obs - 1 against [L]/[S] at low concentrations of added L
(0.75 mM) where about 25% of the Rh must have been present as
H 2(RhClL 2h. which itself reacts rapidly with cyclohexeneY4)
In a system with enough added L present to keep dimeric species
insignificant (above about 0.01 M [L] dimers are less than 2% of the Rh
at 25°), a weakly binding olefin like cyclohexene, and 1 atm H 2, and
assuming that steps 1, 3, and 5 in the mechanism in Figure 2.5 are rapid
and reversible relative to the other steps, and that step 7 is irreversible,
one can use the treatment of Cleland(27) to derive the rate law in
Equation 66.
k- 1 = [Rh]tot/ Rate = [L]/(k 2 [L] + K 1 k 4 )[H2 ] + [L]/k~K3K5[S] (66)

where k~ = k6k7/(k7 + k-6)'


The two terms arise from the two principal species present under
catalytic cyclohexene hydrogenation conditions-RhClL 3 and H 2RhClL 3 ,
respectively. An additional constant term, like that shown in Rate Law
No.4 in Table 2.2, is not necessary unless a significant amount of another
intermediate [H 2RhCIL 2(S)] is also present in the system under conditions
where the rate law applies. t
With cyclohexene or 1-hexene and no dimeric species, the relative
amounts of RhCIL 3 and H 2RhCIL 3 depend on the rate at which RhClL 3
is converted to H 2RhClL 3 by H2 compared to the rate at which H 2RhClL 3
can be converted to RhClL 3 by substrate olefin, indicated schematically in
Figure 2.7. When the two concentrations are equal, the two reactions
contribute equally to the overall resistance of the system. The 1/ k I in
Figure 2.7 can be thought of as the system resistance (reciprocal of the
overall conductivity) and the two terms as the resistances of the two
reactions in series.
The complexity of the Wilkinson hydrogenation system emphasizes
the need to specify the ranges of variables over which a rate law applies,
and it illustrates the power of combining spectroscopic studies with kinetic
studies of the individual steps.
The rate law of Equation 66 becomes inadequate at low [L] concentra-
tions, when [RhCIL 2]2 and H 2[RhCIL 2h become significant, or in the
presence of ethylene, when (C 2H 4)RhClL 2 can be a major species. Each
t Equation (66) reduces to Rate Law 4 in Table 2 if a = 0 and [L] is small enough so that
most of the reaction goes through Step 4 in Figure 2.5 rather than through Step 2.
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 61

RhCLL3

1I k' = 11k'! + l/k~

Figure 2.7. A simplified kinetic mechanism for Wilkinson's catalyst with added L. Here k;
and k; are functions of the true rate constants k;, [S], [H 2 ], and [L]. (See Reference 27).

additional significant species will add a term to the rate law written in
reciprocal form.
The accelerating effect of adding small amounts of L to the dimer
(Figure 2.4) suggests a step such as reaction (67) in a more complete
mechanism, which might also include an additional catalytic cycle based
(67)

on dimer.(14) The question of whether an olefin route exists with ethylene,


with a path involving direct reaction of H2 with (C 2H 4)RhCIL 2, remains
to be resolved.

2.2.3. Extensions to Related Systems

Changing the phosphorus ligand has a marked effect on hydrogenation


rate, as seen in Table 4. Both electronic and steric effects are evident.
Putting an electron-donating methyl group in the para position of PPh3
increases the rate of I-hexene hydrogenation by 2.2. Independent studies
of the rate of reaction of H2 with RhCI[P(pC 6 H 4MehJ3 show that the rate
is 2.4 times as fast as with the PPh 3 analogY4) Thus, better donors increase
k ~ in Figure 2.7. While PEtPh 2 is a better donor than PPh3, the overall
hydrogenation rate decreases by more than a factor of 2. In this case k ~
probably increases while k ~ substantially decreases, owing no doubt to the
decreased equilibrium constantK3 for PEtPh 2 dissociation from H 2 RhCIL 3,
which is probably the major species in solution under those conditions. A
smaller value of K3 is attributable to the reduced steric crowding in the
H 2 RhCIL 3 complex relative to the case where L = PPh3. In the case of
P(oC 6 H 4 Meh the very slow overall reaction may be the result not of a
62 C. A. TOLMAN AND J. W. FALLER

Table 4. Phosphorus Ligand Dependence of


Relative Rates of Hvdrogenation of 1-Hexene bV
RhCIL/ Complexes

Ligand Rate (ml/min)b

P(pC 6H 4OMeh 99.5


P(pC 6H4Meh 85.3
PPh 3 38.9
PEtPh 2 17.5
P(pC 6H 4Clh 1.58
P(oC6~Meh 0.11
P(2, 4, 6-C6H2Me3) 0.09
P(OPhh 0.02

a Catalyst prepared in situ by the addition of L to [RhCl(eyclo-


oetene)]2'
b 1.25 mM Rh, 0.6 M 1-hexene in benzene, and 50 em Hg H2
pressure at 25°, from C. O'Connor and G. Wilkinson, Tetrahedron
Lett. 18, 1375 (1969).

small K 3 , but of a very small Ks-the eqailibrium constant for olefin


complexation to H 2RhClL 2. Indeed RhCI(PCY3h reacts readily with H2 to
form H 2RhCI(PCY3h; however, as a hydrogenation catalyst for cyclo-
hexene, it has only 1/40 the activity of a RhCI(PPh 3h under similar
conditions. (198) RhCI[Ph 2P(CH 2hPPh(CH 2hPPh 2] should be electronically
very similar to RhCI(PEtPh 2h. It forms a very stable dihydride at room
temperature, but does not act as a hydrogenation catalyst. (199) The reason
given by Dubois and Meek(200) is that the coordinatively saturated dihy-
dride, with structure 48 analogous to 44, is unable to dissociate the P trans
to H without breaking two chelate chains. The most active catalyst of the
Wilkinson type appears to be RhCI{PhP[N(CH 2)s]h. The turnover rate of
n-alkenes at 30° is 1.7 S -1 compared to 0.2 s -1 for RhCI(PPh 3h. (201)

48

Rates of olefin hydrogenation also depend on olefin structures, as seen


in Table 5. Steric inhibition is indicated by the much slower rate of
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 63

Table 5. Olefin Structure Dependence of Hydrogen Rates


Using RhCI(PPh 3J3

Olefin

Styrene 93.0
Cyclopentene 34.3
Cyclohexene 31.6
1-Hexene 29.1
cis- 2 -Pentene 23.2
4-Methyl-cis- 2- pentene 9.9
4-Methyl-trans- 2- pentene 1.8
1-MethyIcycIohexene 0.6

a 1.25 in M Rh in benzene at 25', from F. H. Jardine, J. A. Osborn, and G. Wilkinson,


1. Chern. Soc. (A), 1574 (1967).

I-methylcyclohexene compared to the similar rates of cyclohexene and


I-hexene. Styrene hydrogenates considerably more rapidly than either, but
Halpern has found a different rate law in this case and suggests a pathway
involving a H 2RhCIL(Sh intermediate.(188)
Hydrogenation rates also depend on the halogen in RhXL3 complexes,
with rates increasing in the series X = CI < Br < 1. (17) Both steric and
electronic factors are probably involved, but not enough is currently known
about the bromide and iodide systems to say more.
A theoretical study has been carried out on the olefin insertion reaction
(68) by Dedieu.(202)

(68)

It is found that the reaction is best described as an olefin insertion, rather


than a hydrogen migration, and that it is promoted by good 7T-donor ligands
in the coordination sphere. Note that from the rate law (66) one cannot
say whether the slowest step in the inner loop of Figure 2.5 is due to olefin
insertion k6 or reductive elimination k 7 •
The Co and Ir analogues of Wilkinson's catalyst are completely inac-
tive, but for quite different reasons. The Co complex does not react with
H 2,(203) while the Ir analog reacts irreversibly to give H 2IrCI(PPh 3h. This
I8-electron dihydride, however, is stable to ligand dissociation and thus
does not provide a site for olefin coordination. (204) Other iridium complexes
can, however, be extremely active.
64 C. A. TOLMAN AND J. W. FALLER

[Ir(COD)(PMePh2h]PF 6 in CH 2Clz at 0° will give 3800 cycles/hr


of cyclohexene hydrogenation compared to only 70 for RhCl(PPh 3h (in
benzene/ethanol).(205) It has been possible to identify the [H 2Ir(COD)
(PMePh 2h]PF 6 intermediate by 1H NMR at -800. (206)

2.3. The Nickel-Catalyzed Cyclooligomerization of Butadiene

2.3.1. Background

The nickel-catalyzed cyclooligomerization of butadiene (BD) is ()ne


of the most thoroughly studied homogeneous catalytic reactions.(207.20S)
Extensive 13 C NMR studies of intermediates and model compounds have
recently been reported. (209) It provides one of the best examples of the
ability of phosphorus ligands to control both rates and product distributions
in homogeneous catalysis and shows just how complex catalytic systems
can become.
In the absence of added phosphorus ligands, the principal product
(80-90%) is trans, trans, trans-1,5,9-cyclododecatriene (ttt-CDT 49), with
smaller amounts of trans, trans, cis-1,5,9-cyclododecatriene (ttc-CDT 50),
trans, cis, cis-1,5,9-cyclododecatriene (tcc-CDT 51), 1,5-cyclooctadiene
(COD 52), 4-vinylcyclohexene (VCH 53), and cis-1,2-divinylcyclobutane
(DVCB 54). With the addition of phosphorus ligands, especi:llly bulky
phosphites, COD can become 96% of the product, and the rate of butadiene
consumption can be increased by a factor of 10 (at 80°).(211)

~ 49
0 50
C8
51

0 52
U 53
cC 54

Ethylene can be readily cooligomerized with BD to produce cis,


trans-1,5-cyclododecadiene (55) in good yield/ 212) but propylene reacts
only poorly in this way and 2-butene not at all. The strained olefin norbor-
nene, however, gives 56 in high yield,(213) and dimethyl acetylene can be
used to prepare 57.(214)
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 65

55 56 57
A variety of nickel catalysts may be used, including Ni(II) with reducing
agents, (7T-C 3 H shNi, Ni(CODh, Ni(CH 2 =CHCNh, Ni(CO)4, and even
atomic Ni (by metal atom evaporation). t Phosphorus ligands may be added
separately or coordinated in an added nickel complex such as (CDT)NiL,
(COD)NiL 2 , or NiL 4.

2.3.2. Phosphorus Ligand Effects and Intermediates Involved

Rates and product distributions depend on L : Ni ratio, as well as ligand


type, temperature, and BD conversion. Table 6 shows the effect of varying
the ligand: Ni ratio for PPh 3 • The rate of CDT formation (sum of all isomers
49-51) is reduced from 25 cycles/hr:j: with no added PPh3 , to 4 at 1: 1,
and to less than 0.4 with PPh 3 : Ni of 2: 1 or greater. The maximum rate
of Cs products (COD + VCH) is realized at a 3: 1 ratio (not shown) but
decreases from 105 cycles/hr§ to 45 as the L : Ni ratio is increased to 8: 1.

Table 6. Dependence of Product Distribution a on L: Ni Ratio for PPh 3 at 80°

0:1 1:1 2:1 4: 1 8:1

%CDT 87.2 6.0 0.5 0.6


% COD
%VCH } 8.2
64.0
27.0
62.0
36.0
56.0
41.0
5.0
48.0
% >C 12 3.6 2.8 1.9 1.5 1.6
gBD/gNihr 75.0 180.0 185.0 165.0 80.0
Mol BD/mol Ni hr 85.0 200.0 205.0 185.0 90.0

a Data at high conversion, from Reference 211 except for 0: 1 data from reference 210.

t See references cited on p. 136 of Reference 207. Ni(CH 2 =CHCNlz might appear to be a
14-electron complex, but its insolubility in noncoordinating solvents suggests a polymeric
structure with nitrile bridges. Attempts (with L. J. Guggenberger) to determine a single
crystal X-ray structure of the solid failed because of disorder problems.
:j: The turnover'rate for CDT is one-third the moles of BD consumed per mol of Ni per hour,
times the fraction of BD going to CDT, or 1/3(85 x 0.872).
§ The turnover rate for cyclodimers is one-half the moles of BD consumed per mol of Ni per
hour, times the fraction of BD going to cyclodimers, or 1/2(865 x 0.991).
66 C. A. TOLMAN AND J. W. FALLER

Table 7. Dependence of Product Distribution a on L:Ni Ratio for


P(OoC6 H4 Ph)3 at BO°

0:1 0.5: 1 1:1 2:1 3:1

%CDT 87.2 0.4 0.2


% COD

%VCH
} 8.2
96.0

3.1
96.0

3.1
96.0

3.5
96.0

3.7
% >C 12 36.0 0.2 0.3 0.3
gBD/gNi hr 75.0 380.0 780.0 550.0 230.0
Mol BD/mol Ni hr 85.0 420.0 865.0 610.0 255.0

" Data at high conversions. from same sources as Table 6.

The ratio COD /VCH decreases from 2.4 to 1.0 as the PPh 3 : Ni ratio is
increased from 1: 1 to 8: 1-but actually increases again to nearly 2.0 at
100: l.(215)
Table 7 shows the effect of varying the L : Ni ratio for P(OoC 6 H 4 Phh.
In this table, the selectivity to COD is very high at all ratios of 1: 1 or
greater, but the high rate of COD formation (415 cycles/hr) decreases as
more L is added above 1 : 1. With the phosphite, the L is so strongly bonded
to Ni that the trimerization reaction is effectively cut off. The at first puzzling
result that CDT is only a minor product when the L : Ni ratio is 0.5: 1
(when only half the nickel can be tied up by L) is explained by the fact
that the other half is tied up as Ni(CODh; at the temperature of the
experiment and in the presence of excess COD, the BD cannot effectively
compete for coordination. Note that the BD consumption rate in Table 7
in this case is only half the rate at a 1: 1 ratio. COD is a more severe
inhibitor of cyclotrimerization than is CDT, because Ni(CODh is consider-
ably more stable than Ni(CDT). (216)

Table B. Dependence of Product Distribution a on Conversion, with 1: 1


P(OoC6 H4 Ph)3: Ni at 20 0b

% Conversion: 4 8 55 85 95 100

% COD 55.0 56.0 59.0 60.0 84.0 98.0


%VCH 8.7 5.3 1.9 1.8 2.2 2.0
% DVCB 36.0 38.0 39.0 38.0 14.0 0.0

a In 50% benzene, 50% butadiene, from Reference 211. Less than 1% of CDT was observed.
h The rate was about 7 g BD/g Ni (4 mol dimer/mol Ni hr). Less than 1% was >C B •
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 67

Divinylcyclobutane (DVCB) is also formed in these systems as a


kinetically controlled product (as much as 40% of the cyclodimers) but is
isomerized to the more thermodynamically stable COD and VCH. The
isomerization rate is enhanced by higher temperatures and inhibited by
BD, so that isomerization becomes significant under low-temperature
catalytic conditions only at high conversions, as seen in Table 8. (The
results in Tables 6 and 7 were under high conversion conditions, where
DVCB was no longer present.)
The cis orientation of the vinyl groups in DVCB is consistent with its
formation by coupling of carbons C 3 and C 6 in an intermediate of structure
58a (or its anti, anti-isomer, 58b, below).

4 5

If?
2~
(69)
1
'-~iiI6
I 8
L L

58a 59

This cis (head-to-head) orientation of '7T-allyl groups has been established


by X-ray crystallography for ('7T-C3HshNiPMe3, which retains its bis-'7T-allyl
structure in solution, as shown by NMR and Raman spectra.(217) Reversal
of rxn (69) by oxidative addition of Ni(O) in 59 can regenerate the bis-'7T-allyl.
The inhibition of DVCB decomposition by excess BD can be attributed
to a competition of BD and DVCB for coordination to the available Ni(O).
Coupling of BD in rxn (70) provides a pathway to 58, and, by further
coupling of C 1 and Cs, to COD.

i!/ I
L
(70)

58b

Since the C1-C s coupling in 58 fixes the stereochemistry of the double


bonds, the fact that they are both cis in COD suggests that the CH 2
substituents on the '7T-allyl groups, as in 58b, are anti with respect to the
hydrogens on C 2 and C7 at the moment of coupling. NMR studies(217) on
solutions of (713-1MeC3H4hNiPMe3, however, show that the only isomer
observed in solution has both methyl groups syn.
With sufficient steric crowding (and L = phosphine), both '111 and
713-allyls can be observed in solution, for example with 60 and 61.(217)
68 C. A. TOLMAN AND J. W. FALLER

Coupling of C 1 and C s in 61 provides a likely path to VCH, the major C s


product with this ligand. Structure 62, the analog of 61 produced by isoprene
cyclodimerization, has been established by X-ray diffraction.(219)

T(
Ni
Me)P /~

60
D
Cy)p/
Ni

61
#
b
Cy)p/
N'
1

62
#

In the absence of added phosphorus ligand, BD can take the place of


L in intermediate 58. In that case, however, in addition to coupling C 1
and Cs (or C3 and C s), there is the possibility of coupling C 1 with a terminal
carbon of the coordinated butadiene. This leads to a new bis-1T-allyl complex
63, an intermediate in the cyclotrimerization of BD which can be isolated
and is stable at low temperature in the absence of free BD or L. (210) Recent
13C CH] and 1H NMR studies(219) show that 63 exists as two isomers in
solution, both with coordinated trans-olefinic double bonds. The major
one is assigned structure 64, where the double bond is parallel to the planes

63 64

defined by the 1T-allyl groups, while the minor one is obtained by rotating
the double bond 90° about the Ni-olefin bond. A structure analogous to
64 has been found by X-ray diffraction for the RuClz analog 65.(220)

65
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 69

In the absence of a 2-electron donor, such as the double bond in 64


or the L in 58, bis-7T-allyl nickel complexes are found with the 7T-allyl
groups in nearly parallel planes and trans to each other (tail-to-tail), as
shown by the X-ray structures of (T/3 -2MeC3~hNi (66)(221) and
(T/ 3-cyclooctatrienylhNi. (219) In toluene-d 8 solution (T/ 3-C3H6hNi and (T/ 3_

}-
Ni

~
66

2MeC3~hNi are found by 13C NMR to exist as a mixture of cis and trans
isomers-30% of one and 70% of the other.(219)
With addition of a 2-electron donor to 63 at room temperature,
coupling of the terminal carbons occurs to give CDT. With one PEt3 per
nickel, (CDT)NiPEt3 can be isolated.(2lQ) Excess phosphine displaces CDT
from Ni. The X-ray structure of (ttt-CDT)Ni shows the expected ruffled
symmetrical 3 -coordinate structure. (222)
The presence of a trans double bond in trimerization intermediate 63
explains why the observed cyclotrimers 49-51 all contain at least one trans
double bond. The two cis double bonds in 51 or the cis and second trans
bonds in 50 must form during the final C 12 ring closure-implying that the
precursors to 51 and 50 have 7T-allyl groups whose substituents (the chain)
on C 3 and ClQ are anti, anti or anti, syn, respectively. (The possibility that
BD couples with the terminal carbon of an T/ 1-allyl containing a cis double
bond (as 61) is regarded as less likely.) The greater thermodynamic stability
of syn -7T-allyls accounts for the predominance of the ttt-CDT among the
trim eric products.
Cyclooligomerization reactions of d o- and d 6 -BD mixtures have been
carried out and found to give do, d 6 , d 12 , or d 18 CDTs (49-51) and do, d 6 ,
and d 12 cyclodimers (52-54) in nearly statistical ratios-the small deviations
from statistic being attributable to secondary isotope effects. (223) This work
confirms the idea that the mechanism involves C-C-bond formation without
breaking or making C-H bonds.

2.3.3. Ligand-Concentration Control Maps and Mechanisms

Schenkluhn and co_workers(208,215,224) have made extensive studies of


cyclooligomer concentrations as a function of L: Ni ratios. By plotting
percentages of various products against log (LINi), they get "titration
curves,,(224) or "ligand-concentration control maps,,(215) like that shown in
70 C. A. TOLMAN AND J. W. FALLER

100 .....
A

80 ttt-COT

VCH
60
0/0

40
COD

20
tcc- COT

0 ....L-
-co -5 -4 -3 -2 -1 0 2 3
log (LI Ni 1

100 -0-0

80 o
B
60

40

20 b

o
-co -5 -4 -3 -2 -1 o 2 3
log (LlNi 1
Figure 2.8. Ligand control charts or titration curves for butadiene oligomerization by nickel
with L = PPh 3 : A, product distribution among cydotrimers and cydodimers; B, distribution
among the cydotrimers. From Reference 215.

Figure 2.8. it is fascinating to see with either pyridine(224) or PPh 3 (215) that
increasing L: Ni above about 10-4 starts shutting off the formation of
tcc-eDT and that above 10-2 (L at 1% of the Ni concentration) this trimer
is essentially gone. Increasing L : Ni over this range increases the ttt-eDT
but has virtually no effect on the percentage of the ttc-isomer! The explana-
tion of the disappearance of the tcc-isomer is that the added ligand decreases
the steady-state concentration of an anti, anti-bis -7T-allyl e 12 intermediate
to a more thermodynamically stable syn, syn-bis-7T-allyl by accelerating
the rate of isomerization. Kinetic preference for an anti-7T-allyl complex,
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 71

which subsequently isomerizes to a more stable syn isomer, is also seen in


the reaction of BD with RNiL; [L = P(OEth]' The equilibrium constant
Equation 71 is about 20 at 25°.0 75 )

rr~~
NiL3+ NiL3+
(71)

The virtual elimination of eDT between log (L/Ni) = -2 and 0


indicates that the major catalytic trimerization intermediate (63) changes
to one (58) which contains one L per Ni. The fact that COD reaches a
maximum near log (L/Ni) = 0 and then declines at higher L/Ni ratios in
favor of VCR shows that VCR is formed by an intermediate (67) which
has two Ls.

67

The switch from COD to VCR can be understood in terms of the 16-
and 18-Electron Rule. Coupling in an intermediate like 58 (to COD or
DVCB) is allowed, because it converts an 18-electron complex into a 16.
A 16-electron intermediate like 61 would drop to 14 electrons if coupling
occurred in the TJ \ TJ 3 form. On the other hand 67 is an 18-electron
complex and should couple much more rapidly, giving VCR. This probably
accounts for why 62 can be isolated whereas a complex like 67 has not
been reported-though we have identified the analogous (syn-TJ 3_
1MeC3 R 4 )NiL 2CN intermediate in the hydrocyanation of BD.(225) At very
high concentrations of some phosphorus ligands an 18-electron TJ 1, TJ 1
intermediate 68 can become important, as we shall see, providing an
additional route to COD.

68

The thermodynamic stability of the 18-electron complex 64 (in the absence


of added donors) may be the result of the unfavorable trans orientation
across the nickel of the terminal carbon atoms. Addition of another ligand
probably displaces the trans double bond and allows the complex to
rearrange to the more favorable 69, which then leads to ttt-CDT.
72 C. A. TOLMAN AND J. W. FALLER

Ii I
L

69

With a catalytic system as complex as this, no one has attempted to


write a rate law. Using the 16- and 18-Electron Rule one can, however,
write a plausible mechanism for cyclotrimerization, shown in Figure 2.9.
Here S is used to represent substrate BD, bound as a simple olefin (7j2)
when attached to Ni. A bis-7T-allyl intermediate with an 8-carbon chain
(as in 58) is written as 7j 6 -Cs and the one with a twelve carbon chain (as
in 69) as 7j 6 -C 12 . No attempt is made to show the detailed stereochemistries
of the intermediates or the paths leading to isomeric cyclotrimers.
The catalytic cycle in Figure 2.9 takes up three molecules of substrate
BD and produces one CDT. Step 1, in which the first two BDs are coupled,
is a type of oxidative addition, since the oxidation state, the number of
coordinated atoms, and the number of valence electrons all increase by
2. (6) Coupling of the third BD in step 2, a type of olefin insertion reaction,
fixes the stereochemistry of the first formed double bond in the C 12 chain
[the 7j 2 in Ni( 7j 3, 7j 2, 7j 1 -e 12 )] trans. Step 3 gives the 18-electron complex

Figure 2.9. Cyclotrimerization of butadiene (S) to 1, 5, 9-cyclododecatriene (CDT). The


multiplicity of Ni-carbon bonds is shown by the superscripts on 1). C g and C 12 show the
lengths of carbon chains.
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 73

Figure 2.10. Cyciodimerization of butadiene (S) to cyclooctadiene (COD) and vinyl-


cyciohexene (YCH) in the absence of added phosphorus ligand L.

63. Note that this is one of the few examples in this chapter where a
catalytic loop species b.as been sufficiently stable to isolate. In step 4, the
double bond dissociates to provide a coordination position for S in step 5.
Step 6 is the final irreversible ring closure. Steps 7-10 then accomplish the
stepwise replacement of CDT by S to regenerate S3Ni. Reversible steps
11 and 12 give the isoluble Ni(CDT) complex. It is clear that Ni(CDT),
Ni(1/s-Cd, or (1/3-C3H shNi [an analog of Ni(1/6-C12)] can be used as
catalysts for the cyclotrimerization.
In addition to adding a third BD to the chain, SNi(1/ 6-Cs) can undergo
C1-Cs ring closure to give COD, or go to a 16-electron 1/\ 1/3 complex,
written SNi(1/ 4-CS ) in Figure 2.10, where the left-hand loop produces COD
and the right, VCH. Steps 12-14 show how Ni(CODh can be used as a
catalyst, and how the concentration of loop species can be decreased at
high COD concentrations (See Table 7 and the associated discussion.)
In the presence of added phosphorus ligand, loops exactly like
those in Figure 2.8 are possible, except that one S in each loop species is
replaced by L (Figure 2.11). In addition, there is a path (steps 12-17),
where ring closure to VCH occurs in a complex with two Ls (step 13). This
additional path can be expected to contribute at higher L concentrations.
At very high [L] yet another path (steps 18-22 in Figure 2.12) could
74 C. A. TOLMAN AND J. W. FALLER

Figure 2.11. Cyclodimerization of butadiene in the presence of added L.

become important, in which COD forms (step 20) by reductive elimination


in a dialkyl intermediate like 68.
The effect of these multiple cycles on product distribution is seen in
the distribution of dimers, plotted against log (L/Ni) for L = PPh 3 and
PPh(OPhh in Figure 2.13. At very low ligand concentrations, the dimer
distribution (about 48% COD and 42% VCH) is determined primarily by
the ligand-free cycles in Figure 2.10. At log (L/Ni) between -2 and 0, the
NiL cycles (Steps 1-11 in Figure 2.11) dominate. Even with only 1 % of
MECHANISTIC STUDIES OF CA TAL YTiC REACTIONS 75

100 -,-
A Nilo NilS
80

60
°4 a,~__~~~
40 b---~!!tI-.n..~

20
co---~~~~~~~~~
o -'-
-CD -5 -4 -3 -2 -I o 2 3
10Q (LlNi)

100 -,-
B
80

60
°'0 a
40 be:::==~~~
20

o ~
-CD -5 -4 -3 -2 -1 o 2 3
lOQ (l' Ni)

Figure 2.13. Product distribution among the dimers for: A, L = PPh; and B, L = PPh(OPhh:
COD(a), YCH(b), and linear octatrienes(c). The NiL n show the regions where different nickel
species dominate the dimer distribution. From References 208 and 215.

the Ni in the NiL cycles, most of the COD and VCH are produced here,
both because of the more rapid C g ring closure induced by L compared to
S and because of a reduced inhibiting effect of COD in the NiL system.
The crossovers in Figure 2.13 at about log (L/Ni) = 0.5 indicated that
at higher L : Ni ratios more VCH is forming by step 13 in Figure 2.11 than
by step 9. The second crossover in Figure 2.13B with L = PPh(OPhh at
log (L/Ni) = 0.9 indicates that at higher L: Ni ratios more reaction is going
through step 20 in Figure 2.12 than through step 13-once again making
COD the major cyclodimer. That a second crossover does not occur with
PPh 3 but does with PPh(OPhh is probably a consequence of the large cone
angle of PPh 3 ((J = 145°), which precludes a favorable equilibrium to form
68, which contains three Ls.
76 C. A. TOLMAN AND J. W. FALLER

NiLO

NiL4
Figure 2.14. A complete mechanism for butadiene cyclooligomerization at all L : Ni ratios.

The various cycles shown in Figures 2.9-2.12 can be combined to


show the entire BD cyclooligomerization system in a three-dimensional
representation in Figure 2.14. The reactions are shown in different planes
(NiLo, NiLb ... NiL 4 ), according to the number of coordinated phosphorus
ligands on each complex in that plane. Species in the NiLl and NiL2 planes
are the same as in the NiLo, except for replacing one or two Ss by Ls.
Some of the reactions connecting planes are shown. Others, indicated in
Equation (72), where 01 represents an olefinic double bond, no doubt
occur also. Adding L shifts the steady-state distribution of
-01 +L -01
(01)4Ni ~ (OlhNi ~ (OlhNiL ~ (OljzNiL
+01 -L +01

+L +L -01 +L
~ (Ol)NiL 2 ~ (Ol)NiL3 ~ NiL3 ~ NiL4 (72)
-L -L +01 -L
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 77

Table 9. Examples· of (Ol)xNiLy Species

NVE Compd Example Reference

18 (OI)4Ni Ni(COD)z 216


16 (OlhNi (C 2H 4hNi
18 (OlhNiL (CDT)NiPEt3 216
16 (OI)zNiL (CH 2=CHCN)zNiPPh 3
18 (0l)zNiL2 (1- Hexene )zNi[P(Oo- tolylhJ2 233
d
16 (OI)NiL 2 (C214)Ni[P(Oo-tolylhJ2
18 (0l)NiL 3 (C2F)4Ni(Ph2PCH2CH2)3CMe
16 NiL3 Ni[P(Oo- tolylhJ3
18 NiL4 Ni[PMe3J4

a For further examples and for l3C NMR data on several of these, see reference 209, Table IV, p. 270.
b K. Fischer, K. Jonas, and G. Wilke, Angew. Chern. Int. Ed. Engl. 12, 565 (1973).
C G. N. Schrauzer, J. Arner. Chern. Soc. 81, 5310 (1959).
d L. J. Guggenberger, Inarg. Chern. 12,499 (1973).

, J. Browning and B. R. Penfold, J. Chern. Soc., Chern. Carnrnun., 198 (1973).


f L. W. Gosser and C. A. Tolman, Inarg. Chern. 9, 2350 (1970).
g c. A. Tolman, J. Arner. Chern. Soc. 92:2956 (1970).

nickel species to lower and lower planes and eventually cuts off the BD
reactions.
Examples are known of every type of species shown in Equation 72,
beginning with Ni(CODh in Table 9 as an example of (Ol)4Ni. All known
(Ol)xNiLy complexes have 16- or 18-valence electrons.
Calorimetric studies(226) of the reaction of Ni(CODh with phosphorus
ligands show why the coupling reactions are cut off, particularly for small
L. With excess P(OMeh (0 = 107°) COD is quantitatively displaced to give
NiL 4 , in a reaction that is exothermic by 51 kcal/mol. The average Ni-P
bond strength in Ni[P(OMehJ4 (37.5 kcal) is stronger than the average
Ni-Ol bond in Ni(CODh (24.6) by 13 kcal/mol. With the bulkier ligand
P(Oo-C 6 H 4 Meh (8 = 141°), the final product is (COD)NiL 2 (LlH =
-20 kcal/mol), while with P(Oo-C 6 H 4 -t-Buh (8 = 175°), no reaction with
Ni(CODh is observed.

2.3.4. Steric and Electronic Control Surfaces

Both steric and electronic properties of the ligands are very important
in determining product distributions, as seen in the dimer distributions in
Figure 2.13. Figure 2.15 shows the percentage of CDT at 1: 1 L : Ni in a
large series of experiments, (215) where the response surface is plotted as a
78 c. A. TOLMAN AND J. W. FALLER

@[deg] __ 160

Figure 2.15. The response surface showing the effects of varying ligand steric and electronic
character on the percentage of cyclotrimers in butadiene cyclooligomers for L/Ni = 2. The
lowest point on the surface represents the maximum cyclodimer formation. From Reference
215.

function of fJ and X. The minimum CDT (max COD) is found at large X


(Electronegative L favors reductive elimination and C s ring closure) and
moderate fJ. With very large ligands much of the nickel may remain in the
NiLo plane, while with very small ones "disproportionation" to NiLo and
NiL2 planes may occur. By analyzing the surface, Schenkluhn and co-
workers(21S) have concluded that the control of oligomer distribution in
Figure 2.15 is 75% steric and 25% electronic.
The first application of multiple regression analysis to define a response
surface as a function of fJ and X was reported by Schenkluhn and co-workers
using data from a calorimetric study of the reaction of 22 phosphorus
ligands with [(1,3-Mer1/3 -C 3H 4 )NiMeh at 0°. In that case, the heat evolved
in forming (1T-allyl)NiMeL complexes was found to be 40% steric and 60%
electronic, the greatest exotherms being given by small electronegative
ligands. (227)t

2.3.5. Further Extensions

The cyclotrimerization of BD seems to be peculiar to nickel. Thus,


while biS-1T-allyl nickel is an excellent catalyst for producing CDT, the Pd
analog gives only linear n-dodecatetraenes and higher oligomers; bis-1T-allyl

t References 228-230 are recommended for those interested in reading further.


MECHANISTIC STUDIES OF CATAL YTfC REACTIONS 79

Pt shows no reaction with BD up to the temperature of its decomposition


with precipitation of Pt metal.(231) The absence of catalytic activity with Pt
does not appear to be the result of very different Pt-Ol bond strengths
compared to Ni. Calorimetric studies of the reaction of excess P(OMeh
with Pt(CODh gave an exotherm of 51 kcaljmole(232)-the same value
found using Ni(CODh.(226) The formation of linear oligomers does, of
course, require transfer of a hydride at one point in the mechanism.
N-octatrienes form in the nickel system (Figure 2.13B), but they are minor
by -prod ucts.
Additional ring compounds can be prepared using substituted dienes,
or by cooligomerizing olefins or acetylenes with dienes. One intriguing result
involves the stereochemistry of the DVCB derivative (70) prepared from
cis-piperylene and a nickel catalyst containing P(Oo-C 6H 4Phh. With trans-
piperylene only 71 and 72 were isolated. t The regioselectivity-in par-
ticular the absence of head-to-head coupling-suggests that prior to coup-
ling at least one of the olefins is coordinated to Ni by its least-substituted

70 71

double bond. This is in accord with measurements of equilibrium constants


for reactions of ole fins with Ni[P(Oo-tolylhh, where substitution of a methyl
group on the terminal carbon of a I-alkene decreases K in Equation 73
K
01 + NiL3 ~ (01)NiL 2 + L (73)

at 25° by a factor of 200.(233) The stereoselectivity shown in products 70-72


is not in accord with a simple 2 + 2 cycloaddition reaction, but is consistent
with application of the Woodward-Hoffmann rules to a stepwise process
in which one C-C bond forms at a time.(218,234) Formation of DVCB and
its methyl-substituted derivatives from dienes is reversible at ambient
temperature. With trans-piperylene, for example, the K for Equation 74
is 2.3 x 10-2 M- 1 at 30°, giving a maximum conversion to 72 of only 20%
in neat diene:(234)

2 trans-piperylene ~
K:W ~ (74)

72

Cooligomerization of propylene with butadiene gives two derivatives


of cis, trans-cyclodecatriene: 76% 73 and 24% 74.
t See p, 150 in reference 207 and references cited therein.
80 C. A. TOLMAN AND J. W. FALLER

73 74

The relative amounts can be explained in terms of the interaction between


the LUMO of propylene and the HOMO on the 7T-allyl group to which it
couples to give intermediates like 75 and 76.(235)

~ 76

While propylene gives primarily 73, styrene gives exclusively the phenyl-
substituted derivative corresponding to 74. It is noteworthy in this connec-
tion that hydrocyanation of propylene (using a Ni[P(OotolylhJ3 catalyst)
gives primarily n-butyronitrile, while styrene gives predominantly the
branched product 2-phenyl-propionitrile. (225.236)
Notice that 75 and 76 contain a trans-double bond and are strictly
analogous to the Ni( 1/ 3, 1/ 2, 1/ 1 -C 12 ) intermediate for cyclotrimerization
shown in Figure 2.9. The lack of dependence of the ratio 73: 74 on the
type and concentration of added ligand indicates that the coupling occurs
in an intermediate (S)Ni(1/6 -Cg ) in which propylene takes the place of BD
as S.
An intermediate analogous to 75 (or 76) has been isolated from the
reaction of (CDT)NiPPh 3 with butadiene and the diethylester of acety-
lenedicarboxylic acid. Structure 77 has been established by assignment of
the lH NMR spectrum. On treatment with CO at -20°, 78 is produced
along with Ni(COk(237) Note that the syn-7T-allyl in 77 must isomerize to
anti in order to give the unsubstituted cis double bond observed in the
product.

+ Ni(CO)4 (75)

As the reader may imagine, the co-oligomerization of various dienes,


ole fins, and acetylenes can be elaborated to prepare a vast number of ring
compounds, some including heteroatoms. References 208 and 234 are
recommended to the interested reader.
MECHANISTIC STUDIES OF CA TAL YTiC REACTIONS 81

2.4. Rhodium-Catalyzed Olefin Hydroformylation

2.4.1. Background and Rate Law

Hydroformylation, shown in Equation 76 for a terminal olefin, involves


the addition of hydrogen and carbon monoxide to produce

(76)

an aldehyde and is exothermic by about 28 kcal/mol.I239a) The formal


addition of a hydrogen atom and a formyl group across the double bond
provides the origin of the name "hydroformylation." It has also been called
the "Oxo" synthesis. Since its discovery using cobalt catalysts by Rolen (238 )
in 1938, the reaction has been very thoroughly studied with both Co and
Rh catalysts. A review by Comils in Falbe's recent book (239 ) contains nearly
2000 references, and has an extensive discussion of industrial processes
and applications. An excellent recent review has also been published by
Pruett. 1240l Hydroformylation is now one of the largest-scale processes
based on homogeneous catalysis by transition metal complexes. O )
Industrially, the rhodium-catalyzed hydroformylation is normally
operated at about 100°, at pressures up to 50 atm; and in the presence of
a large excess of added phosphorus ligand, it can be carried out in molten
PPh 3. Under these conditions, a terminal olefin can be converted in over
90% yield to linear aldehyde. By-products include branched aldehydes as
well as small amounts of alkanes and isomerized ole fins. Advantages over
the more conventional cobalt catalysts include lower temperatures and
pressures, higher ratios of linear to branched products, and less hydrogena-
tion of aldehyde products to alcohois.1239b)
As in other homogeneous catalyst systems, many different catalysts
can be used. These include Rh metal on carbon,1241l Rh4(CO)17.1242)
Rh 20 3 ,12421 RhCl 3 · 3H 20,12431 (acac)Rh(C012,12441 RhCI(CO)(PPh 312,1243)
and HRh(CO)(PPh 3h. ,L43) In cases where one starts with a higher oxidation
state of Rh, reduction by H2 and CO takes place under reaction conditions.
An induction period may be observed. In the case of RhCI(CO)(PPh 3h.
it can be eliminated by adding a base to the system which removes HCI
in reactions like (77)."'1

With HRh(CO)(PPh 3 h or (acac)Rh(COh, hydroformylation can be carried


out directly at ambient temperature and pressure with no induction

.;. Evans and coworkers'2431 write this reaction without CO to produce HRh(CO)(PPh 3 )2;
however, this species has never been directly observed.
82 C. A. TOLMAN AND J. W. FALLER

period.(243.244) Spectroscopic studies show that (acac)Rh(COh is rapidly


converted to HRh(CO)(PPh3h in the presence of PPh 3, H 2, and CO. (245)
An empirical rate law has been written in the form,

d[aldehydelldt = k[Sj'[Rhy[H 2 l/[COl (78)

where x can be less than 1.0 and depends on the substrate olefin and y
can be less than 1.0. (239c) While CO inhibits the reaction at high pressures
(>40 atm), it can accelerate the reaction at low pressures,<242) and, of
course, is one of the reactants necessary for hydroformylation-illustrating
the limited range of applicability of Equation 78. A similar rate expression
for cobalt has been proposed for high pressures (100 atm), where the rate
is linear in [Co], [S], and [COr 1 .(242)

2.4.2. Spectroscopic Studies

High-temperature and pressure IR studies with the original phosphine-


free cobalt system under conditions for catalytic hydroformylation of cyclo-
hexene (150°, 250 atm) normally show CO 2(CO)s and HCO(CO)4 as the
only spectroscopically detectable species, with the hydride as the major
species. With l-octene or ethylene, the hydride is not observed and the
only detectable species are CO 2(CO)s and RCOCo(COk (246.247) With
excess PBU3 (190°, 80 atm), the dimer species CO 2 (COhL and C02(CO)~2
and hydrides HCo(COhL and HCo(COhL 2 are observed, and there is no
evidence for the alkyl or acyl complexes, RCo(COhL or RCOCo(COhL,
which can be prepared by other routes. (248)
In recent work with ethylene and a phosphine-free rhodium system
using a high-pressure IR cell, King and co-workers have assigned bands
at 2115 m, 2037 s, and 2019 s cm- 1 to C 2H 5Rh(CO)4(249) and bands at
2075 wand 2026 s cm -1 (in n -tetradecane) to HRh(COh(C 2H 4). (250) While
these assignments must be regarded as tentative in the absence of confirming
NMR data, they are very interesting. Many of the experiments were done
at 35° and less than 100 atm; the new species are quite unstable and
decompose on reducing the pressure to 1 atm or removing the solvent. The
studies show that equilibria (79) and (80) lie to the right under the conditions

(79)

(80)

used, in the absence of ethylene, but that the clusters are effectively broken
down to monomeric species by the addition of ethylene. HRh(CO)4 has
only recently been unambiguously characterized, using FTIR.(251) Unfortu-
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 83

nately, there appear to be no spectroscopic studies involving hydroformyla-


tion of other olefins, t or using phosphorus ligands with rhodium under the
catalytic conditions of high pressures, where the rate laws have been
determined. Attempts were made by Wilkinson and co-workers(252) to
determine steady-state concentrations of rhodium species during hydrofor-
mylation at near ambient conditions, using an IR cell and a pump to
circulate solution from the reaction vessel, but the equipment was not good
enough to get useful results. Extensive spectroscopic studies employing
both IR and 1H nmr at ambient conditions using HRh(CO)(PPh 3h have,
however, been made by Wilkinson and co_workers(243,252-254) and provide
considerable insight into individual steps of the mechanism. 31 p and 13C
nmr studies have been reported by Kastrup and co_workers(244,245,45) at
various temperatures with solutions of HRh(CO)(PPh3h containing
added PPh 3, H 2 , CO, ole fins, and chelating diphosphines.
The 31 p spectrum of a solution containing HRh(CO)(PPh3h and added
PPh 3 in a 1 : 6 ratio shows separate slightly broadened resonances of complex
and free ligand at 25° in a 1: 2 intensity ratio, the former appearing as a
doublet due to Rh-P coupling. On heating, the resonances broaden and
at 60° the Rh-P coupling is lost. At 105° a single broad resonance is seen
about 2/3 of the way between the original resonances of HRh(CO)(PPh3h
and PPh 3-consistent with rapid exchange and a very small Kd for
Equation 81 at this temperature.

HRh(CO)(PPh 3h ¢.:.. HRh(CO)(PPh3h + PPh3


K
(81)

The fact that the shape of the HRh(CO)(PPh 3h doublet is independent


of the concentration of added PPh 3 shows that the ligand-exchange process
is indeed dissociative. Fitting of the line shapes gives kd = 100 s - at 25°
and Ed = 20 ± 1 kcal/mol.(45)
The IH nmr spectrum of HRh(CO)(PPh 3 h shows a single broad line,t
which sharpens to the expected quartet of doublets on cooling to -30°.
The trigonal bipyramidal structure indicated in solution by nmr is also
found in the solid state by X_ray.(255) In solution the IR shows only two
bands around 2000 cm- 1 independent of temperature from -45 to +30°:
a stronger one at 2000 cm -1 assigned to v Rh-H and a weaker one at 1920

t King and Tanaka(2S1a) have identified a new species RRh(CO)4 on treatment of 7)3_
C3HsRh(COh with CO at higher pressure; R is either CH 2=CHCH2 - or
CH 2=CHCH 2CO - .
t The observation of only a single broad line without resolvable H-P coupling in the IH
spectrum, at the same temperature at which the 31 p spectrum shows P-Rh coupling, is a
consequence of the fact that J(H-P)« J(P-Rh),
84 C. A. TOLMAN AND J. W. FALLER

assigned to vco;t the higher frequency band is not seen in the RhD
analog.(245) The simple spectrum is consistent with no measurable dissoci-
ation of the complex by Equation 81. Evans and co-workers(321) proposed
extensive phosphine dissociation to HRh(CO)(PPh 3h and even to
HRh(CO)PPh 3 (a 14-electron complex) based on low-molecular weights;+
however, measuring reliable molecular weights is notoriously difficult,
especially with air-sensitive compounds, and extensive dissociation is not
supported by more recent spectroscopic studies. (244.245,45)
On bubbling CO through a solution of HRh(CO)L 3 (L = PPh 3) equi-
librium (82) is rapidly established, the yellow solution becoming pale yellow.

HRh(CO)L 3 + co ;::!: HRh(COhL 2 + L (82)

With further bubbling for about 30 minutes, the color changes to


yellow orange via rxn (83), the CO serving to sweep out H 2.

2 HRh(CO)zL2 ;::!: H2 + [Rh(CO)zL 2l 2 (83)

Reintroducing H2 reforms the hydride. Under 1 atm H2 and CO in a 1: 1


ratio, the equilibrium is heavily on the left-hand side. This can be compared
with reaction (79) in the absence of phosphine where reduced electron
density on the metal favors H210ss. HRh(COh(PPh 3h has not been isolated,
since it is stable only in an atmosphere of H2 and CO. The Ir analog is,
however, well-known, and its X-ray structure has been done.(258)
On bubbling N2 or oxygen through solutions of the yellow dimer, CO
is lost, and red compounds of composition Rh(CO)(PPh 3h(solvent) can
be isolated. These are presumed to be dimeric, showing bridging carbonyl
bands in the IR, but in the absence of other evidence could contain Rh4
or higher clusters. The red compounds absorb H2 in the presence of added
PPh 3, and HRh(CO)(PPh 3h is re_formed.(254)
Treatment of HRh(CO)L 3 with ordinary olefins (even ethylene at
40 atm) does not give detectable concentrations of alkyls (or acyls). C2 F4,
however, gives HCF 2 CF 2 Rh(CO)L 2 , which is stable enough to be isolated
as a crystalline solid; with H2 and CO (or CO alone) HCF 2 CF 2 Rh(COhL 2
and HCF 2 CF 2 Rh(COhL are formed, but no acyl complexes.(252) Allene
and butadiene give stable (7T-allyl)Rh(CO)L 2 complexes, which do not
react with H2 or CO at 25° and 1 atm. (259) In the absence of dienes, normal
t The strange relative intensities are due to kinetic coupling of the two stretching modes,
possible when the H and CO are mutually trans. The coupling also causes 'vca' to increase
from 1920 to 1960 cm-1(254) on replacing H by D.(256)
:j: By believing the MW data, the authors were forced to assume that the various HRh(CO)-
(PPh 3 )n complexes (n = 1 to 3) have accidentally similar IR spectra. We have found
differences of about 10 cm -1 in VCN in HNiL n CN complexes, with higher frequencies in
n =2thaninn =3.12571
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 85

ole fins are isomerized. In one elegant experiment,(Z43) DRh(CO)L 3 in


benzene was treated with 1-pentene. The high field resonance of
HRh(CO)L 3 grew in with a 20 s half-life at 25°, which increased to 45 min
when PPh 3 (added L: Rh = 10: 1) was added. Isomerization of 1-pentene
to 2-pentenes also occurred in these solutions, but was much slower than
exchange (in the absence of added ligand t1/Z > 1 hr), and was still slower
with ligand added (13% isomerization in 1 day). These experiments are
consistent with L dissociation from the 18-electron HRh(CO)L 3 as the first
step preceding olefin coordination in both exchange and isomerization
reactions, which proceed through low concentrations of alkyl rhodium
intermediates. The fact that exchange is about 200 times faster than
isomerization indicates that the linear alkyl RRh(CO)L z in Figure 2.16
forms about 200 times as often (by reversible steps 2 and 6) as loop cycling
which passes through the branched alkyl R'Rh(CO)Lz formed in step 3. S
and S' represent 1-alkene and 2-alkene (cis and trans), respectively. Note
the similarity of Figures 2.16 and 2.2. In the rhodium system, the activity
is gradually lost as an orange compound proposed to be [Rh(CO)(PPh 3hJz
forms, even in the absence of added olefin.
On treatment of HRh(CO)(PPh3h with Hz (even at 30 atm), one does
not observe a reaction. (Z60) Though apparently no H3RhL3 complexes are

R'Rh{COlL2 RRh{COlL2

Figure 2.16. A mechanism for isotope exchange (beginning with DRh(CO)L 3 ) and isomeriz-
ation of terminal olefin (S) to internal olefin (S').
86 c. A. TOLMAN AND J. W. FALLER

R'Rh(CO)L2 HRh(CO)L2(S) RRh(CO)L2


~~
Figure 2.17. A mechanism which includes terminal olefin (S) isomerization and hydrogena-
tion, and conversion of HRh(CO)L 3 to DRh(CO)L 3 (using D2). Rand R' are linear and
branched alkyls.

known, t(262) several of the type H3IrL'L2 are, including H3Ir(CO)-


(PEt 2 Ph)z.(263) With D 2, however, DRh(CO)(PPh3h forms rapidly from
HRh(CO)(PPh 3h, with a half-life of about 2 minutes(254)-presumably via
a HRhD 2(CO)L 2 intermediate. Analogous RRhH 2 (CO)L 2 complexes are
believed to be intermediates in the hydrogenation of l-alkenes catalyzed
by HRh(CO)L 3 • (260)
Figure 2.17 shows a mechanism which includes paths for isomerization
of I-alkene (via steps 2-5), hydrogenation (2 and 6-8), and deuterium
exchange (9), using HRh(CO)L 3 • Of the species shown, only HRh(CO)L 3
is sufficiently stable to be used as the catalyst. (The dimeric complexes
described earlier could also be used but are not shown in the interest of
simplicity.) Detailed studies by Yagupsky and Wilkinson(264) show that
isomerization and hydrogenation of I-pentene occur at about the same
rate (3 turnovers/min at 27° and 1 atm H2 pressure). Both reactions are
suppressed by adding L,t which will decrease the concentrations of species
in both loops. The dependence of rate on catalyst concentration in the
absence of added L is approximately half-order,§ consistent with a larger
fraction of Rh in the loops (via equilibrium 1 in Figure 2.17) as
t H3RhL2 complexes have been prepared with the very bulky electron donating ligands PCY3,
PPh(tBuh, and P(tBuh.(261)
*Though the experiment has not yet been reported, we can anticipate that adding L will also
suppress formation of DRh(CO)L 3 for HRh(CO)L 3 and D z.
§ See Figure 2 in Reference 264.
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 87

the concentration is decreased. In the presence of sufficient added L,


the rate should become first-order in [Rh], since the ratio of
[HRh(CO)L 2]/[HRh(CO)L 3 ] will be fixed by [L] in Equation 6, indepen-
dent of [Rh].
Figure 2.17 shows oxidative addition of Hz only to the linear alkyl
complex RRh(CO)L z. Hydrogenation via R'Rh(CO)L2 could also occur,
but must be very minor because of the high selectivity for hydrogenation
of terminal alkenes. The selectivity could be due to an unfavorable oxidative
addition of H2 to R'Rh(CO)L 2 because of steric crowding in the branched
alkyl complex, t but a more important factor may be a very small steady-
state ratio of R'Rh: RRh due to the lower stability of the more crowded
complex.
With added CO, but not in its absence, ethylene reacts rapidly with
HRh(CO)L 3 solutions to give the acyl complex EtCORh(CO)zL 2.(Z52) With
1 atm of H2 and CO (in a 2: 1 ratio) propionaldehyde forms with a half-life
of 5 minutes. Using a 1: 2 mixture of H2 and CO instead of 2 : 1 increases
the half-life to an hour, suggesting that CO dissociation from the i8-electron
acyl complex is necessary before oxidative addition of H 2. Isomerization
and hydrogenation are much slower under hydroformylation conditions
(with H2 and CO) than under H2 alone. These reactions are typically only
1-2% of the hydroformylation rate. The reason must be high rates of
capture of 16-electron alkyl rhodium complexes by CO compared to rates
of H2 capture or {3 -hydride abstraction to form isomerized hydrido-olefin
complexes.

2.4.3. Catalytic Hydroformylation with HRh(CO)(PPh 3 h

With olefins, CO, and H2 catalytic hydroformylation takes place even


at 25° and subatmospheric pressure. Rates and product distributions depend
on substrate type, [S], [H2], [CO], ligand type, [L], [Rh], and temperature.
Rates with selected olefins are given in Table 10. Note that 2-pentenes
react about 25 times slower than 1-pentene, and that 2-methyl-l-pentene
(a hindered terminal olefin) is slower still. Cyclooctene is much faster than
cyclohexene, presumably because of ring strain effects on olefin coordina-
tion.t Butadiene reacts rapidly with the catalyst to form an inert ('IT-
crotyl)Rh(CO)L z complex and no gas uptake occurs at 25°. 1,5-Hexadiene
can be successfully hydroformylated, because the hydroformylation rate
(to primarily linear dialdehyde) is fast compared to the rate of isomerization
t Decreasing rate and equilibrium constants for oxidative addition of Hz and HCl to
IrCl(CO)Lz complexes with increased steric crowding by L have been reported by Vaska
and coworkers. (265)
t The equilibrium constant for olefin coordination to Ni(O) displacing P(Ootolylh is larger for
cyclooctene than for cyclohexene by a factor of 180.(233)
88 C. A. TOLMAN AND J. W. FALLER

Table 10. Hydroformylation Rates· of Various Olefins Using HRh(CO)(PPh 3J3


as Catalyst at 25°

Gas uptake Turnover rate b


Substtate (ml/min) (cycles/hrJ

Allyl alcohol 7.05 4.5


Styrene 4.32 2.8
1,5-Hexadiene 4.26 2.7
1-Pentene 3.74 2.4
Allyl cyanide 3.27 2.4
Ethylene 4.55 2.4 c
1-Hexene 3.52 2.3
1-Dodecene 3.18 2.0
Cyclooctene 0.26 0.17
2-Pentene(cis & trans) 0.15 0.10
2- Methyl-1-pentene 0.06 0.04
Cyclohexene <0.05 <0.03
1,3-Butadiene 0 Od

C 2 F4 0 0'

a With 2.5 mM HRh(CO)(PPh 3)3' 50 cm Hg gas pressure of H2 and CO (1 : 1), and usually 1 M substrate
in 50 ml benzene. With ethylene 1 : 1 : 1 mixtures of C2H 4 , H2 and CO were used at 60 cm total pressure.
From reference 253, except as noted otherwise.
b Turnover rate = gas uptake (ml/min) x 50 cm Hg x 0.5 x 60 min/hr· [24.5 ml atm/m mol x 760 cm Hg/

atm x 2.5 m mol/I x 0.050 I] = gas uptake x 0.64 except for C2H 4 .
'For ethylene, turnover rate = gas uptake (ml/min) x 60 cm Hg x 0.33 x 60 min/hr· [24.5 ml atm/
m mol x 760 cm Hg/atm x 2.5 m mol/I x 0.0501] = gas uptake x 0.52.
d (,.,3-IMeC 3 H 4 )Rh(CO)PPh 3 h forms, but no hydroformylation occurs at 25'.
, HCF2CF 2Rh(CO)(PPh 3 )2 forms on reaction of C 2 F 4 with the HRh(CO)L 3 complex. Under H, and
CO no hydroformylation occurs, but HCF2CF2 Rh(CO),L 2 and HCF2CF2Rh(CO)3L can be observed.

of the double bonds into conjugation. The fact that the turnover rate for
ethylene is not significantly greater than for many of the other ole fins is
no doubt because the ethylene concentration in solution in these experi-
ments was only that in equilibrium with 20 cm Hg C2 H 4 gas, while the
concentration of liquid ole fins was 1 M. Styrene is unusual in giving both
a high-reaction rate and a predominance of branched aldehyde product.
(Normal/branched ratios observed are in the range of 0.1 to 0.5.)(253) The
reader may recall (Section 2.3.5) that unusual regioselectivities for C-C
coupling at the carbon adjacent to the phenyl ring have been observed in
both the cooligomerization of styrene and butadiene (to give solely structure
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 89

74, with Ph in place of Me) and in styrene hydrocyanation (to give primarily
branched 2-phenyl-propionitrile). We attribute this effect to an enhanced
steady-state ratio 80: 79 in these systems relative to 82: 81.

~M ~M ~M
79 80 81 82
The higher ratio is achieved by TJ 2 electron donation from the benzene
ring to the otherwise 16-electron alkyl complex. The resulting TJ2 bonding
is, however, not nearly as complete as in a normal 1T-allyl, because of the
benzene resonance stabilization effect.
The high rate of hydroformylation of allyl alcohol in Table 10 could
also involve some stabilization of intermediate alkyl complexes by TJ 2
electron donation from the oxygen.
The effects of various reaction variables on rates and product distribu-
tions using RRh(CO)(PPh 3h and I-hexene near ambient temperature and
pressure have been reported by Wilkinson and co-workers, (253) and using
I-butene at higher temperatures (90-160°) and pressures (25-50 atm) by
Kastrup and co-workers;(244) results of both groups are given in Table 11.
The four products observed are linear aldehyde (R CRO), branched aldehyde
(R 'CRO), alkane (R R), and cis and trans -2-alkene (S'). The designations

Table 11. Effects of Reaction Variables on Rates and Product Distributions


in Hydroformylations Using HRh(CO)PPh 3 ) /

Variables Rate RCHO/R'CHO (RH + S')/~ Products RH/S'

[H 2 ] + + + +
[CO] -(+)
[S] + 0
T ++ ±d(_) + 0(+)
[Rh] + +(0)
[L] + - - (0) 0

a Using I-hexene in benzene,(253) or I-butene in 2-ethylhexylacetate (244 ) (in parentheses if different).


For the meanings of the ++, +, 0, -, and - - entries, see the text. Abbreviations: [H2 ] and [CO],
concentrations (or pressures) of H2 and CO; [S], concentration of olefin substrate; T, temperature; [Rh],
initial catalyst concentration; [L] concentration of ligand PPh 3 ; RCHOjR'CHO, ratio of normal to
branched aldehyde; RH, alkane; S', cis and trans-2-alkene.
b In this absence of added CO no appreciable hydroformylation occurs, but isomerization and hydrogenation

can be fast.
, Not reported.
d The RCHOj R'CHO ratio can increase or decrease with temperature. I25 "
90 C. A. TOLMAN AND J. W. FALLER

Figure 2.18. A mechanism for hydroformylation of a I-alkene, involving only RhL2 species.
RCHO and R'CHO are linear and branched aldehyde products.

+ + to - - express both the direction and magnitude of the effects, with 0


meaning no significant effect observed. The differences in the two studies
emphasize the dependence of some of the effects on the region explored.
For example, increasing the [Rh] does not affect product linearity(254) with
a large excess of added L, but does increase linearity without added L. (244)
Figure 2.18 shows a mechanism for the formation of the aldehyde
products which is consistent with the spectroscopic experiments described
earlier and with the 16- and 18-Electron Rule. Steps are numbered starting
from HRh(CohL 2 in step 1, both because this is the principal species in
solution under 1 atm H2 and CO (1: 1) before addition of olefin, and for
the sake of simplicity, one could equally well begin with the dimer because
of the reversibility of step 15. Other dimers, tetramers, and HRh(CO)nL 4 - n
complexes may be present under some conditions. The rapid formation of
RCORh(COhL 2 by steps 1-5 and 13 (or of R'CORh(COhL 2 by similar
steps) observed when I-hexene is added to a HRh(COhL 2solution suggests
that these steps are fast, and that steps 6 and 11 are rate determiningt
This accounts for the positive effect of [H2] on rate in Table 11. The
inhibiting effect of [CO] (at low Peo) is attributable to the large equilibrium
constants K13 and K14 (and also to the small values of K 1), which cause a
decrease in a concentration of loop species as [CO] is increased. Though
CO is taken up in steps 4 and 9, these reactions are so fact that they do
t The alternate possIbility, that 7 and 12 are rate determining but that equilibrium constants
K6 and Kll are small, cannot be excluded.
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 91

not affect the overall rate. Very rapid capture of (alkyl)Rh(CO)L 2 by CO


is consistent with the severe inhibiting effect that [CO] has on the rates of
hydrogenation and isomerization, which pass through the same intermedi-
ates (Figure 2.17). Competition of H2 and CO for these 16 electron alkyls
explains why hydrogenation can become faster than hydroformylation at
high H2 : CO ratios. (254)
Isomerization probably occurs primarily via R'Rh(CO)L 2 prior to
capture by CO in step 9. After R'CORh(COhL 2 forms by further steps
10 and 14, getting back to R'Rh(CO)L 2 by the reverse of steps 14, 10,
and 9 seems unlikely, especially under CO and H2 pressure. This conclusion
is supported by isotopic-labeling studies by PinoY66) If steps 9 and 4 are
essentially irreversible, then the R CHO I R 'CHO ratio will primarily
depend on the steady-state concentration ratio of RRh(CO)L 2 to
R'Rh(CO)L 2. If k4 and k9 are similart, then k3, k~, and the relative
stabilities (K3IKg) of RRh(CO)L2 and R'Rh(CO)L 2 will determine the
aldehyde product distribution. Steric crowding in the branched alkyl com-
plex is no doubt a major factor in destabilizing it compared to the linear
alkyl as early recognized by Evans, Osborn, and Wilkinson;(243) the
increased crowding of phosphines relative to carbonyls accounts for the
increased yields of desired linear products in both the Rh and Co hydrofor-
mylation systems. Electronic factors no doubt also playa role, and in fact
most dominate the product distribution with styrene, where R'Rh(CO)L 2
capture by CO is favored over R Rh(CO)L 2. The question is, to what extent
is the ratio [R Rh(CO)L 2]/[R 'Rh(CO)L 2] kinetically or thermodynamically
controlled under steady-state conditions?
Treatment of HRh(CO)L 3 with styrene and CO (without H 2 ) gave a
mixture of acyl complexes R CORh(COhL 2 (>70%) and R 'CORh(COhL 2
«30%), which could be distinguished by their lH nmr spectra. (254) Bleeding
H2 into the nmr tube gave a mixture of aldehydes in a similar ratio. The
RCHOIR'CHO ratio of >2 in the nmr experiment was much larger than
observed under any catalytic hydroformylation conditions, where the
maximum value with styrene was about 0.6. This indicates that under the
conditions of stepwise addition in the nmr experiments, where the alkyls
and acyls presumably had time to equilibrate, the R Rhl R 'Rh and
R CORhl R 'CORh ratios were higher than at steady state. Thus with styrene
K3 > Kg but kg > k 3 •
Increasing I-hexene concentration has no effect on the product ratio
(Table 11), as expected from Figures 2.18 and 2.17, since all four products
depend on the same step 2. The overall rate does, however, depend on

t The rate constants for steps 4 and 9 are probably very similar. Reduced steric crowding
should slightly favor k4 over k9, while a slightly higher electron density on the metal in
R'Rh(CO)L 2 should favor k 9 • The same can be said for k6 and k l l .
92 C. A. TOLMAN AND J. W. FALLER

[S). Lineweaver-Burke plots(24) of rate -1 against [Sr 1 give straight lines


with positive intercepts on the y -axis, indicating that at very high olefin
concentrations the rate becomes independent of [S], and the composition
at the transition state of the rate-determining step (CTSRDS, Section 1.1)
contains the olefin (as acyl complexes). At low [S] the rate becomes
proportional to [S], and the CTSRDS does not contain olefin, suggesting
that the RDS changes under these conditions to steps 1 or 2.
Not surprisingly, the total rate is strongly temperature dependent, the
rate of 1-hexene consumption increasing by about a factor of 10 on
increasing the temperature from 15° to 30°. t This corresponds to an
activation energy of about 25 kcal/mol. Studies by the Exxon group(244)
gave a similar activation energy for I-butene hydroformylation near 100°,
but the very small effect of temperature between 145° and 160° corresponds
to an apparent activation energy of less than 5 kcal/mole, indicating a
different rate-determining step.+ The fraction of (RH + S')/~ products
increases with temperature, indicating a higher activation energy for hydro-
genation and isomerization than for hydroformylation, but this undesirable
effect can be largely overcome by increasing [L]. Good results can be
achieved by operating near 100° in molten PPh3 • (254) The temperature
effect on R CHO / R 'CHO was positive with 1-hexene under catalytic hydro-
formylation conditions, but negative under stepwise reaction conditions.
Figure 2.18 is drawn with two phosphorus ligands in each one of the
complexes, in what we shall call the RhL2 plane, by analogy with the NiL2
plane of Section 2.3.3. Similar figures could be drawn for RhLo, RhL 2,
and RhL3 planes, which differ from Figure 2.18 only in the number of L
(and CO) ligands in each complex, keeping the total electron count the
same for complexes similarly located in the different planes. Figure 2.19
shows a three-dimensional view of the four planes stacked one above the
other. There are no catalytic cycles in the RhL4 plane, since it contains
only 18-electron species: RCORhL 4 , RRhL 4 , HRhL 4, R'RhL 4 , and
R 'CORhL 4 (from left to right). The figure is oversimplified in not showing
dimer or cluster species, or the paths for olefin isomerization and hydro-
genation indicated in Figure 2.17. It does, however, serve to explain many
of the remaining effects in Table 11, as well as the effects of changing
phosphorus ligands.
Increasing the concentration of HRh(CO)(PPh 3h (in the absence of
excess added L) increases the hydroformylation rate, but in a nonlinear
way, indicating that the turnover rate is greater at higher dilution; the
R CHO / R 'CHO ratio is, however, lower in dilute solutions. This is readily
.;. See Figure 6 in reference 253.
*Great care must be exercised with rapid reactions involving gases to be sure that the reaction
rate is not mass transfer limited. A simple test for chemical rate control is to show that the
rate does not increase with increasing stirring speed.
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 93

RhLO

CO>
CO>

Figure 2.19. A mechanism for rhodium catalyzed hydroformylation at all L : Rh ratios.

explained, since higher dilution will tend to favor species with fewer Ls as
CO and L compete for coordination, i.e., reaction will tend to shift from
the RhL2 to the RhLl and RhLo planes at high dilution. The turnover
rates are higher in the higher planes, but the product distribution is worse
because of diminished steric crowding as L is replaced by CO. In the RhLo
plane-as in a phosphine-free system using Rh 2(CO)s or Rh 4 (COh2 as
catalysts- the product ratio RCHO/R'CHO is typically less than one.
(See Table 12 below.)
94 C. A. TOLMAN AND J. W. FALLER

Table 12. Effects of Olefin Structure on Hydroformylations· Using


AcacRh(CO)2 in 1 M PPh 3 at 1450

[Rh] K/[Rht
Olefin (mM) (min-I M- 1 ) Percent linear aldehyde

CH 3 CH=CH 2 1.0 91.0 88.1


CH 3 CH 2 CH=CH 2 0.25 200.0 91.8
CH 3 (CH 2 hCH=CH 2 0.25 260.0 91.4
CH 3 (CH 2 hCH=CH 2 0.25 416.0 91.8
CH 3 CHCH 2 CH=CH 2 0.6 240.0 90.5
I
CH 3
CH 3 CH 2 CHCH=CH 2 0.5 284.0 95.5
I
CH 3
CH 3 CH 2 C=CH 2 5.0 6.0 100.0
I
CH 3
CH 3 CH=CHCH 3 5.0 5.6 41.9
a Twenty percent by weight olefin in 2-ethylhexylacetate under 5 : 1 H 2 : CO at 25 atm. From Reference 244.
b Observed first-order rate constant for CO (not olefin) consumption divided by the rhodium concentration.

AddingL has effects which are the opposite of diluting catalyst; shifting
more of the reaction to lower planes reduces the turnover rate, but reduces
hydrogenation and isomerization rates even more, while increasing the
RCHOjR'CHO ratio. The negative effect of [CO] on product isomer ratio
can be understood in terms of competition between CO and L for coordi-
nation.
It is unlikely, for ligands as large as PPh 3, that much of the rhodium
will be found under hydroformylation conditions in the RhL3 or RhL4
planes of Figure 2.19. Compounds in these planes, such as HRh(CO)-
(PPh 3h and HRh(PPh3)4, can, however, be used as catalysts. In the presence
of CO, even in a large excess of PPh 3, HRh(PPh 3 )4 no doubt rapidly loses
PPh 3 in favor of CO. The strain energy in this highly crowded molecule is
indicated by the severe distortion of the three equatorial phosphines of
the trigonal bipyramid away from the axial phosphine to give a nearly
tetrahedral arrangement of P atoms in the X-ray crystal structure.(255)
The importance of substrate steric effects on hydroformylation rate
and product linearity is shown by the results in Table 12. There is essentially
no difference in product linearity for the normal alkenes longer than
propylene, or for 4-methyl-1-pentene; the linearity with propylene is a
little lower due to the slightly reduced crowding in its R'Rh intermediate.
(The lower rate with propylene is no doubt due to the reduced solubility
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 95

of this olefin under the high-temperature conditions of the experiment.)


With 3-methyl-1-pentene the additional crowding significantly increases
the product linearity but does not affect the overall rate. The more crowded
2-methly-1-butene reacts very slowly but gives explusively linear aldehyde.
2-Butene reacts very slowly and gives a poor product distribution because
some of the R 'Rh intermediate is trapped by CO; some goes on to isomerize
to 1-butene (detected at the end of the run)(244) and give n-aldehyde.

2.4.4. Ligand Steric and Electronic Effects

With phosphorus ligands smaller than PPh 3 the steric strain will be
less, as will the tendency to be replaced by CO. Studies of the competition
between CO and phosphorus ligands for coordination to Ni(O), starting
with Ni(CO)4, showed that the degree of substitution of CO by L decreases
linearly with increasing cone angle fJ.(4) PPh3 (fJ = 145°) was unable to
displace more than two COs. Smaller ligands like P(OMeh (fJ = 107°) were
able to displace all four. On this basis we can anticipate that the distribution
of Rh among the RhL n planes will depend on fJ, and that systems containing
a very small ligand like P(OMeh, especially if it is present in excess, should
show little catalytic activity.
Important studies of rhodium-catalyzed hydroformylation with ligands
other than PPh 3 were carried out by Pruett and Smith. (241) Table 13 shows
some of their results with varying concentrations of P(OPhh, using rhodium
on carbon as the catalyst charged. The metal dissolved under the reaction
conditions. t The increase in percentage of linear aldehyde with added [L]

Table 13. Dependence of Product Distribution on L:Rh


Ratio for P(OPh)3 in 1-0ctene Hydroformylation 8

Percent linear
Wt. P(OPhh (g) L:Rh b aldehyde

0 0.0 31
5 2.2 74
15 6.6 86
30 13.3 87
60 26.6 89

a 112 g of l-octene and 15 g of 5% RhjC in 200 ml toluene at 90' with


80-100 psig of 1 : 2 H 2 : CO. Data from Reference 241.
b Mole ratio, assuming that all the Rh metal dissolved.

t When L was PPh 3 • the same rates and products were observed starting with HRh(CO)-
(PPh 3 h·
96 C. A. TOLMAN AND J. W. FALLER

Table 14. Dependence of Product Distribu-


tion on Pressure in 1-0ctene Hydroformy-
lation 8

Percent linear
Pressure (psig)b aldehyde

80-100 86
280-300 80
560-600 74
2500 69

a 112 g of l-octene, 15 g of 5% Rh/C, and 15 g P(OPh),


in 200 ml toluene at 90'. L: Rh = 6.6: 1.0. Data from
Reference 241.
b Total pressure of H2 and CO in a 1 : 1 ratio.

is striking, but this trend can be partially reversed by increasing the total
pressure, as shown in Table 14. The authors recognized that the effect was
due to a competition of L and CO for coordination, and wrote Equation 84: t

L L L L
HRh(CO)4 ~ HRh(COhL ~ HRh(CO)zLz ~ HRh(CO)L 3 ~ HRhL4

(84)

theorizing that each species results in an individual reaction rate and a


characteristic product distribution. The species in Equation 84 are represen-
ted in Figure 2.19 by the small circles which are furthest away from the
viewer in each plane. Here we see that each plane has an individual reaction
rate and characteristic product distribution, with the ratio of
R CHO / R 'CHO increasing-due to more steric crowding-as the reaction
shifts to successively lower planes.
Table 15 shows the effects of changing the steric and electronic charac-
ter of L on the product distribution when other variables are kept constant.
For the p-substituted phenyl phosphites, which are sterically similar, the
percentage of linear aldehyde is greatest for p -Cl. Without spectroscopic
data on these systems under hydroformylation conditions, we cannot know
whether the electronic effects are due to changing the distribution among
the planes of Figure 2.19, or to changing the relative rates of the two loops
within a plane, though the latter seems more likely. Though the authors
do not state that the reactions were all carried to approximately the same
t Pruett and Smith(Z41) obtained IR evidence for the middle three of the five species in
Equation 84 by using L = PPh 3 under simulated hydroformylation conditions.
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 97

Table 15. Dependence of Product Distribution on Phosphorus Ligand in


Hydroformylation of 1-0ctene 8

Percent linear Reaction time


Ligand aldehyde (min) (J (deg)b x(cm -l)b

P(OpC 6H 4 Clh 93 55 128 33.2


P(OPhh 86 50 128 29.2
P(OpC 6H 4 Phh 85 70 128 28.9 c
P(OpC 6H 4 OMeh 83 270 128 28.0
PPh 3 82 35 145 12.8
P(OBuh 81 e 60 107 20.2 d
P(OC 6 H 4 Meh 78 52 141 28.0
PBU3 71 225 132 4.2
P(OoC 6H 4 Phh 52 95 132 28.9
P(O-2, 6-Me2C6H3h 47 80 190 27.1

a 112 g l-octene, 10 g 5% Rh/C, and 0.05 mol L (L: Rh = 10: 1) with 80-100 psig of 1: 1 H 2 : CO at 90'
(unless noted otherwise). From Reference 24l.
b Taken from ref. 239b unless noted otherwise. X = v - 2056.1 em -1.
C The value for P(OoC 6 H 4 Ph),.
d The value for P(OEt)3'
'At 110°.

conversion, we assume that they were and that the reaction times reported
give a measure of the relative reaction rates. The best electron donor of
the p-substituted phenyl phosphites gives the lowest rate (perhaps due to
reduced CO dissociation constants from 18-electron acyls). PBU3, the best
donor in Table 15, is also very slow.
Electronically similar 0 -substituted phenyl phosphites show a decreas-
ing percentage of linear aldehyde as the cone angle is increased; this is no
doubt due to a shifting of the Rh to higher and higher planes. Though the
rate of formation of R CHO relative to R 'CHO within a plane probably
increases with 0, t due to increased crowding, the shifting of Rh to higher
planes (through competition of CO and L) dominates, and the net result
is less linear product.

t This hypothesis can be tested by carrying out experiments with 0- substituted phenyl
phosphites, adjusting the L:Rh ratio to maintain the Rh distributions the same among the
RhL n planes (as determined spectroscopically) as (J is varied.
98 C. A. TOLMAN AND J. W. FALLER

The reaction times in Table 15 suggest that hydroformylation of


l-octene is relatively fast using PPh 3. Ogata and co-workers(267) report that
hydroformylation of styrene (at 50°) is 2.5 times as fast using HRh(DBP-
Ph)4 as the catalyst as with HRh(PPh 3k DBP-Ph is 5-phenyl-5-H-diben-
zophosphole (83).

@(JlJ p

© 83

The percentage of linear aldehyde was not reported for these experiments,
but in others at 140° the percentage increased from 23% to 26% on going
from PPh 3 to DBP-Ph. We cannot say to what extent the differences in
behavior are due to changing steric size and to what extent to changing
electron donor character. BDP-Ph has a smaller () but a larger X than
PPh 3. Replacing PPh 3 by Ph2P(CH2)zPPh 2 or Ph 2P(CH2hPPh 2 decreases
both rates and percentages of linear aldehyde in the hydroformylation of
I-butene. (45)
The mechanisms shown in Figures 2.18 and 2.19 are consistent with
the 16- and 18-Electron Rule and are able to account qualitatively for the
behavior described in Table 11. They correspond to the "dissociative
mechanism" of hydroformylation of Evans and co_workers.(243) These
authors also proposed an "associative mechanism" in which the first step
in the reaction of I-alkene with HRh(CO)zL2 was coordination to form a
20 electron HRh(CO)zL 2(S) complex. They were led to this because of the
different selectivities using HRh(CO)L 3 (L = PPh 3) for I-alkene vs 2-
alkene in the hydrogenation and hydroformylation reactions, where the
rate ratios were about 200: 1 and about 25: 1, respectively. They wrote,
"The objection can be raised that for a d 8 species, the effective atomic
number rule is thus exceeded, but we do not regard this as serious for a
short-lived species. ,,(243) The associative pathway has been widely quoted,
including in hydroformylations using cobaltt We regard the associative
pathway as both unlikely and unnecessary. The necessary experiments to
show this would involve studying the kinetics of reaction of I-alkene and
2-alkene with HRh(CO)zL2 to form acyls as a function of CO pressure,
varying [L] to keep HRh(CO)zL 2 as the major initial Rh species in solution.

t See, for example, Figures 1.6. and 1.7. in Reference 239.


MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 99

Table 16. Comparison of Cobalt and Rhodium in Position to which Formvl


Group is Bounds

Catalyst P eo (atm) Percent distribution of formyl placement

CD 3 -CH 2-CH 2-CH=CH 2


CO 2(CO)8 100 9 3 3 12 72

Rh 4(COh2 100 0 4 6 35 54
4-5 8 19 12 25 36
CD 3 -CH 2-CH=CH - CH 3
CO 2(CO)8 400 29 19 12 20 39
100 25 7 7 14 47
Rh 4(COh2 100 16 28 48 7
4-5 11 24 14 26 25

a At 100° with 80-100 atm H 2 , from Reference 270.

The 16- and 18-Electron Rule predicts a rate law given by Equation 85

~d[HRh(COhL2]/dt = k[HRh(COhL 2 ][S]/[CO] (85)


while a combination of dissociative and associative paths would give (86).

~d[HRh(COhL2]/dt = k[HRh(COhL 2][S](a/[CO] + b) (86)

2.4.5. Comparison of Rhodium with Cobalt and Iridium Catalysts

Cobalt and iridium complexes can also function as hydroformylation


catalysts, but of course the cobalt catalysts are much better known. t The
reduced activity of Co and Ir compared to Rh has been attributed to the
greater stability of their 18-electron complexes to ligand dissociation. (5)
The Co and Ir systems do provide some interesting comparisons with
Rh, (269) both in the behavior of the catalytic systems and in the identification
of catalytic intermediates.
Table 16 shows some interesting work by Pino and co-workers(270)
comparing the position of formyl placement by Co and Rh in the hydro-
formylation of isotopically-labeled d 3 -pentenes. The greater percentage of
linear aldehyde products with phosphine-free Co than with Rh has been
attributed to greater steric crowding in Co complexes as a consequence of
its smaller atomic radius. (266) In addition to the tendency for terminal

t A review of HCO[CO]4 chemistry has recently been published by Onchin.(268)


100 C. A. TOLMAN AND J. W. FALLER

formylation with Co, there is an increased tendency to isomerization, seen


in the relatively large amounts of formylation at the labeled end of the
molecule in reactions at 100 atm CO. Even when the starting olefin is
2-pentene, the major product is linear aldehyde. With Rh there is a much
greater tendency to add the formyl to the carbons which were originally
on the double bond. Increased CO pressure, with either metal, has the
effect of increasing the extent of formylation of the original double-bond
carbons. This is clearly a consequence of trapping by CO of the alkyl
complexes first formed by metal hydride addition across the double bond,
before f3 -hydride abstraction can occur to move the double bond down
the chain.
High-temperature and pressure IR spectroscopic studies by Penninger
and co-workers(271l have accurately defined equilibrium constants for
Equation 87:

H2 + Co 2(CO)s ¢ 2HCo(CO)4 (87)

At 100°, they report K = 0.20. For the forward reaction, the pseudo
first-order rate constant K = 8.8 X 10- 3 min- 1t at PH2 = 25 atm (and Peo =
25 atm). AH = 6.6 kcal/mole and EA = 11.3 kcal/mol. The fact that both
the equilibrium constant and forward rate constant increase with increasing
temperature explains why high temperatures are used with cobalt to achieve
high rates. Under these conditions, however, higher CO pressures are also
required to stabilize the carbonyls against decomposition to metallic Co.
Ungvary and Marko(272l have more recently reported that the kinetics
of (87) are complex, and that the back reaction, while second-order in
[HCO(CO)4] as expected, is inhibited by CO and accelerated by CO2(CO)s,
with a concentration dependence of [C0 2(CO)S]O.5. They propose a mechan-
ism involving 'CO(CO)4 radicals.
Penninger and co-workers(273l studied the behavior of a system of
CO2(CO)s and HCO(CO)4 at equilibrium at 125° and 100 atm of H2 and
CO (1: 1) to which excess 1-pentene was injected. They observed a rapid
decrease ((1/2 < 2 min) in the concentration of HCO(CO)4 from about
3.8 mM to a new steady-state concentration of about 1.3 mM, with forma-
tion of a mixture of (acyl)Co(CO)4 complexes,:j: (about 1.3 mM) and an
increase in CO 2(CO)s from about 0.25 to 1.5 mM. They explained the fact
that [C0 2(CO)s] was higher at steady state than at equilibrium in terms of
Equation 88. Hydrogenolysis of acyls was said to be by reaction with

t The authors(27\) report both forward and reverse rate constants in units of min-t. This is
okay for a pseudo first-order rate constant for the forward reaction (if [H 2] is essentially
constant), but cannot be correct for the reverse reaction.
t Characterized by a new band at 2010 cm -t.
MECHANISTIC STUDIES OF CA TAL YTiC REACTIONS 101

HCO(CO)4, rather than with H 2, implying that CO2(CO)s is a loop species.


RCOCo(COh,4 + HCO(CO)4 --> RCHO + Co 2 (COh. 8 (88)
The extent to which hydrogenolysis takes place by reaction with HCO(CO)4
rather than with H 2, as in the conventional Heck and Breslow mechan-
ism(274) and under what conditions, and whether it plays a role at all in
rhodium hydroformylations, remains to be established by kinetic and
spectroscopic experiments. The situation with cobalt is also complicated
by the occurrence under some conditions of odd-electron pathways.
Ungvary and co_workers(27S,276) have shown that reactions of HCO(CO)4
with ole fins can be catalyzed by CO 2(CO)s, implying a significant role for
·CO(CO)4.
While iridium is less active than rhodium as a catalyst, it does provide
insight into the mechanism, because several intermediates are stable enough
to be isolated or observed when the rhodium analogs cannot. Examples
include HIr(COh(PPh 3 h, whose X-ray structure(2SS) was mentioned earlier,
and the Etlr(COhL2 (alkyl), EtCOIr(CO)L 2 (16-electron acyl), and
EtCOIr(COhL (tricarbonyl acyl) with L = PPh 3 reported by Yagupsky
and co_workers.(2S2) Formation of propionaldehyde from EtCOIr(COhL 2
or EtCOIr(COhL with H2 is inhibited by CO pressure, indicating that the
oxidative additions proceed via the 16-electron acyls as shown for Rh in
Figure 2.18. While the acyldihydrides have not been seen, HCI and
EtCOIr(CO)L 2 give the EtCOIrHCI(CO)L 2 analog, whose formation is
also inhibited by CO.
While HRh(CO)L 3 does not react with C 2H 4 under pressure,
HIr(CO)L3 reacts under 10 atm C 2H 4 at 35° in 15 minutes to give about
15% conversion to a new species with IH nmr signals at 8 1.26 (triplet)
and 1.78 (broad) assigned to CH3 and CH 2 protons in Etlr(CO)L 2. While
HRh(CO)zL 2 (formed from HRh(CO)L 3 under CO) reacts rapidly and
quantitatively with C 2H 4 to give EtCORh(CO)zL 2, HIr(CO)zL2 under-
goes a complex series of reactions over 1-2 h at 35°, giving resonances
assigned to EtIr(CO)L 2, EtCOIr(CO)zL 2, EtCOIr(COh(C 2H 4 )L, and
EtCOIr(CO)L 2.(2S2)
By successive treatment of HIr(COhL' [L' = PUPrh] with C 2H 4, CO,
and H2 in a high pressure IR cell, Whyman(277) has been able to identify
C2H sIr(COhL', C2H sCOIr(COhL', and C 2H sCOH + HIr(COhL', and
follow the stepwise conversion of one to the next-each reaction requiring
about half an hour at 50°. He thus has direct spectroscopic evidence for
many of the steps shown in Figure 2.18, except in the IrL~ plane. The
absence of IrL; species is no doubt due to the large size of PUPrh
(0 = 160°)(4)
The greatly reduced activity of Ir compared to Rh can be seen in the
following experiments. Using 1-pentene and HIrCO(PPh 3 h in a 540: 1
102 C. A. TOLMAN AND J. W. FALLER

ratio in benzene under 100 atm H2 and CO (1: 1), less than 10% conversion
to aldehyde occurred in 18 h at 70° «3 cycles/h). Under similar conditions
at 70° even RhCI(CO)(PPh 3h gave complete conversion.(243) HRh(CO)-
(PPh 3h is faster still and gives nearly 3 cycles/h at 25° (Table 10).

ACKNOWLEDGMENTS

We are indebted to Drs. H. Schenkluhn, P. Heimbach, J. Halpern, R.


Kastrup, and R. B. King for sending information prior to its publication,
to Ms. Ruth McFarlane for technical information assistance, and to Ms.
Carol Farber and Ms. Sue Koblitz for painstaking typing of the manuscript,
and to the Journals Department of the American Chemical Society for
permission to reproduce some of the figures.

REFERENCES

1. G. W. Parshall, Homogeneous Catalysis, The Applications and Chemistry of Catalysis


by Soluble Transition Metal Complexes (Wiley-Interscience of John Wiley and Sons,
New York, 1980).
2. J. P. Collman and L. S. Hegedus, Principles and Applications of Organotransition Metal
Chemistry (University Science Books, Mill Valley, California, 1980) a) p. 316.
3. C. Masters, Homogeneous Transition-Metal Catalysis-A Gentle Art (Chapman and
Hall, London, 1981).
4. C. A. Tolman, Chern. Rev. 77, 319 (1977).
5. R. B. Woodward and R. Hoffman, Angew. Chern. Internat. Ed. Engl. 8,781 (1969).
6. C. A. Tolman, Chern. Soc. Revs. 1, 337 (1972).
7. C. A. Tolman and J. P. Jesson, Science 181,501 (1973).
8. J. P. Birk, J. Halpern, and A. L. Pickard, I. Arner. Chern. Soc. 90,4491 (1968).
9. C. A. Tolman, W. C. Seidel, and D. H. Gerlach, I. Arner. Chern. Soc. 94,2669 (1972).
10. C. A. Tolman, I. Arner. Chern. Soc. 94,2994 (1972).
11. S. Otsuka, T. Yoshida, M. Matsumoto, and K. Nakatsu, I. A mer. Chern. Soc. 98,5850
(1976).
12. W. L. M. Van Gaal and F. L. A. Van Den Bekerom, I. Organornet. Chern. 134, 237
(1977).
13. J. Halpern and C. S. Wong, I. Chern. Soc. Chern. Cornrnun. 629, (1973).
14. C. A. Tolman, P. Z. Meakin, D. L. Lindner, and J. P. Jesson, I. A mer. Chern. Soc. 96,
2762 (1974).
15. G. M. Whitesides, J. F. Gaasch, and E. R. Stedronski, I. A mer. Chern. Soc. 94,5258
(1972).
16. Y. W. Yared, S. L. Miles, R. Bau, and C. A. Reed, I. Arner. Chern. Soc. 99, 7076 (1977).
17. J. A. Osborn, F. H. Jardine, J. F. Young, and G. Wilkinson,!. Chern. Soc. A, 1711 (1966).
18. M. H. J. M. de Croon, P. F. M. T. van Nisselrooij, H. J. A. M. Kuipers, and J. W. E.
Coenen, I. Mol. Catalysis 4,325 (1978).
19. T. L. Brown, Anal. N. Y. Acad. Sci., edited by D. W. Slocum and O. R. Hughes, 333,
80 (1980).
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 103

20. J. Halpern, Fundamental Research in Homogeneous Catalysis, edited by M. Tsutsui


(Plenum, New York, 1978), Vol. 3, pp. 25-40.
21. A. A. Frost and R. G. Pearson, Kinetics and Mechanism, 2nd ed. (John Wiley and
Sons, New York, 1961) a) p. 2.
22. J. W. Moore and R. G. Pearson, Kinetics and Mechanism, 3rd ed. (John Wiley and
Sons, New York, 1981).
23. F. Basolo and R. G. Pearson, Mechanisms of Inorganic Reactions, 2nd ed. (John Wiley
and Sons, New York, 1967).
24. H. Lineweaver and D. Burke, 1. Amer. Chern. Soc. 56, 658 (1934).
25. E. L. King and C. Altman, 1. Phys. Chern. 60, 1375 (1956).
26. W. W. Cleland, Biochem. Biophys. Acta. 67, 104 (1963).
27. - - , Biochem. 14, 3220 (1975).
28. D. F. Cook and W. W. Cleland, Biochem. 20, 1790, 1797, 1805 (1981).
29. I. H. Segel, Enzyme Kinetics (John Wiley and Sons, New York, 1975).
30. J. H. Sullivan, 1. Chern. Phys. 46, 73 (1967).
31. K. J. Laidler, Chemical Kinetics (McGraw-Hill Book Co., New York, 1965), pp.
400-406.
32. M. J. Pilling, Reaction Kinetics (Clarendon Press, Oxford, 1975).
33. J. E. Baeckvall, B. Akermark, and S. O. Ljunggren, 1. Am. Chern. Soc. 101,2411 (1979).
34. P. W. Atkins, Physical Chemistry, 2nd ed. (W. H. Freeman & Co., San Francisco, 1982).
35. R. K. Harris and B. E. Mann (Eds), NMR and the Periodic Table (Academic Press,
London, 1978).
36. P. S. Pregosin and R. W. Kunz, NMR Basic Principles and Progress (Springer-Verlag,
Heidelberg, 1979) Vol. 16.
37. R. G. Kidd, Annu. Rep. NMR Spectrosc. lOA, 1 (1980).
38. S. Aime and D. Osella, L. Milone, G. E. Hawkes, and E. W. Randall, 1. Am. Chern.
Soc. 103, 5920 (1981).
39. C. Brown, B. T. Heaton, L. Longhetti, W. T. Povey, and D. o. Smith, 1. Organometal.
Chern. 192, 93 (1980).
40. K. D. Brueninger, A. Schwenk, B. E. Mann, 1. Magn. Reson. 41,354 (1980).
41. O. A. Gansow, D. S. Gill, F. J. Bemis, J. R. Hutchinson, J. L. Vidal, R. C. Shoening,
1. Am. Chern. Soc. 102, 2449 (1980).
42. G. A. Morris and R. Freeman, J. Am. Chern. Soc. 101,760 (1979).
43. C. Brevard, G. C. van Stein, and G. van Koten, J. Am. Chern. Soc. 103, 6746 (1981).
44. B. T. Heaton, J. Jonas, T. Eguchi, and G. Hoffman, 1. Chern. Soc., Chern. Commun.,
313 (1981).
45. R. V. Kastrup, J. S. Merola, and A. A. Oswald, Advances in Chemistry 196,78 (1981),
edited by E. C. Alyea and D. W. Meek, American Chemical Society.
46. A. A. Oswald, R. V. Kastrup, J. S. Merola, E. J. Mozeleski, and J. C. Reisch. ACS
Symp. Ser. 171, 503 (1981).
47. J. Norton, Accts. Chern. Res. 12, 139 (1977).
48. L. M. Jackman and F. A. Cotton, eds., Dynamic Nuclear Magnetic Resonance Spectros-
copy (Academic Press, New York, 1975).
49. J. W. Faller, Adv. Organometal. Chern. 16,211 (1977).
50. A. Moor, P. S. Pregosin, and L. M. Venanzi, Inorg. Chim. Acta 48, 153 (1981).
51. K-H. A. O. Starzewski and P. S. Pregosin, Angew. Chern., Int. Ed. Engl. 19, 316 (1980).
52. J. W. Faller and G. N. LaMar, Tetrahedron Lett. 16, 1381 (1973).
53. M. A. Bennett, F. A. Cotton, A. Davison, J. W. Faller, S. J. Lippard, and S. M.
Morehouse, 1. Am. Chern. Soc. 88, 4371 (1966).
54. J. W. Faller, Determination of Organic Structures by Physical Methods, edited by F. C.
Nachod and J. I. Zuckerman (Academic Press, New York, 1973), Vol. 5, p. 75.
55. R. Bruce Calvert and J. R. Shapley, 1. Am. Chern. Soc. 100, 7726 (1978).
104 C. A. TOLMAN AND J. W. FALLER

56. J. W. Faller, M. E. Thomsen, and M. J. Mattina, 1. Arn. Chern. Soc. 93,2642 (1971).
57. J. W. Faller, Y. Shvo, K. Chao, and H. H. Murray, 1. Organornetal. Chern. 226, 251
(1982).
58. R. A. Hoffman and S. Forsen, Progr., Nucl., Magn., Reson., Spectrosc. 1, 15 (1966).
59. F. W. Dahlquist, K. J. Longmuir, and R. B. DuVernet, 1. Magn. Resonance 17,406
(1975).
60. B. E. Mann, Prog. NMk. Spectros. 11,95 (1977).
61. G. A. Morris and R. Freeman, 1. Magn. Resonance 29,433 (1978).
62. L. Melander, Isotope Effects on Reaction Rates (Ronald Press, New York, 1960).
63. L. Melander and W. H. Saunders, Jr., Reaction Rates of Isotopic Molecules (John Wiley
and Sons, New York, 1980).
64. K. B. Wiberg, Physical Organic Chernistry (John Wiley and Sons, New York, 1964),
pp.351-363.
65. F. H. Westheimer, Chern. Rev. 61,265 (1961).
66. R. P. Bell and D. M. Goodall, Proc. Roy. Soc., Ser. A. 294,273 (1966).
67. R. L. Sweany and J. Halpern, 1. Arn. Chern. Soc. 99,8335 (1977).
68. A. J. Leusink, H. A. Budding, and W. Drenth, 1. Organornetal. Chern. 9, 295 (1967).
69. T. E. Nalesnik and M. Orchin, 1. Organornetal. Chern. 199, 265 (1980).
70. - - , Organornetallics 1,223 (1982).
71. J. P. Collman, R. G. Finke, P. L. Matlock, R. Wahren, R. G. Komoto, and J. 1. Brauman,
1. Arn. Chern. Soc. 100, 1119 (1978).
72. R. P. Bell, Chern. Rev. 3,513 (1974).
73. M. Saunders, M. H. Jaffe, and P. Vogel, 1. Arn. Chern. Soc. 93, 2558 (1971).
74. M. Saunders, L. Telkowski, and M. R. Kates, 1. Arn. Chern. Soc. 99,8070 (1979).
75. J. W. Faller, H. H. Murray, and M. Saunders, 1. Arn. Chern. Soc. 102, 2306 (1980).
76. N. J. Cooper and M. L. H. Green, 1. Chern. Soc., Chern. Cornrnun., 761 (1974).
77. R. C. Brady and R. Pettit, 1. Arn. Chern. Soc. 103, 1287 (1981).
78. J. G. Ekerdt and A. T. Bell. 1. Catat. 62, 19 (1980).
79. L. E. McCandlish, 1. Catal., to be published.
80. G. M. Dawkins, M. Green. A. Guy Orpen, and F. G. A. Stone, 1. Chern. Soc., Chern.
Cornrnun. 41 (1982).
81. R. B. Calvert, J. R. Shapley, A. J. Shultz, Jack M. Williams, S. L. Sluib, and G. D.
Stucky, 1. Arn. Chern. Soc. 100, 6240 (1978).
82. A. J. Schultz, J. M. Williams, R. B. Calvert, J. R. Shapley. and G. D. Stucky, Inorg.
Chern. 18, 319 (1979).
83. R. B. Clavert and J. R. Shapley, 1. Arn. Chern. Soc. 99,5226 (1977).
84. J. C. Huffman and W. E. Streib, 1. Chern. Soc., Chern. Cornrnun., 911 (1971).
85. J. Holton. M. F. Lappert, D. G. H. Ballard, R. Pearce, J. L. Atwood, and W. E. Hunter,
1. Chern. Soc., Dalton Trans., 54 (1979).
86. R. J. Jablonski and E. 1. Snyder, Tetrahedron Lett. 1103, (1968).
87. R. G. Weiss and E. 1. Snyder, 1. Chern. Soc., Chern. Cornrnun., 1358 (1968).
88. R. J. Jablonski and E. 1. Snyder, 1. Arn. Chern. Soc. 91, 4445 (1969).
89. G. M. Whitesides and D. J. Boschetto, 1. Arn. Chern. Soc. 91,4313 (1969).
90. P. L. Bock, D. J. Boschetto, J. R. Rasmussen, J. P. Demers, and G. M. Whitesides, 1.
Arn. Chern. Soc. 96, 2814 (1974).
91. J. A. Labinger and J. A. Osborn, Inorg. Chern. 19,3230 (1980).
92. G. Zweifel and C. C. Whitney, 1. Arn. Chern. Soc. 89, 2753 (1967).
93. J. A. Labinger, D. W. Hart, W. E. Seibert, and J. Schwartz, 1. Arn. Chern. Soc. 97,
3851 (1975).
94. D. A. Slack and M. C. Baird, 1. Arn. Chern. Soc. 98, 5539 (1976).
95. P. L. Bock and G. M. Whitesides, 1. Arn. Chern. Soc. 98, 5539 (1976).
96. J. W. Faller and A. S. Anderson, 1. Arn. Chern. Soc. 91,1550 (1969).
MECHANISTIC STUDIES OF CA TAL YTIC REACTIONS 105

97. W. N. Rogers, J. A. Page, and M. C. Baird, Inorg. Chem. 20, 3521 (1981).
98. D. Dong, B. K. Hunter, and M. C. Baird, f. Chem. Soc., Chem. Commun., 11 (1978).
99. D. Dong, D. A. Slack, and M. C. Baird, f. Organometal. Chem. 153, 219 (1978).
100. - - , Inorg. Chem. 18, 188 (1979).
101. T. C. Flood and F. J. DiSanti, f. Chem. Soc. 18 (1975).
102. T. Majima and H. Kurosawa, f. Chem. Soc., Chem. Commun., 610 (1977).
103. F. R. Jensen, V. Madan, and D. H. Buchanan, f. Am. Chem. Soc. 92, 1414 (1970).
104. K. Stanley and M. C. Baird, f. Am. Chem. Soc. 97,6598 (1975).
105. R. S. Cahn, C. Ingold, and V. Prelog., Angew. Chem. Int. Ed. Eng. 5, 385 (1966).
106. T. G. Attig, R. G. Teller, S-M. Wu, R. Bau, and A. Wojcicki, f. Am. Chem. Soc. 101,
619 (1979).
107. J. W. Faller and A. S. Anderson, f. Am. Chem. Soc. 92, 5852 (1970).
108. C. Dodd and M. D. Johnson, f. Chem. Soc. D, 571 (1971).
109. F. Calderazzo and K. Noack, Coord. Chem. Rev. I, 118 (1966).
110. H. M. Walborski and L. E. Allen, f. Am. Chem. Soc. 93, 5465 (1971).
111. L. Verbit, Progr. Phys. Org. Chem. 7,51 (1970).
112. J. K. Stille and S. Y. Lau, Accts. Chem. Res. 10, 434 (1977).
113. A. I. Scott and A. D. Wrixon, Tetrahedron 27,2339 (1971).
114. J. W. Faller and M. T. Tully, f. Am. Chem. Soc. 94, 2676 (1972).
115. H. Brunner, Adv. Organometal. Chem. 18, 151 (1980).
116. C. K. Chou, D. L. Miles, R. Bau, and T. C. Flood, f. Am. Chem. Soc. 100, 7271 (1978).
117. J. A. Dale, D. L. Dull, and H. S. Mosher, f. Org. Chem. 34,2543 (1969).
118. M. D. McCreary, D. W. Lewis, D. L. Wernick, and G. M. Whitesides, f. Am. Chem.
Soc. 96, 1038 (1974).
119. G. R. Sullivan, Topics in Stereochemistry; edited by E. L. Eliel and N. L. Allinger
(Wiley-Interscience of John Wiley and Sons, New York, 1978), p. 288.
120. T. J. Marks, J. S. Kristoff, A. Alich, and D. F. Shriver, f. Organometal. Chem. 33, C35
(1971).
121. J. W. Faller and B. V. Johnson, f. Organometal. Chem. 96,99 (1975).
122. P. A. MacNeil, and N. K. Roberts, and B. Bosnich, f. Am. Chem. Soc. 103,2273 (1981).
123. M. D. Fryzuk and B. Bosnich, f. Am. Chem. Soc. 99, 6262 (1977).
124. 1. D. Morrison, W. F. Masler, M. K. Neuberg, Adv. Catal. 25,81 (1976).
125. D. S. Valentine, Jr and J. W. Scott, Synthesis, 329 (1978).
126. R. R. Schrock and J. A. Osborn, f. Am. Chem. Soc. 98, 2134 (1976).
127. - - , f. Am. Chem. Soc. 98,4450 (1976).
128. J. Halpern, T. Okamoto, and A. Zakhariev, 1. Mol. Catal. 2, 65 (1977).
129. C. Rousseau, M. Evard, and F. Petit, f. Mol. Catal. 3,309 (1978).
130. M. H. 1. M. DeCroon, P. F. M. T. van Nisselrooij, H. J. A. M. Kuipers, and J. W. E.
Coener, f. Mol. Catal. 4, 325 (1978).
131. F. A. Cotton and G. Wilkinson, Advanced Inorganic Chemistry (John Wiley and Sons,
New York, 1980), a) p. 1266.
132. J. Halpern, D. P. Riley, A. S. C. Chan, and J. J. Pluth, f. Am. Chem. Soc. 99, 8055 (1977).
133. J. M. Brown and D. Parker, f. Chem. Soc., Chem. Commun. 342 (1980).
134. J. M. Brown and P. A. Chaloner, f. Chem. Soc., Chem. Commun., 344 (1980).
135. - - , f. Am. Chem. Soc. 102, 3040 (1980).
136. A. S. C. Chan, J. J. Pluth, and J. Halpern, Inorg. Chim. Acta 37, L477 (1979).
137. A. S. C. Chan, and J. Halpern, f. Am. Chem. Soc. 102, 838 (1980).
138. A. S. C. Chan, J. J. Pluth, and J. Halpern, f. Am. Chem. Soc. 102,5952 (1980).
139. J. K. Kochi, Accts. Chem. Res. 7, 351 (1974).
140. J. A. Osborn, Organotransition-Metal Chemistry, edited by Y. Ishii and M. Tsutsui
(Plenum Press, New York, 1975), p. 65.
141. J. A. Labinger, 1. A. Osborn, and N. J. Coville, Inorg. Chem. 19,3236 (1980).
106 C. A. TOLMAN AND J. W. FALLER

142. M. F. Lappert and P. W. Lednor, Adv. Organometal. Chem. 14,345 (1976).


143. T. L. Hall, M. F. Lappert, and P. W. Lednor, 1. Chem. Soc., Dalton, 1448 (1980).
144. B. H. Byers and T. L. Brown, 1. Am. Chem. Soc. 97, 947 (1975).
145. H. D. Murdock and E. A. C. Lucken, He/v. Chim. Acta 47. 1517 (1964).
146. H. J. Keller and H. Wawersik, Z. Naturforsch. 20b, 938 (1965).
147. H. J. Keller, Z. Naturforsch. 23b, 133 (1968).
148. J. P. Fawcett and A. Poe, 1. Chem. Soc., Dalton Trans., 2039 (1976).
149. P. M. Treichel, K. P. Wagner, and H. J. Much, 1. Organometal. Chem. 86, C13 (1975).
150. A. M. Bond, R. Colton, and J. J. Jackowski, Inorg. Chem. 14,2526 (1975).
151. R. E. Dessy and L. A. Bares, Accts. Chem. Res. 5, 415 (1972).
152. J. A. Connor and P. 1. Riley, 1. Chem. Soc., Chem. Commun., 634 (1976).
153. M. Gargano, P. Giannocaro, M. Rossi, G. Vassapollo, and A. Sacco, 1. Chem. Soc.,
Dalton,9 (1975).
154. P. J. Krusic, J. San Filippo, Jr., B. Hutcins, R. L. Hance, and L. M. Daniels, 1. Am.
Chem. Soc. 103,2129 (1981).
155. P. J. Krusic, 1. Am. Chem. Soc. 103, 2131 (1981).
156. J. Chatt, R. A. Head, and G. J. Leigh, 1. Chem. Soc., Dalton Trans., 1638 (1978).
157. H. Felkin and B. Meunier, Nouv. 1. Chimie 1,231 (1977).
158. P. J. Krusic, P. J. Fagan, and J. San Filippo, 1. Am. Chem. Soc. 99,250 (1977).
159. P. J. Krusic, U. Klabunde, C. P. Casey, and T. F. Block, 1. Am. Chem. Soc. 98, 2015
(1976).
160. C. Lagercrantz, 1. Phys. Chem. 75, 3466 (1971).
161. E. G. Janzen, Accts. Chem. Res. 4, 31 (1971).
162. S. Terabe and R. Konaka, 1. Chem. Soc., Perkin 11,2136 (1972).
163. B. G. Gowerlock and J. Trotman, 1. Chem. Soc., 4190 (1955).
164. S. Terabe, K. Kuruma, and R. Konaka, 1. Chem. Soc., Perkin II, 1252 (1973).
165. A. R. Lepley and G. L. Closs, Chemically Induced Magnetic Polarization (John Wiley
and Sons, New York, 1973).
166. H. R. Ward, Free Radicals, edited by J. K. Kochi (John Wiley and Sons, New York,
1973), Vol. 1, Chapt. 8.
167. A. V. Kramer and J. A. Osborn, 1. Am. Chem. Soc. 96, 7832 (1974).
168. F. D. Green, M. A. Berwick, and J. C. Stowell, 1. Am. Chem. Soc. 92,867 (1970).
169. D. J. Carlsson and K. U. Ingold, 1. Am. Chem. Soc. 90, 7047 (1968). .
170. P. D. Bartlett and J. M. McBride, Pure Appl. Chem. 15,89 (1967).
171. J. A. Labinger, J. A. Osborn, and N. J. Coville, Inorg. Chem. 19, 3236 (1980).
172. F. R. Jensen, L. H. Gale, and J. E. Rodgers, 1. Am. Chem. Soc. 90, 5793 (1968).
173. J. S. Filippo, Jr., J. Silberman, and P. J. Fagan, 1. Am. Chem. Soc. 100, 4834 (1978).
174. C. A. Tolman and L. H. Scharpen, 1. Chem. Soc. Dalton Trans., 584 (1973).
175. C. A. Tolman, 1. Amer. Chem. Soc. 92, 6777 (1970).
176. C. A. Tolman, W. C. Seidel, and L. W. Gosser, 1. Amer. Chem. Soc. 96, 53 (1974).
177. J. C. Marriott, J. A. Salthouse, M. J. Ware, and J. M. Freeman, 1. Chem. Soc., Chem.
Commun., 575 (1970).
178. V. Albano, P. L. Bellon, and V. Scatterin, 1. Chem. Soc., Chem. Commun., 507 (1966).
179. M. Meier, F. Basolo, and R. G. Pearson, Inorg. Chem. 8, 795 (1969).
180. C. A. Tolman, unpublished results.
181. - - , 1 . Am. Chem. Soc. 92, 4217 (1970).
182. P. Meakin and J. P. Jesson, 1. Amer. Chem. Soc. 96,5751 (1974) and references therein.
183. A. J. Deeming, B. F. G. Johnson, and J. Lewis, 1. Chem. Soc. (D) 1848 (1973).
184. C. A. Tolman, 1. Am. Chem. Soc. 92, 6785 (1970).
185. B. R. James, Homogeneous Hydrogenation (Wiley-Interscience of John Wiley and Sons,
New York, 1973).
MECHANfSTIC STUDIES OF CATAL YTIC REACTIONS 107

186. R. S. Coffey, British Patent No. 1121642 (18 February 1965).


187. S. Siegel and D. Ohrt, Inorg. Nucl. Chern. Lett. 8,15 (1972).
188. J. Halpern, Trans. Arn. Cryst. Assoc. 14,59 (1978).
189. P. B. Hitchcock, M. McPartlin, and R. Mason, 1. Chern. Soc., Chern. Cornrnun., 1367
(1969).
190. M. J. Bennett, P. B. Donaldson, P. B. Hitchcock, and R. Mason, Inorg. Chirn. Acta
12, L9 (1975).
191. T. E. Nappier, D. W. Meek, R. M. Kirchner, and J. A. Ibers, 1. Arn. Chern. Soc. 95,
4197 (1973).
192. R. Mason and D. W. Meek, Angew. Chern. Int. Ed. Engl. 17,183 (1978).
193. A. S. Hussey and Y. Takeuchi, 1. Arner. Chern. Soc. 91, 672 (1969).
194. G. C. Bond and R. H. Hillyard, Discuss. Farad Soc. 46,20 (1968).
195. S. Siegel and D. W. Ohrt, 1. Chern. Soc. Chern. Cornrnun., 1529 (1971).
196. P. Meakin, J. P. Jesson, and C. A. Tolman, 1. Arn. Chern. Soc. 94, 3240 (1972).
197. J. Halpern, T. Okamoto, and A. Zakhariev, 1. Mol. Cat. 2,65 (1976).
198. H. L. M. Van Gaal, F. G. Moers, and J. J. Steggerda, 1. Organornet. Chern. 65, C43
(1974).
199. T. E. Nappier, Jr., Ph.D. Dissertation, The Ohio State University, 1972.
200. D. L. Dubois and D. W. Meek, Inorg. Chirn. Acta 19, L29 (1976).
201. R. Ugo, Chern. Ind. (Milan) 58,631 (1976), quoting results in L. Sajus, Rev. Inst. Franc.
Petrole 24, 1477 (1969).
202. A. Dedieu, Inorg. Chern. 20,2803 (1981).
203. N. Aresta, M. Rossi, and A. Sacco, Inorg. Chern. Acta 3,227 (1969).
204. M. A. Bennett and D. A. Milner, 1. Arn. Chern. Soc. 91, 6983 (1969).
205. R. H. Crabtree, Accounts Chern. Res. 12, 331 (1979).
206. R. H. Crabtree, H. Felkin, and G. E. Morris, 1. Chern. Soc., Chern. Cornrnun., 716 (1976).
207. P. W. Jolly and G. Wilke, The Organic Chernistry of Nickel (Academic Press, London,
1975) Vol. II, Chapt. III.
208. P. Heimbach and H. Schenkluhn, Topics in Current Chernistry (Springer-Verlag, Berlin
Heidelberg, 1980), Vol. 92, p. 46.
209. P. W. Jolly and R. Mynott, Adv. Organornet. (hern. 19,257 (1981).
210. B. Bogdanovic, P. Heimbach, M. Kroner, G. Wilke, E. G. Hoffmann, and J. Brandt,
Liebigs Ann. Chern. 727, 143 (1969).
211. W. Brenner, P. Heimbach, H. Hey, E. W. Muller, and G. Wilke, Liebigs Ann. Chern.
727, 161 (1969).
212. P. Heimbach and G. Wilke, Liebigs Ann. Chern. 727,183 (1969).
213. P. Heimbach, R. V. Meyer, and G. Wilke, Liebigs Ann. Chern. 743 (1975).
214. W. Brunner, P. Heimbach, and G. Wilke, Liebigs Ann. Chern. 727,194 (1969).
215. P. Brille, P. Heimbach, J. Kluth, and H. Schenkluhn, Angew. Chern. Int. Ed. Engl. 18,
400 (1979). a) H. Schenkluhn and P. Heimbach, lecture notes used in Essen, 1978.
216. B. Bogdanovic, M. Kroner, and G. Wilke, Liebigs Ann. Chern. 699, 1 (1966).
217. B. Henc, P. W. Jolly, R. Saiz, S. Stobbe, G. Wilke, R. Benn, R. Mynott, K. Seevogel,
R. Goddard, and C. Kruger, 1. Organornet. Chern. 191, 449 (1980).
218. B. Barnett, B. Bussemeier, P. Heimbach. P. W. Jolly, C. Kruger, I. Tkatchenko, and
G. Wilke, Tetrahedron Lett. 15, 1457 (1972).
219. B. Henc. P. W. Jolly, R. Salz, G. Wilke, R. Benn, E. G. Hoffmann, R. Mynott, G.
Schroth, K. Seevogel, J. C. Sekutowski, and C. Kruger, 1. Organornet. Chern. 191,425
(1980).
220. J. E. Lydon, J. K. Nicholson, B. L. Shaw, and M. R. Truter, Proc. Chern. Soc. (London),
421 (1964).
221. R. Uttech and H. H. Dietrich, Z. Kristalloger 122,60 (1965).
108 C. A. TOLMAN AND J. W. FALLER

222. H. Dietrich and H. Schmidt, Naturwissenchaften 52,301 (1965).


223. G. Schomburg, D. Henneberg, P. Heimbach, E. Janssen, H. Lehmkuhl, and G. Wilke,
Liebigs Ann. Chern., 1667 (1975).
224. F. Brille, J. Kluth, and H. Schenkluhn, 1. Mol. Catalysis 5,27 (1979).
225. C. A. Tolman et al., Organornetallics, accepted for publication.
226. C. A. Tolman, D. W. Reutter, and W. C. Seidel, 1. Organornet. Chern., C30 (1976).
227. H. Schenkluhn, W. Scheidt, B. Weimann, and M. Zahres, Angew. Chern. Int. Ed. Engl.
18,401 (1979).
228. R. Berger, H. Schenkluhn, and B. Weimann, Trans. Met. Chern. 6, 272 (1981).
229. H. Schenkluhn, R. Burger, B. Pinttel, and M. Zahres, Trans. Met. Chern. 6, 277 (1981).
230. H. Schenkluhn, H. Bandmann, R. Berger, and E. Hubinger, Trans. Met. Chern. 6, 287
(1981).
231. G. Wilke, B. Bogdanovic, P. Hardt, P. Heimbach, W. Kleim, M. Bruner, W. Oberkirch,
K. Tamaka, E. Steinrucke, D. Walter, and H. Zimmerman, Angew Chern. Int. Ed. Engl.
5,151 (1966).
232. A. Butler and C. A. Tolman, unpublished results.
233. C. A. Tolman, 1. Arn. Chern. Soc. 96, 2780 (1974).
234. P. Heimbach, Angew. Chern. Int. Ed. Engl. 12,975 (1973).
235. P. Heimbach, A. Roloff, and H. Schenkluhn, Angew. Chern. Int. Ed. Engl. 16, 252 (1977).
236. C. M. King, W. C. Seidel, and C. A. Tolman, United States Patent No.3 925 445 (1975).
237. B. Bussemeier, P. W. Jolly, and G. Wilke, 1. Arn. Chern. Soc. 96, 4726 (1974).
238. O. Rolen, German Patent No. 849548 to Ruhrchemie A. G. (1938).
239. B. Comils New Syntheses with Carbon Monoxide edited by J. Falbe, (Springer-Verlag,
New York, 1980), pp. 1-225. a) p. 156. b) p. 174. c) p. 17.
240. R. L. Pruett, Adv. Organornet. Chern. 17, 1 (1979).
241. R. L. Pruett and J. A. Smith, 1. Org. Chern. 34,327 (1969).
242. P. Pino, 1. Organornet. Chern. 200, 223 (1980) and references therein.
243. D. Evans, J. A. Osborn, and G. Wilkinson, 1. Chern. Soc. (A), 3133 (1968).
244. A. A. Oswald, D. E. Hendriksen, R. V. Kastrup, J. S. Merola, and J. C. Reisch, presented
at the Lubrizol Award Sympsoium of the 1982 Spring ACS meeting, Las Vegas, Nevada.
245. A. A. Oswald, J. S. Merola, E. J. Mozeleski, R. V. Kastrup, and J. C. Reisch, Adv.
Chern. Series 104,503 (1981).
246. R. Whyman, 1. Organornet. Chern. 66, C23 (1974).
247. R. Whyman, 1. Organornet. Chern. 81, 97 (1974).
248. F. Piacenti, M. Bianchi, and E. Benedetti, Chirn. Ind. Milan 49,245 (1969).
249. R. B. King, A. D. King, Jr., M. Z. Iqbal, 1. Am. Chern. Soc. 101, 4893 (1979).
250. R. B. King, A. D. King, Jr., M. Z. Iqbal, and K. Tanaka, Ann. N. Y. Acad. Sci. 333,
74 (1980).
251. J. L. Vidal and W. E. Walker, Inorg. Chern. 20, 249 (1981).
251a. R. B. King and K. Tanaka, 1. Indian Chern. Soc., in press.
252. G. Yagupsky, C. K. Brown, and G. Wilkinson, 1. Chern. Soc. (A) 1392 (1970).
253. c. K. Brown and G. Wilkinson, 1. Chern. Soc. (A), 2753 (1970).
254. D. Evans, G. Yagupsky, and G. Wilkinson, 1. Chern. Soc. (A), 2660 (1968).
255. S. J. LaPlaca and J. A. Ibers, Acta Cryst. 18,511 (1965).
256. L. Vaska, 1. Arner. Chern. Soc. 88,4100 (1966).
257. J. D. Druliner, A. D. English, J. P. Jesson, P. Meakin, and C. A. Tolman, 1. Am. Chern.
Soc. 98, 2656 (1976).
258. M. Ciechanowicz, A. C. Skapski, and P. G. H. Troughton, Acta Cryst. 832, 1676 (1976).
259. C. K. Brown, W. Mowat, G. Yagupsky, and G. Wilkinson, 1. Chern. Soc. (A), 850 (1971).
260. C. O'Connor and G. Wilkinson, 1. Chern. Soc. (A), 2665 (1968).
261. T. Yoshida, T. Okano, D. L. Thorn, T. H. Tulip, S. Otsuka, and J. A. Ibers, 1. Organornet.
Chern. 181, 183 (1979).
MECHANISTIC STUDIES OF CATAL YTIC REACTIONS 109

262. J. P. Jesson Transition Metal Hydrides, edited by E. L. Meutterties (Marcel Dekker,


New York, 1972), p. 129.
263. B. E. Mann, C. Masters, and B. L. Shaw, f. Chern. Soc., Chern. Cornrnun., 846 (1970).
264. M. Yagupsky and G. Wilkinson, f. Chern. Soc. (A), 941 (1970).
265. R. Brady, W. H. deComp, B. R. Flynn, M. L. Schneider, J. D. Scott, L. Vaska, and M.
F. Werneke, Inorg. Chern. 14, 2669 (1975).
266. P. Pino, f. Organornet. Chern. 200,223 (1980).
267. T. Hayashi, M. Tanaka, and J. Ogata, f. Mol. Catal. 6, (1979).
268. M. Orchin, Accounts Chern. Res. 14,209 (1981).
269. F. E. Paulik, Cat. Rev. 6, 49 (1972).
270. D. A. von Bezard, G. Consiglio, F. Morandini, and P. Pino, f. Mol. Cat. 7,431 (1980).
271. N. H. Alemdaroglu, J. M. L. Penninger, and E. Oltay, Monats. Chern. 107, 1043 (1976).
272. F. Ungvary and L. Marko, f. Organornet. Chern. 193,303 (1980).
273. N. H. Alemdaroglu, J. M. L. Penninger, and E. Oltay, Monats. Chern. 107, 1153 (1976).
274. R. F. Heck and D. S. Breslow, f. Arn. Chern. Soc. 83,4023 (1961).
275. J. Csizmadia, F. Ungvary, and L. Marko, Trans. Met. Chern. 1, 170 (1976).
276. F. Ungvary, Acta Chirn. Hung, 111, 117 (1982).
277. R. Whyman, 1. Organornet. Chern. 94, 303 (1975).
3
Structurally Characterized
Transition-Metal Phosphine
Complexes of Relevance to
Catalytic Reactions
Nancy L. Jones and James A. Ibers

1. INTRODUCTION

A catalytic reaction that occurs under homogeneous conditions can be


probed by spectroscopic and kinetic methods, but hardly by crystallographic
techniques. Why then include in a book concerned with homogeneous
catalysis a chapter on the structures of isolable transition-metal phosphine
complexes? There are several answers. One is that such structures provide
details, albeit indirectly, on the substrate-metal interaction that occurs in
such a reaction. How this interaction is affected by the nature of the metal
and its ligands can be probed macroscopically by kinetic measurements; it
can also be probed at the molecular level in a qualitative way by spectro-
scopic methods. But if one seeks metrical details about the interaction,
then one must turn to diffraction studies on stable, isolable analogues of
the original system. Of course, one must employ spectroscopic methods to
establish that the analogue is indeed a good model for the temporal species
in the catalytic reaction. A second answer is that such structural information
is an aid to the formulation of reaction schemes, some of which are based

Nancv L. Jones and James A. Ibers • Department of Chemistry, Northwestern University,


Evanston, Illinois 60201.

111
112 NANCY L. JONES AND JAMES A. IBERS

on minimal data. In such schemes there frequently are proposed intermedi-


ates of new or unusual structures. Known structures may serve as a basis
for such speculation. Sometimes, of course, a scheme may precede the
isolation and characterization of a new structural type. In such instances
the ensuing characterization of a useful analogue supports the credibility
(or perhaps minimizes the incredibility) of the scheme. As an example,
transition-metal allyl hydride complexes have been proposed as intermedi-
ates in a number of catalytic reactions, including olefin isomerization, (1)
olefin metathesis,(2) and rearrangements of strained rings.(3) But it was not
until 1978 that a stable transition-metal allyl hydride was synthesized and
characterized.(4) This isolation and characterization lends credibility to the
various reaction schemes.
In this chapter we tabulate structurally characterized transition-metal
phosphine complexes that are analogues of some of the intermediates sug-
gested in various catalytic schemes. While a tabulation of metrical details
may prove useful for a given class of compounds, e.g., metallacycles, the
range of metal phosphine complexes and the types of reactions in which
they enter are so vast that tabulations of metrical information appear to
serve little purpose in general. Rather, we have chosen to tabulate com-
pounds by their type and, except for metallacycles, to provide minimal
metrical details. We believe that these tabulations are reasonably complete.
They are derived from computer searches of the Cambridge Crystallo-
graphic Data File (January 1981 version) along with manual searches of
the more recent literature up to January 1982. The literature is indeed
vast, there being approximately 2500 known structures of transition-metal
phosphine complexes. Nevertheless, it will be apparent from the tabulations
that there is a critical lack of concerted, systematic, high-quality studies
on metal phosphine complexes of relevance to catalytic reactions. Perhaps
this chapter will encourage such studies.

2. ALKENE HYDROGENA TION

The most active metal phosphine complexes for the selective catalytic
hydrogenation of alkenes and alkynes in the presence of other functional
groups are RhCI(PPh 3 h,(5) RuCh(PPh 3 h,(6) and RhH(CO)(PPh 3 h.(7) The
structures of these three complexes are known.(8-10) Scheme 1(5.6) displays
typical reaction pathways proposed for these catalytic reductions. Although
the specific details of the reactions are not known in all cases, and are
particularly sketchy for the Ru system, there is general agreement that the
key intermediates are Tr(P)(H)(olefin) and Tr(P)(H)(alkyl), where P =
phosphine and Tr = transition-metal (Sc-Zn, Y-Cd, Hf-Hg, La-Lu, Ac- ).
STRUCTURALL Y CHARACTERIZED COMPLEXES 113

RuCh(PPh 3 h

H'~'I,{H'
RuCI(H)(PPh 3 h

,ik'" ") / \ ("'"


A2 ~
RuCI(alkyl)(PPh 3 h .-c- - - - - - RuCI(H)(olefin)(PPh 3 h + PPh 3

H
"c=c /
Ph 3 P, I
Rh
/H

/~ Cl: I ':' PPh 3


/
C=C,

S = Solvent

Scheme 1. Alkene Hydrogenation

It is presumed that formal hydride migration to the 17'- bound olefin to


afford the (J'- bound alkyl is facilitated if hydride and olefin ligands are
initially cis to one another.
What can we learn from the structural literature on isolable compounds
related to these intermediates? It turns out that very little can be learned.
~

Table 1. Tr(P)(Hj(olefinj and Tr(P)(Hj(7T-ally/)"

Compound Geom. d" CN Description Reference

1. [MoH(HzC=CHz)z(PhzPCH=CHPPhz)zl pent bp d4 7 H not located 11


[CF 3 CO zl one olefin in eq
plane, one ~ to plane

I'p
P, IA
Mo
Y PI 'P
0
2. Ir(H)[(CNlzC=C(CNlzl(CO)(PPh 3 lz tbp d8 5 H not located, but 12 ~
cis to olefin; olefin
<
C")
-<:
lies in eq plane
r-
t....
0 a
p'~-JC m
p/I :t>
H <
C:l

dB
:t>
3. Ir(H)(CHz=CH-CH=CHz)(P(i-Prhlz dist tbp 5 H located, Ir-H 1.8 A, 13 s::
'Y/ 4 -butadiene m
:t>
4. [Ru(H)(CH 2 =CH-CH=CH z)(PMe zPhhl[PF 6 ] oct d6 6 H not located, P's facial, 14
'Y/ 4 - butadiene [Xi

sq pi dB ~
U)
5. [Pt(H)(CH z=C=C(CH 3lz)(PCY3h][PF 6 l 4 H not located, but trans to allene 15
6. IrCl(H)(1J 3-H2C~CH~CHPh)(PPh3h oct d6 6 H located, Ir-H 1.5 A 4 tI)

H Ph
~
~
P "ir~~
/1 ? ~
P CI ~
r-
r-
-.::
C')
"In this and ensuing tables the following conventions and abbreviations are used: ~
d" represents the d-electronic configuration of the metal.
CN is the coordination number of the metal. In defining CN a 7T-bound olefin or 7T-allyl is counted as a monodentate ligand and a cyclopentadienyl as a tridentate ligand.
~
C')
pent pentagonal eq equatorial ill::Q
bp bipyramid fac facial
tbp trigonal bipyramid dist distorted
N
oct octahedral sq pi square planar ~
reet rectangular sq py square pyramid
Ph = C6 H 5
8
i-Pr = CH(CH 3h ~
Cy = cyclohexyl
Me = CH 3 ~
mnt = S(CN)C=C(CN)S m
P3 N = N(CH 2CH 2PPh 2)3
n-Bu = CH 2CH 2CH 2CH 3
Et = C 2H 5
p-tol = p-C 6 H.CH3
dmpe = (CH3)2PCH2CH2P(CH3)2
PPW = [Ph 3P=N=PPh 3J+

~
116 NANCY L. JONES AND JAMES A. fBERS

There appear to be no known structures of Tr(P)(H)(alkyl) complexes.


Table 1 tabulates Tr(P)(H)(olefin) complexes. Even for these the tabula-
tion is very small and the results are unsatisfactory. There are only two
complexes involving a mono-olefin, 1 and 2. In neither of these structures
was the position of the hydride ligand located. But on the basis of the
stereochemistry it is clear that H is cis to olefin in 2. In fact, compound 2
represents the best (and only) model for the generally proposed intermedi-
ate. That it contains an activated olefin, tetracyanoethylene, and Ir rather
than Rh may be the reason for its stability.
Also included in Table 1 are known structures that involve diolefins.
Complexes 3 and 4 both have coordinated butadiene. Compound 4 is rather
poorly defined: the butadiene ligand appears to be 1/ 4 coordinated; the
hydride ligand was not located. Another diolefin, l,l-dimethylallene, is
coordinated to a PtHP2 center in compound 5. The H ligand, though not
located, is presumed to be trans to the allene, which is coordinated to the
Pt center through the C(2)-C(3) bond.
Several transition-metal phosphine complexes, including
RuClz(PPh 3 h, are effective for the selective reduction of dienes to mono-
olefins. (16) It is generally believed that this selectivity arises from the
exceptional stability of an intermediate 1T-allyl complex. The key intermedi-
ate in this process may be a 1T-allyl hydride complex. Complex 6, as we
indicated in the Introduction, is the only known example of such a complex.
The hydride ligand is cis to the allyl group.

3. HYDROFORMYLA TlON

Hydroformylation, the addition of Hand CHO to an olefinic double


bond, is the oldest and largest industrial process that involves homogeneous
transition-metal catalysts. The original catalyst was derived from
CO2(CO)8. (17,18) Later, phosphine modification of the system was introduced
to afford better product selectivity and lower operating pressuresY9) Still
later, related Rh systems were introduced.(20)
The considerable mechanistic work on these systems leads, for
example, to Scheme 2 for RhH(CO)(PPh J )2.(2oa) With minor modifications
a similar scheme applies to the CO 2(CO)8 and phosphine-modified Co
systems.(17-19) Thus the important intermediates are: (i) M(H)(CO)L 2 ; (ii)
M(H)(CO)(0Iefin)L 2 ; (iii) M(CO)(R)L 2 ; (iv) M(CO)(COR)L 2 ; and (v)
MH 2 (COR)L 3 , where M = Co or Rh and L = CO or phosphine. We will
now point out the paucity of structural analogues for these various species.
Table 2 summarizes the known structures of Tr(P)(H)(CO) com-
plexes, among which could occur (i) M(H)(CO)L 2 • In fact, the presumed
STRUCTURALL Y CHARACTERIZED COMPLEXES 117

co
y

lrll.
H
OC I
'Rh-II
Ph 3 P/ I --R
CO

//
OC" /CO
Rh,
Ph 3P/ CH 2 CH 2 R
PPh 3

Scheme 2. Hydroformylation

precursors to M(H)(CO)L 2 do occur as the dB 5-coordinate species


M(H)(CO)L 3 , M = Co (7),(21) Rh (8),(10) and Ir (9),(22) L = CO or PR3.
In each of these structures the position of the hydride ligand was deter-
mined. As may be seen from 7 and 9, two different trigonal bipyramidal
arrangements about the metal center occur, depending upon L. The sensitiv-
ity of trigonal bipyramidal coordination to ligand type is well-known;(35)
it may be that the sensitivity of the geometries of some of the proposed
intermediates (Scheme 2) to phosphine substitution is an important aspect
of such substitution. Presumably loss of the fifth ligand (either CO or P)
in these M(H)(CO)L 3 systems results in all instances in the desired dB
ex;

Table 2. Tr(P) (H){CO)

Compound Geom. d" CN Description Reference

0
P C
7. Co(H)(CO)(PPh 3 h tbp d8 5 "I
P's equatorial, Co-H 1.4 A Co-P 21
/1
P H
~
8. Rh(H)(CO)(PPh 3 h tbp d8 5 P's equatorial, Rh-H 1.6 A 10 ~
C)
-<::
r-
OC H <.....
"I Q
9. Ir(H)(COh(PPh 3 h tbp d8 5 H, P axial, Ir-H 1.6 A Ir-P 22 ~
OC/I m
P
:b.
10. Os(H)(N2 C 6 Hs)(CO)(PPh 3 )3 disttbp dB 5 Os-H = 1.2 A, P's trans 23
~
11. Ir(H)(Br)(C6 H s )(CO)(PEt 3 h oct d6 6 P's trans, H trans to Br 24 ~
12. Ir(H)(Si(CH3 hOSi(CH 3 h)(CO) oct d 6 6 H not located, P's cis, eq, H trans 25
s::
-(PPh 3 h to CO
m
~
13. Os(H)(Br)(CO)(PPh 3 h oct d6 6 P's meridional, H not 26
05
located, Br trans to CO
~
C/)
C/)
14. Os(H)[ C(S )SCN (Me )(p-tol)] oct d6 6 P's trans, H assumed trans 27
-(CO)(PPh 3 h toCS
il
c:
C')
15. Os(H)(CSzCH 3 )(COh(PPh 3h oct d6 6 Os-H 1.6 A, P's trans, H, co trans 28 2
16. Os(H)(CI)(CO)(SOZ)(PCY3h oct d6 6 H trans to SOz, CO trans to Cl 29 ~
r-
r-
17. Ru(H)[N(p-tol)NN(p-tol)] dist oct d6 6 P's axial, Ru-H (disordered), 30 -.::
C')
-(CO)(PPh 3 h 1.8 A
~
18. Ru(H)[N (p-tol)CHN (p-tol)] dist oct d6 6 P's axial, H, CO cis 31
~
C')
-(CO)(PPh 3h
dist oct 6 32 ~
19. Mn(H)(COh(PMePhzh d6 P's trans, Mn-H, 1.5 A ~
N
20. [Ir(H)(CO)(P(OMeh»)z rect py d7 5 P axial, Ir-H, 1.7 A (assumed) 33 t:
-(IL-SMeh C')
P /S P Q
H-Ir- S - 1r - H
" "/ s::
/ ;:!!
OC "CO ~
21. Ta(H)(COh(dmpejz dist d4 7 H capping 34 m
capped oct

co
~

11012 Table 3. Tr(P) (CO)(ole fin j

Compound Geom. d" CN Description Reference

22. Ir(Br)[(CNhC=C(CNh](CO)(PPh 3 h tbp d8 5 Olefin in eq plane with P's 36


23. Ir[(CNhC=C(CNh] tbp dB 5 Olefin in eq plane, CO, N apical 37
-(N =C=C(CN)CH(CNh](CO)(PPh 3 h
24. CO[1] 2-CH 2 =CHCH 2 CF(CF 3 )CF(CF 3 )] tbp d8 5 Olefin lies in eq plane 38
-(COh(P(OMeh) ~
( C CO <:
1/ ("")
~-Co -.::
1 "CO r-
P
<.....
Q
25. Fe[PF 2N (Me )(:(POF 2)=C(Me )CH=CHPhl 3-legged d 8 5 4C chain nearly planar 39 <:
-(CO)[PF 2 N(Me)(PF 2 )] piano m
~C-C~ ~
stool <:
CY'" /~C
CJ
Fe "
/1" /N i;:
PF2 C P
/ 0 ~
N m
PF2/ "Me ~
OJ
26. Fe(1] 4-PhCH=CHCH=OHCOh(PPh 3 ) sq py d8 5 P apical, 2 CO's and 1]4-diene 40
~
CJ)
in basal plane
27. Fe(COh(PPh 3 )(/L-CH z=C=C=CH 2) dist oct d6 6 2 CO, P facial, P trans to Fe-Fe (f)
41
-Fe(CO), 7T- allyl bonding to each Fe :il
c::
C")
C, C Fe
C/-::::- / CO 2
/C, / /.c;:.J::--C ~
P-Fe--Fe-CO l"-
c' !,C I"-
~
OC /1C " CO
Fe C")
o
~
28. Ru(Clh(H2C=CH2)(CO)(PMe2Ph)z oct d6 6 Cl trans to olefin, trans P's, 42 ~
C")
olefin parallel to P's i;:j
J;)
o N
OC C p J § J ~
29. Mb('1 2 -CH 3 CH=CHC 6 H 4 PPh z)(CO)4 oct d6 6 "IMoI 0 43
OC/ ~ \C
o 'I!
c ~~
30. [Mo(COh(P(n-BuhhJz oct d6 6 Olefin ~ Mo(COh plane 44 m
-(/L-CH 2 =CHCNh
OC P N-C-C P CO
"1/ 11,,1/
~1o C C Mo
/1 "II /1 "
OC P C-C-N P CO

I\.)
--
122 NANCY L. JONES AND JAMES A. IBERS

4-coordinate M(H)(CO)Lz (i) species. The other Tr(P)(H)(CO) com-


plexes of Table 2 have no direct relevance to hydroformylation.
The general structural type Tr(P)(H)(CO)(olefin) encompasses
intermediate (ii), M(H)(CO)(olefin)L z. There is only one such example,
namely Ir(H)[(CN)zC=C(CN)z](CO)(PPh3 )z, 2, a structure known since
1969! This complex has the H and olefin ligands cis, as is proposed in
Scheme 2.
Not only is the structural class Tr(P)(H)(CO)(olefin) extremely
limited (one example!), but the less restricted class Tr(P)(CO)(olefin) is
also very limited, with only nine examples. For the sake of completeness
we tabulate this class in Table 3. Among these only 22, a dB 5-coordinate
species that contains Br rather than H, is even vaguely related to the
proposed intermediates in hydroformylation.
There are no known structural analogues of (iii), M(CO)(R)L z, M =
Co, Rh. The structural class Tr(P)(CO)(R), Table 4, is very limited. The
three known examples are all d 6 6-coordinate complexes with R = Me.
Intermediate (iv), M(CO)(COR)L 2 is represented by Table 5, which
tabulates complexes of the type Tr(P)(CO)(COR). But as can be seen,
none of the known structures involves Co, Rh, or Ir. Even if we turn to a
less restricted class, namely Tr(P)(COR) compounds (Table 6), we find a
suprisingly small tabulation in view of the interest in transition-metal acyl
complexes. Although no dB 4-coordinate complexes of the Co triad are
found, several d6 complexes of the triad occur, compound 44 being perhaps
the closest (but distant) analogue of intermediate (iv).
There are no known structures of the type Tr(P)(H)(COR) and
hence no analogues of intermediate (v), MH 2 (COR)L 2 • However, a number
of d 6 6-coordinate complexes of the Co triad are found in the various
tables. These complexes are generally unremarkable; the propensity of the
Co triad to form such complexes is well known.

4. OLEFIN OLIGOMERIZA TlON

Although olefin oligomerization, especially polymerization, is an


exceedingly high-volume industrial process, most of the commercial
catalysts do not contain phosphines. However, one of the most active
catalysts is obtained by treating NiX2L2 (X = anionic ligand, L =
phosphine) with an aluminum alkyl. (64.65) Such systems are believed to
involve Ni(H)( Y)L, where Y is a complex Lewis-acid anion of the type
X AIR xX 3 - x • A possible reaction scheme is shown in Scheme 3.(66) A key
intermediate is of the type Tr(P)(R)(olefin) where olefin and R are cis
to one another. Table 7 tabulates the known structures of this type. The
relevance to the Ni-mediated catalysis is slight. The three compounds do
CI)

:il
c::
C)

2
~
r-
r-
-<:
C)

~
~
Table 4. Tr(P) (CO)(R) C)

ill~
Compound Geom. d" CN Description Reference l\i
~
31. [Rh(C 5 H 5 )(CH 3 )(CO)[(S) - (P(C 6 H 5 h oct d6 6 CH 3, CO, P fac 45 C)
-NHCH(CH 3)(C 6 H 5 )]][BF4] 0

0 ~
r-
P C X X Me ~
32. [Rh(wX)(X)(CH3)(CO)(PMe2Ph)]2, oct d6 6 ,,-1/,,-1/ 46
Rh Rh m
X = CI,Br /1"-/1"-
Me X X C P
0

33. Mn(CH 3)(CO)4(PPh 3) oct d6 6 P, CH 3 cis 47

~
Table 5. Tr(P)(CO)(COR) ~
Compound Geom. d" CN Description Reference

OOC P cO
34. [PPN][Fe(COCH 3)(COh sq py dO 6 II "~P~ /
CH -C-Fe-Fe-CO 48
-(/L- PPh 2hFe(COh] 3 / "
OC Co

35. [Na][Fe(COCH 3)(COh(/L-PPh 2hFe(COh] sq py dO 6 48


36. Fe(1) 5 -1-Me-3-Ph-CsH3)(COCH3)(CO) oct dO 6 CO, P, acyl fac 49
-(PPh 3 )
37. Fe[C(OCH 3)CH(C0 2CH 3)CH zCO] oct dO 6 Cyclopentenone, acyl, CO in ring 50
(COh(PPh 3) plane, 3 CO's fac
OMe ~
o P 1 ~
C" 1 /C""=c/ R
Fe I "'"'
r-
/1 \
OC C C~CHz c....
o II a
~
o m
38. Ru(I)( 1)2 -COCH 3)(CO)(PPh 3h sq py dO 5 P's trans, 1) 2-acyl apical 51 :t.
39. [Mo(/L-CI)(1)2-0QCH2SiMe3)) oct d4 6 1)2-acyl CO, P trans to 1)2-acyl, 52 ~
-(COh(PMe3)lz 2 CO's, 2 /L-CI'S in eq plane
~
40. Mo(C s H s )(COCH 3)(COh(PPh 3) 4-legged d4 7 P trans to acetyl, 53
piano CsHs is stool seat ~
stool :t.

41. V(1)2_0C(C3H2Ph3))(COh oct d3 6 1)2-acyl CO, 3 CO's facial 54


OJ
-(Ph 2PCH 2CH 2AsPh 2) ~
CJ')
Table 6. Tr(P)(COR)
CI)

Compound Geom. d" CN Description Reference :il


c::
\)
42. Rh(l)( COCH3)(pMe2Ph)(1L -SCH 3hRh( CO) sq py d6 5 Acyl apical, sq py Rh joined to 55 ~
-(PMe2Ph) sq pI Rh through S ... Sedge ~
r-
r-
43. Rh(COCH 2CH 2CH 3)(mnt)(PEt3h sqpy d6 5 Acyl apical 56 -<:
\)
44. [Rh(CI)(COCH3)(PMe2Phh][PF6] sq py d6 5 Acyl apical 57
~
45. [AsPh4 ][Rh(l)(mnt)(COCH2CH 3)(PPh 3)] sqpy d6 5 Acyl apical 58 ~
\)
46. [[Rh(COCH3)(PMe2Phh]2(1L- Clh][PF6] dist oct dO 6 3 Cl's fac 59 ill::Q
47. Rh(C sH s)(I)(COCH 3) oct d 6 6 CsHs occupies fac sites 45 N
-[(S)-(P(C 6HshNHCH(Me)(C 6Hs)] gj
\)
48. Ir(CsHs)[IL-C(C6Hs)0][wC(CH3)0] Mnoct d4 6 60 Q
-[IL- P(C 6H s h]Mn(COh Ir oct d6 5 s:
P ;!!
OC" / ~ ~
OC-Mn-O-C-Ir~ m
OC
/ \ o - c ''7M e
,
Ph
49. Ni(CI)(COCH 3)(PMe3h sqpl dB 4 P's trans 61
50. Pd(CI)(COCH2CH 2CH 3)(PPh 3h sqpl dB 4 P's trans 62
51. Pt(CI)(COCH 2CH2CH 3)(PPh 3h sq pI dB 4 P's trans 62
52. [Ni(COCH 3)(P3N)][BPh4 ] sq pI dB 5 P's eq, N trans to acyl 63

("N)+
P£~i-P
P I ~
Ol
",C,
0"" CH 3
126 NANCY L. JONES AND JAMES A. IBERS

CH 2 =CH 2
I

l
H-MX(L)

')
CH 2 =CH 2

Scheme 3. Oligomerization of Olefins

not contain Ni. However, each has the olefin and R ligands cis to one
another.

5. METALLACYCLES

Metallacycles have been suggested as intermediates in many


transition-metal catalyzed reactions of ole fins, acetylenes, and cyclo-
propanes. Metallacyclobutane complexes are invoked in olefin and
cyclopropane isomerization schemes(3) (Scheme 4(4)), as well as in olefin
metathesis schemes. (70,71) Metallacyclopentane, -pentene, and -pentadiene
complexes can all be invoked in olefin and acetylene dimerization and
polymerization(72-78) (Scheme 5(77)). Many of these involve early transition
metals and do not include phosphine ligands.
Several platinacyclobutane complexes have been isolated from the
reaction of platinum-phosphine complexes with cyclopropanes. Those
structurally characterized are listed in Table 8. No other structures of
metallacyclobutane-phosphine complexes have been reported, but an
(I)

:ti
~
C:!
~
l"-
I"-
-<:
C)

~
~
C)

Table 7. Tr(P)(R)(olefin)
n1::I)
N
Compound Geom. d" CN Description Reference
~

dB
8
53. Co(C6H5)(H2C=CH2)(PMe3h tbp 5 Olefin and C6H5 cis and coplanar, 67
2 P's axial ~
54. ir1P(CH(CH~)CH2)(i-Prhl(C2H4h tbp dB 5 Olefin and CH 2 cis, olefins and 68
~
(P(i-Prhh metallated P in eq plane
m
55. cis-Pt(CH3h[P(O-C6H4CH~CH2)(Phhl sq pI dB 4 Olefin and CH 3 cis, olefin .1 to sq pI Pt 69

I \)
'-I
-
128 NANCY L. JONES AND JAMES A. IBERS

Scheme 4. Olefin and Cvclopropane Isomerization

Scheme 5. Acetvlene Trimerization


STRUCTURALL Y CHARACTERIZED COMPLEXES 129

iridiacyclobutane arsine complex, Ii(H)(CH2C(CH3)zC'~H2)(AsMe3h, is


known. (94) As may be seen in Scheme 4, migration of the f3 -hydrogen atom
from the ring to the metal is believed to occur upon isomerization of the
metallacycle to the corresponding transition-metal allyl hydride complex.
It may be that transfer of the f3 -hydride atom from ring to metal is facilitated
as the pucker angle increases. (73,74) For the platinacyclobutanes this angle
ranges from 22.40 to 49.7 0 and generally increases as the substitution on
the ring increases. Of course, in these solid state structures the pucker
angle may be merely a manifestation of packing forces. In solution the
barrier to ring inversion is small(71) and greater puckering may be possible.
The formation of a metallacyclopentane from two moles of ethylene
has been observed with nickel phosphine complexes. (72) Metallacyclopen-
tanes are often included in mechanisms for the dimerization of ole fins.
Table 8 contains four Pt complexes, as well as one each of Co, Ni, and Ir.
Metallacyclopentene and -pentadiene phosphine complexes that have
been structurally characterized are also included in Table 8. These may be
relevant to the dimerization of two acetylenes or acetylenes with 0Iefins(75,76)
and to the cyclotrimerization of acetylene with an olefin. (78)

6. SUMMARY

This chapter has presented a tabulation of transition-metal phosphine


complexes bearing some relation to proposed intermediates in a number
of catalytic reactions. It is an interesting commentary on structural inorganic
chemistry that, despite the determination of approximately 2500 crystal
structures of such complexes, there are so few structures of direct relevance
to such processes. For example, there is but one direct structural analogue
(compound 2) for any of the intermediates proposed in the hydroformyla-
tion reaction, the most important and oldest industrial process based on a
homogeneous transition-metal phosphine complex. Similarly, the structure
of only one transition-metal allyl hydride is known,(4) despite the fact that
allyl hydrides are frequently proposed intermediates. This paucity of struc-
tural information does not necessarily result from the inherent reactivity
of such intermediates-for example the synthesis of suitable analogues for
the hydroformylation reaction would appear to be straightforward. There
appear to be several potential routes to complexes of the type
MH 2 (COR)L 3 , and one would expect that complexes of the type M(H)-
(CO)L 2 would be stable, especially for L, a bulky phosphine.
There are several examples of metallacycle structures of direct
importance to catalytic processes, but such structures only serve to
emphasize another problem, that of the relation of structure to reactivity.
~

Table 8. Tr(P)(metallacycle)

Compound Geom. d" CN M-C(1)" M-C(3) C(1)-C(2) C(2)-C(3) C(3)-C(4) Description b Reference c

A, Metallacy~lobl1tanes M-C(1)-C(2)-C(3)

56. Pt[C(CN)2CH2C(CN),](PPh 3)2 sq pi d' 4 2.137(6)d 2.139(6) 1.545(9) 1.584(9) 24.4° 79

57. Pt[CH(C0 2CH 3)C(O)CH(C0 2CH 3)] sq pi d' 4 2.149(6) 2.128(6) 1.496 1.456 49.7° 80
-(PPh3)2

SR. Pt[CICN)(C0 2CH 2CH,)CHIC oH,)CICN),l sqpl d8 4 2.158(14) 2.200(14) 1.556(19) 1.509(14) 29.7° 81
-(PPh')2

59. Pt[C(CN),CH(C 6H s )C(CN),](PPh 3)2 sq pi d8 4 2.137(6) 2.159(6) 1.557(9) 1.5.48(10) 28.6° 81 ~


<:
C)
60. Pt[CH2C(CH3)2CH2](PEt3)2 sq pi d8 4 2.086(6) 2.080(6) 1.535(9) 1.536(9) 22.4° 82*

13· Metallacyclop"n1.alles M-C( 1 )-Q2J=-CD~ill)


"""
:-
<-
Q
61. Pt(IJ,(CH2CH2CH2CH,)(PMe2Ph)2 dist oct db 2.15(1) 2.15(1) 1.54(2) 1.49(4) 1.54(2) 83
<:
62. Pt[CH(COCH 3)CH 2CH(COCH 3)CH,] sqpl d8 4 2.139(7) 2.094(9) 1.540(11) 1.579(23) 1.430(20) 84 m
-(PPh 3)2 ~
63. Pt[CH(CH=CH 2)CH,CH 2CH(CH=CH 2)] sqpl d8 4 2.142(8) 2.131(5) 1.52(1) 1.52(1) 1.54(1) 85*
<:
t:)
-(PMe 3),
~
64. Pt(CH2CH2CH2CH2)(PPh31z sq pi d8 4 2.12(2) 2.05(2) 1.57(3) 1.45(3) 1.48(3) 86 s::
65. Ni[CH 2C(CH 2)C(CH 2)CH 2] sq pi d8 4 1.975 d 1.965 1.47 1.47 1.48 Ring 87 m
-[(C6H,,),P(CH2IzP(C6H,,),] planar ~
66. Co(CsHs)(CH2CH2CH2CH2)(PPh3) dist oct db 6 2.02516) 2.024(7) 1.509 d 1.466 1.561 88 iii
67. Ir(Cs(CH3)s)(CH2CH2CH2CH,) dist oct d6 6 2.10(4) 2.14(5) 1.30 d 1.45 1.67 88 gj
(PPh 3) en
CI)
C. Metallacyclobuten<' LM-Ci1E=Cl2J,Cl3J

68. sq pi d" 4 2.09(4) 2.08(61 1.31(8) 1.45(7) 89


:iI
Pt[C(C6HsI=C(C6HsJC(OI)(PPh3h c::C")
!:LM«!~lIacyclol!ent-2-enes M -Clll-Ci21-Ci31-C(4j 2
69. CO(C,H,)[C(C 6H s )=C(C0 2CH 3)CH distoct d6 6 1.947(11) 2.079(12) 1.375(17) 1.527(13) 1.513(15) Co 0.583 A 90
~
r-
-(C02CH3)CH(C02CH3»)(PPh3) Qutot r-
CoC 3
C")
""
plane

2.001(4) 2.005(4) 1.349(5) 1.499(5) 1.509(5)


~
70. Fe[C(OCH 3)=C(C0 2CH 3)CH 2CO)(CO)3 oct d6 6 50
-(PPhMe 2) ~
C")

E.Metallacyclol!ent-3-eneM -Cll )-Ci21-031-CJ±j


n1
:l)

71. RhCl[C(O)C(Cl)=C(Cl)ttO))(H 2O) oct d6 6 1.970(6) 1.970(61 1.50(1) 1.32(1) 1.52(1) P's trans 91
N
-(PMe2Ph),
~

F. Metallacy~_ntadienes M-Ci 1 I=C(2)-ClJl=C(4)


8
72. Co(CsHS)[C(C6Fs)=C(C6Fs)C(C6Fs)=C distoct d6 1.995(11) 1.993(11) 1.326(15) 1.467(16) 1.335(15) Co 0.203 A 92 ~
-(C6F s )](PPh 3) Qutot ~
C 4 plane
m
73. Rh(CsHs)[C(C6Fs)=C(C6Fs)C(C6Fs)=C dist oct d6 6 2.060(12) 2.067(11) 1.457(16) 1.354(15) Rh 0.237 A 93
-(C6F s »)(PPh 3) Qutof
C4 plane

a Distances in A.
b The pucker angle given for metallacyclobutanes is the dihedral angle between C(1)-C(2)-C(3) and C(I)-M-C(3).
C An asterisk on the reference indicates a low-temperature structure determination (-70°C or below).
d Estimated standard deviation, if available, is in parentheses.

t..l
--
132 NANCY L. JONES AND JAMES A. IBERS

From the structures of the platinacyclobutanes and the one iridia-


cyclobutane it is not at all evident why the synthesis of the Pt complex,
60, requires forcing cortditions, (82) whereas the Ir analogue is formed
easily.(94) Moreover, even among the platinacyclobutanes the relation, if
any, of solid-state structure, especially the pucker angle, to reactivity is
unclear. Of course, structure-reactivity relationships are at the very founda-
tion of our potential understanding of such catalytic processes. At least in
the area of transition-metal phosphine complexes it seems clear that such
an understanding will not emerge without considerably increased effort in
the isolation and structural characterization of carefully tailored, model
complexes.

ACKNOWLEDGMENTS

We thank Ms. Jan Goranson for skillful typing of the manuscript, Ms. Jean
Wisner for thorough proofreading, and Dr. Paul Swepston for assistance
with the structural search. JAI thanks the Sherman Fairchild Distinguished
Scholars Program of the California Institute of Technology under whose
auspices a portion of this review was written.

REFERENCES

1. For reviews of transition metal-catalyzed olefin isomerizations see (a) C. A. Tolman,


Transition Metal Hydrides, edited by E. L. Muetterties (Marcel Dekker, New York, 1971)
pp. 271-312; (b) A. J. Hubert and H. Reimlinger, Synthesis 1, 405-430 (1970); (c)
N. R. Davies, Rev. Pure Appl. Chern. 17,83-93 (1967); (d) M. Orchin, Adv. Catal. 16,
1-47 (1966).
2. M. Ephritkhine, M. L. H. Green, and R. E. MacKenzie, I. Chern. Soc., Chern. Cornrn.
1976,619-621.
3. For a review, see K. C. Bishop III, Chern. Rev. 76,461-486 (1976).
4. T. H. Tulip and J. A. Ibers, I. Am. Chern. Soc. 101,4201-4211 (1979).
5. J. A. Osborn, F. H. Jardine, J. F. Young, and G. Wilkinson, I. Chern. Soc. A 1966,
1711-1732.
6. P. S. Hallman, B. R. McGarvey, and G. Wilkinson, I. Chern. Soc. A 1968,3143-3150.
7. C. O'Connor and G. Wilkinson, I. Chern. Soc. A 1968, 2665-2671.
8. P. B. Hitchcock, M. McPartlin, and R. Mason, I. Chern. Soc., Chern. Cornrn. 1969,
1367-1368; M. J. Bennett and P. B. Donaldson, Inorg. Chern. 16,655-660 (1977).
9. S. J. La Placa and J. A. Ibers, Inorg. Chern. 4,778-783 (1965).
10. - - , Acta Crystallogr. 18, 511-519 (1965).
11. J. W. Byrne, J. R. M. Kress, J. A. Osborn, L. Ricard, and R. E. Weiss, I. Chern. Soc.,
Chern. Cornrn. 1977, 662-663.
12. K. W. Muir and J. A. Ibers, I. Organornet. Chern. 18, 175-187 (1969).
STRUCTURALL Y CHARACTERIZED COMPLEXES 133

13. G. Del Piero, G. Perego, and M. Cesari, Gazz. Chirn. Ital. lOS, 529-537 (1975).
14. T. V. Ashworth, E. Singleton, M. Laing, and L. Pope, J. Chern. Soc., Dalton Trans.
1978,1032-1036.
15. H. C. Clark, M. J. Dymarski, and N. C. Payne,J. Organornet. Chern. 165, 117-128 (1979).
16. (a) J. Tsuji and H. Suzuki, Chern. Lett. 1977, 1083-1084; (b) A. Andreetta, F. Conti,
and G. F. Ferrari, Aspects on Hornogeneous Catal. I, 203-267 (1970).
17. R. F. Heck and D. S. Breslow, I. Arn. Chern. Soc. 83,4023-4027 (1961).
18. N. H. Alemdaroglu, J. L. M. Penninger, and E. Oltay, Monatshefte fiir Chernie 107,
1153-1165 (1976).
19. M. van Boven, N. H. Alemdaroglu, J. L. M. Penninger, Ind. Eng. Chem., Prod. Res.
Dev. 14, 259-264 (1975); L. H. Slaugh and R. D. Mullineaux, I. Organornet. Chern. 13,
469-477 (1968); E. R. Tucci, Ind. Eng. Chern., Prod. Res. Dev. 7, 32-38 (1968).
20. (a) C. K. Brown and G. Wilkinson, J. Chern. Soc. A 1970, 2753-2764; (b) R. L. Pruett
and J. A. Smith, J. Org. Chern. 34,327-330 (1969).
21. J. M. Whitfield, S. F. Watkins, G. B. Tupper, and W. H. Baddley, I. Chern. Soc., Dalton
Trans. 1977,407-413.
22. M. Ciechanowicz, A. C. Skapski, and P. G. H. Troughton, Acta Crystallogr., Sect. B 32,
1673-1680 (1976).
23. M. Cowie, B. L. Haymore, and J. A. Ibers, Inorg. Chern. 14, 2617-2623 (1975).
24. U. Behrens and L. Dahlenburg, J. Organornet. Chern. 116, 103-111 (1976).
25. M. D. Curtis, J. Greene, and W. M. Butler, J. Organornet. Chern. 164,371-380 (1979).
26. P. L. Orioli and L. Vaska, Proc. Chern. Soc., London 1962, 333.
27. G. R. Clark, T. J. Collins, D. Hall, S. M. James, and W. R. Roper, I. Organornet. Chern.
141, C5-C9 (1977).
28. J. M. Waters and J. A. Ibers, Inorg. Chern. 16,3273-3277 (1977).
29. R. R. Ryan and G. J. Kubas, Inorg. Chern. 17, 637-641 (1978).
30. L. D. Brown and J. A. Ibers, fnorg. Chern. 15,2788-2793 (1976).
31. L. D. Brown, S. D. Robinson, A. Sahajpal, and J. A. Ibers, Inorg. Chern. 16, 2728-2735
(1977).
32. M. Laing, E. Singleton, and G. Kruger, J. Organornet. Chern. 54, C30-C32 (1973).
33. J. J. Bonnet, A. Thorez, A. Maisonnat, J. Galy, and R. Poilblanc, J. Am. Chern. Soc.
101,5940-5949 (1979).
34. P. Meakin, L. J. Guggenberger, F. N. Tebbe, andJ. P. Jesson, Inorg. Chern. 13, 1025-1032
(1974).
35. B. A. Frenz and J. A. Ibers, "M.T.P. International Review of Science, Physical
Chemistry," Series One, 11, 33-72 (1972).
36. J. A. McGinnety and J. A. Ibers, J. Chern. Soc., Chern. Cornrn. 1968,235-237.
37. J. S. Ricci and J. A. Ibers, J. Arn. Chern. Soc. 93,2391-2397 (1971).
38. M. Bottrill, R. Goddard, M. Green, and P. Woodward, J. Chern. Soc., Dalton Trans.
1979,1671-1678.
39. M. G. Newton, R. B. King, M. Chang, and J. Gimeno, I. Arn. Chern. Soc. 101, 2627-2631
(1979).
40. M. Sacerdoti, V. Bertolasi, and G. Gilli, Acta Crystallogr. Sect. B 36,1061-1065 (1980).
41. J. N. Gerlach, R. M. Wing, and P. C. Ellgen, Inorg. Chern. 15,2959-2964 (1976).
42. L. D. Brown, C. F. J. Barnard, J. A. Daniels, R. J. Mawby, and J. A. Ibers, Inorg. Chern.
17,2932-2935 (1978).
43. H. Luth, M. R. Truter, and A. Robson, I. Chern. Soc. A 1969, 28-41.
44. F. Hohmann, H. T. Dieck, C. Kruger, and Y.-H. Tsay, J. Organornet. Chern. 171,
353-364 (1979).
45. S. Quinn, A. Shaver, and V. W. Day, J. Arn. Chern. Soc. 104, 1096-1099 (1982).
46. M. J. Doyle, A. Mayanza, J.-J. Bonnet, P. Kalck, and R. Poilblanc, I. Organomet. Chem.
146,293-310 (1978).
134 NANCY L. JONES AND JAMES A. fBERS

47. A. Mawby and G. E. Pringle, J. Inorg. Nucl. Chern. 34,877-883 (1972).


48. R. E. Ginsberg, J. M. Berg, R. K. Rothrock, J. P. Collman, K. O. Hodgson, and L. F.
Dahl,!. Am. Chern. Soc. 101, 7218-7231 (1979).
49. T. G. Attig, R. G. Teller, S.-M. Wu, R. Bau, and A. Wojcicki, J. Am. Chern. Soc. 101,
619-628 (1979).
50. T.-A. Mitsudo, T. Sasaki, Y. Watanabe, Y. Takegami, K. Nakatsu, K. Kinoshita, and Y.
Miyagawa, J. Chern. Soc., Chern. Cornrn. 1979, 579-580.
51. W. R. Roper, G. E. Taylor, J. M. Waters, and L. J. Wright, J. Organornet. Chern. 182,
C46-C48 (1979).
52. E. C. Guzman, G. Wilkinson, J. L. Atwood, R. D. Rogers, W. E. Hunter, and M. J.
Zaworotko, J. Chern. Soc., Chern. Cornrn. 1978, 465-466.
53. M. R. Churchill and J. P. Fennessey, Inorg. Chern. 7,953-959 (1968).
54. U. Franke and E. Weiss, J. Organornet. Chern. 165, 329-340 (1979).
55. A. Mayanza, J.-J. Bonnet, J. Galy, P. Kalck, and R. Poilblanc, J. Chern. Res. 146,
2101-2133 (1980).
56. C.-H. Cheng, D. E. Hendriksen, and R. Eisenberg, J. Organornet. Chern. 142, C65-C68
(1977).
57. M. A. Bennett, J. C. Jeffery, and G. B. Robertson, Inorg. Chern. 20, 323-330 (1981).
58. C.-H. Cheng, B. D. Spivack, and R. Eisenberg, J. Am. Chern. Soc. 99,3003-3011 (1977).
59. M. A. Bennett, J. C. Jeffery, and G. B. Robertson, Inorg. Chern. 20, 330-335 (1981).
60. J. R. Blickensderfer, C. B. Knobler, and H. D. Kaesz, J. Am. Chern. Soc. 97, 2686-2691
(1975).
61. G. Huttner, O. Orama, and V. Bejenke, Chern. Ber. 109, 2533-2536 (1976).
62. R. Bardi, A. M. Piazzesi, G. Cavinato, P. Cavoli, and L. Toniolo, J. Organornet. Chern.
224,407-420 (1982).
63. P. Stoppioni, P. Dapporto, and L. Sacconi, Inorg. Chern. 17,718-725 (1978).
64. G. LeFebvre and Y. Chauvin, Aspects of Homogeneous Catal. 1, 107-201 (1970).
65. R. G. Miller, P. A. Pinke, R. D. Stauffer, H. J. Golden, and D. J. Baker, J. Am. Chern.
Soc. 96, 4211-4220 (1974); P. A. Pinke and R. G. Miller, J. Am. Chem. Soc. 96,
4221-4229 (1974); P. A. Pinke, R. D. Stauffer, and R. G. Miller, J. Am. Chern. Soc. 96,
4229-4234 (1974); H. J. Golden, D. J. Baker, and R. G. Miller, J. Am. Chem. Soc. 96,
4235-4243 (1974).
66. B. Bogdanovic, B. Henc, H.-G. Karmann, H.-G. Niissei, D. Walter, and G. Wilke, Ind.
Eng. Chem. 62 (12), 34-44 (1970).
67. H.-F. Klein, R. Hammer, J. Gross, and U. Schubert, Angew. Chem. Int. Ed. Eng. 19,
809-810 (1980).
68. G. Perego, G. Del Piero, M. Cesari, M. G. Clerici, and E. Perrotti, J. Organomet. Chern.
54, C51-C52 (1973).
69. M. A. Bennett, H.-K. Chee, J. C. Jeffery, and G. B. Robertson, Inorg. Chern. 18,
1071-1076 (1979).
70. J.-L. Herisson and Y. Chauvin, Makromol. Chern. 141, 161-176 (1970).
71. R. J. Puddephatt, Coord. Chem. Rev. 33, 149-194 (1980).
72. R. H. Grubbs, A. Miyashita, M. Liu, and P. Burk, J. Am. Chern. Soc. 100,2418-2425
(1978); R. H. Grubbs and A. Miyashita, J. Am. Chern. Soc. 100, 1300-1302 (1978);
R. H. Grubbs and A. Miyashita, J. Am. Chern. Soc. 100,7416-7418 (1978).
73. J. X. McDermott, J. F. White, and G. M. Whitesides, J. Am. Chem. Soc. 98, 6521-6528
(1976).
74. J. X. McDermott, M. E. Wilson, and G. M. Whitesides, J. Am. Chern. Soc. 98, 6529-6536
(1976).
75. J. P. Collman, J. W. Kang, W. F. Little, and M. F. Sullivan, Inorg. Chern. 7, 1298-1303
(1968).
STRUCTURALL Y CHARACTERIZED COMPLEXES 135

76. A. J. Chalk, J. Arn. Chern. Soc. 94, 5928-5929 (1972).


77. D. R. McAlister, J. E. Bercaw, and R. G. Bergman, J. Arn. Chern. Soc. 99, 1666-1668
(1977).
78. L. D. Brown, K. Itoh, H. Suzuki, K. Hirai, and J. A. Ibers, J. Arn. Chern. Soc. 100,
8232-8238 (1978).
79. D. J. Yarrow, J. A. Ibers, M. Lenarda, and M. Graziani, J. Organornet. Chern. 70,
133-145 (1974).
80. D. A. Clarke, R. D. W. Kemmitt, M. A. Mazid, M. D. Schilling, and D. R. Russell, J.
Chern. Soc., Chern. Cornrn. 1978, 744-745.
81. J. Rajaram and J. A. Ibers, J. Arn. Chern. Soc. 100,829-838 (1978).
82. J. A. Ibers, R. DiCosimo, and G. M. Whitesides, Organornetallics 1, 13-20 (1982).
83. A. K. Cheetham, R. J. Puddephatt, A. Zalkin, D. H. Templeton, and L. K. Templeton,
Inorg. Chern. 15, 2997-2999 (1976).
84. M. Green, J. A. K. Howard, P. Mitrprachachon, M. Pfeffer, J. L. Spencer, F. G. A.
Stone, and P. Woodward, J. Chern. Soc., Dalton Trans. 1979, 306-314.
85. G. K. Barker, M. Green, J. A. K. Howard, J. L. Spencer, and F. G. A. Stone, J. Chern.
Soc., Dalton Trans. 1978,1839-1847.
86. C. G. Biefeld, H. A. Eick, and R. H. Grubbs, Inorg. Chern. 12,2166-2170 (1973).
87. P. W. Jolly, C. Kriiger, R. Salz, and J. C. Sekutowski, J. Organornet. Chern. 165, C39-C42
(1979).
88. P. Diversi, G. Ingrosso, A. Lucherini, W. Porzio, and M. Zocchi, J. Chern. Soc., Chern.
Cornrn. 1977, 811-812.
89. W. Wong, S. J. Singer, W. D. Pitts, S. F. Watkins, and W. H. Baddley, J. Chern. Soc.,
Chern. Cornrn. 1972, 672-673.
90. Y. Wakatsuki, K. Aoki, and H. Yamazaki, J. Arn. Chern. Soc. 101,1123-1130 (1979).
91. P. D. Frisch and G. P. Khare, J. Arn. Chern. Soc. 100, 8267-8269 (1978).
92. R. G. Gastinger, M. D. Rausch, D. A. Sullivan, and G. J. Palenik, J. Arn. Chern. Soc.
98,719-723 (1976).
93. - - , J. Organornet. Chern. 117, 355-364 (1976).
94. T. H. Tulip and D. L. Thorn, J. Arn. Chern. Soc. 103, 2448-2450 (1981) and private
communication.
4
Asymmetric Hydrogenation
Reactions Using Chiral
Diphosphine Complexes of
Rhodium
John M. Brown and Penny A. Chaloner

1. INTRODUCTION

The archetypal reaction in asymmetric homogeneous hydrogenation is


reduction of a z-acyldehydroamino acid to the corresponding amino acid
derivative, which represents one of the most efficient routes to introduction
of a chiral center, with optical yields often in excess of 98 %. Soon after
Wilkinson's discovery that tris(triphenylphosphine)rhodium chloride was
a homogeneous catalyst for alkene hydrogenation,(l) several authors real-
ized the potential for asymmetric induction. Early attempts using simple
resolved chiral phosphines were not impressive, (2) and the first major
development was due to Kagan and Dang. (3) They demonstrated that the
use of a chiral chelating biphosphine, (1) (see Figure 4.1), bearing the
now-famous acronym of mop, gave good optical yields, particularly in the
reduction of dehydroamino acid derivatives. The most startling aspect of

Dr. John M. Brown • Dyson Perrins Laboratory, South Parks Road, Oxford OXl 3QY,
England.
Dr. Penny A. Chaloner • School of Molecular Sciences, University of Sussex, Falmer,
Brighton, u.K.

137
138 JOHN M. BROWN AND PENNY A. CHALONER

1 5

Me,(p L:.
Ph 2 Ph 2 Ph 2
PPh2 C6H1 P
Ph,(:
PPh 2
P P
Q. Ph 2 9 Ph 2 Ph2
~ 1Q

Figure 4.1. Some of the commoner biphosphines which have been successfully employed in
asymmetric catalysis.

this result is that the chiral center in the catalyst which controls asymmetric
induction is three bonds removed from the metal and spatially remote from
it. This leads to an immediate suspicion that the chelate ring is in some
way controlling the orientation of the P-Ph rings, which in turn determines
the stereochemical outcome of the rate-determining stage in catalysis. A
second development, one of considerable commercial significance, was due
to W. S. Knowles and co-workers at Monsanto.(4) Their chelating biphos-
phine, DIPAMP, (2), whose synthesis was inspired by Mislow's pioneering
work on phosphine oxide resolution,<5) effected the reduction of (3) to (4)
OMI!

o~
~H O~
1.&
Ac HN C02H AcHN :
I
co2H 4-
H -

in 95% optical yield and thus afforded a viable catalytic synthesis of L-DOPA.
Many catalysts of comparable efficiency have been developed subsequently,
all based on the principle of a rigid, cationic rhodium cis- chelated biphos-
phine complex. Complexes of CHIRAPHOS (5)(6) and BINAP (6/ 7) are among
the most successful, but none surpass Knowles's contribution (Figure 4.1).
The simplicity and utility of asymmetric hydrogenation has inevitably
attracted a number of reviews(8-14) in which different aspects are empha-
sized. The article by Marko and Bakos is particularly useful, as it is intended
to be a comprehensive source of all applications of asymmetric hydrogena-
tion up to the end of 1978. Because of this background, the present chapter
ASYMMETRIC HYDROGENA TfON REACTIONS 139

is not intended to be comprehensive, but will draw as far as possible on


developments within the last two or three years which have not yet appeared
in other reviews. Our intention is to survey the field, present its successes
and limitations, and show how mechanistic studies have helped explain the
origins of stereoselection.

2. THE VARIA TION OF REACTION CONDITIONS

The success of asymmetric hydrogenation as a route to amino acid


derivatives does not find more general applicability. To appreciate why
this is so, an analysis of the various factors involved is required. The
limitations will only be overcome when a full appreciation of mechanism
and of the source of stereoselectivity has been assimilated.

2.1. Changing the Catalyst


Almost without exception(15) monophosphines are ineffective in asym-
metric hydrogenation, and flexible chelate biphosphines(16) give rise to
poor optical yields. The Monsanto phosphine, DIPAMP, is an apparent
exception, but the lack of rigidity in the chelate backbone may be counter-
manded by a preferred orientation of the ortho- methoxy groups during
key stages of the catalytic cycle. There are many variables within a rigid
cis- chelate, and these give rise to subtle changes in characteristics.
The simplest case is that of a 5-ring chelate based on 1,2-
bis(diphenylphosphino)ethane. Inorganic stereo chemists have long
appreciated the chiral conformations adopted by 2,3-diaminobutane
derivatives, but it was not until 1977 that Fryzuk and Bosnich applied this
knowledge to phosphine chemistry. The synthesis of CHIRAPHOS and its
cationic rhodium complex (7) was most fruitful, and optical yields of nearly

100% were obtained in the reduction of z-acylaminocinnamic acids and


esters. In X-ray analysis(17) of complex (7), it was shown that the methyl-
groups occupy pseudoequatorial sites in a nonplanar chelate that maintains
C2 symmetry. Once this had been recognized, an obvious experiment was
to prepare the lower homologue PROPHOS, (8), (17) and to determine whether
a single methyl group was capable of maintaining conformational rigidity.
140 JOHN M. BROWN AND PENNY A. CHALONER

I""""l-r OM.
~/Ph
P

~
78 ~
( 94
P
Ph/~
MeO~

o PhXH M.· '


X PPh 2
89 ~
AN (0 H M. P Ph 2
H 2
R. H 49 l!
R.M. 89 ~

84 ~

93 ~
91 ~

95 ~
96 ~

Figure 4.2. Typical optical yields in asymmetric hydrogenation of z-dehydroamino acids by


typical catalysis.

Optical yields were only slightly lower (Figure 4.2), and comparable results
have subsequently been obtained using the ligands PHEPHOS, (9),08.19) and
CYPHOS, (10).(20) Chiral 1,2-bis(diphenylphosphino)-ethane units may be
derived by resolution procedures as in the case of NORPHOS, (11),(21·22)

dfPh 2 U PPh2
Ph P

J:(Ph 2 ,,
PPh 2

PPh 2

1.1 12 ,Jl

where Brunner and co-workers exploited the selective crystallization of


one diastereomer of the complex dibenzoyl tartrate with the bis- phosphine
oxide of (11). The synthesis of PHELLANPHOS, (12), involved reaction of
trans-l,2-bis(diphenylphosphino)ethylene disulfide with (-)-phellan-
drene,(23) whereas NOPAPHOS, (13), was prepared by the same Diels Alder
reaction with (+ )-nopadiene.
ASYMMETRIC HYDROGENATION REACTIONS 141

The chair-form of a 6-ring biphosphine chelate has an achiral local


orientation of PPh 2-groups. This makes it predictable that phosphine (14)
should be a poor ligand in asymmetric hydrogenation, (24,25) but the efficiency
of (15) is more surprising at first sight. Bosnich and co-workers attribute
this to the existence of a preferred C2 twist conformation in complexes
of biphosphine (15), which is thus dubbed SKEWPHOS. They support
their conclusions by showing that square-planar rhodium complexes of
(15) are five times more CD-active than comparable complexes of (14). The
catalyst apparently sits on a conformational knife-edge, and z-dehydro-
isoleucine derivative (16) is out of line in giving a poor optical yield on
hydrogenation.

Me,
Me,
(PPh 2
~PPh2
PPh 2
Me
PPh 2 ";X:02H
14 15 16
-

Seven-ring chelates were the subject of much early effort following the
initial successes of Kagan and co-workers. It should suffice to comment
that the nonrigid chelate, (17), does not effect asymmetric hydrogenation,
that a range of meta- and para- substituted arylphosphine analogues of mop
and a range of stereochemically equivalent bisphosphinobutanes, such as
(18), have comparable efficiency to the parent, (26) but that the ortho-
methoxyphenyl analogue, (19), gives lowered reactivity and reversed chiral-
ity in hydrogenation of z-dehydroamino acids. (27)

Meo
'j~-~PPh2
0 t ' PAr2
MeO-~n
PPh 2
>(

D
'0 PAr2 OMe
17
~ Ar =

A number of crystal structures of mop complexes have been


described(28,29,30) (vide infra), from which it is apparent that the geometry
may vary from an ideal C 2-twist-chair to a twist-boat in accord with
conformational variations within bis (mop )platinum. (31) Kagan (32) has car-
ried out a detailed conformational analysis of the ligand in its chelate
complexes, Figure 4.3 representing a simplified version.
The basic framework of mop, which is a C2 chiral 1,4-
bis(diphenylphosphino)butane derivative, is subject to wide variation. The
142 JOHN M. BROWN AND PENNY A. CHALONER

/h
!.!
Ph- p
\-i[0Y ~h
Ni

Ih 0
Pd Pt
(twist-boat)

Figure 4.3. Conformational analysis of DIOP complexes.

heteroatom analogues, (20) and (21), are examples of a range studied by


Italian, (33.34,35) Japanese, (36.37) and East German (36) workers. Chiral
diaminobutanes are readily available and potentially modified with wide
variation by N-substitution. While the rhodium complex of (20) gives high
optical yields in reduction of z-dehydrophenylalanines, the simple change
by N-methylation to (21) causes reduction of reaction rate and inversion
in the configuration of product, This provides an interesting clue towards
the origin of stereoselectivity.

20 R=H
II R= Me

Two further variations on the theme of 7-ring chelates deserve com-


ment. The ingenious synthesis of a series of chiral biphosphines based on
4-hydroxyproline provided first Achiwa(39) and then Ojima and co-
workers(40} with an efficient and reactive catalyst. The N -t- butyloxycarbonyl
derivative BPPM, (22), has been most widely used. If the disadvantage of
7 -ring chelates is their flexibility and the availability of alternative confor-
mations of comparable energy, then BINAP complexes deserve attention.
The axial chirality in (6) ensures that the chelate is rigid, and it does indeed
ASYMMETRIC HYDROGENA TlON REACTIONS 143

provide very high optical yields in de hydro amino acid reduction. The
simplicity of its structure makes it an almost ideal model for mechanistic
studies.

24 ---- S=MeOH - - -,25

Larger-ring chelate biphosphines are rare and may suffer from a


number of disadvantages in catalysis. The tendency to trans- chelation
means that DIOXOP (23) only functions as an effective catalyst under certain
conditions(41) where the reaction sequence may involve (24) as well as (25).
In the simple 8-ring rhodium chelate derived from 1,5-bis(diphenylphos-
phino)pentane, ready C-H insertion occurs under hydrogenation condi-
tions, so that complex (26) is observed. The cis- chelate biphosphine unit
is apparently essential to successful asymmetric hydrogenation, and purely
trans-chelating chiral biphosphines, such as (27), show poor selectivity.(42)

2.2. Changing the Substrate


The previous section summarizes how various biphosphines catalyze
asymmetric hydrogenation as their cationic rhodium complexes, illustrated
by reduction of z-acetamido- or benzamidocinnamic acids. Generally
speaking, 5-ring chelate complexes offer predictable results, and a range
of z-dehydroamino acid derivatives give optical yields in excess of 85%
with catalysts based on (2), (5), or (8)-(13). On lesser evidence, the same
may be true of rigid chelates based on (6). With 7-ring chelates, the reaction
144 JOHN M. BROWN AND PENNY A. CHALONER

Rl NHCOR2
Table 1. Asymmetric Hydrogenation of "--/ by R,R-
OIOP Rhodium Complexes a ~
H COOR2

Enantiomer
Rl R2 R3 excess Configuration Reference

H Me H 73 R 16
H Me Me 60 R 27
IPr Me H 74 R 43
lPr Me Me 21 R 43
Ph Me H 82 R 16
Ph Me Me 69 R 43
Ph Ph H 64 R 16
Ph Ph Me 55 R 16
Ph IPr H 57 R 43
Ph IPr Me 15 R 43
Ph tBu H 52 R 43
Ph tBu Me 0 43
Ph Me Et 72 R 43
Ph Me Wr 76 R 43
Ph Me tBu 77 R 43

a The catalyst used was formed in situ by reaction of R. R-DIOP with (Rh(cyciooctene)2Cl)2
in benzene/ethanol.

is very sensitive to the precise pattern of substitution in the de hydro amino


acid derivative, as the results in Table 1 indicate. Glaser and co-workers(43)
have made a systematic study of variations in the ester and amide moiety
and find that the latter posseses a much more powerful stereodirecting
effect, explicable if there are conformational variations in the biphosphine
chelate induced by alteration of the substrate.
Changing the stereochemistry of the dehydroamino acid has a profound
effect. While Z-a- benzamidocinnamic acid usually gives optical yields com-
parable to a- acetamidoacrylic acid, the E-isomer gives greatly lowered
optical yields, and isomerization(44.45) may compete with hydrogenation.
Not so for DIPAMP in benzene solution(45) (Table 2), nor for BINAP, both
of which give high optical yields in the reduction of E-a- benzamidocinnamic
acid. While z-acids and esters frequently reduce with equal facility, there
is no case of an E-acylaminocinnamic acid ester giving effective asymmetric
ASYMMETRIC HYDROGENATION REACTIONS 145

Table 2. Enantiomer Excesses in Hydrogenation of a-Benzamidocinnamic


Acids and Methyl Esters

Phosphine Solvent z-Acid z-Ester E-Acid E-Ester

R,R-DIPAMP EtOH 94s 30 s

C6 H6 16 s 80 s
R, R-DIOP 3: 1 EtOH:C 6 H 6 a 70 R 37.5 R 25 R 5R
C6 H/ 62R
s-BINAP EtOH 96R 93 R

a Neutral RhCl (OIOP) used as catalyst.


b Cationic Rh (norbornadiene) OIOP used as catalyst.
C Reaction carried out in THF.

reduction. A recent and very extensive paper(45a) has shown that both E-
and z-acylamino-3-alkylacrylates are hydrogenated with high stereo-
specificity to s-amino acid derivatives with RR DIS AMP-derived catalysts.
The fundamental and apparently indispensable structural unit in asym-
metric hydrogenation is the enamide group (vide infra for structural
studies), and this led several authors to examine related reactants. Kagan
showed that simple enamides such as (28) were reduced in the presence
of DIOP complexes in up to 90% optical yield. (46) Reduction of itaconic
acid and its esters has been extensively studied and two groups of work-
ers(47.48) report high optical yields. DIPAMP complexes are effective catalysts
for the reduction of a range of itaconate derivatives (29),(47) the ~-ester
being the least successful, while the proline-derived complexes(48) (such
as BPPM) only work well in the presence of triethylamine, a point to which
we shall return later. Enol esters possess the same relative disposition of
carbonyl and olefinic groups and so would be expected to be reduced with
high stereoselectivity. The Monsanto group(49) showed that a-aryl vinyl
acetates, (30), are reduced with moderate optical efficiency by DIPAMP

H Me H)(Me H~H
)( Ph (02R'
Ph NHCOPrl NHA( RO (
2

complexes, the enantiomer excess increasing with the electron withdrawing


power of the aryl residue. The example provided by Bosnich and Fryzuk
demonstrates a simple synthesis of (chiral-methyl) lactic acid (Figure
4.4).(17)
A final example of the chelating substrate principle is given by reduc-
tion of enol phosphates, an interesting feature being the high (and not
146 JOHN M. BROWN AND PENNY A. CHALONER

BrXD Pdt PPh~


AcO COlt

2
~ +

NaOH
TXD fit H Rh complex
of IBi
then HCl
AcO ~ COp

Figure 4.4. The synthetic route to CHrchiral lactic acid.

readily predicted) specificity afforded by phosphine (31) in reduction of


(32),(SO) where DIOP complexes are both chemically and optically less
efficient.

c6'r-PPh 2
H H
1'Y
Fe OH
Me
Ph)(O/~Ph2
~PPh2
31

Enamide hydrogenation has found application in dipeptide synthesis,


since dehydrodipeptides are quite readily available.(SI) Three groups of
workers(S2-S5l have examined the reduction of dehydrophenylalanyldipep-
tides, with an interesting divergence of results. The asymmetric center in
dehydrodipeptide (33) (see Figure 4.5) has very little effect on the configur-
ation of the new asymmetric center when the catalyst is derived from
DIPAMP,(52) DIOP,(S4) or the prolinebiphosphine (34),(S2) and the major
diastereomer formed is that predicted by analogy with simple enamides.
Thus high yields of either RS- or ss-dipeptides may be obtained at will.
With DIoxop-derived catalysts, the existing chiral center plays an important
part, (S5) and high optical yields are only obtained when the chiral amino
acid residue in the starting material has the s-configuration (Figure 4.5).
ASYMMETRIC HYDROGENA TlON REACTIONS 147

R -RS + RR 5 -55 + 5R

90 10 95 5 010 P 1

95 5 01 PAMP 2

27 73 • 01 OX 0 P fl

04 996
.. 99
Ph P

2~PPh2
J 34
O~NHPh -

36 62 62 38 Phl(CH2)4PPh2

.. Ligand 22

Figure 4.5. Diastereomer yields in the reduction of dehydroamino acid derivatives.

Prochiral a~-unsaturated carboxylic acids present an anomaly, since


some of the best examples of their asymmetric hydrogenation require
monophosphine-derived catalysts. An early example using Morrison's
neomenthyldiphenylphosphine, (35),(56) is shown and in a wide-ranging

y
study(57) respectable optical yields were obtained. In reduction of (36) and
'-.,/
M'

PPh 2

other carboxylic acids, triethylamine seems to be necessary, so that the


true substrate is likely to be the a~- unsaturated carboxylate, which binds
more efiectively(58) to the rhodium center.
The reader is referred to Reference 12 for a more exhaustive range
of substrates, but one class is worth mentioning since it has received little
attention. If the chiral center is in the substrate rather than the catalyst,
then internal asymmetric induction may occur during hydrogenation. A
simple example is in the reduction of dehydrovaline(59) by HT catalyzed
by tris(triphenylphosphine)rhodium chloride, where the tritium is intro-
duced with ~90% stereoselectivity giving (chiral-methyl)valine. Allylic
alcohols possessing an asymmetric center (e.g., 37 in Figure 4.6) are reduced
with high stereoselectivity by a cationic rhodium complex derived from
148 JOHN M. BROWN AND PENNY A. CHALONER

Me ,H
HifH

Ph~OH
XX /OH
Ph/"
Me H Me H

~Ph2
Ph P- rl/]i-c'.
1
HMe H Me

2 /1 H
H 0
I
~
Ph~OH
H/ Me
s

Figure 4.6. Internal asymmetric induction in homogeneous hydrogenation of allylic and


homoaIIyIic alcohols.

l,4-bis(diphenylphosphino)butane. (60) The course of reduction is inverted


for the homolog (38) (Figure 4.6), and stereoselectivity appears to be
controlled by nonbonded interactions within a chela ted substrate.

2.3. Changing the Solvent

Most reductions of z-dehydroamino acids are carried out using cationic


biophosphine rhodium diene complexes in polar media, the most popular
being methanol. The course of reduction is not much altered by changing
the solvent to other polar coordinating media, such as higher alcohols or
tetrahydrofuran, although E-dehydroamino acids seem to give much better
results in aprotic solvents, particularly benzene.
Some early results by Kagan(46) suggest that the nature of the complex
involved in reduction of simple enamides may change with solvent polarity.
Substrate (39) gives s-product on reduction by "mop rhodium chloride"

HXH 39
Ph NHAc

in ethanol, but increasing amounts of benzene in the medium lead to an


increased amount of R-enantiomer, which eventually predominates. This
implies that the true catalyst is cationic in ethanol, but hydrogenation by
a neutral complex (which exhibits opposite selectivity) becomes increasingly
important as the medium polarity is reduced.
ASYMMETRIC HYDROGENATION REACTIONS 149

2.4. Changing the Pressure

Most asymmetric hydrogenation is carried out at ambient temperature


and pressure, but several workers have explored the relationship between
optical yield and overpressure of hydrogen. The most thorough studies are
due to Ojima(61) and to Sinou,(62) representative data being presented in
Table 3. The general trend is to lower optical yields at higher pressures,
but the sensitivity is greatly dependent on both catalyst and substrate. It
should be possible in principle to derive mechanistic information from
these observations; either the increased hydrogen pressure leads to a change
in rate-determining stage, or, alternatively, there is a second mechanism
which comes into play at high hydrogen pressures, possibly involving prior
coordination of hydrogen to the metal, so that the order of addition (and
hence the stereoselectivity) is reversed. Despite one authoritative statement
in the literature, (63) there seems no basis for preferring one alternative over
the other on present evidence.

2.5. Summary of the Features of Effective Catalysts

So many examples of the asymmetric hydrog.enation of dehydroamino


acid derivatives are known which have sufficiently similar features that a
common mechanism is probable. It seems that the rhodium complex must
be cationic, preferably in a weakly coordinating solvent, and carry a single
chelating biphosphine occupying adjacent coordination sites throughout
the catalytic cycle. Only rhodium complexes are genuinely successful,
although ruthenium mop complexes(64) show some promise in the asym-
metric hydrogenation of af3-unsaturated carboxylic acids. cis-Chelate
iridium complexes would be likely to be deactivated by JL- hydride formation
in the course of catalysis. (65)
Enamides are successful substrates on account of their closely defined
coordination geometry, since both the olefin and amide carbonyl group
are available to bind at cis- related sites. This must be necessary during the
rate-determining stage, for species lacking the amide group (or a closely
related functionality similarly sited) react much more slowly and with lower
stereoselectivity.

3. STUDIES OF REACTION MECHANISM

It is probably true to say that more is known about the mechanism of


asymmetric hydrogenation than any other process in homogeneous cataly-
sis. This is partly because the complexity of the reaction and its
Table 3. Variation in Optical Yield in Reduction with Hydrogen Pressure
~
pH2

Substrate Catalyst Et3N/Rh Solvent 1 S 20 SO 100

Z-a -Benzamido
cinnamic acid (BPPM)Rh <±lcod EtOH 83.8 R 21,2 R 4.7 s 8.4 s

Cinnamic acid (BPPM) RhEllcod 2 EtOH 93.3 R 78.7 R 66.2 R 64.2 R

a-Acetamido
acrylic acid (BPPM) RhEllcod MeOH 9S.2 R 2l.8s
a-Acetamido C5
acrylic acid (BPPM) RhEllcod 2 MeOH 98.S R 7.3 s ~
Itaconic acid (BPPM) RhEllcod MeOH 9l.3s 8S.3 s ~

Itaconic acid (BPPM) Rh cod Cl MeOH/C6 H 6 92.3 s 79.1 s ~


Q

Itaconic acid (BPPM) Rh cod CI 2 MeOH/C6 H 6 94.S s 91.3 s ~


:t>
z-a- Acetamido-
cinnamic acid DIOXOP RhEllcod EtOH 13.0s 10.0R 32.0R 30.0R
~
Z-a- Acetamido-
cinnamic acid DIOXOP RhEll cod 3 EtOH 84.0s 69.0s 43.0s 3S.0s ~
-<:
~
Methyl Z-a- C")
acetamidocinnamate DIOXOP RhEll cod EtOH l.Os lO.OR 23.0R 33.0R ~
r--
Methyl z-a- Q
acetamidocinnamate DIOXOP RhEll cod 3 EtOH 6.0s lO.Os 6.0s S.Os ~
::\]
ASYMMETRIC HYDROGENA TION REACTIONS 151

stereochemical component provide many factors that can be systematically


varied and partly because several intermediates on the reaction pathway
may be isolated or defined in solution, and the stereochemical course of
each step thus established.

3.1. Reaction Rates

There are only two controlled kinetic studies of asymmetric hydrogena-


tion, one of which(66) was carried out using neutral OlOP complexes before
the significance of ionization in polar solvents was fully appreciated, and,
hence, the pathway studied is not necessarily the most efficient one. In the
second, Halpern and Chan(67.68) demonstrate that the kinetic form of the
hydrogenation of methyl z-a-acetamidocinnamate catalyzed by the 1,2-
bis(diphenylphosphino)-ethane rhodium cation in methanol is:
dP
- = k 2 [H 2][Rh,o,]
dt

with k2 = 10 2 mol L -1 S-1 at 298 K for a formal hydrogen concentration


of M/22.4 at STP. The kinetic isotope effect is 1.22.(69) This corresponds
to a turnover number of about 0.07 M s-1, which is 60 times faster than
the corresponding CHIRAPHOS complex. Most of the catalyst under these
conditions is bound to the substrate with K = 5 X 103 . These experiments
are not informative of the rate-determining stage, but it has recently been
shown that this particular reaction occurs irreversibly, since ortho- and
para- hydrogen are not equilibrated in the course of hydrogenation by a
para-enriched sample.(70) In contrast, reaction in the presence of Wilk-
inson's catalyst leads to equilibration of nuclear spin-states as reaction
proceeds.
A more general qualitative survey(71) of reactivity shows that 5-ring
chelate complexes are much less reactive than 6- or 7-ring chelates where
turnovers in excess of 0.3 M S-1 may be observed. The nature of the
substrate is also important, and in OlPAMP complex catalyzed hydrogena-
tions almost any structural change in the substrate leads to a depression
of reactivity. This includes N-methylation (which also profoundly affects
the stereoselectivity) conversion of the carboxylic acid moiety into a nitrile
and changing from z- to E-stereochemistry about the double bond. The
latter further indicates the stringent structural requirements of asymmetric
catalysis.

3.2. X-Ray Structural Analvsis of Catalvsts and Reaction


Intermediates

A fortunate feature of asymmetric catalysis is that chiral biphosphine


rhodium diene complexes, especially with norbornadiene, crystallize well
152 JOHN M. BROWN AND PENNY A. CHALONER

4Q exl

lU ex ~

Figure 4.7. Approximate crystal structures of chelating biphosphine rhodium complexes


(other ligands omitted for clarity).

and are amenable to X-ray study. The structures will be discussed in detail
elsewhere in the book (see Chapter 2), but it is useful to analyze some
general features here. Pertinent structures are shown in Figure 4.7; the
DIPAMP complex, (40), appears in the original work of Knowles and co-
workers while the CHIRAPHOS complex structure, (41), was published
shortly after the description of its use in catalysis. (72) The 1,2-
bis(diphenylphosphinomethyl) cyclobutane structure (42) is due to Town-
send(73) and the DIOP structure (43) again due to Knowles and co-
workers.(74) The BINAP complex (44) with its highly constrained geometry
was solved by Ito and co-workers. (75)
All of these structures with C 2 symmetry have a special feature, first
recognized by Knowles. If the P-phenyl groups are viewed from the remote
side of the coordination plane, then the axial pair appear edge-on, and the
equatorial pair are approximately face-on. Viewed from above, a pair of
the ortho- hydrogens belonging to opposed axial phenyl groups is in proxim-
ity to the metal, within 3 A. Whether this plays any role in stabilizing the
structure is unclear but the P-phenyl orientation is sufficiently well defined
to permit prediction of the sense of asymmetric hydrogenation. If the
relationship is drawn in (40), then the reaction will give rise to an s-amino
ASYMMETRIC HYDROGENATION REACTIONS 153

ex CHIRAPHOS ex 0 I PH OS

Figure 4.8. Crystal structures of dehydroamino acid complexes.

acid derivative. All catalysts which are structurally defined conform to this
rule.
A further important development is due to Halpern and co-
workers. (63.67) They managed to crystallize two en amide complexes, one
derived from 1,2-bis(diphenylphosphino)ethane and the other from
CHIRAPHOS. They are almost isostructural with regard to the binding of
the de hydro amino acid ester, as indicated in Figure 4.8. Complex (46)
shows the olefin 1T'- bonded and rotated somewhat from orthogonality with
respect to the coordination plane. The Rh-C distance to the heteroatom-
substituted carbon is less than that to the benzylidene carbon, and the
N -acetyl group is bonded to rhodium through a (T- interaction with the lone
pair of the carbonyl group. The ligand backbone is skewed so that it is
chiral. Interestingly, the CHIRAPHOS complex, (47), has the opposite
configuration to that of the amino acid ester produced on its hydrogenation
by cis-transfer of H2 from rhodium. The solid-state structure has been
shown to persist in solution by analysis of its CD spectrum. (76)

3.3. Structure and Dynamics in Solution

3.3.1. Solvates

In a classic paper(77) Tolman and co-workers at Dupont demonstrated


that the dihydride (48) was observable by lH and 31 p nmr when
tris(triphenyl-phosphine) rhodium chloride was sealed in solution under
154 JOHN M. BROWN AND PENNY A. CHALONER

hydrogen. The stereochemistry of this intermediate was thus established,


and at that time it was suggested that stereospecific PPh 3 loss preceded
alkene coordination in hydrogenation. Later work 1781 suggests that at high
concentrations of alkene and low concentrations of PPh 3, (48) may be
bypassed in the true catalytic cycle. Thus, early attempts to explain the
mechanism and stereoselectivity of asymmetric hydrogenation were
influenced by these results and modeled on them, notwithstanding the fact
that H 2Rh(PPh 3hCl has two trans- related PPh3 ligands and efficient chiral
catalysts possess a cis-chelated biphosphine. Observations by Halpern(79)
and Baird(80) and co-workers clarified the issue. It was shown, (79) for
example, that complex (49) reacted with three moles of hydrogen in
methanol solution giving (50), whereas complex (51) reacted with two
moles of hydrogen in methanol solution giving (52), which had no apparent
affinity for hydrogen. Further work has established that the division between
monophosphines (dihydride formation) and chelating biphosphines (solvate
formation) is not quite so clear-cut.(81) Direct evidence for the binding of
methanol to rhodium in (52) is lacking, but the structure is consistent with
spectral properties, and 31 p nmr indicates that the phosphorus nuclei are
trans to an electronegative ligand. In the corresponding DIPAMP solvate,
(53), EXAFS measurements(82) indicate that oxygen ligands are trans to
phosphorus in the inner coordination sphere. Attempts to crystallize com-
plex (52) led to the formation of an arene-bridged dimer, characterized as
(54) by X-ray structure determination. (79) Dimers of a different structure
may be formed on hydrogenation of the PHEPHOS complex, (55), since it
gives a monomeric and two dimeric solvates (four in the case of racemic
phosphine) on hydrogenation in methanol.(8!)
Related solvate complexes have been formed from many other asym-
metric chelate biphosphine complexes and with other coordinating solvents.

H H Ph 2 H
PhP

PhP/ -
+/~
3 ......... Rh
I
P~P"
~ Rh
I+/OMe
I
H/ "PPh
C :~Rh~~
p/ "'-V C,+ P

P
/Rh
/OMe

"0 Me
3 X X- HOMe 3 Ph 2 Ph 2 H

50 ~

Q-;,oMe
" P,h H

C P,- +~E'
/Rh
"'a--1e

'-b
P
~h - H
MeO ~ h
55
ASYMMETRIC HYDROGENA TlON REACTIONS 155

The apparent formation of arene complexes on hydrogenation in benzene


(e.g., in Reference 61) seems to be an anomaly, since their solution nmr
spectra are very different from authentic rhodium biphosphine tetraphenyl-
borates. (81)

3.2.2. Substrate Complexes

Under an inert atmosphere, solvate complexes of the type described


above react with de hydro amino acid derivatives to produce new species,
the solution structure of which is comparable to (46) or (47). Binding of
the olefin and amide groups was established by 13 C j31p nmr employing
specifically labeled de hydro amino acids, which additionally demonstrated
that the carboxylate group is not involved in coordination in most cases.
A range of enamide complexes derived from rhodium bis(diphenylphos-
phino)ethane all exhibit very similar 31 p nmr spectra which are sharp at
room temperature.(83) With larger ring chelates, the corresponding com-
plexes give 31 p spectra which may be broadened by dynamic processes at
room temperature and require cooling before the characteristic sharp 8-line
spectrum is observed. The nature of this dynamic process is revealed by
saturation transfer experiments employing the DANTE pulse sequence.(84)
This demonstrates that an intermolecular exchange is involved (Figure 4.9),

if PAr2 chiral

Figure 4.9. The mechanism of ligand exchange in en amide complexes.


156 JOHN M. BROWN AND PENNY A. CHALONER

which is fast in relation to the time-scale of catalysis but strongly affected


by the nature of the amide-group, with the pivalamide exchanging an order
of magnitude more slowly.
Of a number of enamide complexes studied by the present authors,
those derived from DIPAMP proved to be the most rewarding. (81.83.85) If the
solvate, (53), is reacted with Z-a- benzamidocinnamic acid or its methyl
ester, then two enamide complexes of similar structure are formed in a
ratio of ca 10: 1 at room temperature. Since 13C nmr of labeled enamides
shows that these have very similar solution structures, they are presumed
to be the two diastereomeric complexes (56) and (57), although at this
stage the experiment gives no information on their configuration. The
DANTE experiment(86) provides the intriguing result that their
intra molecular exchange (by coordination/ decoordination of the olefinic
bond) is considerably faster than dissociative exchange and again rapid on
the time-scale of catalysis.
Ph Ph
An., , ;<Ph A.~ , Ph
Me
'p 0

=:xn
'P o=( 0/
( " . / Ph ( ,,/. NH
//R~}NH P
/Rh /
"-
An

Ph ~ ,/ \Meoif~Ph
An COMe Ph A
2 n R

~ maj or -57 -
minor
-
The simplicity of solution equilibria observed here is typical of 5-ring
chelate rhodium complexes of z-enamides, where complexation constants
are high (-10 4 M- 1 ) and the geometry well-defined. Under catalytic condi-
tions, the enamide complex represents the major resting-state, and its
bright scarlet color is characteristically observed. In other cases, particularly
CHIRAPHOS, only a single diastereomer is observed with most substrates.
Although this pattern of complexation is frequently comparable with larger-
ring chelate complexes, anomalies are often observed. Under appropriate
conditions, z-enamide complexes with bound carboxylate, (87) with 2: 1
stoichiometry, (88) with tridentate binding(89) and possible (1'- bonding to the
benzylidene carbon(27) have all been observed and characterized in solution
by 13 C/31 P nmr spectroscopy, structures being shown as (58)-(61).
OH NHAc

;h, ~CO'H
Ph 2
H,r P 0
c('p >ha:/~ OMe Cp?~xHph
!\.-p HI\I
H Ph2 Ac H Ph
Ph 2 NH Ac

~
ASYMMETRIC HYDROGENATION REACTIONS 157

Complexation of af3-unsaturated acids and ami des has been exten-


sively studied. (58) In such cases, both diastereomers of a species involving
coordination of both olefin and carboxylate group are observed, usually in
comparable proportions. The extent of complexation is enhanced and the
rate of exchange of free and bound substrate reduced when base is added,
or a preformed tetra alkyl ammonium carboxylate is employed. This
suggests that structures (62) and (63) best describe the complex. The spectra
observed are almost identical to those derived from E-dehydroamino acid
derivatives, implying that the latter have structure (64) in accord with
labeling studies. (83.88)

Xr~(Rh/~O
o--f-..- P /' '\
H Ph 2 0
Ph

+/r
62 63

'. jJr.~~~2Rh Ph
1\
O
64( di as t e reome rs)
o p/ '-..: ""'0
H Ph
2
NHCO Ph

3.3.3. Rhodium Alkyls

If en amide complexes such as (46) are exposed to hydrogen in methanol


solution then the color quickly fades, and 31 p nmr observation suggests
direct conversion into solvate by hydrogenation of the bound substrate
without the intervention of an intermediate. At low temperatures, (46) is
converted into (65) which is stable at -80° but decomposes on warming
to -50°. (68) The presence of a solvent molecule trans to the rhodium-bound
hydride is indicated by a quantitative reaction with one equivalent of
CH3 C 15 N giving (65a) with the anticipated changes in nmr spectra. Since
158 JOHN M. BROWN AND PENNY A. CHALONER

Ph H
P':.I+/oyMe
(
p/RtyNH
Ph Z 5 I"CHlh
COZMe
65 S=MeOH
~a S=CH~5CN
addition of hydrogen to a square-planar complex is second-order, but the
decomposition of the alkylrhodium hydride, (65), is first-order with ill'" =
17.1 kcal mole -1 and AS'" = +6 cal mole -1 deg -t, its stability at low tem-
perature is explicable and reflects a change in the rate-determining stage
of reaction.
If the enamide complexes (56) and (57) are prepared normally and then
cooled, the minor diastereomer becomes increasingly disfavored and rep-
resents no more than 2% of the species present at 220 K. Under these
conditions the solution is unreactive to hydrogen, and warming leads to a
reversion to solvate complex with concomitant hydrogenation of the sub-
strate. If the same solvate complex is mixed with a slight excess of methyl
Z-a- benzamidocinnamate at 200 K and sealed under hydrogen, then the
enamide complex mixture which forms under kinetic control is rich in (57).
Warming this solution to 220 K now leads to a reaction in which the minor
diastereomer disappears selectively, and a new set of signals appear.(90)
The 1H and 31 p nmr of (66) suggest that it is closely similar to (65).
CH Ph
\ 2 ____ ,

/
- H
C:, I ..' 0
H02e
' /
Rh-+ An
PhP~/
~Me1>Ph

66 Q Q]

In the course of characterization of this new alkylrhodium hydride it


was demonstrated that the amide remained bound and that the double
bond had been reduced because of the loss of *C 1*C3 coupling in a doubly
labeled sample of enamide on going from (57) to (66). It is possible that
the carboxylate group occupies the coordination site trans to hydride,
because it experiences a down field 13 C chemical shift of 12 ppm relative
to the free acid or ester. This may be due to the change in its environment
combining the effects of saturation and {3- rhodium deshielding, and an
alternative structure, (67), cannot be ruled out on current evidence. Cer-
tainly the intermediate is different in detail from the 1,2-bis (diphenylphos-
phino)ethane case; the 31 p chemical shifts and 31P_Rh, 31 p _31 p coupling
ASYMMETRIC HYDROGENA TlON REACTIONS 159

constants are distinct, and it does not react with CH3 CN at low tem-
peratures. The critically important fact is that only the minor diastereomer
is reactive to hydrogen, and extrapolating this observation to ambient
temperature requires that (57) is a true intermediate in catalysis, whereas
(56) must undergo diastereoisomerism before it can react. Since delivery
of hydrogen from rhodium to alkene is known to be cis- selective and
stereospecific(91) the absolute configuration of (56) and (57) is defined. t
It is incidentally the case that regiospecific formation of (65) and (66)
requires that the amide-olefin chelate ring remains intact during the catalytic
cycle. Formation of the alternative structure is then stereoelectronically
precluded as diagram (68) indicates.
Me
P=\NH
C
Ph 2
p, H,
""+~OH
/ Rh 2
P I H
Ph 5 H
2 Ph

3.4. Conclusions

The combination of crystallographic and nmr experiments gives a


very clear indication of intermediates on the reaction pathway of asym-
metric hydrogenation. They leave one major gap in our understanding,
since the rate-determining stage involves addition of hydrogen and may
generate another intermediate which cannot be observed. There are three
possibilities for the rate-determining stage, namely:
(a) concerted reaction which generates the alkylrhodium hydride in
a single step;
(b) rate-determining addition of hydrogen followed by collapse of
dihydride to alkylrhodium hydride; and
(c) Reversible addition of hydrogen followed by rate-determining
formation of alkylrhodium hydride.
No single experiment can discriminate between these three
possibilities. Reversible addition of hydrogen is ruled out by the complete
t A preliminary experiment in which the CHIRAPHOS enamide complex of Z-a- acetamide-4-
methyl-2-pentenoic acid was prepared gave two diastereomeric complexes in the ratio 2: 1.
When hydrogen was admitted at 220 K, reaction was extremely slow. After 24 hours at
230 K both diastereomers had reacted to some extent to give a mixture of products which
could not be well-defined. While this result may not be strictly comparable with the data
reported in the literature, it does suggest that conclusions based on comparisons between
rate data from different phosphine complexes should be treated with some caution.
160 JOHN M. BROWN AND PENNY A. CHALONER

lack of ortho- ~ para- hydrogen interconversion during reduction of z-a-


benzamidocinnamic acid by DIPAMP, 1,2-bis(diphenylphosphino)ethane or
1,4-bis (diphenylphosphino) butane rhodium complexes. (70) Distinction
between concerted pathway (a) and stepwise pathway (b) is more exacting.
The deuterium isotope effect (Hz versus D 2) is 1.22 for reduction of the
same substrate by the methanol solvate of 1,2-bis(diphenylphos-
phino)ethane under ambient conditions.(69) Addition of HD, although
reported to give equipartition of hydrogen and deuterium in reductions by
Wilkinson's catalyst(92) and in asymmetric hydrogenation(44) does so in the
former but not the latter case. With a range of catalysts, careful integration
of the IH nmr of the reduction product (69) demonstrates that hydrogen

PhCH 2D (H, D)

~
°(Hl
PhCOHN----Kco H
2

prefers the benzylic position and deuterium the a- position by a factor of


1.33: 1 to 1.38: 1 depending on the catalyst. Thus, the intramolecular
discrimination (HD) is greater than the intermolecular discrimination
(H 2 vs. D z), which suggests that they cannot arise by a single process. The
consequential requirement for an intermediate rules out a concerted
mechanism and establishes that rate-determining addition of hydrogen
followed by rapid transfer of one rhodium hydride (mechanism (b)) occurs.
Gives the very considerable structural constraints involved in the key
intermediates, it becomes apparent why the efficient asymmetric hydroge-
nation has been limited to enamides and a few related substrates. Both
the substrate and the biphosphine must maintain chelation throughout,
and the rigidity of the ligand is critical. Trans- chelating phosphines seem
quite ineffective with any substrate, and perhaps here it will be necessary
to introduce groups capable of secondary binding.
The energetics of asymmetric hydrogenation are summarized in Figure
4.10, insofar as they are known. If only the hydrogenation step is
stereoselective, then it becomes clear why changing the rate-determining
step (e.g., by making the substrate more bulky and inhibiting binding, or
by increasing hydrogen pressure) can only lead to a reduction in optical
yield.

4. THE ORIGIN OF STEREOS£LECTIVITY

Given detailed knowledge of two intermediates in enamide hydrogena-


tion, both close to the rate-determining transition-state, it should surely
ASYMMETRIC HYDROGENATION REACTIONS 161

70 + X

Approximate energies
-1
in kJ mol

Substrate = 1 M

HZ = 1 atm

(DIPAMP)

1_100
Figure 4.10. Energetics of asymmetric hydrogenation by Rhodium DIP AMP complexes.

be possible to define why the reaction is so stereoselective. In the enamide


complex, reference to the X ray structures of Halpern and co-workers
helps rationalize the stereoselectivity. The main steric interaction is between
an equatorial phenyl-ring and the carbonyl group, which is minimized in
the preferred diastereomer (Figure 4.11). It is not at all clear, however,
why one diastereomer reacts with hydrogen so much more readily. The
steric argument might be that addition of hydrogen in the productive
direction causes a change in coordination sphere geometry in which the
phosphine trans to an amide carbonyl-group rotates to form the octahedral
dihydride (Figure 4.11), this being more favorable for the minor isomer of
enamide complex. While this might be correct, the argument carries a
touch of sophistry and cannot be tested by experiment as it stands.
It is puzzling why rigidity in the biphosphine should lead to a reduction
in the overall rate of enamide hydrogenation, a point which has been made
more than once.(1l·25) By itself, this suggests that the path to the rate-
de~ermining transition-state requires some distortion of the biphosphine
chelate, perhaps a geometrical change brought about by hydrogen addition.
162 JOHN M. BROWN AND PENNY A. CHALONER

o~

Figure 4.11. Diastereoselection in enamide complexes and the hydrogen-addition process.

Each isomer of enamide complex possesses two diastereotopic faces which


give distinct products on hydrogen addition. Only one direction of addition
is productive since the Rh-H bond should be syn-coplanar with the olefin
C=C for maximum overlap in the transfer stage. This means that
stereoselectivity is controlled by the relative free-energies of addition along

Figure 4.12. The torsional changes involved in hydrogen-addition.


ASYMMETRIC HYDROGENA TION REACTIONS 163

paths A and B in Figure 4.12. These are distinguished by the torsional


changes taking place in the rigid chelate ring. Addition B to the minor
enamide complex is much softer, since the direction of twist in the chelate
is reinforced, while the torsional change in A is harder. If the idea is correct
(and it is subject to experimental test!) then it suggests a more general
method for stereochemical control in transition-metal catalysts.

REFERENCES

1. Reviewed by F. H. Iardine, Prog. Inorg. Chern. 28, 63-202 (1981).


2. For example, L. Horner, H. Siegel, and H. Blithe, Angew. Chern., Int. Ed. Engl., 7, 942
(1968).
3. T. P. Dang and H. B. Kagan, I. Chern. Soc., Chern. Cornrnun. 1971,481.
4. B. D. Vineyard, W. S. Knowles, M. 1. Sabacky, G. L. Bachman, and D. 1. Weinkauff,
I. Arn. Chern. Soc. 99, 5946-5952 (1977).
5. O. Korpiun, R. A. Lewis, 1. Chickos, and K. Mislow, I. Arn. Chern. Soc. 90,4842-4846
(1968).
6. M. D. Fryzuk and B. Bosnich, I. Arn. Chern. Soc. 99, 6262-6267 (1977).
7. A. Miyashita, A. Yasuda, H. Takaya, K. Toriumi, T. Ito, T. Souchi, and R. Noyori, I.
Arn. Chern. Soc. 102, 7932-7934 (1980).
8. 1. D. Morrison, W. F. Masler, and M. K. Neuberg, Adv. Catal. 25, 81-124 (1976).
9. D. Valentine, Ir. and 1. W. Scott, Synthesis 1978, 329-356.
10. V. Caplar, G. Comisso, and V. Sunjic, Synthesis 1981,85-116.
11. M. D. Fryzuk and B. Bosnich, Top. Stereochern. 12,119-154 (1981).
12. L. Marko and 1. Bakos, Aspects Hornogeneous Catal. 4,145-202 (1981).
13. 1. M. Brown, P. A. Chaloner, B. A. Murrer, and D. Parker, ACS Syrnp. Ser., 119,
169-194 (1980).
14. P. A. Chaloner and D. Parker, Reactions of Coordinated Ligands, edited by P. S.
Braterman, (Plenum Press, New York, in press).
15. W. S. Knowles, M. 1. Sabacky, and B. D. Vineyard, Ann. New York Acad. Sci. 214,
119-124 (1973).
16. H. B. Kagan and T. P. Dang, I. Arn. Chern. Soc. 94, 6429-6433 (1972).
H. Kagan and T. P. Dang, Ger. Offen. 2,161,200 (1972); Chern. Abstr. 77, 114567k
(1972).
17. M. D. Fryzuk and B. Bosnich,l. Arn. Chern. Soc. 100, 5491-5494 (1978); 101,3043-3049
(1979).
18. R. B. King, 1. Bakos, C. D. Hoff, and L. Marko, I. Org. Chern. 44, 1729-1731 (1979).
19. 1. M. Brown and B. A. Murrer, Tetrahedron Lett. 1979,4859-4862.
20. D. P. Riley and R. E. Shumate, I. Org. Chern. 45, 5187-5193 (1980).
21. H. Brunner and W. Pieronczyk, Angew. Chern. Int. Ed. Engl. 18,620-621 (1979).
22. H. Brunner, W. Pieronczyk, B. Schiinhammer, K. Streng, 1. Bernal, and 1. Korp, Chern.
Ber. 114, 1137-1149 (1981).
23. O. Samuel, R. Couffingal, M. Lauer, Z. Y. Zhang, and H. B. Kagan, Nouv. I. Chirn. 5,
15-21 (1981).
24. H. B. Kagan, 1. C. Fiaud, C. Hoonaert, D. Meyer, and 1. C. Poulin, Bull. Soc. Chirn.
Belg. 88, 923-931 (1979).
164 JOHN M. BROWN AND PENNY A. CHALONER

25. P. A. MacNeil, N. K. Roberts, and B. Bosnich,J. Arn. Chern. Soc. 103, 2273-2280 (1981).
26. T. P. Dang, J-c. Poulin and H. B. Kagan, J. Organornet. Chern. 91,105-115 (1975).
27. J. M. Brown and B. A. Murrer, Tetrahedron Lett. 21,581-584 (1980).
28. V. Gramlich and G. Consiglio, Helv. Chirn. Acta 62, 1016-1024 (1979).
29. V. Gramlich and C. Salomon, l. Organornet. Chern. 73, C61-C63 (1974).
30. S. Brunie, J. Mazan, N. Langlois and H. B. Kagan, l. Organornet. Chern. 114,225-232
(1976).
31. J. M. Brown and P. A. Chaloner, l. Arn. Chern. Soc. 100,4307-4309 (1978).
32. G. Balavione, S. Brunie, and H. B. Kagan, l. Organornet. Chern. 187,125-139 (1980).
33. M. Fiorini and G. M. Giongo, l. Mol. Catal. 7, 411-413 (1980).
34. M. Fiorini, F. Marcati, and G. M. Giongo, l. Mol. Catal. 4, 125-134 (1978); 3,385-387
(1977 /8).
35. M. Fiorini, G. M. Giongo, F. Marcati, and W. Marconi,!. Mol. Catal. 1,451-453 (1975/6).
36. K-i. Onuma, T. Ito, and A. Nakamura, Tetrahedron Lett. 1979,3163-3166.
37. idern, Bull. Chern. Soc. lpn. 53, 2016-2019 (1980).
38. G. Pracejus and H. Pracejus, Tetrahedron Lett. 1977,3497-3500.
39. 1. Ojima, T. Kogure, and K. Achiwa, Chern. Lett. 1978, 567-568.
40. 1. Ojima, T. Kogure, and Y. Yoda, Chern. Lett. 1979,495-498.
41. D. Lafont, D. Sinou, and G. Descotes, l. Organornet. Chern. 169, 87-95 (1979).
42. J. M. Brown and F. M. Dayrit, to be published.
43. R. Glaser, S. Geresh, J. Blumenfeld, and M. Twaik, Tetrahedron 34, 2405-2408 (1978);
R. Glaser, S. Geresh, M. Twaik, and N. L. Benoiton, Tetrahedron 34,3617-3621 (1978);
R. Glaser and S. Geresh, Tetrahedron 35,2381-2387 (1979). J. M. Brown, P. A. Chaloner,
R. Glaser and S. Geresh, Tetrahedron 36,815-825 (1980); R. Glaser, S. Geresh, and
M. Twaik, Isr. l. Chern. 20,102-107 (1980). D. Lafont, D. Sinou, G. Descotes, R. Glaser,
and S. Geresh, l. Mol. Catal. 10,305-311 (1981) and earlier papers.
44. C. Detellier, G. Gelbard, and H. B. Kagan, l. Arn. Chern. Soc. 100, 7556-7561 (1978).
45. K. E. Koenig and W. S. Knowles, l. Arn. Chern. Soc. 100,7561-7564 (1978).
45a. J. W. Scott et al. l. Org. Chern. 46,5086-5093 (1981).
46. H. B. Kagan, N. Langlois, and T. P. Dang, 1. Organornet. Chern. 90, 353-365 (1975);
D. Sinou and H. B. Kagan, l. Organornet. Chern. 114,325-337 (1976).
47. W. C. Cristopfel and B. D. Vineyard, l. Arn. Chern. Soc. 101, 4406-4408 (1979).
48. 1. Ojima, T. Kogure, and K. Achiwa, Chern. Lett. 1978,567-568.
49. K. E. Koenig, G. L. Bachman, B. D. Vineyard, l. Org. Chern. 45, 2362-2365 (1980).
50. T. Hayashi, K. Kanehira, and M. Kumada, Tetrahedron Lett. 22,4417-4420 (1981).
51. O. Pieroni, G. Montagnoli, A. Fissi, S. Merlino, and F. Ciardelli, l. Arn. Chern. Soc. 97,
6820-6826 (1975).
52. I. Ojima, T. Kogure, N. Yoda, T. Suzuki, M. Yatabe, and T. Tanake, l. Org. Chern. 47,
1329-1334 (1982).
53. K-i. Onuma, T. Ito, and A. Nakamura, Chern. Lett. 1980,481-482.
54. D. Meyer, J-c. Poulin, H. B. Kagan, H. Levine-Pinto, J. L. Morgat, and P. Fromageot,
l. Org. Chern. 45, 4680-4682 (1980).
55. D. Sinou, D. Lafont, G. Descotes, and A. G. Kent, l. Organornet. Chern. 217,119-127
(1981).
56. J. D. Morrison, R. E. Burnett, A. M. Aguiar, C. J. Morrow, and C. Phillips, l. Arn.
Chern. Soc. 93, 1301-1303 (1971). A. M. Aguiar, C. J. Morrow, J. D. Morrison, R. E.
Burnett, W. F. Masler, and N. S. Bhacca, l. Org. Chern. 41, 1545-1547 (1976).
57. D. Valentine, Jr., J. F. Blount, and K. Toth, l. Org. Chern. 45, 3691-3698 (1980); D.
Valentine, Jr., K. K. Johnson, W. Priester, R. C. Sun, K. Toth, and G. Saucy, l. Org.
Chern. 45, 3698-3703 (1980). D. Valentine, Jr., R. C. Sun, and K. Toth, l. Org. Chern.
45,3703-3707 (1980).
ASYMMETRIC HYDROGENA TlON REACTIONS 165

58. 1. M. Brown and D. Parker, I. Chern. Soc., Chern. Cornrnun. 1980,342-344 and I. Org.
Chern. 47,2722-2730 (1982).
59. D. H. G. Crout, M. Lutstorf, P. 1. Morgan, R. M. Adlington, 1. E. Baldwin, and M. J.
Crimmin, I. Chern. Soc., Chern. Cornrnun. 1981, 1175-1176.
60. J. M. Brown and R. G. Naik, I. Chern. Soc., Chern. Cornrnun. 1982,348-350.
61. I. Ojima, T. Kogure, and N. Yoda, I. Org. Chern. 45, 4728-4739 (1980).
62. D. Sinou, Tetrahedron Lett., 22,2987-2990 (1981).
63. A. S. C. Chan, J. J. Pluth, and 1. Halpern, I. Am. Chern. Soc. 102, 5952-5954 (1980).
64. B. R. James and D. K. W. Wang, Can. I. Chern. 58,245-250 (1980).
65. R. H. Crabtree, Acc. Chern. Res. 12, 331-338 (1979).
66. J. Vi lim and 1. Hetfiejs, Collect. Czech. Chern. Cornrnun. 43, 122-133 (1978).
67. A. S. C. Chan, 1.1. Pluth, and 1. Halpern, Inorg. Chirn. Acta 37, L477-L479 (1979).
68. A. S. C. Chan and 1. Halpern, I. Am. Chern. Soc. 102, 838-840 (1980).
69. 1. M. Brown and D. Parker, Organornetallics 1,950-957 (1982).
70. J. M. Brown, L. R. Canning, A. 1. Downs, and A. M. Forster, I. Organornet. Chern., 1983, in
press.
71. G. Descotes, D. Lafont, D. Sinou, J. M. Brown, P. A. Chaloner, and D. Parker, Nouv.
I. Chirn. 5,167-173 (1981).
72. R. G. Ball and N. C. Payne, Inorg. Chern. 16, 1187-1191 (1977).
73. J. M. Townsend and 1. F. Blount, Inorg. Chern. 20,269-271 (1981).
74. W. S. Knowles, B. D. Vineyard, M. 1. Sabacky, and B. R. Stults, Fundarn. Res.
Homogeneous Catal., edited by M. Tsutsui (Plenum Publishing, 1979), Vol. 3, pp.
531-548.
75. K. Toriumi, T. Ito, H. Takaya, T. Souchi, and R. Noroyi, Acta Crystallogr. 838,807-812
(1982).
76. P. C. Chua, N. K. Roberts, B. Bosnich, S. J. Okrasinski, and J. Halpern, I. Chern. Soc.,
Chern. Cornrnun. 1981, 1278-1280.
77. C. A. Tolman, P. Z. Meakin, D. L. Lindner, and J. P. lesson, I. Am. Chern. Soc. 96,
2762-2774 (1974).
78. J. Halpern, T. Okamoto, and A. Zakhariev, I. Mol. Catal. 2,65-68 (1977). M. H. 1. M.
de Croon, P. F. M. T. van Nisselrooij, H. 1. A. M. Kuipers, and 1. W. E. Coenen, I.
Mol. Catal. 4, 325-335 (1978).
79. J. Halpern, D. P. Riley, A. S. C. Chan, and 1. J. Pluth, I. Arn. Chern. Soc. 99, 8055-8057
(1977).
80. D. A. Slack, I. Greveling, and M. C. Baird, Inorg. Chern. 18,3125-3132 (1979).
81. 1. M. Brown, P. A. Chaloner, A. G. Kent, B. A. Murrer, P. N. Nicholson, D. Parker,
and P. 1. Sidebottom, I. Organornet. Chern. 216,263-276 (1981).
82. B. R. Stults, R. M. Friedman, K. Koenig, and W. Knowles, I. Am. Chern. Soc. 103,
3235-3237 (1981).
83. J. M. Brown and P. A. Chaloner, I. Am. Chern. Soc. 102, 3040-3048 (1980).
84. G. A. Morris and R. A. Freeman, I. Mag. Res. 29,433-462 (1978).
85. 1. M. Brown and P. A. Chaloner, I. Chern. Soc., Chern. Cornrnun. 1979,613-615.
86. J. M. Brown, P. A. Chaloner, and G. A. Morris, I. Chern. Soc., Chern. Cornrnun., 1983,
in press.
87. G. Descotes, D. Lafont, D. Sinou, J. M. Brown, P. A. Chaloner, and R. Glaser, I. Chern.
Soc., Chern. Cornrnun. 1979,611-613.
88. J. M. Brown and P. A. Chaloner, 1. Chern. Soc., Perkin 111982,711-719.
89. J. M. Brown and B. A. Murrer, I. Chern. Soc., Perkin II 1982,489-497.
90. 1. M. Brown and P. A. Chaloner, I. Chern. Soc., Chern. Cornrnun. 1980, 344-346.
91. G. W. Kirby and J. Michael, I. Chern. Soc., Perkin 11973,115-120.
92. F. H. Jardine, J. A. Osborn, and G. Wilkinson, 1. Chern. Soc. A 1967,1574-1579.
5
Binuclear, Phosphine-
Bridged Complexes:
Progress and Prospects
Alan L. Balch

NOTATION

Bu' tertiary butyl


Cp (1)S = CsHs)
dam bis(diphenylphosphino)arsine, Ph 2 AsCH 2 AsPh 2
dba dibenzylideneacetone, PhHC=CHC(O)CH=CHPh
dpm bis( diphenylphosphino )methane, Ph 2PCH 2 PPh 2
Me methyl
Ph phenyl
Ph 2Ppy,PN 2( diphenylphosphino )pyridine
PNP methylaminobis(difluorophosphine), F 2 PN(CH 3 )PF2
rP generalized diphosphine ligand

1. INTRODUCTION

As the existence of this volume attests, transition metal complexes with


phosphine ligands have proven to be useful and versatile catalysts for
homogeneous reactions. Recently, binuclear metal complexes, particularly
those with bridging phosphine ligands, have begun to attract interest
because of their potential as catalysts and because of their novel structural

Alan L. Balch • Department of Chemistry, University of California, Davis, California 95616.

167
168 ALAN L. BALCH

and reactive features. The interest in binuclear complexes arises because


of anticipation that they should allow for increased versatility in catalyst
design. Complexes with two (or even more) metal atoms can have several
advantages over a catalyst containing only a single metal. Small molecules,
particularly those like dinitrogen and nitriles which are notoriously difficult
to reduce, may be more readily activated by attachment to several metal
centers. (1) In catalysts containing two metal centers, one may act to bind
the substrate, while the second acts to feed or remove electrons from
the first site. The presence of two metal atoms may facilitate multi-
electron redox reactions which could not be handled by only a single metal
atom.(2)
There already exist examples of significant chemical reactions which
demonstrate the unusual reaction characteristic resulting from the presence
of two metal centers. (3) Here we examine two, neither of which involves
phosphine bridging ligands. The autoxidation of triphenylphosphine
catalyzed by four-coordinate iron(II) porphyrins in a noncoordinating sol-
vent like toluene occurs by the mechanism shown in Figure 5.1. (4,5) Two
iron(II) porphyrins bind dioxygen sequentially to give the peroxobridged
intermediate PFe(III)OOFe(III)P (P is a porphyrin dianion). This is inert
to the substrate, triphenylphosphine, but when it spontaneously fragments
it forms PFe(IV)O, which is capable of transferring an oxygen atom to

zeFe N 0

Figure 5.1. The cycle for the iron (II) porphyrin catalyzed oxidation of triphenyl phosphine
to triphenylphosphine. oxide. P is a porphyrin dianion.
BINUCLEAR. PHOSPHINE -BRIDGED COMPLEXES 169

triphenylphosphine. In formal terms, two iron(II) centers are oxidized to


iron(IV). This provides the four electrons necessary to break the oxygen-
oxygen bond of dioxygen. This is a spontaneous reaction. No bridging
ligand (other than the dioxygen) is necessary to facilitate the formation of
PFeOOFeP.
A second example of a unique transformation in a bimetallic system
is the observation of facile carbon-carbon bond making and breaking
reactions in the binuclear complexes shown in reaction 1: (6)

R
I

_
0-
::::
OCR
~/ ~c/
I
M, /MSJ_-
I/CO (1)
---' C :
" ',
O ' .. -

The complex involved is formed by photolysis of Cp2M2(CO)4 (M =


Ru, Fe) in the presence of an acetylene. The fluxional process shown in
reaction 1 occurs so rapidly that the two cyclopentadienyl resonances
coalesce at 56°C, and at 90°C the 13C resonances due to the ketonic and
terminal carbonyl ligand are merged.
Along with the interest in binuclear complexes, there has been con-
siderable development of the chemistry of metal carbonyl clusters. The
analogy between metal clusters and metal surfaces as catalytic centers has
been drawn. (7.8) A number of clusters have been shown to act as efficient
catalysts or catalyst precursors. (9) The preparation of phosphine-bridged,
polynuclear complexes can be seen as complementary to the synthesis of
cluster catalysts. Whereas metal carbonyl clusters are held together
primarily by metal-metal bonds, phosphine-bridged, polynuclear com-
plexes have the phosphine bridges, as well as metal-metal bonds to maintain
the integrity of the polynuclear unit. Thus, we find that selective metal-
metal bond breaking (as opposed to cluster degradation) is relatively rare
for carbonyl clusters. (An example exists in the oxidation of
[OslOC(CObr- by iodine to give [OslOC(CObIr and OslOC(COh4h (10))
However, metal-metal bond breaking and formation is a fairly common
reaction for phosphine-bridged binuclear complexes (vide infra). Addi-
tionally with phosphorus ligand-bridged complexes there is an added
dimension of synthetic flexibility that occurs because the bridging ligand
can now be tailor-made to meet certain requirements. For example, we
shall encounter cases where bridging ligands have been constructed with
different sites capable of binding different metal ions.
170 ALAN L. BALCH

2. STRUCTURAL ASPECTS

A variety of phosphorus containing ligands have been utilized in the


preparation of binuclear complexes. Among the most versatile are the
diphosphines of the type Ph2P(CH2)nPPh2. These are readily available (at
the present time from commercial sources) and are easily handled, air-stable
solids. (11,12) The length of the methylene chain can be varied to give different
separations between the donor atoms and hence different spacings between
metal centers. Many of the diarsine analogs of these ligands are known
and form similar complexes. The formation of binuclear complexes with
these and other ligand types is not immediately assured, for all of them
have the potential for forming monodentate and chelating complexes, as
well as diphosphine-bridged complexes. Of particular relevance in this
context are the observations of Sanger, who showed that the reaction of
Ph2P(CH2)nPPh2 with [Rh(COh(IL-CI)h produced binuclear complexes
(IL-Ph2P(CH2)nPPh2hRh2(COhCh for n = 1,3,4 (and this has later been
extended to n = 5 and 6), whereas the chelated monomer
Ph 2P(CH 2hPPh 2Rh(CO)CI was formed when n = 2.0 3 ) While these results
are not entirely general for all metal ions,(14) they serve to emphasize the
fact that bis(diphenylphosphino)ethane and, of course, cis-
bis(diphenylphosphino)ethylene generally have strong preferences to bind
as chelating rather than bridging ligands. Additionally, it should be noted
that the presence of a diphosphine ligand in a binuclear complex does not
imply that the phosphine acts as a bridge. For example,
(Bu~P(CH2hPBu~hPt2 is a binuclear species with an unbridged metal-
metal bond and a chelating diphosphine bound to each metal. (15.16) Related
to these diphenyl phosphino ligands are analogs bearing other groups as
the terminal phosphorus substituents. Those in the t-butyl groups have
been extensively investigated by Shaw and his group. (17)
The aminobis(difiuorophosphines), RN(PF 2 h, form another group of
bridging ligands that has received considerable attention. (18) These do not
have the capacity for variable spacing of the phosphino groups, but other
relatives, F 2 PXPF 2 where X = 0, S, (CH 2)n, could offer that possibility.
The presence of the fiuoro substituents makes this class of ligand a good
7T-acceptor. Consequently, these ligands stabilize low-metal oxidation
states. Moreover, the small size of the fiuoro groups makes such ligands
considerably less bulky than the ligands bearing diphenylphosphino groups.
Ligands of the type R 2 0PXPOR 2 where X can be 0, S, NR, (CH 2)n
also are potential candidates for bridging ligands. Of these, tetraethyl-
diphosphite, (EtOhPOP(OEth, (19-22) has received the most attention,
while some work with (ROhPNR'P(ORh has been reported. (23.24)
Tetraethyldiphosphite has been observed to function as either a bridging
or a terminal ligand. It is not known to function as a chelating ligand, and
BINUCLEAR, PHOSPHINE-BRIDGED COMPLEXES 171

it has been speculated that the favored P-O-P bond angle of ca. 120-150°
precludes chelation. (22)
Phosphide anions, R 2 P- are also known to function as bridging
ligands. (25,26) In this regard, they resemble the more familiar bridging halide
and chalconide ligands. We have, somewhat arbitrarily, excluded these
from this review.
Within this group, the number of bridging phosphorus ligands in a
binuclear complex can vary from one to four. However, the most extensive
studies have been carried out on species containing two bridging phos-
phines. Two phosphine ligands are sufficient to hold the two metal centers
together while allowing the reactants to have reasonably good access to
the metal centers. They also accommodate a significant variation in the
metal-metal separation even for a single ligand. This aspect of flexibility
appears to be a significant benefit of these diphosphine ligands.
Three and four bridging phosphorus ligands have only been observed
with the relatively small ligand CH 3 N(PF 2b (PNP). With four such ligands,
the two metal ions are effectively enclosed in a cage. This can be seen in
Figure 5.2 where the structure of ({.L- PNP)4M02Cl2 is shown. (27) The conges-
tion about the two metal ions would appear to offer a significant impediment

FlO

CII
FlO'
~~""'1'.Pf

Figure 5.2. The structure of (Il-PNP)4Mo2CI2 taken from Reference 27 by permission.


172 ALAN L. BALCH

(II)

Figure 5.3. The structure of (/-L-PNPhCo 2(COh taken from Reference 28 by permission.

Br(2)

Figure 5.4. The structure of (/-L-PNPhCo2Br4 taken from Reference 29 by permission.


BINUCLEAR, PHOSPHINE-BRIDGED COMPLEXES 173

to catalytic activity where substrate needs to bind to the metal ions involved,
unless bonding only to the axial sites was required. Compounds with three
bridging phosphorus ligands suffer from similar although less extensive
crowding. The structure of (~-PNPhC02(COh is shown in Figure 5.3.(28)
The related complex (~-PNPhC02(PF2NHMeh shows similar structural
features. (28) Oxidation of the latter complex with bromine yields (~­
PNPhC02Br4 which is shown in Figure 5.4.(29) Here, the three phosphorus
ligands have moved together to accomodate two additional ligands. In
addition to retaining the three PNP ligands, the complex also retains its
Co-Co bond.
With two phosphorus ligands acting to support the binuclear com-
pounds, there is an extensive body of data available. Generally, these
bridging ligands adopt trans arrangements at each metal ion as shown in
A of Figure 5.5. In most cases, this probably occurs for steric reasons,
since it places the phosphorus atoms and their bulky substituents as far
apart as possible. However, there are now cases known where the phosphine
ligands bridges are both cis on each metal (C of Figure 5.5) and where
the phosphine bridges are cis on one metal and trans on another (B of
Figure 5.5). Examples of types Band C are the platinum methyl complexes
(~-dpmhPt2(CH3)4 shown in Figure 5.6(30) and (~-dpmhPt2(CH3); shown
in Figure 5.7.(31)
Within the large group of molecules containing trans phosphorus
ligands on both metals, several groups of ligand arrangements can be
identified. These are shown in Figure 5.8. The face-to-face dimers have
the greatest flexibility in that there are no other connections between the
metals. With long chain diphosphines, this can lead to the formation of

A B

c
Figure 5.5. Orientation of two diphosphine ligands about two metal ions.
174 ALAN L. BALCH

Figure 5.6. The structure of (lL-dpm)zPt 2 (CH 3 )4 taken from Reference 30.

complexes in which the two metal centers are so remote that they are
effectively isolated from one another. An example is given by the struc-
ture of (~-Ph2P(CH2hO(CH2hPPh2hRh2(COhClz as shown in Figure
5.9.(32) The two square planar rhodium ions are separated by 7.95 A.
Similar structures with large gaps between metal ions are seen for the

C(46)

C(39) C(45)
C(40)

C(44)

C(20)

C(22)

Figure 5.7. The structure of (lL-dpm)zPt 2 (CH 3 h + taken from Reference 31.
BINUCLEAR, PHOSPHINE -BRIDGED COMPLEXES 175

",r-'~~r/
p~p

I ,
--M M-- M M
I
p~p
I I"'-x/ I
p~p

Side-by-Slde A-FrcJlle

p~p

--M"
I ' ,y" "M---
I
I~x/I
p~p

Face-to-Face Double A-Frome

Figure 5.8. Idealized geometries of binuclear-trans-diphosphine complexes.

Figure 5.9. The structure of (iL-Ph2P(CH2hO(CH2hPPh2)Rh2(CO)zCI2 taken from Refer-


ence 32 by permission.
176 ALAN L. BALCH

Rh' Rh

0' ((1)' (I' CI ((I) 0

Figure 5.10. The structure of (/-L-dpm}zRh 2 (COhCh taken from Reference 35 by permission.

26-membered rings in (p,-Burp(CH 2h oPBurhRh 2(COhCh and (p,-


Bu~P(CH2hoPBu~hPd2CI4(33) and in the 16-membered ring in (p,-
Bu~P(CH2hCMeH(CH2hPBu~hPd2CI4.(34) With fewer methylene (or
analogous groups) in the chain, the two metal centers can be brought into
very close proximity without necessarily entailing the formation of a classical
metal-metal bond. For example, in (p,-dpmhRh 2(COhCh shown in Figure
5.10 the Rh···Rh separation is only 3.24 A.(3S) The structurally similar
complex (#J.-dam)zRh2(COhCh shows the plasticity of molecules of this
sort which are not supported by a metal-metal bond. The structures of
two different crystalline forms of this complex show significant variation
in the Rh .. · Rh separation. In the unsolvated form, this separation is
3.396(1),(36) while in the dichloromethane solvate it is 3.236(2).(37)
Other variations of the face-to-face structure can be identified. An
additional ligand can be present on one or both metals to give five-
coordinate species. The two metals may actually be directly bonded. Such
is the case for (#J.-dpmhM0 2C14 where a Mo-Mo quadruple bond is present,
and the Mo-Mo separation is only 2.138 A.(38)
BINUCLEAR, PHOSPHINE -BRIDGED COMPLEXES 177

Figure 5.11. The structure of (lL-dpmhPd 2 CI(SnCI 3 ) taken from Reference 39 by permission.

The side-by-side dimers have a direct metal linkage and the square
planar coordination about each metal is completed by the presence of two
terminal axial ligands. A typical example is (M-dpmhPd 2CI(SnCh) shown
in Figure 5.11. (39)
The dimeric structures can also contain additional bridging ligands.
These include the A-frame structure, as well as the doubly-bridged A-frame
structure. Figure 5.12 shows the structure of (M-dpmhPd 2(M-S02)Ch, a
typical A-frame,(40) Figure 5.13 shows the structure of a doubly-bridged
A-frame. (41) Of some interest is the observation that the structure of these
A-frames is amazingly consistent from molecule to molecule. The CP 2M 2X
rings generally adopt boat, rather than chair, conformations, and the
orientation of phenyl rings appears to be common to many of the complexes.
While complexes of the long-chain diphosphines are largely limited
to the face-to-face arrangement of metal centers, the smaller bridging
ligands, particularly dpm and dam, have the capacity of stabilizing all four
structural types and of allowing for the interconversion between these
various forms. (42) The flexibility of these ligands in spanning a variety of
metal-metal separations is considerable. These range from cases where
178 ALAN L. BALCH

11--"'--- Q0
CI 2

Figure 5.12. The structure of (/L-dpmhPd 2(/L-S02)CI 2 taken from Refenrence 40 by per-
mission.

Figure 5.13. The structure of (/L-dpmhRh2(wCO)(/L-Cl)(COh + taken from Reference 41


by permission.
BINUCLEAR. PHOSPHINE-BRIDGED COMPLEXES 179

Table 1. Structural Parameters for Bis(diphenylphosphino)methane (dpm)


Bridged Complexes

Compound M'''M(A) p ... p (A) p-c-p (0) Reference

(/-t-dpmhMo 2CI4 2.138(2) 106.2(3) 38


(/-t-dpmhMo 2(NCS)4 2.167(3) 107(1) 38
(/-t-dpmhMn2(/-t-CO)(CO)4 2.934(6) 3.05(1) 115(2) 43
3.03(1) 111(2)
(/-t-dpmhMn2(/-t-CNC6H4Me-p )(CO)4 2.936(2) 3.129(3) 115(1) 44
3.103(3) 113(1)
(/-t-dpmhRe2C1s 2.263(1) 107.5(11) 45
106.3(10)
(/-t-dpmhRe2(CO)6(Hh 2.893(2) 46
(/-t-dpm)Fe2(/-t-CO)(CO)6 2.709(2) 2.999(4) 109(1) 47
(/-t-dpm)Ru3(COh(/-t3-PPh)(/-t-CHPPh2) 2.843 48
(/-t-dpm)Co 2(/-t-C 2 Ph 2)(CO)4 2.459(2) 2.946(4) 106(1) 49
(/-t-dpmhRh4(CO)8 2.696(1) 113.8(7) 50
2.671(1) 111.0(7)
[(/-t-dpmhRh(/-t-H)(/-t-CO)(COhr 2.732(2) 51
(wdpmhRh2(/-t-C4F6)CI2 2.744(9) 2.959(3) 108.4(4) 52
2.981(3) 110.1(4)
(/-t-dpmhRh 2 (/-t-CO)Br2 2.7566(9) 3.026 111.0(4) 53
2.992 109.9(4)
(/-t-dpm)2Rh2(/-t-S02)CI 2 2.7838(8) 3.018(3) 111.6(4) 54
3.012(3) 110.5(5)
(/-t-dpmhRh 2(/-t-C 2S4)CI 2(CO) 2.811 (3) 3.016(9) 109(1) 55
3.094(11) 113(1)
[(wdpmhRh 2(/-t-CO)(/-t-Cl)(CO)2] 2.810(3) 3.024(10) 110(1) 41
3.096(10) 115(1)
[(/-t-dpm)zRh 2(/-t-CO)(/-t-Cl)(CO)2r 2.838(1) 3.060(4) 114.2(6) 56
(wdpmhRh 2(wS)(COh 3.154(2) 57
[(/-t-dpmhRh 2(/-t-Cl)(COhr 3.1520(8) 3.088(2) 114.7(3) 58
(/-t-dpmhRh 2 (COhCI 2 3.2386(5) 3.130(1) 116.8(2) 35
(/-t-dpmhRh 2(/-t-CO) 3.3542(9) 59
(/-t-C2{Co 2 Me}z)CI 2
(/-t-dpm)zIr 2(/-t-S )(/-t-CO)( CO h 2.843(2) 3.045(6) 11.6(7) 60
3.028(6) 111.7(7)
(/-t-dpmhPd 2(SnCI 3)Cl 2.644(2) 2880(6) 102(1) 39
2.960(6) 108(1)
180 ALAN L. BALCH

Table 1. (Continued)

Compound M"'M(A) p···p(A) P-C-P (0) Reference

(wd pmhPd2Br2 2.699 105(2) 61


103(2)
[(J.L-dpmhPd 2(J.L-I)(CH 31It 2.976(6) 3.15(2) 62
3.12(2)
[(J.L-dpmhPd 2(J.L-CNCH 3)(CNCH 3)2t 3.215(2) 3.066(6) 114(1) 63
(J.L-dpmhPd 2(J.L-S)C1 2 3.258(2) 3.059(7) 112.0(10) 40
3.508(7) 113.6(9)
(J.L-dpmhPd 2(J.L-S02)C12 3.221(2) 3.111(5) 119(1) 40
3.383(2) 3.170(5) 115(1)
(wdpmhPd(J.L-C4 F 6 )C12 3.492(1) 3.115(4) 114.6(6) 64
3.113(4) 115.4(6)
(J.L-dpmhPt 2(CO)Cl+ 2.620 107 65
106
(J.L-dpmhPt 2Cl 2 2.652 66
(J.L-dpmhPt 2H(dpm) 2.770(2) 67
(J.L-dpmhPt 2(CH 3h + 2.769(1) 115.0 31
111.5
(J.L-dpmhPt 2(J.L-H)(CH 3h + 2.933(1) 68
(J.L-dpmhPt 2(J.L-CS 2)C1 2 3.094 69
(J.L-dpmhPt 2(CH 3)4 4.36 30
(J.L-dpmhCu3Cl2 2.678(6) 70
(J.L-dpm)Cu3(J.L-dpm-Hh 2.836 71

(J.L-dpmhCu 4 I 4 2.908(4) 106(1) 72

(J.L-dpmhCu 4 Br4 2.939(6) 112(1) 72

(J.L-dpmhCu 4 Cl4 3.110(6) 114(1) 73

(J.L-dpmhCu3I3 2.916(4) 74

[(wdpmhAg 2f+ 3.041 117.5 75

(wd pmhAg3Br3 3.303 110 76

(J.L-dpmhAu2C12 2.962(1) 114.2(8) 77

(J.L-dpm)Au2C12 3.341 116 78


BINUCLEAR, PHOSPHINE -BRIDGED COMPLEXES 181

Table 2. Structural Parameters for Bis(diphenylarsino)methane (dam)


Bridged Complexes

Compound M···M (A) As···As (A) As-C-As(O) Reference

(l-t-dam)zW 2(I-t-C 2Me2)(CO)sBr 2.937(1) 79


(l-t-dam)zRe2(I-t-CI)z(CO)4 3.806(1) 3.320(2) 115(1) 80
3.814(1) 3.394(2) 113(1)
(l-t-dam)zRu 3(1-t3- 0 )(CO)6 2.750(2) 81
2.723(2)
(l-t-damjzCo 2(I-t-C2Ph 2)(COjz 2.518(4) 3.165(3) 105(1) 82
3.246(3) 108(1)
(l-t-dam)zRh 2(COhCIz 3.396(1) 3.298(1) 114(1) 36
(l-t-damhRh 2(COhCI 2·CH2CI 2 3.236(2) 3.272(2) 114.4(5) 37
(l-t-dam)2Pd2(I-t-CO)CI 2 3.274(8) 3.22(1) 111 (3) 83
(l-t-damhPt 2(I-t-CO)CI 2 3.162(4) 3.22(1) 107(3) 84

there are no metal-metal bonds but other bridging ligands that require
rather large metal-metal separations to cases where there are metal-metal
multiple bonds. Some idea of the degree of flexibility can be gained from
Tables 1, 2, and 3, which show some structural parameters for a variety
of complexes of dpm, dam, and PNP, respectively. Factors responsible for

Table 3. Structural Parameters for Methylaminobis(difluorophosphine)


(PNP) Bridged Complexes

Compound M"'M(A) P-N-P (0) Reference

(I-t-PNP)4Mo2CI2 2.457(1) 111.9(2),118.4(2) 27


114.8(2), 109.2(2)
(I-t-PNP)zFe2(I-t-CO)(CO)4 2.661(1) 85
(I-t-PNPhFe2(I-t-PF2)(CO) 2.725(2) 86
(PF 2NMe)

(I-t- PNPhFe2 (I-t- PF 2) 3.646 87


(I-t-PF2NMe)Cp2
(I-t-PNPhFe2(CO)6 3.90 85
(w PNPhCo 2(COh 2.716(1) 28
(I-t-PNPhCo 2Br4 2.717(5) 29
(I-t-PNPhCo 2(PF2NMeHh 2.769(1) 28
182 ALAN L. BALCH

Figure 5.14. The structure of (/.L-PhzP(Chz)4PPhz)Rhz(COjzCpz taken from Reference 88


by permission.

the flexibility of the dpm ligand include changes in the P-C-P angle,(41)
changes in the conformation of the ring formed by the metal(s) and the
ligand,(47) and changes in the P-M-M-P torsional angles.(76)
Compounds with only one phosphorus ligand linking the two metal
centers are also known. These have the least constraints placed on the
interaction between the two metals. One such molecule, (J.L-
Ph2P(CH2)4PPh2)Rh2Cp2(COh, is shown in Figure 5.14.(88) Other species
with a single phosphine linking two essentially independent metal centers
include (J.L-Ph2P(CH2)4PPh2){Rh(CO)(Ph2P(CH2)4PPh2)h 2+, (89) ciS-(J.L-
Ph2PCHCHPPh2)Au2Ch, (90) and J.L-dpm{Ru(bipyhClh 2+. (91)
Recently the problem of constructing bifunctional phosphine ligands
with two different binding sites, each capable of bonding a different type
of metal ion, has received attention. Two such ligands are 2(diphenylphos-
phino)pyridine, (92) (Ph 2Ppy) 1, and [dimethyl( diphenylphosphino-
methyl)silyl]cyclopentadienyllithium, 2.(93)

Ph
Ph-P
D
" I N~
1 2

The behavior of Ph 2Ppy as a bridging ligand will be described in a succeeding


section.
BINUCLEAR, PHOSPHINE-BRIDGED COMPLEXES 183

3. NUCLEAR MAGNETIC RESONANCE SPECTROSCOPIC PROBES

NMR spectroscopy has been useful in structural characterization of


binuclear phosphine-bridged complexes. The most extensive data available
involved complexes involving two trans dpm ligands, and this brief section
will be devoted to a review of some general features which have been found.
The lH-NMR spectra of dpm ligands are informative in establishing
the mode of coordination. In dpm itself the methylene resonance occurs
as a triplet with 3J (PH) = 1.5 Hz at 2.8 ppm.(94) When dpm acts as a
chelating ligand, the methylene resonance is shifted downfield to 4-4.5 ppm,
and the resonance appears as a triplet with J(P, H) == 10 Hz. (94.95)
For bridged complexes, two types of spectra are observed. For com-
plexes like (wdpm)zPdzCh, where rapid conformational change in the
CPzPd z ring can render the two methylene protons equivalent, the methyl-
ene resonance appears as a 1 : 4: 6: 4: 1 quintet with the apparent J(P, H) =
5 Hz. (95.40) The quintet pattern arises due to virtual coupling of the phos-
phorus nuclei, where the trans P-Pd-P coupling is expected to be - 300 Hz.
While such a complex belongs to the AA'XX'X"X"'A"A'" (A = H, X = P)
spin system, the spectra can be more readily analyzed as an AXX'X"X'"
system since J(A, A") and J(A, A"') are expected to be very small. Some
spectra simulations are shown in Figure 5.15. The values chosen for J (2, 3),
J(4, 5), J(2, 5), and J(2, 6) are in the normally observed range. Once the
trans P-P coupling constant exceeds 100 Hz, the spectral pattern appears
as a symmetrical quintet and the apparentJ(P, H) is the average of lJ(p, H)
and 3J (p, H).
For molecular A-frames, no motion of the CPzMz ring can render the
two methylene protons of a dpm ligand equivalent. However, for most
A-frames the two dpm ligands are generally equivalent by symmetry.
Hence, there the methylene group of dpm appears as an AB quartet with
J(H, H) in the range of 12-15 Hz. Frequently, the chemical shift difference
between the two types of protons reaches 0.5 to 1 ppm. Superimposed on
this AB quartet is the phosphorus-proton coupling which further splits
each resonance into a quintet. Again virtual coupling is responsible for the
splitting pattern. As a result of these considerations, lH-NMR spectra are
capable of differentiating between binuclear A-frames and other dimers
in which the two methylene protons are equivalent.
Proton-decoupled(31) P-NMR spectroscopy has also been useful in
structural characterization. The most frequent use has involved attempts
to simulate the observed and often complex spectral patterns on the basis
of an assumed structure and reasonable values for the appropriate coupling
constants and chemical shifts. It is the coupling constants, more than the
chemical shifts, that are structurally useful, and a number of coupling
constants are now known that are unique to binuclear complexes. The
184 ALAN L. BALCH

I Iii iii iii i


'i.2 'i.0
I
3.8 'i.2 3.8
I
PPM

Figure 5.15. Simulations of the lH_NMR spectra of methylene protons of a model (f.L-
dpm),M2 complex with trans-phosphine ligands. The model is based on a AX2 Y 2 spin system.
(Taken from the Ph.D. thesis of C. T. Hunt. University of California Davis 1981.) Sets, A-H,
of coupling constants, I ij , were entered into a simulation routine and produced spectra A-H.
These spectra show the phenomenon of virtual coupling (see text).
1
H
2 / C ....... 3
P P
I
M ~
I
p
I
P
4 ....... C / 5

Sets Iij (Hz) 12 13 14 15 23 24 25 34 38 45

A 10 10 0 0 0 0 0 0 0 0
B 10 10 0 0 30 0 0 0 0 30
C 10 10 0 0 30 0 5 5 0 30
D 10 10 0 0 30 5 5 5 5 30
E 10 10 0 0 30 150 5 5 150 30
F 10 10 0 0 30 300 5 5 300 30
G 5 5 5 5 30 300 5 5 300 30
H 20 20 -10 -10 30 300 5 5 300 30
BINUCLEAR, PHOSPHINE -BRIDGED COMPLEXES 185

question arises as to how much structural information is contained within


these parameters.
Most of the available information can be discussed by reference to
structure 3:

which encompasses a (~-dpmhM2 unit that may contain two additional


ligands Ps and P6 (but these need not be present). Coupling constants
between trans ligands [i.e., J(l, 2)] are known to be large (ca. 300 Hz) for
Group VIII metal ions, while cis coupling constants [i.e., J(l, 5)] are smaller
(ca. 10-52 Hz). Major questions to be considered are: do the other coupling
constants which are more or less unique to binuclear species reflect the
presence and strength of metal-metal bonding, and do they reflect the
geometry of the unit?
Pertinent data are given in Table 4. Some attempts at analysis of this
data have been made. (101.108) The P-P coupling constant in an unsym-
metrical diphosphine R 2 PCH 2 PR ~ is generally about 110 Hz. (It is 115 Hz
for Ph 2PCH 2PMePh.)(109) In the complexes, J(l, 3) is always of smaller
magnitude but otherwise offers little structural information. It has been
proposed that J(l, 4) is sensitive to the presence of a metal-metal bond.
Low magnitudes of J(l, 4) correspond to cases where there is no metal-
metal bond, and larger magnitudes occur in compounds which are expected
to have metal-metal single bonds. (l08) Within Table 4, (~-dpmhPt2(~­
CI)H 2+ appears to be the exception, since this compound should have no
Pt-Pt bond. Although originally proposed for platinum compounds, this
r~lationship also is maintained for all of the palladium and rhodium com-
pounds for which it can be tested. Similarly, it has been proposed that, for
platinum, the two-bond coupling J(1, 8) is also effected by metal-metal
bonding. (l08) For those compounds which have Pt-Pt single bonds, J(l, 8)
is negative with values of ca.-100 Hz, while with compounds lacking a
Pt-Pt bond J(l, 8) is positive. The two-bond P-Pt coupling constant also
appears to give geometric data. When the P-Pt-Pt unit is linear, 2J (Pt, P)
is positive and falls in the range 150-750 Hz, while for cases where the
P-Pt-Pt angle is -90°, 2J (Pt, P) is negative and smaller in magnitude
(85-120 Hz). (107)
Two technical problems in dealing with 31 p _NMR spectra of binuclear
complexes are worthy of special notice. Selective population transfer
~
Table 4. Coupling Constants from the 31ptH} NMR Spectra of (p.-dpm)2M2 Complexes a

CouQJin~nstants (Hll
Compound 12 13 14 15 16 17 18 56 57 58 78 Reference

(/J.-dpm)zRhz(/J.-Cl)(/J.-CO)(CO)z + -355 68.4 21.5 93.4 0.8 0.1 96


(/J.-dpm)zRhz(wBr)(/J.-CO)(CO)z + -363 71.6 22.0 92.6 0.8 0 96
(/J.-dpm)zRhz(/J.-I)(/J.-CO)(CO)z + -548 74.3 19.9 90.4 2.3 0.2 96
(/J.-dpm}zRhz(/J.-CI)(/J.SOz)(CO}z + -350 46.9 17.5 90.8 0.5 0 96
(/J.-dpm}zRhz(/J.-Cl)(wCO)(CNBu'}z + -390 82.4 21.9 106.9 0.8 0 97
(/J.-dpm}zRh z(/J.-CO)(CNBu')4 +z -327 78.2 21.9 88.9 -1.6 0 97
(/J.-dpm}zRh z (/J.-SOz)(CNBu')4 +Z -329 58.6 27.2 91.0 -1.2 0 97
(/J.-dpmhPd z(/J.-CHCH 3 )I z ±363.6 ±104.5 ± 5.5 98
±44.9
(/J.-dpmhPd 2{/J.-HC z(Co zH)}CIz 44.7 0.8 99
(/J.-dppmhPd 2{/J.-HC 2 Co 2 Me}CI 2 48.5 1.0 99
(/J.-dpmjzPdzCIBr 43.2 35.5 100
(/J.-dpmhPd 2 BrI 39 39 100
(/J.-dpmjzPd 2 ClI 39 39 100 h
+
....
h
(/J.-dpmhPtz(/J.-Cl)Mez 33 3 3030 30 101
<:
(/J.-dpmhPtz(/J.-CI)Ph z+ 29 ,,;3 3037 39 102 :-
(/J.-dpmhPtz(wCI)(COPhh + 28 ,,;3 3354 48 102 ....~
(/J.-dpmhPt 2 (/J.-CH 2 )CI 2 23 6 3388 73 103 2
(lL-dpmhPtz(IL-S)Cl z 43 0 3588 273 103
~
c:
(lL-dpmhPtz(~-CI)Hz + 52.9 15.5 2665 6.8 104
p
(lL-dpmhPtz(IL-CH)Cl z+ 52.0 21.2 2317 -40.4 104 ~
?l
(lL-dpmhPtz(IL-H)H z+ 56.0 18.2 2769 16.6 104
(lL-dpmhPtz(CO)CI+ 46 43 2706 62 105
~
0
(J)
2573 92
(lL-dpmhPtzCl z 62.5 26.4 2936 -136 103
~
~
(lL-dpmjzPtz(PPhMezh z+ 48 22 24 2840 2.100 1938 650 106
~
(lL-dpmjzPtz(PPhzMeh 2+ 50 20 22 2870 -120 1914 750 106 6C)
(lL-dpmhPt 2(C sH zNjz 2+ 58 30 2828 -113 4940 106 ~
C)
(lL-dpmhPtz(COjz z+ 46 33 2390 -96 4810 106 0
:s::
(lL-dpm)pt z(IL-S)(PPh 3 h 37.9 14.7 7.6 +3537.5 -108.6 174.8 3191.6 1231.2 107 ;:E
~
a The numbering system is: m
~
PI P3
, ,
S 7
P -M - - Ms -P 6
, ,
P2 " - . / P4

~
188 ALAN L. BALCH

experiments have been shown to be valuable in analyzing the complex


spectra that result when several isotopomers (isotope isomers) are
present. (107) This is particularly common with platinum where the spin
one-half isotope, 195pt, is 33.4% naturally abundant. Secondly, the tem-
perature dependence of 31 p chemical shifts can produce unusual effects in
the spectra. We have encountered cases in (~-dpmhPd2XY (X = Cl,
Y = Br or 1) where the chemical shifts of the two chemically distinct
phosphorus atoms approach each other with increasing temperature and,
in one case, they actually merge and then move away again. (100)

4. REACTIVITY PA TTERNS FOR BRIDGED,


BINUCLEAR COMPLEXES

Behind the interest manifested in binuclear complexes stands the


expectation that they will display fundamentally new modes of reactivity.
Naturally, we also expect that they can show the patterns of reactivity
known for mononuclear complexes. These include Lewis base associ-
ation/dissociation, Lewis acid association/dissociation, ligand migration
(insertion/deinsertion), oxidative addition/reductive elimination, and oxi-
dative ligand coupling/reductive ligand uncoupling, as well as electron-
transfer. (110.111) While these reaction patterns do occur with binuclear
complexes, it is the new reactions that are emphasized in this section. Not
surprisingly, the bulk of these involve either the formation or breaking of
metal-metal bonds or the formation or loss of bridging groups other than
the phosphorus ligands. Such reactions generally require that the two metal
ions involved get relatively close to one another. As a consequence, ligands
which allow the metal to maintain their distance from one another do not
induce particularly interesting modifications. Conversely, those ligands
which force the metals into close association, also induce the more interest-
ing new chemistry.
The dinuclear rhodium(I) complex [~-Ph2P(CH2)4PPh2]Rh2(Cph­
(COh offers a good example of a case where the two metal centers are so
loosely linked that each acts essentially independently of the other to give
reactions which are characterisitc of R 3PRh(Cp)CO.(88) For example,
(PhMe2P)Rh(Cp)CO reacts with bromine to form the salt [(PhMe2P)-
Rh(Cp)(CO)Br]Br, and [~-Ph2P(CH2)4PPh2](Rh(Cp)(CO)h reacts
similarly to produce ([~-Ph2P(CH2)4PPh2](RhCp(CO)Brh}Br2.
The doubly bridged complex [~-Ph2P(CH2hPPh2]2Rh2(COhClz shows
a different sort of behavior. (112) The mononuclear Rh(I) complex
(Ph 3PhRh(CO)CI adds one molecule of sulfur dioxide or tetracyano-
ethylene to give the adducts (Ph3PhRh(CO)CI(S02) and
BINUCLEAR, PHOSPHINE -BRIDGED COMPLEXES 189

(Ph3PhRh(CO)CI(C6N4), respectively. While both rhodium ions initially


reside in identical environments in [1L-Ph2P(CH2hPPh2J2Rh2(CO)ZCh, the
addition of one mole of substrate to form [1L-Ph 2P(CH 2hPPh 2h(Rh-
(CO)CI)(Rh(CO)CI(S02)) or [1L-Ph 2P(CH 2hPPh 2h(Rh(CO)Cl)(Rh(CO)-
CI(C6N4)) effectively alters the reactivity of the second (unreacted) site so
that further addition of substrate does not occur even when the substrate
is in one-hundredfold excess. The details of the origin of this effect are not
known. There is no evidence that the substrate has bound to the one
rhodium in any usual fashion, and it certainly does not bridge the two sites.
However, it does appear probable that there is a steric interaction between
the coordination spheres of the two metal ions that prohibits the four-
coordinate rhodium from reacting.
In the remainder of this section, we deal with cases in which the two
metal centers are in much closer proximity.

4.1. Insertions of Small Molecules into Metal-Metal Bonds

The metal-metal bonds in the dimeric Pd(I) and Pt(I) complexes,


(lL-dpmhM2X 2 (X = CI, Br, I, N3), show an unusual propensity to insert
small molecules. The overall reaction converts a side-by-side dimer to an
A-frame as shown in Equation 2 and results in a substantial increase in
the metal-metal separation:
p~p p~p

I I
X-M-M-X+A ---+ ~t J~ (2)
I I
p~p
IVI
p~p

The metal-metal separations in (lL-dpmhM2X 2 are about 2.6 A. Such


distances are comparable to those in unbridged Pd(I) and Pt(I) dimers (e.g.,
Pd 2(CNCH 3)6 2+, 2.5310(9);(113) Pd 3(CNCH 3)g(PPh 3h 2+, 2.5921(5);(114)
and Pt2(CO)zCL/-, 2.584(2).0 15 ) On the other hand, the products of
insertion have metal-metal distances in excess of 3.1 A, which puts them
outside the range of full metal-metal bonds. In electron-counting terms,
both the reactant complex and the product have 16 electrons per metal,
and both exhibit nearly square planar coordination. A theoretical,
molecular orbital interpretation of the bonding in these A-frames is
available. (116)
The range of substances which insert into the metal-metal bonds of
(lL-dpmhPd 2Ch and (lL-dpmhPt 2Ch is summarized in Figure 5.16. Note
that both a one-atom bridge (CO,42,63,84 CNR, (42) S, (40,103) S0 2, (40,103)
CH2, (103) N 2Ph +(117) insertion) and a two-atom bridge (acetylene, (64,99)
CS 2(69) insertion) can be placed between the two metal atoms. In the case
+
~
X",,-' /X
•.........r/x x~ ("'.1/ X X~ 1
...........1
'!"t A M M My M
I{-S I '="1 I' f "I
~p
Ip~p p~p

h
~ r· •• +;7
~ ? ....... ,

r P p/ "p
X",,-
" A1/ X < RC=CR X-M
1
M-X..--
1 ~ X~("r/X
M M
"c=c'RI n I I "co l'c"l
p. 0 p
p~p p~p
'¥'
HI • I " ' - - R.,
~ ("NI)

-
.. / r
r~p p~p ..........P ~
r-
x~ I/x x~ I 1/ x X~, I/x ~
M M <!
!"'"
,'s'"
p~p
i'CHfi
p~
i'~"i
p~p ~
r-
2
Figure 5.16. Reaction chemistry of (/L-dpmhPdzCl z and (/L-dpmhPtzCl z.
BINUCLEAR. PHOSPHINE -BRIDGED COMPLEXES 191

of acetylene insertion, this two-atom insertion places the two metal ions
(Pd) 3.492 A apart, the largest separation yet found in a dpm bridged
complex. A number of other small molecules including dioxygen,
dinitrogen, and nitrile are unreactive toward (p,-dpm)zPd 2Clz.
In some cases the insertion reactions are reversible. This is the case
with the palladium complexes for carbon monoxide and sulfur dioxide.
With acetylenes, preliminary studies indicate that the insertion products
can be made to eliminate the acetylene on photolysis. The remaining
A-frames shown in Figure 5.16 are formed irreversibly.
Qualitatively, the chemical behavior of palladium and platinum in
(p,-dpm)zM2Clz is similar. Both metals appear to undergo similar reactions.
Detailed comparisons of equilibrium constants, reaction rates, and struc-
tural parameters are generally not possible because relevant data are
lacking. The one place where a comparison can be made involves the solid
state structures of (p,-dam)zPd 2(p,-CO)Clz (83) and (p,-dam)zPt 2(p,-
CO)Clz. (84) These molecules have closely similar A-frame structures which
are compared in Figure 5.17. Despite the similarities, the metal-metal
separation and M-C(O)-M angle are noticeably lower for (p,-dam)zPt2(p,-
CO)Clz than for (p,-dam)zPd 2(p,-CO)Clz. Since it is known that these
parameters are subject to external pressures on the molecule, the differences
could be simply due to packing forces. However, the observation of
decidedly different carbonyl stretching frequencies for (p,-dam)zPd 2(p,-
CO)Clz) (1723 cm- I ) and for (p,-dam)zPt2(p,-CO)Clz (1638 cm- I ) suggests
that the structural variations may have an electronic origin. The difference
in carbonyl stretching frequencies of these two molecules originally misled
some workers into suggesting that (p,-dam)zPt 2(p,-CO)Clz had a four-
electron donating carbonyl group similar to that of (p,-dpmhMn2(p,-CO)-
(COk(84)

The effects of the terminal ligands X on these insertion reactions have


received little attention. In the only systematic study available, it has been
shown that the equilibrium for carbon monoxide insertion is dependent on

Figure 5.17. A comparison of structural features of 1I.LdamhPd 2 (/L-Co )C1 2 and (/L-damhPt 2 (/L-
CO)CI 2 •
192 ALAN L. BALCH

X, with the stability of the carbon monoxide adduct increasing with X in


the order 1< Br < N3 - Cl.(42) In the A-frame products themselves, the
anionic X ligands can frequently be replaced by neutral ligands such as
isocyanides, monodentate phosphines, or pyridine to give cationic deriva-
tives.
While the observations on insertion in metal-metal bonds are most
extensive for Pd(I) and Pt(I) dimers, other rather isolated examples of
similar behavior are known. Although structural data are lacking
(dpm)zRh 2(CO)z(vco, 1915 cm- 1), which is isoelectronic with (f.,t-
dpm)zPd 2Cl2, adds one equivalent of carbon monoxide to form
(dpm)zRh 2(f.,t-CO)(CO)z(vCQ, 1940, 1920, 1835 cm- 1).(51) The product is
valence isoelectronic and, presumably, structurally similar to (f.,t-
dam)zPd 2(f.,t-CO)CL2. Structural data are available for another example
of insertion of carbon monoxide into a Rh-Rh bond. Carbon monoxide
reacts with (f.,t-dpm)zRH 2(C 2R 2)C!z(Rh-Rh, 2.744(9) A)(52) according to
reaction 3:

(3)

Although an additional bridging group has been added, the Rh-Rh separa-
tion has increased to -3.35 A in the product and the Rh-C-Rh angle is
-116. (59) This reaction type is not limited to phosphine-bridged complexes.
Consider reaction 4:
Ph 3 P" /0" /PRh 3 +
+ PhN/ --+ Ir Ir
ON/ "N/ "NO (4)
N
""Ph
Here, addition of diazonium cation results in an increase in the Ir-Ir
separation from 2.555 A (a reasonable value for a single bond) to
3.063(6) A. (118.119)
The reactivity of the unique bridging ligand in these A-frames has
received some attention, and further exploration in this area may be
anticipated. The reactivity of the unique carbonyl groups with wide M-C-M
angles and large M .. ·M separations and a comparison with the more
standard bridging carbonyl ligands and organic ketones bear exploration.
Such studies with (f.,t-dpm)zPd2(wCO)X2 have been thwarted because of
the ease with which carbon monoxide dissociates. The oxidation of bridging
BINUCLEAR, PHOSPHINE-BRIDGED COMPLEXES 193

sulfide to bridging sulfur dioxide, (40) the alkylation and protonation of


sulfide, (57) and the protonation of a bridging methylene ligand in (f,L-
dpm)zPd 2(f,L-CH 2)!z to give a terminal methyl group in {(f,L-dpm)zPd 2(f,L-I)-
(CH 3 )It (62) are unusual transformations of these bridges which have been
documented. Additionally, the catalytic activity of (f,L-dpm)zPd 2Ch in the
cyclotrimerization of dimethylacetylene dicarboxylate (vide infra) involves
.. 0 f t he umque
reactIvity . bn'd' . (6499)
gmg site. '

4.2. Introduction of a Bridging Ligand with Contraction of the


Metal-Metal Separation

Some 16-electron, four-coordinate dimers add a ligand to a bridging


site according to equation 5:
p~p p~p

M/
ILI M/
L I} I}
+A---.L -M-M-L (5)
L/ I L/ I I~/I
p~p p~p

(In structural changes this reaction is the opposite of the previous class.)
The reactant can either be a face-to-face dimer, as shown, or an A-frame.
The metal-metal separation is decreased in forming the product to less
than 2.9 A. In electron-counting terms it is convenient to write a metal-
metal bond in the product. This gives the product an 1S-electron count at
each metal and readily accounts for the contraction of the metal-metal
distance.
Some examples of the reaction involving the addition of carbon
monoxide and sulfur dioxide to rhodium(I) and iridium(I) A-frames are
shown in equation 6:(41,60,120,121)

p~p n
p~p
I....x.", I I ...x." I
OC-M--M-CO OC-M-'-'M-CO (6)
/\11/ I'c/ I
pOp
'-.../
p~p

M = Rh, X = Cl, n = + 1 M = Rh, X = Cl, n = + I M = Rh, X = Cl, n = + 1


M = Rh, X = S, n = 0 M = Rh, X = S, n = 0 M = Ir, X = S, n = 0
M = Ir, X = S, n = 0 M = Ir, X = S, n = 0

The structural data for (f,L-dpm)zRH 2(f,L-Cl)(CO)z +(Rh .. ,Rh,


31520(S) A)(58) and for (f,L-dpm)zRH 2(f,L-CI)(f,L-CO)(COh + (Rh-Rh,
2.S3S(1) A)(41) clearly establish the structural changes that are probably
194 ALAN L. BALCH

typical of the other examples in Equation 6. Another set of related reactions


is summarized in Equation 7: (97)

~ 2+ ~ 2+
p P p P

L"L ~h
[ ..... 1 L ..... I
L
+ CO ~
L"I

L
1 "L
7,h-Rh'
l"c"""'-I
'L
p p pOp
'-..../ '-..../

n so, (7)
p~p 2+

L, 1 I, L
\11\
"Rh--Rh

L ..... \ 'L

pOp
'-..../

These involve addition to face-to-face Rh(I) dimers. Detailed structural


characterization of the product is not as yet available, but it should prove
interesting. The electronic absorption spectra of the face-to-face reactants
show that there is significant steric interaction between the t-butyl groups
in the reactant. (122,123) Consequently, ligand addition to an already crowded
molecule is somewhat surprising. It appears likely that the isocyanide
substituents must bend away from one another in the product.
Reactant ligands for this series of reactions appear to be limited so
far to carbon monoxide and sulfur dioxide, It would be interesting to learn
what other potential ligands (isocyanides, acetylenes, etc.) can also be added
to the A-frame and face-to-face reactants in this fashion,
The addition of carbon monoxide to (JL-dpmhRh 2(JL-Cl)(COh +
(Equation 6) is one of the few reactions of these binuclear complexes where
any mechanistic observations have been made. (120) By following the
introduction of l3 CO into this complex by infrared and 13 C_NMR
spectroscopy, Mague and Sanger(120) have shown that the entering labeled
carbon monoxide initially resides in a terminal position. They propose the
mechanism shown in Equation 8 and specifically suggest initial attack of
the incoming ligand on a site which is exo to the final bridging site, In
support of this mechanism, it is reported that t-butyl isocyanide reacts with
(JL-dpmhRh 2(JL-CI)(COh + to form (JL-dpmhRh2(JL-Cl)(wCO)(CO)-
(t-BuNC). On the other hand, addition of sulfur dioxide in reaction 6,
as well as the additions of carbon monoxide and sulfur dioxide to
BINUCLEAR, PHOSPHINE-BRIDGED COMPLEXES 195

(lL-dpm)zRh 2 (t-BuNC)/+, appears to place the added molecule in the


bridging site with no evidence as yet presented for exo attack.

,
p~p + p~p +

,~l,,' , p, , ',3,CO (exo)


Rh 'Rh Rl( "Rh'
oC''', ''co oC''', , 'co
p~p p~p

H (8)

p~p +
, "Cl, ,
RIi--'Rh
oC'''' "c/' ,l3co
o
p~~p

j[
etc,

4.3. Loss of Carbon Monoxide Accompanied bV Carbon Monoxide


Bridge Formation

Extrusion of carbon monoxide, as shown in reaction 9, has been seen


from a few diphosphine bridged metal complexes. This sort of reaction is
also found when simple metal carbonyls condense to form polynuclear
clusters (e.g., the conversion of Fe(CO)s to Fe2(CO)9). In electron-counting
terms, all of the examples available involve loss of carbon monoxide from
two 18-electron centers. If a metal-metal bond is accepted as part of the
p~p p~p

, , , ,
(CO)nM M(CO)n --+ co + (CO)n-1M-M(CO)n-l (9)
I I 1'8/ ,
p~p p~p

/N'-..
P P
oc., , "co
'Fe--Fe' (10)
I 'c/ I
OC ..... 'co
pOp
'---N/
196 ALAN L. BALCH

product, then an 18-electron count is readily established there as well.


Examples of this behavior are shown in reactions 10, (85) 11, (124) and 12: (121)

--+ CO + (11)

P~P

--+ co
I I
+ (NC)(OC )Rh--Rh(CO)(CN) (12)
I"c/I
pOp
'--../

The first example shows the drastic shortening of the iron-iron separation
that accompanies carbon monoxide loss. The Fe .. ·Fe separation decreases
from 3.90 A in the reactant to 2.661(1) in the product. (85) The product of
reaction 11 has a Ni-Ni separation of only 2.508 A. The geometric details
for the other compounds in these reactions are not known. The loss of
carbon monoxide is reversible for the nickel complex.
In the reactions in this and the two proceeding sections, a number of
compounds with bridging carbonyl ligands are formed. In these, the metal-
metal separation can vary significantly. As can be seen from Table 5, there
is no simple correlation between the carbonyl stretching frequency of these
molecules and the metal-metal distance. Thus, while infrared spectroscopy

Table 5. Properties of Complexes with Bridging Carbonyl Ligands

M .. ·M
Compound distance (A) v(CO) cm- I Reference

(/J.-dpm)zPd 2(/J.-CO)CI 2 1705 42


(/J.-dam)zPd 2(/J.-CO)CI 2 3.274(8) 1723 42,83
(wdam)zPt2(/J.-CO)CI2 3.162(4) 1638 84
[(/J.-dpm)zRh 2(/J.-CO)(/J.-CI)(CO)2t 2.838(1) 1865 41,56
(/J.-dpm)zRh 2(/J.-CO)Br2 2.7566(8) 1745 53
(/J.-dpm)zRh2 (/J.-CO )(/J.-C 2{C0 2Me h)C1 2 3.3542(9) -1700 59
(/J.-dpmh1r2(/J.-CO)(/J.-S)(CO)2 2.843(2) 1760 60
(/J.- Ph 2Ppy hRh 2(/J.-CO)CIz 2.612(1) 1797 92
BINUCLEAR, PHOSPHINE -BRIDGED COMPLEXES 197

is useful in detecting the bridging carbonyl group, it does not give direct
evidence about the other significant geometric aspects of the M 2 CO unit.

4.4. Two-Center, Two-Fragment Oxidative Addition


A number of low-valent, diphosphine bridged dimers undergo oxida-
tive addition of conventional substrates, such as the halogens according to
reaction 13.

(13)

In this case the stoichiometry of oxidant uptake is one-half mole of


X 2 per metal ion rather than the one mole per metal ion that is characteristic
of the monomeric analogs with the same metal ions. As a result, the metal
ion is oxidized (formally) by one electron rather than by two electrons,
and a metal-metal bond is formed.
This reaction has been characterized for a wide variety of Rh(I)
complexes. The behavior of (lL-dpmhRh2(CNR)4 z+ shown in reaction 14
is typical. (122) Oxidants capable of performing this reaction include chlorine,
bromine, iodine, and triftuoromethyl disulfide. However, dihydrogen does
not add to these Rh(I) compounds. The dimeric Rh(II) products are quite
stable. Treating these with excess oxidant does not result in the formation
of the more usual Rh(III) complexes. However, some oxidants, including
diphenyl disulfide, pentaftuorophenyl disulfide, and diphynel diselenide,
react with (lL-dpmhRh2(CNR)/+ to form Rh(I1I) products,
(dpm)Rh(CNRh(X)z 2+, exclusively. (123) The difference in behavior of these
oxidants has not been explained.

p~p 2+ p~p 2+

1 /L 1 /L 1 /L 1 /L
Rh Rh + X 2 --+ X-Rh--Rh-X (14)
L/I L/I L/I L/I
p p p p
~ ~

These oxidation reactions are readily monitored by following


isocyanide stretching vibrations by infrared spectroscopy. The energy of
this vibration increases by about 40 cm -1 for oxidation of Rh(I) to Rh(II).
For comparison, the oxidation of mononuclear Rh(I) complexes to Rh(III)
species results in an 80 cm- 1 increase in the energy of the isocyanide
stretching vibration.
198 ALAN L. BALCH

Similar oxidative addition reactions of halogens with the neutral


complexes (JL-dpmhRh 2(COhCh have been reported to form (JL-
dpmhRh 2(COhChX2 and (JL-damhRh 2(COhChX2. (122)
Other ligands are also capable of sustaining similar oxidative behavior.
Among these the bridged isocyanide complexes of Gray and co-workers
offer the widest range of examples. The bridging isocyanide is most
frequently 1,3-disocyanopropane (bridge) and the reactions of
(bridge )4Rh/+ are summarized by reaction 15: (125)

2+

(15)

Here XY can be not only the halogens, "Chlorine, bromine and iodine, but
also methyl iodide and Mn2(COho. (126) The latter forms a linear Mn-Rh-
Rh-Mn chain. The Rh .. · Rh separation in (Bridge )4Rh/+ is 3.262 A, (127)
a value which is simlar to those found for the phosphine- and arsine-bridged
analogs (JL-dpm)zRh2(COhCh and (JL-damhRh 2(COhCh. These are non-
bonded distances. On oxidation the Rh-Rh distance contracts to give a
normal Rh-Rh single bond. In (bridge)4Rh 2Ch 2+ the Rh-Rh distance is
2.837(1) A. (128) It should be noted that bridging ligands are not absolutely
required to form Rh(II) dimers. With simple isocyanides The Rh(II) dimers
Rh 2(CNR )sX2 2+ are readily formed. (129,130) However, in solution these are
subject to fairly facile disproportionation (reaction 16), a reaction which
is inhibited when bridging ligands are present.

2+ + +
L L L L L L L L
1/ 1/ 1/ 1/
X-Rh--Rh-X ~ X-Rh-X + Rh (16)
L/I L/I L/I L/I
L L L L

L = RNC

Oxidative addition to molecular A-frames appears to be little studied.


Eisenberg and co-workers(57) have shown that alkyl halides and protons
react at the bridging sulfur in (JL-dpm)zM2(wS)(COh(M = Rh, Ir) to
give (JL-dpmhM2(JL-SR)(COh rather than undergoing oxidative addition
T

to the metals. However, (JL-dpmhIr2(JL-S)(COh does add dihydrogen to


yield two hydrido complexes which may have any of the structures 4-6: (60)
BINUCLEAR, PHOSPHINE -BRIDGED COMPLEXES 199

P~P P~P P~P


I
Ir~Ir
.,S. I I .. I
S ...
Ir'-H-'Ir
HJ /S .. ..I /H
Ir--Ir
OC/ I 'H H/I OC CO/I '-W'/I 'co OC/I I 'CO

P~P P~P P~P

4 5 6
These species are probably involved in the homogeneous hydrogenation
activity of (~-dpmhIr2(~-S)(COh.
However, the behavior of Ir2(~-SRh(COh(PR~h shows that molecular
A-frames and similarly shaped molecules can undergo two-center, two-
fragment oxidative addition. This complex adds diiodine and dihydrogen
· to reactIOn
accor d mg . 17 : (131 ' 132)
Bu' Bu'
I I
R,P S
.,,/,,/
CO x , , /S, , /X
Ir Ir + Xl ----. P P·--Ir Ir--'CO (17)
J /"/"
OC / " S/ "PR 3 OC ~ PR 3
I
Bu' Bu'

In the process the Ir-Ir separation shrinks from 3.216 in Ir2(~­


SButh(COh{P(OMehh to 2.703 in Ir2(~-SButh(COh{PPhMe2hI2'
All the preceding reactions have involved oxidative additions to dB
metal complexes. Similar reactions are also known for d 10 metal complexes.
The Pd(O) complex, (dpmhPd 2, whose detailed structure is unknown, also
undergoes oxidative addition according to reaction 18(95):
(dpm h Pd 2 + X2-~ (wdpm)lPd2X2 ~

The initial step is a two-center, two-fragment addition which offers a new


route to Pd(I) complexes. These, however, unlike the Rh(II) dimers, are
subject to further oxidation, which eventually produces the Pd(II) chelates,
(dpm)PdX2, by way of intermediates (~-dpmhPd2X4' which have not been
isolated. In related chemistry with different bridging ligands, the d,lO Au(l)
complexes are reported to undergo two-center, two-fragment oxidative
additions as shown in reaction 19:
X
I
/CH 2 - AI u-CH 2 "
R 2 P" /PR 2 (19)
CH 2 -Au-CH l
I
y

where XY can be Clz, Br2, Iz or methyl iodide. (133,134)


200 ALAN L. BALCH

4.5. Two-Center, Three-Fragment Oxidative Addition


In this reaction a dihalo compound undergoes oxidative addition to
two metal centers as represented in reaction 20:

p~p p~p

In+ In+ K" Mn-+2 M'/X


+'1 (20)
,'A",
H

, ,
p~p p~p

Each of the metal centers undergoes a net two-electron oxidation as in a


conventional oxidative addition. However, the two substrate bonds are
boken to produce three litigating fragments: two terminal halides, and a
bridging ligand. The known reactions of this class all involve a single
complex, (dpmhPd 2 , and one diphosphine ligand is lost from it in the
process. (98) The reaction chemistry is summarized in Figure 5.18. The
reaction with methylene dihalides offers a new route to methylene bridged
complexes. Methylene bridged species have yet to be formed starting with
mononuclear metal complexes. In that case, halomethyl complexes are
formed. (135-137) The same holds for the addition of a-diiodobenzene. With
Pd(PPh3h this yields (Ph 3P)zPd(0-C5H 4I)I as the sole product, while the
(dpmhPd 2 , the phenylene bridged complex is formed in high yield. The
addition of oxalyl chloride and phenyl isocyanide dichloride to (dpmhPd 2
provides alternate routes to molecular A-frames that were originally
synthesized by the addition of carbon monoxide and phenyl isocyanide,
respectively, to (Jt-dpm)zPd 2 Clz. In regard to the behavior of the
isocyanide dichloride, it should be noted that three-fragment, one-
center oxidative addition has been previously reported. For example,
addition of phenyl isocyanide dichloride to (Ph3P)4Pt and to
(Ph 3PhRhCl yields (Ph 3P)Pt(CNPh)Clz and (Ph 3P)zRh(CNPh)Ch,
. I y. (138)
respective
The methylene bridged complex, (Jt-dpm)zPd 2 (Jt-CH 2 )X2 , and its
platinum analog are isoelectronic and probably isostructural with (Jt-
dpm)zPd 2 (Jt-S)Clz. Consequently, they are expected to have a large metal-
metal separation and to differ appreciably from the larger class of
dimetallacyclopropanes. (139) Some reaction chemistry of (Jt-dpm)zPd 2 (Jt-
CH 2 )X2 has been explored. While the methylene bridge is resistant to
insertion of carbon monoxide, sulfur dioxide, and isocyanides, it is readily
protonated. (98) The product is the A-frame (Jt-dpm)zPd 2 (Jt-I)-
(CH 3)I+, which has been isolated as the ftuoroborate salt. It contains a
terminal methyl group that resists further protonation.
BINUCLEAR, PHOSPHINE -BRIDGED COMPLEXES 201

r PhNCel,

l I
Q I

Figure 5.18. Two center, three-fragment oxidative addition reactions of (dpmhPd 2 .

4.6. The Formation or Stabilization of Other Novel


Bridging Ligands

This classification involves a catchall for the formation of unusual


bridging ligands that have not been accounted for in previous sections.
Clearly, some of the preceding sections involve reactions that create unusual
bridging ligands. Here we are concerned with some other rather singular
reactions that do not fall in the previous categories.
202 ALAN L. BALCH

Reaction of dpm with dimanganese decacarboryl produces the substitu-


tion product (J,L-dpmhMn2(COk On further heating to 1400 (J,L-
dpmhMn2(CO)6 undergoes carbon monoxide loss. Some pertinent reaction
chemistry is summarized in Equation 21: (43,44,140,141)
p~p p~p
/CO I
/CO I -co /CO /CO I I
l'c~/ I
OC-Mn-Mn-CO ~ CO+OC-Mn-Mn-CO
OC/ I OC/ I KO
p~p P~P

R =:
-RNC
C'H,
PhCH,
or II RNC (21)

This sequence leads to the formation of the unusual bridging isocyanide


and carbon monoxide ligands, which act as normal two-electron donors to
one Mn and donate a pair of 11' electrons to the second Mn. A drawing of
the in-place portion of the structure of (J,L-dpmhMn2(J,L-p-CH3C6H4NC)-
(CO)4 is shown in Figure 5.19.(43) This gives details of the novel mode of
isocyanide coordination.
The fluxional behavior of both (J,L-dpmhMn2(CO)6 and of (J,L-
dopmhMn2(J,L-CO)(CO)4 has been examined by both 31 p and l3C NMR
spectroscopy. (142,143) For (J,L-dpmhMn2(CO)6, global scrambling of carbon
monoxide ligands over all sites occurs, presumably by pairwise terminal-
bridge interchange. In contrast, limited scrambling occurs when the novel
bridging carbonyl ligand is present. The proposed mechanism for this
intramolecular reaction is shown in reaction 22;

l
3 4 3 3
o 0 /0
C
I
Mn--Mn-CO 5
C
I
2
C
oc-~n-M~
0
;::0
4l"-== /C \
2 OC - Mn - - Mn - CO 4
/....... /f
I CO I I
OC C~ C C C
2 0 1 o o 0
1 5 5

(22)
BINUCLEAR, PHOSPHINE -BRIDGED COMPLEXES 203

Figure 5.19. A view of a planar portion of (lL-dpm)zMn2(IL-p-CH3C6H4NC)(CO)4, which


contains the bridging isocyanide ligand. The bridging dpm ligands lie above and below this
plane and are not shown.

The exact role of the diphosphine ligand in stabilizing the novel 4-electron
donor carbonyl and isocyanide ligands is not clear. Other examples of
similar carbonyl groups include Cp2NbMo(COhCp, Cp2Zr(OC)-
(OCMe)Mo(COhCp, and (C5Me5hZrCo(COhCp. None of these
requires a bridging phosphine ligand to stabilize the bridging carbonyl
group.
In a different vein, the reaction of carbon disulfide with (I-L-
dpmhRH 2(I-L-CO)CL 2 produces 7 as the final product. (55) In 7 two
molecules of carbon disulfide have condensed to form a tridentate
ligand which chelates one metal and also forms a bridge to the second
metal.
204 ALAN L. BALCH

p~p

/7 r-
CI" II/CO
c- t CI

C-s p P
S/ ~
7

4.7. Capture of a Second Metal Induced by


2(Diphenylphosphino)pyridine

The bifunctional ligand 2(diphenylphosphino)pyridine can be used to


construct binuclear metal complexes in a stepwise fashion. (92) With Group
VIII metal ions, this ligand appears to preferentially bind via the phosphorus
atom. From this reaction, a host of metal complexes, very similar to those
of triphenylphosphine, are available. All have the potential for binding a
second metal ion through the uncoordinated pyridine nitrogen. Some of
the reactions that result in the formation of binuclear complexes are shown
in Figure 5.20. (92.144-146) The structure of one example, (f.L-
Ph 2Ppy hRhPd( CO )Ch, is shown in Figure 5.21. In all cases so far examined,
the formation of the binuclear complex results in the formation of a
metal-metal bond as well. For the crystallographically characterized
molecules these are all of similar length: (f.L-Ph2PpyhRh2(f.L-CO)Ch,
2.612(1) A, (145) (f.LPh2PpyhRhPd(CO)CI3 , 2.594(1) A, (92) and (f.L-
Ph 2PpyhRuPd(COhCh, 2.660(1) A. (146) This probably reflects the limited
flexibility of Ph2Ppy.
In all but the first example the reactions involve the oxidative addition
of a metal-halogen bond to another metal center. Also, in all cases shown
the bridging ligands have rearranged themselves so that a head-to-tail
orientation of the Ph 2Ppy ligands is formed. In other words, one metal-
phosphorus bond has broken and that phosphorus has become bound to
the second metal atom. This behavior emphasizes the ability of phosphine
ligands to dissociate from metal complexes, a fact that is frequently over-
looked in viewing the reactions of diphosphine bridged complexes. For
example, with the dpm bridged complexes, the diphosphine-dimetal unit
generally appears as a sturdy framework about which reactions occur at
all of the other coordination sites.
None of the metal-metal bonds shown on the binuclear complexes in
Figure 5.20 undergoes facile insertion of small molecules. For example
(f.L-Ph2PpyhPd2Ch, which is isoelectronic with (f.L-dpmhPd2Ch, does not
react with carbon monoxide or sulfur dioxide. This lack of reactivity has
been attributed to the limitations that the bridging Ph 2Ppy places on the
BINUCLEAR, PHOSPHINE-BRIDGED COMPLEXES 205

o P~
___ C
I ___ I I
[Rh(CO)2CIJ2 + N P-Rh-P N CI--Rh /Rh--{ I
I
CI I '----c 1
N 0 P
~

o ~P
N""\ C I I CO
I -----.
(PhCN)2PdCI2 + P-Rh-P N CI---Pd Rh~1
I
CI I c(1
P N
~

~
r-N N P
P
I
CI-Pd-P N
.--... I
CI---Pd
I
Pd--{I
I
CI I
N
I
P
~

rN ~N
P
I ,...--.,. I I
CI-Pd-P N CI---Pt Pd--{I
I
CI I I
~P

~N ~
P
I ___ II/co
[Rh(CO)2CIl2 + CI-Pt-P N Cl---Pt /Rh--{I
I
CI I CI 1
N~P

Figure 5.20. Reactions leading to the formation of binuclear complexes through the use of
2-(diphenylphosphino)pyridine as a bridging ligand.
206 ALAN L. BALCH

C1S
Figure 5.21. The structure of (IL-Ph 2 PpyhRhPd(CO)CI 3 taken from Reference 92 by per-
mission.

metal-metal separation and its inability to span metal-metal separations


of 3 A or greater.

5. DEMONSTRATED CATAL YTIC ACTIVITY OF


BINCULEAR COMPOUNDS

A number of binuclear phosphine-ligand bridged complexes have been


shown to function as catalysts or catalyst precursors. Here we review these
cases briefly. M'Jre detailed coverage of hydrogenation and hydroformyla-
tion using rhodium catalysts will appear in the next chapter. In all cases
of catalytic activity shown by binuclear complexes, there is a serious
question about the true identity of the catalytically active species. As with
catalysis begun by metal carbonyl clusters, the possibility exists that a small
amount of highly active mononuclear compound is the true catalyst.
BINUCLEAR, PHOSPHINE-BRIDGED COMPLEXES 207

(IL-DpmhPd 2 Ch is a catalyst for the conversion of dimethylacetylene


dicarboxylate to hexamethyl mellitate (reaction 23) (99):

3RC=,=, ~R -4
R

R
¢
~
R
7'
I
R

R
0
II
R = CH 3 0C (23 )

During the process, (lL-dpmhPd 2(IL-C 2{C0 2 CH 3 }z)Ch forms. This complex,
which has been isolated in crystalline form, can also serve as a catalyst.
Other related complexes, dpmPdCh and (dpmhPd 2, are much less effective.
The structure of (lL-dpmhPd 2(IL-C 2{C0 2CH3 }z)Ch shows that, on insertion
into the Pd-Pd bond, the acetylene has been reduced to a dimetalated
olefin. The central C-C bond distance has elongated to 1.338 A, and all
the bond angles about the central two carbon atoms are near 120°. The
catalytic cycle may involve further insertions of the acetylene into the Pd-C
bonds of this A-frame, but direct eivdence of the mechanism is lacking.
However, an analysis of the molecular orbitals involved indicates that such
a process is feasibleY 16)
(DpmhPd 2 is a catalyst for the hydrogenation and isomerization of
ole fins and diolefins (both conjugated and nonconjugated). (147,148) It
catalyzes the dihydrogen reduction of methyl acetylene to propylene (which
is then more slowly reduced to propane) and of propionaldehyde to ethane
and ethylene. As a hydrogenation catalyst, it is effected by dioxygen
pretreatment and by triphenylphosphine and carbon monoxide, which
inhibit activity. After exposure to dioxygen it catalyzes exchange between
H2 and D 2 • There is no information available that bears on the question
of what the form of the catalytically active species actually is whether it
contains two palladium atoms. So far hydride complexes of palladium with
bridging dpm ligands have not been characterized, although a number of
dinuclear hydrides of platinum have been isolated and characterized. (67,68)
A brief report of the ability of (dpmhPt 2 to catalyze reactions 24 and
25 has appeared, (149) (DpmhPt 2 is also reported to be less effective

(24)

(25)

in the hydrogenation of butadiene than is (dpmhPd 2 • (147,148)


(IL-DpmhRh 2(COh is a catalyst for the reduction of acetylene to
ethane by dihydrogen. The rate of acetylene reduction exceeds that of
ethylene reduction and no cyclotrimerization of acetylene is observed. (51)
Treatment of (lL-dpmhRh 2(CO)z with carbon monoxide and NaBH(OMeh
208 ALAN L. BALCH

yields (~-dpmhRh2(~-CO)(~-H)(COh +, which is an unusually active


catalyst for the water gas shift reaction at 90° and 1 atm pressure. (51)
(~-Dpmhlr2(~-S)(COh is a catalyst for the hydrogenation of acety-
lene, ethylene, and propylene and can be recovered quantitatively from
the reaction. (60)

6. CONCLUDING REMARKS

The previous sections have documented the fact that binuclear com-
plexes can be obtained with a variety of novel structural features and that
these undergo, in some cases, unprecedented chemical reactions. We can
expect that work along these lines will continue to point out new structural
types and new classes of reactions. Progress so far can be said to have laid
a foundation for the basic coordination chemistry of the most readily
obtained binuclear systems. It remains to be seen whether these reaction
types can be clearly utilized in the design of new catalysts or whether
unusual forms of catalytic activity for binuclear complexes will be discovered
serendipitously.
Several areas for future development can be easily identified. Strategies
for the incorporation of two different metal ions into the same complex
have only recently begun to be considered. (26,92,93,144-146) The range of
possibilities for hetero-binuclear complexes is enormous, and we can expect
considerable development of new ligands and new combinations of metals.
With two different metals, each can perform an individual, separate func-
tion, or both could bind to a bridging ligand and induce unusual polarization.
The opportunity afforded by the ability to design phosphine ligands
to a variety of geometric specifications allows us to consider templated-
based construction of metal clusters with several metal ions, possibly difficult
metal ions. This has considerable potential for extending the now popular
field of cluster chemistry. With polyphosphine backbones, the cluster need
no longer rely solely on metal-metal bonding to determine its shape and
stability. Some inroads into this area have begun to appear. The triphos-
phine, tripod HC(PPh 2h, has been used not only to cap triangular faces
of tetrahedral metal clusters(159) but, more importantly, to construct the
new triangular array 8. (151)
Bifunctional ligands such as Ph 2Ppy, which have the ability to capture
a second metal ion, also have the potential to facilitate the transfer of
ligands from one metal to another. For example, in Figure 5.20 most of
the metal-metal bond formation is accompanied by the transfer of a chloride
ligand from one metal to another. It remains to be seen whether the transfer
BINUCLEAR, PHOSPHINE-BRIDGED COMPLEXES 209

Ph ~ Ph

O,Ph-~J
C,! "-~-Ph
~~!
Ni Ni-C-O

\1 ic,
!Ph-P-Ph!
/C
o ""'Ni 0
I
C
I
o
8
of other ligands, particularly reactive organometallic ligands (alkyl, aryl,
vinyl groups, hydrides), can be facilitated by bifunctional ligands.
The design of new ligands to use in constructing binuclear and polynu-
clear clusters does not have to be restricted to phosphine ligands.
2(Diphenylphosphine) pyridine is an example of a hybrid ligand with two
different metal binding sites. Further elaboration of this concept should
allow for the development of a variety of ligands capable of binding hard
metal ions at one site and soft metal ions at another. The diisocyanide,
1,3-diisocyanopropane, resembles the diphosphine described, here, and we
can expect to see the development of further aspects of its chemistry. Some
work on the properties of complexes of sulfide ligands, such as PhSCH 2SPh,
has been reported, (152) but this type of ligand appears to be less promising
than others.
Finally, the ability of diphosphine ligands to form stable binuclear
complexes as well as chelating complexes raises new interpretations of
observations made using these ligands. For example, the activity of
platinum(I) chloride/tin(II) chloride mixtures as hydroformylation
catalysts depends on the chain length (n) of the diphosphine
Ph2P(CH2)nPPh2 employed. (153) While this was initially interpreted as an
effect of chelate ring size, which may well be the case, it is also just as
possible that a binuclear complex is formed and that at n = 4 (the phosphine
that produces the fastest rate) the optimal spacing in a binuclear catalyst
is obtained.

A CKNOWLEOGMENTS

I thank Rich Eisenberg, Alan Sanger, and George Staneiy for sharing
unpublished data, and the NSF for the support of those aspects of original
research done at the University of California, Davis.
210 ALAN L. BALCH

REFERENCES

1. E. L. Muetterties, Bull. Soc. Chern. Belg. 85,451 (1976).


2. J. P. Collman, P. Denisevich, Y. Konai, M. Marrocco, C. Koval, and F. C. Anson, J.
Arn. Chern. Soc. 102, 6027 (1980).
3. M. H. Chisholm, ed., "Reactivity of Metal-Metal bonds," ACS Syrnposiurn Series 155,
Washington, D.C., 1981.
4. D.-H. Chin, G. N. La Mar, and A. L. Balch, J. Arn. Chern. Soc. 102, 5945 (1980).
5. - - , J. Arn. Chern. Soc. 102,4344 (1980).
6. A. F. Dyke, S. R. Finnimore, S. A. R. Knox, P. J. Naish, A. G. Orpen, G. H. Riding,
and G. E. Taylor in "Reactivity of Metal-Metal Bonds," M. H. Chisholm, ACS
Syrnposiurn Series 155, 1981, p.259.
7. E. L. Muetterties, Bul. Soc. Chern. Belg. 84, 959 (1975).
8. - - , Science 196,839 (1977).
9. R. Whyman, Transition Metal Clusters, edited by B. F. G. Johnson (John Wiley and
Sons, New York, 1980), p. 545.
10. D. H. Farrar, P. G. Jackson, B. F. G. Johnson, J. Lewis, W. J. H. Nelson, M. D. Vargas,
and M. McPartlin, J. Chern. Soc. Chern. Cornrn., 1009 (1981).
11. J. Chatt and F. A. Hart, J. Chern. Soc. 1378 (1960).
12. W. Hewertson and H. R. Watson, J. Chern. Soc. 1490 (1962).
13. A. R. Sanger, J. Chern. Soc. Chern. Cornrn. 893 (1975).
14. - - , J. Chern. Soc., Dalton Trans. 1971 (1977).
15. T. Yoshida, T. Yamagata, T. H. Tulip, J. A. Ibers, and S. Otsuka, J. Arn. Chern. Soc.
100, 2063 (1978).
16. A. Dedieu and R. Hoffmann, J. Arn. Chern. Soc. 100,2074 (1978).
17. C. Crocker, R. J. Errington, R. Markham, G. J. Moulton, K. J. Odell, and B. L. Shaw,
J. Arn. Chern. Soc. 102,4373 (1980).
18. R. B. King, Acc. Chern. Res. 13,243 (1980).
19. A. L. du Preez, I. L. Marais, R. J. Haines, A. Pidcock, and M. Safari, J. Organornet.
Chern. 141, CI0 (1977).
20. R. J. Haines, A. Pidcock, and M. Safari, J. Chern. Soc., Dalton Trans., 830 (1977).
21. F. A. Cotton, R. J. Haines, B. E. Hanson, and J. C. Sekutowski, Inorg. Chern. 17, 2010
(1979).
22. A. L. du Preez, I. L. Marais, R. J. Haines, A. Pidcock, and M. Safari, J. Chern. Soc.,
Dalton Trans., 1918 (1981).
23. R. J. Haines, E. Maintjies, and M. Laing, Inorg. Chirn. Acta 36, L403 (1979).
24. R. J. Haines, M. Laing, E. Meintjies, and D. Sommerville, J. Organornetal. Chern. 215,
C17 (1981).
25. E. Keller and H. Vahrenkamp, Chern. Ber. 112, 1626 (1979).
26. R. Miiller and H. Vahrenkamp, Chern. Ber., 113, 3517 (1980).
27. F. A. Cotton, W. H. Ilsley, and W. Kaim, J. Arn. Chern. Soc. 102,1918 (1980).
28. M. G. Newton, R. B. King, M. Chang, N. S. Pantaleo, and J. Gimeno, J. Chern. Soc.,
Chern. Cornrn., 531 (1977).
29. M. G. Newton, N. S. Pantaleo, R. B. King, and T. J. Lotz, J. Chern. Soc., Chern. Cornrn.
514 (1978).
30. R. J. Puddenphatt, M. A. Thomson, Lj. Manojlovic-Muir, K. W. Muir, A. A. Frew,
and M. P. Brown, J. Chern. Soc., Chern. Cornrn. 805 (1981).
31. M. P. Brown, S. J. Cooper, A. A. Frew, Lj. M~nojlovic-Muir, K. W. Muir, R. J.
Puddephatt, K. R. Seddon, and M. A. Thompson, Inorg. Chern. 20,1500 (1981).
32. N. W. Alcock, J. M. Brown, and J. C. Jeffery, J. Chern. Soc., Dalton Trans., 888 (1977).
BINUCLEAR, PHOSPHINE -BRIDGED COMPLEXES 211

33. F. C. March, R. Mason, K. M. Thomas, and B. L. Shaw, 1. Chern. Soc., Chern. Cornrn.
584 (1975).
34. N. A. AI-Salem, W. S. McDonald, R. Markham, M. C. Norton, and B. L. Shaw, 1.
Chern. Soc., Dalton Trans., 59 (1980).
35. M. Cowie and S. K. Dwight, Inorg. Chern. 19,2500 (1980).
36. 1. T. Mague, Inorg. Chern. 8, 1975 (1969).
37. M. Cowie and S. K. Dwight, Inorg. Chern. 20,1534 (1981).
38. E. H. Abbott, K. S. Bose, F. A. Cotton, W. T. Hall, and 1. C. Sekutowshi, Inorg. Chern.
17,3240 (1978).
39. M. M. Olmstead, L. S. Benner, H. Hope, and A. L. Balch, Inorg. Chirn. Acta 32, 193
(1979).
40. A. L. Balch, L. S. Benner, and M. M. Olmstead, Inorg. Chern. 18, 2996 (1979).
41. M. M. Olmstead, C. H. Lindsay, L. S. Benner, and A. L. Balch, 1. Organornetal. Chern.
179,289 (1979).
42. L. S. Benner and A. L. Balch, 1. Arn. Chern. Soc. 100,6099 (1978).
43. c. 1. Commons and B. F. Hoskins, Aust. 1. Chern. 28,1663 (1975).
44. L. S. Benner, M. M. Olmstead, and A. L. Balch, 1. Organornetal. Chern. 159,289 (1978).
45. F. A. Cotton, L. W. Shive, and B. R. Stults, Inorg. Chern. 15, 2239 (1976).
46. M. 1. Mays, D. W. Prest, and P. R. Raithby, 1. Chern. Soc., Chern. Cornrn., 171 (1980).
47. F. A. Cotton and 1. M. Troup, 1. Arn. Chern. Soc. 96, 4422 (1974).
48. G. Lavigne and 1.-1. Bennet, Inorg. Chern. 20,2713 (1981).
49. P. H. Bird, A. R. Fraser, and D. N. Hall, Inorg. Chern. 16,1923 (1977).
50. F. H. Carre, F. A. Cotton and B. A. Frenz, Inorg. Chern. 15, 380 (1976).
51. C. P. Kubiak and R. Eisenberg, 1. Arn. Chern. Soc. 102,3637 (1980).
52. M. Cowie and R. S. Dickson, Inorg. Chern. 20, 2682 (1981).
53. M. Cowie and S. K. Dwight, Inorg. Chern. 19,2508 (1980).
54. --,Inorg. Chern. 19, 209 (1980).
55. M. Cowie and S. K. Dwight, 1. Organornetal. Chern. 214,233 (1981).
56. M. Cowie, Inorg. Chern. 18, 286 (1979).
57. C. Kubiak and R. Eisenberg, Inorg. Chern. 19,2726 (1980).
58. M. Cowie and S. K. Dwight, Inorg. Chern. 18, 2700 (1979).
59. M. Cowie and T. G. Southern, 1. Organornetal. Chern. 193, C46 (1980).
60. C. P. Kubiak, C. Woodcock, and R. Eisenberg, Inorg. Chern. 19, 2733 (1980).
61. R. G. Holloway, B. R. Penfold, R. Colton, and M. 1. McCormick,l. Chern. Soc., Chern.
Cornrn., 485 (1976).
62. M. M. Olmstead, 1. P. Farr, and A. L. Balch, Inorg. Chirn. Acta 52,47 (1981).
63. M. M. Olmstead, H. Hope, L. S. Benner, and A. L. Balch, 1. Arn. Chern. Soc. 99, 5502
(1977).
64. A. L. Balch, c.-L. Lee, C. H. Lindsay, and M. M. Olmstead, 1. Organornetal. Chern.
177, C22 (1979).
65. Lj. Manojlovic-Muir, K. W. Muir, and T. Solomun, 1. Orgnornetal. Chern. 179, 479
(1979).
66. Lj. Manojlovic-Muir, K. M. Muir, and T. Solomun, Acta Cryst. B35, 1237 (1979).
67. Lj. Manojlovic-Muir and K. W. Muir, 1. Organornrnetal. Chern. 219, 129 (1981).
68. M. P. Brown, S. 1. Cooper, A. A. Frew, Lj. Manojlovic-Muir, K. W. Muir, R. 1.
Puddephatt, and M. A. Thomson, 1. Organornetal. Chern. 198, C33 (1980).
69. T. S. Cameron, P. A. Gardner, and K. R. Grundy, 1. Organornetal. Chern. 212, C19
(1981).
70. N. Bresciani, N. Marsich, G. Nardin, and L. Randaccio., Inorg. Chirn. Acta 10, L5 (1974).
71. A. Camus, N. Marsich, G. Nardin, and L. Randaccio, 1. Organoetal. Chern. 60, C39
(1973).
212 ALAN L. BALCH

72. A. Camus, G. Nardin, and L. Randaccio, Inorg. Chirn. Acta 12,23 (1975).
73. G. Nardin and L. Randaccio, Acta Cryst. 830, 1377 (1974).
74. G. Nardin, L. Randaccio and E. Zangrando, 1. Chern. Soc., Dalton Trans., 2566 (1975).
75. H. H. Karsch and V. Schubert, unpublished results quoted in reference 76.
76. U. Schubert, D. Neugebauer, and A. A. M. Aly, Z. anorg. aUg. Chern. 464, 217 (1980).
77. H. Schmidbaur, A. Wohlleben, U. Schubert, A. Frank, and G. Hattner, Chern. Ber.
110,2751 (1977).
78. H. Schmidbaur, A. Wohlleben, F. Wagner, O. Orama, and G. Huttner, Chern. Ber.
110, 1748 (1977).
79. E. O. Fischer, A. Ruhs, P. Friedrich, and G. Huttner, Angew. Chern. Int. Ed. Eng. 16,
465 (1977).
80. C. Commons and B. F. Haskins, Aust. 1. Chern. 28, 1201 (1975).
81. G. Lavigne, N. Lugan, and J.-J. Bonnet, Nov. 1. Chirn. 5, 423 (1981).
82. P. H. Bird, A. R. Fraser, and D. N. Hall, Inorg. Chern. 16, 1923 (1977).
83. R. Coiton, M. J. McCormick, and C. D. Pannan. Aust. 1. Chern. 31,1425 (1978).
84. M. P. Brown, A. N. Keith, Lj. Manojlovic-Muir, K. W. Muir, R. J. Puddephatt, and
K. R. Seddon, Inorg. Chirn. Acta 34, L223 (1979).
85. M. G. Newton, R. B. King, and M. Chang, and J. Gimeno, 1. Am. Chern. Soc. 99, 2802
(1977).
86. - - , 1 . Am. Chern. Soc. 100,326 (1978).
87. - - , 1 . Am. Chern. Soc. 100, 1632 (1978).
88. F. Faraone, G. Bruno, G. Tresoldi, G. Faraone, and G. Bombieri, 1. Chern. Soc., Dalton
Trans., 1651 (1981).
89. L. H. Pignolet, D. H. Doughty, S. C. Nowicki, M. P. Anderson, and A. L. Casalnuovo,
1. Organornetal. Chern. 202, 211 (1980).
90. P. G. Jones, Acta Cryst. 836, 2775 (1980).
91. B. P. Sullivan and T. J. Meyer, Inorg. Chern. 19,752 (1980).
92. J. P. Farr, M. M. Olmstead, and A. L. Balch, J. Am. Chern. Soc. 102, 6654 (1980).
93. N. E. Schore, J. Am. Chern. Soc. 101,7410 (1979).
94. C. H. Lindsay, L. S. Benner, and A. L. Balch, Inorg. Chern. 19, 3503 (1980).
95. C. T. Hunt and A. L. Balch, Inorg. Chern. 20,2267 (1981).
96. A. Sanger and J. T. Mague, unpublished results, A. Sanger, personal communication.
97. J. T. Mague and S. H. deVries, Inorg. Chern. 19, 3743 (1980).
98. A. L. Balch, C. T. Hunt, C.-L. Lee, M. M. Olmstead, and J. P. Farr, J. Am. Chern.
Soc. 103, 3764 (1981).
99. c.-L. Lee, C. T. Hunt, and A. L. Balch, Inorg. Chern. 20, 2498 (1981).
100. c. T. Hunt and A. L. Balch, Inorg. Chern. 21, 1641 (1982).
101. S. J. Cooper, M. P. Brown, and R. J. Puddephatt, Inorg. Chern. 20, 1374 (1981).
102. G. K. Anderson, H. C . Clark, and J. A. Davies, 1. Organornetal. Chern. 210, 135
(1981).
103. M. P. Brown, J. R. Fisher, R. J. Puddephatt, and K. R. Seddon, Inorg. Chern. 18,2808
(1979).
104. M. P. Brown, R. J. Puddephatt, M. Rashidi, and K. R. Seddon, 1. Chern. Soc., Dalton
Trans., 516 (1978).
105. - - , 1 . Chern. Soc., Dalton Trans., 1540 (1978).
106. M. P. Brown, S. J. Franklin, R. J. Puddephatt, M. A. Thomson, and K. R. Seddon, 1.
Organornetal. Chern. 178, 281 (1979).
107. c. T. Hunt, G. B. Matson, and A. L. Balch, Inorg. Chern. 20, 2279 (1981).
108. M. P. Brown, J. R. Fisher, S. J. Franklin, and K. R. Seddon, 1. Orgnornetal. Chern. 161,
C46 (1978).
109. S. O. Grim and J. D. Mitchell, Inorg. Chern. 16, 1770 (1977).
110. C. A. Tollman, Chern. Soc. Rev. I, 337 (1972).
BINUCLEAR, PHOSPHINE-BRIDGED COMPLEXES 213

111. J. P. Collman and L. Hegedus, Principles and Applications of Organotransition Metal


Chemistry (University Science Books, Mill Valley, CA, 1980).
112. A. L. Balch and B. Tulyathan, Inorg. Chern. 16,2840 (1977).
113. D. J. Doonan, A. L. Balch, S. Z. Goldberg, R. Eisenberg, and J. S. Miller, I. Am.
Chern. Soc. 97,1961 (1975).
114. A. L. Balch, J. R. Boehm, H. Hope, and M. M. Olmstead, I. Am. Chern. Soc. 98, 7431
(1976).
115. A. Modinos and P. Woodward, I. Chern. Soc., Dalton Trans., 1516 (1975).
116. D. M. Hoffman and R. Hoffmann, Inorg. Chern. 20, 3543 (1981).
117. A. D. Rattray and D. Sutton, Inorg. Chirn. Acta 27, L85 (1978).
118. G. S. Brownlee, P. Carty, D. N. Cash, and A. Walker, Inorg. Chern. 14, 323 (1975).
119. F. W. B. Einstein, D. Sutton, and P. L. Vogel, Inorg. Nucl. Chern. Lett. 12,671 (1976).
120. J. T. Mague and A. R. Sanger, Inorg. Chern. 18, 2060 (1979).
121. A. R. Sanger, I. Chern. Soc., Dalton Trans., 228 (1981).
122. A. L. Balch, I. Am. Chern. Soc. 98, 8049 (1976).
123. A. L. Balch, J. W. Labadie, and G. Delker, Inorg. Chern. 18,1224 (1979).
124. G. G. Stanley and P. H. Bird, unpublished results; G. G. Stanley, personal communi-
cation.
125. N. S. Lewis, K. R. Mann, J. G. Gordon, II, and H. B. Gray, I. Am. Chern. Soc. 98,
7461 (1976).
126. D. A. Bohling, T. P. Gill, and K. R. Mann, Inorg. Chern. 20, 194 (1981).
127. K. R. Mann, J. A. Thich, R. A. Bell, C. L. Coyle, and H. B. Gray, Inorg. Chern. 19,
2462 (1980).
128. K. R. Mann, R. A. Bell, and M. B. Gray, Inorg. Chern. 18,2671 (1979).
129. A. L. Balch and M. M. Olmstead, I. Am. Chern. Soc. 98,2354 (1976).
130. M. M. Olmstead and A. L. Balch, I. Organornetal. Chern. 148, CIS (1978).
131. A. Thorez, A. Maisonnat, and R. Poilblanc, I. Chern. Soc. Chern. Cornrn., 518 (1977)
132. J. J. Bonnet, P. Kalck, and R. Poilblanc, Angew. Chern. 92, 572 (1980).
133. H. Schmidbaur and R. Franke, Inorg. Chirn. Acta 13,85 (1975).
134. H. Schmidbaur, J. R. Mandl, A. Frank, and G. Huttner, Chern. Ber. 109,466 (1976).
135. J. R. Moss and J. C. Spiers, I. Organornetal. Chern. 182, C20 (1979).
136. N. J. Kermode, M. F. Lappert, B. W. Skelton, A. H. White, and J. Holton, I. Chern.
Soc., Chern. Cornrn., 698 (1981).
137. H. Werner, R. Feser, W. Paul, and L. Hofmann, I. Organornetal. Chern. 219, C29 (1981).
138. W. P. Fehlhammer, A. Mayr, and B. Olgemiiller, Angew. Chern. Internal. Ed. 14, 369
(1975).
139. W. A. Herrmann, Adv. Organornetal. Chern. in press.
140. R. Colton and C. J. Commons, Aust. I. Chern. 28, 1673 (1975).
141. A. L. Balch and L. S. Benner, I. Organornetal. Chern. 135,339 (1977).
142. K. G. Caulton and P. Adair, I. Organornetal. Chern. 114, Cll (1976).
143. J. A. Marsella and K. G. Caulton, submitted for publication.
144. A. Maisonnat, J. P. Farr, and A. L. Balch, Inorg. Chirn. Acta 53, L217 (1981).
145. J. P. Farr, M. M. Olmstead, C. H. Lindsay, and A. L. Balch, Inorg. Chern. 20, 1182 (1981).
146. J. P. Farr, A. Maisonnat, M. M. Olmstead, and A. L. Balch, unpublished observations.
147. E. W. Stern and P. K. Maples, I. Catalysis 27,120 (1972).
148. - - , I. Catalysis 27,134 (1972).
149. c.-S. Chin, M. S. Sennett, P. J. Wier, and L. Vaska, Inorg. Chirn. Acta 31, L443 (1978).
150. A. A. Arduini, A. A. Bahsoun, J. A. Osborn, and C. Voelker, Angew. Chern. Int. Ed.
Engl. 19, 1024 (1980).
151. J. A. Osborn and G. G. Stan ely, Angew. Chern. Int. Ed. Engl. 19, 1025 (1980).
152. A. R. Sanger, C. G. Lobe, and J. E. Weiner-Fedorak,Inorg. Chern. Acta 53, L123 (1981).
153. Y. Kawabata, T. Hayashi, and I. Ogata, I. Chern. Soc., Chern. Cornrn., 462 (1979).
6
Hydrogenation and
Hydroformylation Reactions
Using Binuclear
Diphosphine-Bridged
Complexes of Rhodium
Alan R. Sanger

NOTATION

dppm Ph 2PCH 2PPh 2


dpam Ph 2AsCH2AsPh 2
dppe Ph 2P(CH 2hPPh 2
dppp Ph 2P(CH 2 hPPh 2
dppb Ph2P(CH2)4PPh2
tdpme (Ph 2PCH 2hCCH 3
triphos (Ph 2PCH 2CH 2hPPh
H
o : /CH 2 PPh 2
/ -c
(-)-diop (CH 3 hC I
\ .... C
o l "CH 2 PPh 2
H
(P) a tertiary phosphine ligand

Dr. Alan R. Sanger. Alberta Research Council, 11315-87 Avenue, Edmonton, Alberta,
Canada T6G 2C2.
215
216 ALAN R. SANGER

a di(tertiary phosphine)
a di(tertiary phosphine) coordinated to a metal by one donor atom

a dppm or dpam ligand bridging two metal centers

1. INTRODUCTION

Modification of transition-metal complexes by complexation with Lewis-


base ligands, especially phosphines or phosphites, has led to major improve-
ments in catalytic reactivity and selectivity. 0.2) Further, by permitting
operation of production facilities at more moderate temperatures and
pressures, considerable reductions in both capital and operating costs have
been realized. For example, this is true of recent installations for the
production of butyraldehyde using rhodium hydroformylation catalysts.(3.4)
The majority of industrial hydrogenation reactions are performed using
heterogeneous catalysts.(S) Nevertheless, the homogeneous hydrogenation
catalysts are of importance for laboratory-scale reactions, the synthesis of
thermally unstable or otherwise sensitive products (especially those of
biological interest), applications of prochiral catalysts for the production
of optically-active products, notably L-dopa (see Chapter 4 ), and mechanis-
tic studies. (1.2.6.7)
The majority of studies involving phosphine-modified homogeneous
catalysts have concentrated on the variation of activity with steric effects(S)
or the Lewis basicity of the modifying ligands. (1.2) The effect of introducing
a bidentate ligand into a catalyst system is frequently considered as merely
the sum total of the effects of two separate ligands maintained in a rigid
geometry with respect to one another. The assumption that chelation will
occur has frequently and sometimes erroneously been made. Chelation is
but one possibility; some alternative effects are suggested in Figure 6.1.
The variety of phosphine complexes of the transition metals is exten-
sive, (9) and the variety of complexes of diphosphines is also impressive. (0)
However, the effects of bridging two or more metal centers with a diphos-
phine ligand have been relatively neglected. In the case of tri(tertiary
phosphine) or higher systems, a greater variety of bridging and chelating
bonding modes are available and many have been realizedY 1 a-lld) Extreme
examples of such schemes are "heterogenized homogeneous catalysts":
transition metal catalysts bonded to flexible polymer supports, especially
those with phosphine anchor-sites. (12) Such systems are discussed in Chapter
14.
In the following sections, binuclear complexes of rhodium bridged by
one or two diphosphine ligands, which are active for the hydrogenation or
HYDROGENA TlON AND HYDROFORMYLA TlON REACTIONS 217

C7
Diphosphine ligand

Bridging
Mo"od,,""

Mononuclear complexes Binuclear complexes

Similar behavior to
monophosphine ligands

/
/M."M~
Sho~ ~ong ~
"Hinge" complexes "Template" complexes Binuclear activation Binuclear elimination
(M - Z - M or M - M - Z) (M - X + M'H ~ M - M' + HX)

Figure 6.1. Potential bonding modes of diphosphines.

hydroformylation of alkenes or alkynes, will be described. When it is


informative to do so, comparison with complexes of other metals, notably
cobalt or iridium, will be made.

2. HYDROGENA TlON OF ALKENES OR ALKYNES

The use as hydrogenation catalysts of neutral or cationic complexes


of rhodium with one or two chelating diphosphine ligands is well estab-
· h ed ..
1IS (1 2 ' 13-15) Th e mech '
amsms 0 f reactIOn
. h ave b een mtense
. 1y stud'Ied
(see Chapter 2), especially in relation to asymmetric hydrogenation using
complexes of chiralligands (see Chapter 4). In contrast to the above systems,
the series .of complexes [RhCi(CO) (PP)]x, where (P-P) is
Bu~P(CH2)nPBu~ (16) or Ph2P(CH2)mPPh2,(17) unexpectedly were found to
be binuclear (x = 2) for a variety of values of n or m, (I), with the sole
exception of the dppe complex, (11).0 7)

p--------P
I ,.cJ I ",CO
Rh Rh
oc~1 cl~1
P-------P

II

The structures of binuclear, diphosphine-bridged complexes of various


metals, including rhodium, are discussed in more detail in Chapter 5.
218 ALAN R. SANGER

Complexes of type (I) are of special interest when the ligating phos-
phorus atoms are separated by one atom as in, for example, a methylene
group. The rhodium atoms are sufficiently close to interact in a repulsive
manner, as demonstrated by the electronic spectrum of
[Rh2Ch(COh(dppmh](8,9) and the separation between the metal atoms
(3.239 A), which is larger than that between the phosphorus atoms of a
bridging dppm ligand (3.130 A).(20) However, the repulsion between the
rhodium centers is not large, as the Rh-Rh separation in the arsine analog
is reduced from 3.396 A (21) to 3.236 A when the crystal contains
dichloromethane of crystallization colinear with the Rh-Rh vector. (22.23)
The flexibility of the structure allowed by the dppm (or dpam) ligands is
considerable. The Rh-Rh bond length in [Rh2Br2(~-CO)(dppmh], (III),
is 2.757 A.(24) Similarly, the oxidative addition of halogens to dppm com-
plexes of structure (I) gives Rh-Rh-bonded complexes of rhodium (II) ,
(IV).(18.19) It is, however, noteworthy that no analogous oxidative addition
complex was observed when [Rh 2Ch(COh(dppmh] was treated with
hydrogenY9)

CI .
OC -Rlr----Rh -co
x

x"~
III IV

Results, using cluster complexes of nickel or iridium, support the


proposal that hydrogenation of triple-bonded substrates can be more readily
achieved by the use of polynuclear homogeneous catalysts than by use of
mononuclear catalysts. (25) The requirement for multinuclear centers to
heterogeneously catalyze the hydrogenation of CO has also been pro-
posed.(26) Consequently, complexes of transition metals in which the metal
atoms are constrained to be proximate are potentially of interest as catalysts.
However, the only neutral complexes of structure (V) that are catalyti-
cally active are the cyano-complex and the diarsine analog (see Table
I), (27,28) which are formed in situ by loss of CO from the tetracarbonyl
precursor, (VI). (29)

(xi:~o ~FO
oc'liJ
Rh Rh NC-Rh Rh-CN
II ~jl
OC~

V VI
HYDROGENA TION AND HYDROFORMYLA TlON REACTIONS 219

Table 1. Hydrogenation Catalytic Activity of A-Frame Complexes·

Concentration
Catalyst b (x10- 3 M) Rate C

For 1-Hexene ~ n-Hexane


[Rh 2 CI(COh(dpmht 0.40 2.3
[Rh 2 C1(COh(damht 0.40 3.4
[Rh 2 Br(COh(dpmht 0.40 1.1
[Rh 2Br(COh(damht 0.40 0.17

For Phenylacetylene ~ Styrene


[Rh 2 CI(COh(dpmht 0.39 0.31
[Rh 2 CI(COh(damht 0.40 0.20
[Rh 2 Br(COh(dpmht 0.40 0.23
[Rh 2 Br(COh(damht 0.42 0.45
[Rh 2 (CNh(CO)4(dpmh] 0.40 3.0
[Rh 2 (CNh(CO)4(damh] 0.40 0.31
[Rh 2 (NCOh(COh(dpmh] 0.40 0.03
[Rh 2 (NCOh(COh(damh] 0.41 0.0001
[Rh 2 (NCSh(COl4(dpmh] 0.38 0.002
[Rh 2 (NCSh(CO)4(damh] 0.39 O.
[Rh 2 C1(N 3 )(COh(dpm)2] 0.42 0.12

a Reference 28. For dpm, read dppm. For dam, read dpam.
b The formulation of the catalytically active species may differ from that of the "catalyst" orginally dissolved
in methanol under an atmosphere of dihydrogen.
C Moles of substrate hydrogenated per minute per mole of catalyst, at 25°C.

The first rhodium complex of the structure now commonly called


"A-frame" to be described was the neutral sulphido-bridged complex
(VII, X = S, n = 2). (30.31) Unlike the iridium analog, (32) this complex is a
very poor catalyst for the hydrogenation of ethylene.(31) In contrast, both

W
the rhodium and iridium cationic complexes (VII, X = Cl or Br, n = 1)

II
(2-n)+

Rh 'Rh OC-Rh-Rh - co
OC~U'CO
VII (X is an anionic
U VIII
ligand of charge n- )
220 ALAN R. SANGER

Table 2. Alkenes and Alkynes Hydrogenated 8 • b

Reagent Initial product Subsequent products

1-Pentene n-Pentane
1-Hexene n-Hexane
2-Hexene n-Hexane
c-Hexene c-Hexane
Vinyl-cyclohexane Ethylcyclohexane
Styrene Ethylbenzene
cis -Stilbene 1,2-Diphenylethane
trans- Stilbene 1,2-Diphenylethane
Phenylacetylene Styrene Ethylbenzene
1-Phenylpropyne cis -1-Phenylpropene 1-Phenylpropane
1-Phenylpropyne C cis -1-Phenylpropene trans -1-Phenylpropene,
1-Phenylpropane
Diphenylacetylene cis -Stilbene 1,2-Diphenylethane

a Catalyst: [Rh 2 (I'-Cl)(CO),(dppmlzr (2.0 x 1O- 3 M) in methanol (50 em'), at 22°C and 1 atmosphere
total pressure.
b References 27, 28.
c Catalyst: [Rh 2 (I'-Cl)(CO),(dpam)'1+; for this catalyst hydrogenation of the alkene was very slow, and
concurrent cis to trans isomerization occurred. Reference 35.

are active catalysts for the hydrogenation of a variety of alkenes or alkynes


(see Tables 1 and 2).(27,28,33-36)
The rhodium (0) complex, (VIII), reacts with an alkyne to form a
1 : 1 addition complex which is catalytically active for the hydrogenation
of alkynes and (weakly) alkenes. (37) Complexes of types (III), but with
chloro-ligands, (38) and (VII, X = CI)(33) also form complexes with acety-
lenes, which can be subsequently hydrogenated. A series of complexes of
structure (VII) with RO-,(31) RS-,(35,39) RC0 2-,(35.39) HS0 4-,(39) or
ClHO- (39) ligands, and [Rh 2(CH 3C(CH 2hh(dppmh](40) have also been
prepared, but their catalytic activity has not yet been reported.
Cationic complexes (VII, X = CI or Br) hydrogenate alkynes success-
ively to the corresponding alkene and alkane (Figures 6.2 and 6.3). (27,33-35)
Even though hydrogenation of an alkene may occur at a higher rate than
hydrogenation of an alkyne, it does not occur while alkyne is present. In
contrast, the corresponding iridium complex, generated in situ e1Pnmr)
from [Ir2CI(COMdppmh][IrC}z(COh], (36) hydrogenates both the alkyne
and alkene simultaneously (Figure 6.4). (28.34)
HYDROGENA TION AND HYDROFORMYLA TlON REACTIONS 221

100
/
/
80 /
I
c I
.,g 60 /
';;;
8.
~ 40
U
*-
20

400 800 1200 1500 2800


Time(min.)
Figure 6.2. Hydrogenation of phenylacetylene catalyzed by a solution of
[Rh 2 Cl(COh(dppm)z][RhCI 2 (CO)z] in methanol (2.00 x 10~3 M; 22°C).

200

180

160

140

c 120
Q
~

iii
0 100 6
0.
E
0 80 o
0 0

60

40

20

0 50 100 150 200 250


Time (min)
Figure 6.3. Hydrogenation of a second aliquot of phenylacetylene catalyzed by a solution of
[Rh 2 Br(COh(dppmlz]BPh4 in methanol (0.70 x 1O~3 M; 25°C). The ethylbenzene present
at time = 0 is the product from hydrogenation of the first aliquot of phenylacetylene. 0 =
PhCCH; [;] = PhCHCH 2 ; 6 = PhCH 2 CH 3 . Reproduced. with permission, from Reference 28.
222 ALAN R. SANGER

100
I
I
80 I
PhC=CH I
c:
.~ 60
/ I

1g 40
U
~

20

O~---=~~---------L------ __
- L_ _ _ _~_ _~_ _~.~~

o 400 800 1200 1600 2800


Time (min.)

Figure 6.4. Hydrogenation of phenylacetylene catalyzed by a solution of


[Ir2Ci(CO)4(dppm)zJ[IrCi2(CO)ZJ in methanol (0.27 x 10- 3 M; 22°C). Reproduced, with
permission, from Reference 28.

This contrasting behavior may be due to the difference between the


abilities of the rhodium and iridium systems to react with hydrogen to form
oxidative addition products. Hydrogen reacts with [Ir2(~- S)
(COh(dppmhJ to form a dihydride,(32) and with [Ir2CI(COh(dppmhJBPh4
to form an unidentified hydrido-complex. (36) In contrast, no oxidative
addition product was observed when complexes (V, X = Cl),(19) (VI;
X = Cl),(33) or (VII, X = S)(30.31) were treated with hydrogen.
It is tempting to speculate that only for the dinuc1ear complexes of
iridium is there initial formation of a dihydrido-complex, which sub-
sequently reacts with an alkene to form an alkyl intermediate, or with an
alkyne to form an alkenyl intermediate. If such is the case, the activity of
dinuc1ear rhodium complexes must depend on initial formation of an alkyne
or alkene complex, which would then react with hydrogen. There exists
some evidence for such a scheme. The successive hydrogenation of alkynes
and alkenes(28.34) suggests that activation of an alkene is inhibited by an
alkyne, probably by preferential coordination of the latter. Further, com-
plexes (VII, X = H) or (IX) do not alone react with hydrogen, but do so
after reaction with an alkyne (acetylene or phenylactylene). (37)

OC"U
r1J
Rh Rh
"co

IX
HYDROGENA TION AND HYDROFORMYLA TlON REACTIONS 223

Table 3. Formation of Carbon Monoxide Bridged Complexes 8

Complex Product

(i) Association of Carbon Monoxide


[Rh 2(/-L -CI)(COlz(dpmht [Rh 2(/-L-Cl)(COh(/-L-CO)(dpmht
[Rh 2(/-L-CI)(COh(damht [Rh 2(/-L-Cl)(COh(/-L-CO)(damht
[Rh 2(/-L-Br)(COlz(dpmht [Rh 2(/-L -Br)(COh(/-L -CO)(dpmht
[Rh 2(/-L -Br)(COlz(damht [Rh 2(/-L-Br)(COlz(/-L-CO)(dam)st
[Rh 2(CNh(COh(dpmhJ [Rh 2(CNlz(COh(/-L -CO)(dpm}ZJ
[Rh 2(CNh(COh(dam}zJ [Rh 2(CNh(COlz(/-L -CO)(damhJ
[Rh 2(NCOh(COh(dpmhJ none detected
[Rh 2(NCOh(COh(damhJ none detected
[Rh 2CI(N 3 )(COh(dpmhJ

(ii) Dissociation of Carbon monoxide


[Rh 2(CNh(CO)4(dpmhJ [Rh 2(CNh(COh(/-L -CO)(dpm}zJ
[Rh 2(CNh(CO)4(dam2J [Rh 2(CNh(COlz(/-L -CO)(dam}zJ
[Rh 2(NCSh(CO)4(dpmhJ none detected
[Rh 2(NCSh(COMdamhJ none detected

" Reference 29, 33. For dpm, read dppm. For dam, read dpam.
b This product is present in low equilibrium amounts in solution; the solid product has not yet been isolated.

It is also of interest to compare the reactivity of both active and inactive


dinuclear complexes with CO (see Table 3).(29,33) The only complexes that
are of significant activity for the hydrogenation of alkynes are the chloro-
or bromo-complexes of structure (VII), and the cyano- or, weakly, azido-
complexes of structure (V). (27,28) It is these complexes, and only these
complexes, which react readily and reversibly to form carbonyl-bridged
complexes. (29,33) Reaction of 13CO with complex (VII, X = CI) occurs
initially at a terminal site, with subsequent migration of an orginally terminal
CO to the bridging position, (X) (Equation 2.1).(33) The neutral, tricar-
bonyl cyano-complex, (XI), is unique within the series of dinuclear com-
plexes of rhodium(I) described to date (Equation 2.2).(29) The neutral azido
complex (V, X 2 = CI, N3 ) forms weakly conducting solutions of the cationic
complex (X, X = N 3 ) (Equation 2.3).(29)
224 ALAN R. SANGER

w*
+

VII OC-Rh-"Rh-CO (2.1)

QJ X

~
VI ""C()
NC."W . .CN
Rh-Rh
alkyne,

~V
H,
(2.2)
co
O C " U 'CO

XI

The cationic complex (VII, X = Cl) forms a symmetrical e1Pnmr; i.r.)


complex with diphenylacetylene, which is very labile. (33) The complexes
[Rh 2Clz(lL- CO)(IL -acet)(dppm)zJ, where "acet" is hexaftuoro-2-butyne
or acetylenedicarboxylic acid dimethyl ester, are also symmetrical, and the
"acet" ligand is coplanar with the rhodium atomsY8) The C-C bond length
is increased to 1.32 A, and so the complex is truly described as a cis-
dimetallated alkene. (38) Comparison with the reactions with CO suggests
that the activation of an alkyne (or alkene) to hydrogenation may occur
when the alkyne is initially coordinated to the catalyst at a terminal position,
e.g., (XII),(28) with subsequent oxidative addition and hydrogen transfer
steps. A complex of the type (XII) may instead be only a minor but active
component of a solution of which the alkyne-bridged complex is the major
. (4142)

w+
but less actIve component. '

OC-Rl)J'~c;Rh- alkyne
II
o
XII

It must be stressed that the above arguments are conjecture, and that
the possibility of the presence of an undetectably small but finite concentra-
tion of an active hydrido-complex cannot be excluded on the basis of
available data. the above systems are not simple, and the presence of more
than one parallel "mechanism" is possible. Comparison of the activities of
complexes with different anionic ligands (see Table 1) was made by compar-
ing methanol solutions. (27,28) Acetone and other strong donor solvents form
HYDROGENA TION AND HYDROFORMYLA TION REACTIONS 225

stable, less active complexes with many of the catalysts. (29,32) The activities
of both the BPh4- and [RhC}z(COhr salts of [Rh2CI(COh(dppmht were
similar at low-to-intermediate (ca. 0.4 x 10-3M) concentrations, as is
appropriate for fully dissociated complexes of which the cation is the active
catalyst. However, at greater concentrations the molar activity of the
[RhC}z(COhr salt decreased markedly, indicating that significant changes
in the nature of the solutions were occurring. The cause of these phenomena
has not yet been determined.

2.2. Complexes of Diphosphines with Longer Chain Lengths

The chelated complex [HRh( (-)-diop h] is active for the hydrogenation


of terminal ole fins or unsaturated carboxylic acids. (43-45) From kinetic data
and spectroscopic evidence an active intermediate containing a monoden-
tate (- )-diop ligand has been postulated, (44) analogous to a known carbonyl
complex. (17) Polymeric species containing bridging diphosphine ligands
were postulated to explain inhibition of catalytic activity in the presence
of added (-)-diop or dppb, but not added PPh 3.(44) However, proof of
existence of such bridged complexes was not obtained.
The activity of complexes prepared in situ containing one mole of
diphosphine Ph 2P(CH2)nPPh2 (n = 1-6) per mole of rhodium(46) varied
with the diphosphine chain length in a similar manner to the cationic
complexes [Rh(P-Pht. (45) In each case only structures containing chelat-
ing diphosphine ligands were consideredY3-15,43-49) In one instance(49) a
number of the complexes formed have independently been shown to be
of a bridged structureY7) Kinetic studies of the different diphosphine
complexes gave results so varied as to virtually exclude a common mechan-
ism. For example, whereas no interaction of hydrogen with [Rh(dppeht
has been observed directly, the dihydrido-complex [H 2Rh(dpppht is
stable. (45) Bridged complexes analogous to the binuclear carbonyl-complex
dictation [Rh2(CO)4(dppbh(~-dppb)]2+,(50) complexes of type (1),(17) or the
diop-bridged, catalytically active, ruthenium complex [Ru2CI4(dioph(~­
diop)],(51) have not been considered as possible intermediates. However,
the ligands that form the most active complexes are ligands that are known
to also form complexes containing a monodentate diphosphine or bridged
complexes.
The ligand p-C6H4[CH2P(CH2CH2PPh2hM =L) is able to chelate two
separate metal atoms, thereby combining chelation and bridging
features. (52) The catalytic activity of the binuclear complex [RhCI(L )RhCI]
is said to be due to the independent activity of each metal center, maintained
at a distance from one another. (52) Comparison with the mononuclear
analog, [RhCl(triphos)], was not attempted.
226 ALAN R. SANGER

The mechanisms of hydrogenation by neutral or cationic mononuclear


complexes of rhodium are reasonably well understood(l,2,6,41,42,S3.S4) and
are discussed in detail in Chapter 2. The system is, however, sufficiently
complex that new information is constantly being obtained. For example,
it has recently been shown that a solvated complex is kinetically important
when benzene is the solvent. (SS) This may have major implications in studies
of complexes of aryl-phosphines or diphosphines, such as Ph2P(CH2)nPPh2
or diop. (1S) It should be noted that for complexes of chelating chiral
diphosphines (see Chapter 4), the enantioselective step involves the least
stable of two possible catalytic intermediates; the complex which is stable,
isolable, and present in highest concentration is not as important in deter-
mining the reaction rate and does not yield the major product.(41,42)
Homogeneous and heterogeneous hydrogenation catalysts have been
compared, (S6) but multinuclear homogeneous processes have frequently
not been considered. This is surprising when one considers the often
proposed multicenter mechanism for heterogeneous catalysis(S):
H H H H C2H4
7/77// ~ 77/77
I 71)7/7777/77

1 H,
1
H C2HS
77/7777///// ///7///)/777
Similar multinuclear processes have been noted for homogeneous
hydrogenation by complexes of metals other than rhodium. (7) For example,
successive reactions of an alkene with two [HCo(CN)s]3- anions give the
corresponding alkyl complex anion and alkane. (S7,S8) Similarly,
stoichiometric hydrogenation of a terminal alkene by [HM(COh(CsH s)]
(M = Cr, Mo, W) occurs with formation of [M2(COMCsH sh].(S9) For
M = Cr, the binuclear complex product reacts with hydrogen to regenerate
the hydrido-complex, and the system is cyclic. (S9)
Dinuclear elimination reactions of metal alkyls (Equation 2.4, L = all
other ligands) have been studied less intensely(60) than, for example, similar
reactions of metal amides (Equation 2.5). (61) The following characteristics
are common to known dinuclear elimination reactions of metal alkyls: (i)
a site of un saturation must be present cis to the alkyl ligand; and (ii) the
other reagent must be a hydrido-complex. (60)
(2.4)
(2.5)
The existence of a step similar to Reaction 2.4 in the hydrogenation
of alkenes or alkynes has not been proven, or even considered, for most
HYDROGENA TlON AND HYDROFORMYLA TlON REACTIONS 227

systems. If the rate-determining step is either formation of an alkene


complex, or hydrogen migration to form an alkyl complex, the reaction
will be first-order in hydrido-metal complex, even if a binuclear alkane
elimination step occurs as a major or minor reaction. On the available
evidence such a binuclear process cannot be totally disregarded. When the
metal centers are held in proximity to one another, as with binuclear
complexes bridged by a diphosphine ligand, or a polymer with phosphine
anchor-sites, such a step may well be significant.

3. HYDROFORMYLA TlON OF ALKENES

Improvements in rate, selectivity, and catalyst lifetime may be achieved


when cobalt hydroformylation catalysts are modified by addition of a
tertiary phosphine. (62,63) When the phosphine added is dppe, the complex
formed, [HCo(COh(dppe)],(64) is a weak catalyst for the hydroformylation
of 1-pentene, exhibiting poor selectivity to formation of the unbranched
product (Equation 3.1a).(62)

(3.1)

(b) RCH(CH 3 )CHO

In contrast, addition of Ph2P(CH2)nPPh2 (n = 4 or 5) to solutions of


[HCO(CO)4] gives complexes of unknown structure that are of high activity
and good selectivity.(62)
When the substrate to be hydroformylated is methyl acrylate, the
above trend is reversed. The most active catalyst is that formed by addition
of dppe. (63) The greatest activity was found for a catalyst with dppe: Co =
0.5. (63) However, direct comparisons of the two systems are not reasonable
because the conditions of reaction, proportions of diphosphine to cobalt,
and the substrate differ markedly. In each case hydrogenation is a competi-
tive reaction.
The mononuclear catalyst [Ru(COh(dppe)] is of lower activity than
monophosphine ruthenium complexes, but of higher selectivity. (65)
Platinum complexes promoted with SnCh are also of low activity if the
phosphine ligand is dppe, but are much more active with dppb, or a related
his (phosphinomethyl)cycloalkane ligand, (66) which are known to form
bridged structures. (17b)
Reaction of Ph2P(CH2)nPPh2 (n = 2-5), or corresponding cyclo-
hexylphosphines, with [Rh2Ch(CO)4] (PP: Rh = 1) gave solutions active
228 ALAN R. SANGER

for the hydrogenation of a, l3-unsaturated esters, but of low hydroformyla-


tion activity.(49) Under these conditions, bridged complexes of type (I)
(p.217) are formed, except for the chelate complex of dppe, (11).(17)
Although chelation of rhodium by the reaction of diop(67) or cyclo-1,2-
(Ph 2POhC6H 10 with [Rh 2Ch(CO)4] has been assumed, formation of
bridged complexes occurs under the conditions reported. (17) At greater
P-P : Rh ratios, the hydroformylation activity and selectivity to formation
of the unbranched aldehyde both increased.(49,67) The structures of com-
plexes with (P-P: Rh > 1) vary with the chain-length of P-P; dppe forms
[Rh(dppeht, but other diphosphines form carbonyl complexes, some of
· h con tam
whIC . mono d entate or bn'd' gmg d'lp hosp h'me I'19ands. (17485068)Th
' " e
variation in catalytic activity of chlororhodium complexes with diphosphine
ligand chain-length is not, therefore, a simple correlation.
Addition of Ph2P(CH2)nPPh2 (n = 1_4),(70-74) diop,<67,69) triphos-
phines,(71,72) or soluble polymers with phosphine anchor-sites(72) to sol-
utions containing [HRh(CO)(PPh 3h] gives complexes active for the hydro-
formylation of a variety of alkenes under mild conditions. For complexes
of diop, asymmetric induction is observed in appropriate cases.(67.69) Com-
plexes of rhodium containing both a monophosphine and a diphosphine
are efficient, unbranched product-selective catalysts for hydroformylation
of 1-hexene,(73) as are complexes [{HRh(CO)(PP)h(~-P-P)].(74) At
higher temperatures (105°C) the maximum activity for the dppb system
occurs at the ratio dppb : Rh = 1.5. (74)
The predominant complex in a solution prepared by the dissolution
of [HRh(CO)(PPh 3h] under an atmosphere of CO and H2 is
[HRh(COh(PPh3h] (Equation 3.2/75 -78 ):

(3.2)

Comparison of spectroscopic data with the catalytic activity of solutions


containing various proportions of a polyphosphine and [HRh(COh(PPh3h]
indicate both the existence of bridged complexes and their relatively high
activity. (70-72)
Addition of a small amount of a diphosphine (Figure 6.5) or triphos-
phine (Figure 6.6) to a solution containing [HRh(COh(PPh 3h] increased
the hydroformylation catalytic activity. (70-72i However, addition of
equivalent or greater amounts of the same polyphosphines gave solutions
of lower activity. For the diphosphine dppe, the following equilibria were
proposed (Equations 3.3 and 3.4)(71):

2HRh(COh(PPh 3 h + dppe 2PPh 3 + {HRh(COh(PPh3 )}z(dppe) (3.3)


(XIII) (XIV)
HYDROGENA TION AND HYDROFORMYLA TlON REACTIONS 229

30,-------,--------,--------.-------~------__,

20
!
0
cc:
.2

QI
.~
.]
QI
cc:

0.25 0.50 0.75 1.00 1.25


Moles of Ph 2 P(CH 2 )n PPh 2 per
Mole of Rhocli~m

Figure 6.5. Relative rates of hydroformylation of I-hexene catalyzed by benzene solutions


containing [HRh(CO)(PPh 3 h] and various amounts of the diphosphine Ph2P(CH2)nPPh2
(n-2-4). Reproduced, with permission, from Reference 72.

(XIV) + dppe ¢ 2PPh 3 + {HRh(COh(dppe)}n (3.4)


(XV)

Infrared spectra of solutions of mixtures of [HRh(CO)(PPh3 hJ and dppe


in molar ratios up to and exceeding 1 : 2 under hydroformylation conditions
show, in each case, bands due to complex (XIII) and two substituted species
in equilibria (Figure 6.7, Table 4). The bands assigned to (XIV) were
predominant for solutions of composition dppe/Rh = 0.3-0.5. This ratio
is close to but slightly higher than that of the solution of maximum catalytic
activity (Figure 6.5). (71,72) The bands assigned to (XV) were the most intense
for solutions of relatively greater amounts of dppe, which were also solutions
of lesser catalytic activity. No other infrared bands (2300-1700 cm -1) were
observed. (71)
230 ALAN R. SANGER

~,---------------,---------------,---------------,--------------.
o

o
o
]
:E
">
.~
o
-;;
"'1

~~------------~~~2S~----------~O~.~~------------~O~.7~S----------~~1.00

CH 3ClcH 2PPh 2 )3' Rh

Figure 6.6. Relative rates of hydroformylation of l-hexene catalyzed by benzene solutions


containing [HRh(CO)(PPh 3 hJ and various amounts of (Ph 2 PCH 2 hCCH3 . Reproduced, with
permission, from Reference 71.

Figure 6.7. Infrared spectra of solutions in benzene of the products of reaction of [HRh(CO)-
(PPh 3 hJ with CO and various amounts of Ph 2 P(CH2 lzPPh 2 under hydroformylation condi-
tions.
HYDROGENA TION AND HYDROFORMYLA TION REACTIONS 231

Table 4. Infrared Spectra (2300-1700cm 3 ) of Solutions


of HRh(COMPPh 3)2 and Ph 2PCH2CH2PPh 2 or
(Ph2PCH2)3CCH3

Complex v(RhHt v(cot

(XIII) 2040(w) 1977(vs),1943(sh)


(XIV) 2031(w) 1954(vs)b
(XV) 2015(w) 1971(vs)b
(XVI) 2004(w) 1954(vs)b
(XVII) 1977(w) 1909(vs)b

a Values in wave numbers.


b Nonoverlapping bands; shoulders or weak bands overlapping bands due to
other complexes may also be present.

For solutions prepared from [HRh(CO) (PPh 3 h] and the tripod


triphosphine, tdpme, the following equilibria (Equations 3.5 and 3.6) were
postulated(71) :
3HRh(COh(PPh3 h + tdpme ¢ 3PPh 3 + {HRh(COh(PPh 3 )h(tdpme) (3.5)
(XVI)

2(XVI) + 4tdpme 6PPh 3 + 3CO + 3{HRh(CO)(tdpme)h (3.6)


(XVII)

The infrared spectrum of a solution containing equimolar amounts of


rhodium and tdpme showed only bands due to the binuclear complex,
(XVII) (Figure 6.8, Table 4).(71) Solutions containing lesser amounts of
tdpme show, in addition to bands due to (XIII) or (XVII), bands at 2004
and 1954 cm-\ assigned to the monosubstituted, tdpme-bridged complex,
(XVI).(71) These bands were most intense for solutions of composition close
to tdpme/Rh = 0.3, which is close to but slightly higher than the ratio for
the solution of maximum catalytic activity (Figure 6.6). (71) Further evidence
for the nature of (XVI) has been obtained from the 1Hnmr spectra at
different temperatures of a solution of composition tdpme/Rh = 0.33. At
303 K no coupling of phosphorus to the hydrido-ligand was discernable.
At 233 K a broad doublet (718.0 p.p.m., 2J (P-H) - 4 Hz) was observed,
and at 213 K a doublet of doublets eJ(p-H) - 2J (P'_H) - 4 Hz] was re-
solved. (72) The resolution of the coupling constants at different temperatures
confirms that the two phosphine ligands bonded to rhodium are different.
The linear triphosphine, triphos, behaves similarly. (72)
The mechanism of cobalt-catalyzed hydroformylation (see Chapter 2)
has been studied intensively. Early postulates that the rate-determining
232 ALAN R. SANGER

2100 2000 1900 em' l


" " " 0=0

~:2~2
~~
~tr// 0.58
0.72
0.91

1.06(5)

Figure 6.8. Infrared spectra of solutions in benzene of the products of reaction of [HRh(CO)-
(PPh 3 hJ with CO and various amounts of (Ph 2 PCH 2 hCCH 3 under hydroformulation condi-
tions.

step is hydrogen addition to an unsaturated, mononuclear, acyl-cobalt


complex(79) have been challenged on the basis of in situ infrared spectro-
scopic studies. (80-82) The mechanism has been shown to include a binuclear
reaction (Equation 3.7) as the major or only reductive elimination step. (80.81)

(3.7)

It has been suggested(8o.81) tr..at the acyl complex is [RCO·CO(CO)4], or


[RCO'Co(COhPBu31 for modified catalysts, but a site of unsaturation cis
to the acyl ligand may be required (c.f., Reference 60). A more probable
formulation is therefore [RCO·Co(COh], or [RCO·Co(COhPBu3]. A
binuclear, free-radical mechanism for the cobalt-catalyzed hydroformyla-
tion of styrene or other conjugated substrates has also been proposed.(84)
These studies are far-reaching/ 83 ) especially because similar binuclear
elimination steps have not received much consideration in studies of
rhodium hydroformylation catalysts. (72.83)
Cobalt complexes with diphosphine ligands have shown features con-
sistent with the above proposals. With an equimolar or excess amount of
HYDROGENA TION AND HYDROFORMYLA TlON REACTIONS 233

dppe, which forms a mononuclear, chelated complex,(64) poor activity is


obtained.(62) However, with Ph2P(CH2)nPPh2 (n = 4,5), which show a
greater tendency to form bridged cobalt complexes, (85) good activity and
selectivity are obtained. (62) The maximum activity of the dppe-modified
cobalt catalyst for hydroformylation of methyl acrylate occurs at a ratio
dppe/Rh = 0.5, suggestive of formation of a binuclear system. (63) Similarly,
for PtCh/SnCh/diphosphine hydroformylation catalysts it was found that
dppe depressed the activity of the platinum complex, but that dppb or
related diphosphino-cycloalkane ligands considerably improved the
activity.(66) For complexes containing no SnCh, dppe chelates platinum,
but dppb forms bridged complexes. (17b)
The reaction of [HRh(COh(PPh 3hJ with a diphosphine (P-P /Rh =
0.5(70-72) or 1.5(74)) to give a solution containing predominantly a binuclear,
diphosphine-bridged complex has been established spectroscopically, as
discussed above. A binuclear elimination step in the mechanism of hydro-
formylation using mononuclear rhodium catalysts cannot be excluded on
the basis of data available. (86) The coincidence between the maximum
activity of solutions containing rhodium and a diphosphine, and the forma-
tion of bridged complexes in maximum concentration under the same
conditions suggests that binuclear elimination may indeed be both possible
and important. Maintenance of two rhodium, or other metal, centers in
close proximity would favor such a reaction (Equation 3.8):

HRhL n ·pp. RhL mCOR ~ RCHO + Rh2Ln+m (P~P)


H2,L~ {HRhLnh(P~P) (3.8)

Reaction of binuclear complexes of rhodium (0) with hydrogen to


regenerate hydridorhodium (I) complexes under hydroformylation condi-
tions has been established. os .77 ) The extension of these arguments to
complexes with triphosphines or polymers with a high density of phosphine
anchor-site has been made for complexes of cobalt(12) and rhodium.(12.72)

4. CONCLUDING REMARKS

In Sections 2 and 3, I have attempted to show that when two metal


centers are bridged by a diphosphine they do not necessarily react as
independent catalytic centers, but may react in a binuclear manner.
The relationship of such systems to polymers,(12) or clusters,(87,89) both
of which are important in catalysis, is strong, but comparisons with reactions
between neighboring sites at surfaces are more tenuous.(S,87,88,90)
234 ALAN R. SANGER

Important developments in the field of bridged binuclear catalysts will


undoubtedly come from studies of unsymmetric diphosphines, e.g.,
R2PCH2PR~,(91) phosphines also containing nitrogen,(92·93) sulphur,(94) or
other donor atoms, and other innovations yet to be described.
The importance of bridged, binuclear complexes for decarbonylation
reactions (see Chapter 11), the water-gas shift reaction (see Chapter 5),
and other catalytic applications is only now beginning to be investigated.

Notes Added in Proof

The cationic complex reported to be [Rh2CI(COh(dppmh-


(~-PhCCPh)t has been reinvestigated, and the spectroscopic evidence
reevaluated. The differences in 31 p nmr parameters between solutions
containing [Rh 2CI(CO)z(dppm)zt alone and those with added PhCCPh
may be due to differences in conditions alone, such as solvent effects.
However, the cationic complex (XVIII) has been synthesized. (95)
+

XVIII, R = CH 3 C0 2

The use of 1,1'-bis(diphenylphosphino)ferrocene as a bidentate ligand


in rhodium hydroformylation catalysis has been studied in detail.(96) Com-
plexes containing this ligand as a bridge are remarkably stable and effective
as catalysts. Although a mechanism involving independent reactions of
each metal center as catalyst was proposed, the authors allowed for the
possibility of an internuclear interaction.
The diversity of results of studies of hydrogenation and hydroformyla-
tion catalysis have prompted a report on the factors governing the different
mechanisms of binuclear elimination reactions of complexes leading to
carbon-hydrogen bond formation. (97)

ACKNOWLEDGMENTS

I am grateful to Drs. J. T. Mague, A. L. Balch, O. R. Hughes, M. D.


Fryzuk, and K. R. Grundy for communication of results prior to publication,
and to these colleagues and Dr. J. P. Collman for helpful and stimulating
discussions.
HYDROGENA TION AND HYDROFORMYLA TlON REACTIONS 235

REFERENCES

1. G. W. Parshall, Homogeneous Catalysis (Wiley-Interscience of John Wiley and Sons,


New York, 1980).
2. J. P. Collman and L. S. Hegedus, Principles and Applications of Organotransition Metal
Chemistry (University Science Books, Mill Valley, CA 1980).
3. R. Fowler, H. Connor, and R. A. Baehl, Hydrocarbon Proc. 1976, 247-249.
4. J. Falbe and E. G. Hancock, Propylene and Its Industrial Derivatives (Halstead, New
York, 1973), Chapt. 9.
5. P. N. Rylander, Catalytic Hydrogenation in Organic Syntheses (Academic Press, New
York, 1979).
6. F. J. McQuillin, Homogeneous Hydrogenation in Organic Chemistry (Reidel, Dordrecht,
1976).
7. R. E. Harmon, S. K. Gupta, and D. J. Brown, Chern. Rev. 73, 21-52 (1973).
8. c. A. Tolman, Chern. Rev. 77,313-348 (1977).
9. c. A. McAuliffe and W. Levason, Phosphine, Arsine, and Stibine Complexes of the
Transition Metals (Elsevier, Amsterdam, 1979).
10. W. Levason and C. A. McAuliffe, Adv. Inorg. Chern. Radiochern. 14, 173-253 (1972).
11. a) R. B. King, P. N. Kapoor, and R. N. Kapoor, Inorg. Chern. 10, 1841-1850 (1971);
b) R. B. King and J. C. Cloyd, Jr., Inorg. Chern. 14, 1550-1554 (1975); c) R. B. King,
J. A. Zinich, and J. C. Cloyd, Jr., Inorg. Chern. 14, 1554-1559 (1975); d) M. M. Taqui
Khan and A. E. Martell, Inorg. Chern. 13,2961-2966 (1974).
12. D. C. Bailey and S. H. Langer, Chern. Rev. 81, 109-148 (1981).
13. a) D. A. Slack and M. C. Baird, J. Organornetal. Chern. 142, C69-C72 (1977); b) D. A.
Slack, I. Greveling, and M. C. Baird, Inorg. Chern. 18,3125-3132 (1979).
14. R. R. Schrock and J. A. Osborn, J. Arner. Chern. Soc. 93, 2397-2407 (1971).
15. J. Halpern, D. P. Riley, A. S. C. Chan, and P. L. Pluth, J. Arner. Chern. Soc. 99, 8055-8057
(1977).
16. F. C. March, R. Mason, K. M. Thomas, and B. L. Shaw, J. C. 5., Chern. Cornrn. 1975,
584-585.
17. a) A. R. Sanger, J. C. 5., Dalton Trans. 1977,120-129; b) A. R. Sanger, J. C. 5., Dalton
Trans. 1977, 1971-1976.
18. A. L. Balch, J. Arner. Chern. Soc. 98, 8049-8054 (1976).
19. A. L. Balch and B. Tulyathan, Inorg. Chern. 16,2840-2845 (1977).
20. M. Cowie and S. K. Dwight, Inorg. Chern. 19, 2500-2507 (1980).
21. J. T. Mague, Inorg. Chern. 8, 1975-1981 (1969).
22. M. Cowie and S. K. Dwight, Inorg. Chern. 20, 1534-1538 (1981).
23. A. L. Balch, unpublished results.
24. M. Cowie and S. K. Dwight, Inorg. Chern. 19, 2508-2513 (1980).
25. E. L. Muetterties, Pure Appl. Chern. 50,941-950 (1978).
26. A. Brenner and D. A. Hucul, J. Arner. Chern. Soc. 102, 2484-2487 (1980).
27. A. R. Sanger, Preprints, 7th Canadian Symposium on Catalysis (Edmonton, 1980), pp.
67-74.
28. - - , Canad. J. Chern. 60,1363-1367 (1982).
29. - - , J. C. 5., Dalton Trans. 1981, 228-233.
30. C. P. Kubiak and R. Eisenberg, J. Arner. Chern. Soc. 99,6129-6131 (1977).
31. - - , Inorg. Chern. 19, 2726-2732 (1980).
32. C. P. Kubiak, C. Woodcock, and R. Eisenberg, Inorg. Chern. 19,2733-2739 (1980).
33. J. T. Mague and A. R. Sanger, Inorg. Chern. 18, 2060-2066 (1979).
34. A. R. Sanger, Preprints, Sixth Canadian Symposium on Catalysis (Ottawa, 1979), pp.
34-43.
236 ALAN R. SANGER

35. J. T. Mague, unpublished results.


36. A. R. Sanger, unpublished results.
37. C. P. Kubiak and R. Eisenberg, J. Arner. Chern. Soc. 102, 3637-3639 (1980).
38. M. Cowie and T. G. Southern, J. Organornetal. Chern. 193, C46-C50 (1980).
39. T. S. Cameron, S. P. Deraniyagala, K. R. Grundy, and K. Jochem, unpublished results.
40. M. D. Fryzuk, Inorg. Chirn. Acta, in press.
41. J. M. Brown and P. A. Chaloner, J. C. S., Chern. Cornrn. 1980, 344-346.
42. A. S. C. Chan, J. J. Pluth, and J. Halpern, J. Arner. Chern. Soc. 102,5952-5954 (1980).
43. W. R. Cullen, A. Fenster, and B. R. James, Inorg. Nuclear Chern. Letters 10, 167-170
(1974).
44. R. G. Ball, B. R. James, D. Mahajan, and J. Trotter, Inorg. Chern. 20, 254-261 (1981).
45. B. R. James and D. Mahajan, Can ad. J. Chern. 57, 180-187 (1979).
46. J. C. Poulin, T. P. Dang, and H. B. Kagan, J. Organornetal Chern. 84, 87-92 (1975).
47. J. M. Brown and D. Parker, J. C. S., Chern. Cornrn. 1980,342-344.
48. M. Tanaka and I. Ogata, J. C. S., Chern. Cornrn. 175, 735.
49. M. Tanaka, T. Hayashi, and I. Ogata, Bull. Chern. Soc. Japan 50,2351-2357 (1977).
50. L. H. Pignolet, D. H. Doughty, S. C. Nowicki, M. P. Anderson, and A. L. Casalnuovo,
J. Organornetal. Chern. 202, 211-223 (1980).
51. a) B. R. James, D. K. W. Wang, and R. F. Voigt, J. C. S., Chern. Cornrn. 1975, 574-575;
b) B. R. James, R. S. McMillan, R. H. Morris, and D. K. W. Wang, Adv. Chern. Series
167,122-135 (1978).
52. M. M. Taqui Khan, M. Ahmed, and B. Swamy, Indian J. Chern. 20A, 359-362 (1981).
53. J. Halpern, Inorg. Chirn. Acta 50, 11-19 (1981).
54. J. Halpern, T. Okamoto, A. Zakhariev, J. Molecular Catal. 2,65-68 (1976).
55. M. H. J. M. de Croon, P. F. M. T. van Nisselrooij, H. J. A. M. Kuipers, and J. W. E.
Coenen, J. Molecular Catat. 4, 325-335 (1978).
56. S. Siegel, J. Catal. 30, 139-145 (1973).
57. J. Kwiatek, Catal. Rev. 1, 37-72 (1968).
58. J. Halpern and L. Y. Wong, J. Arner. Chern. Soc. 90,6665-6669 (1968).
59. a) A. Miyake and H. Kondo, Angew. Chern. Internat. Ed. Eng. 7, 631-632 (1968); b) A.
Miyake and H. Kondo, Angew. Chern. Internat. Ed. Eng. 7, 880-881 (1968).
60. J. Norton, Accounts Chern. Res. 12,139-145 (1979).
61. M. F. Lappert, P. P. Power, A. R. Sanger, and R. C. Srivastava, Metal and Metalloid
Arnides (Wiley, New York, and Ellis Horwood, Chichester, 1980), Chapt. 12.
62. L. H. Slaugh and R. D. Mullineux, J. Organornetal. Chern. 13, 469-477 (1968).
63. K. Murata and A. Matsuda, Bull. Chern. Soc. Japan 53,214-218 (1980).
64. T. Ikariya and A. Yamamoto, J. Organornetal. Chern. 116, 231-237 (1976).
65. R. A. Sanchez-Delgado, J. S. Bradley, and G. Wilkinson, J. C. S., Dalton Trans. 1976,
399--404.
66. Y. Kawabata, T. Hayashi, and I. Ogata, J. C. S., Chern. Cornrn. 1979, 462--463.
67. C. Salomon, G. Consiglio, C. Botteghi, and P. Pino, Chirnia 27,215-217 (1973).
68. B. R. James and D. Mahajan, Canad. J. Chern. 58,996-1004 (1980).
69. G. Consiglio, C. Botteghi, C. Salomon, and P. Pino, Angew. Chern. Internat. Ed. Eng.
12,669-670 (1973).
70. A. R. Sanger, Preprints, Fifth Canadian Syrnposiurn on Catalysis (Calgary, 1977), pp.
281-287.
71. A. R. Sanger, J. Molecular Catal. 3, 221-226 (1977-1978).
72. A. R. Sanger and L. R. Schallig, J. Molecular Catal. 3,101-109 (1977-1978).
73. O. R. Hughes, Hydroformylation Catalyst, United States Patent No.4 201728 (1980).
74. O. R. Hughes and J. D. Unruh, J. Molecular Catal. 12,71-83 (1981).
75. D. Evans, G. Yagupsky, and G. Wilkinson, J. Chern. Soc. (A) 1968,2660-2665.
76. G. Yagupsky, C. K. Brown, and G. Wilkinson,!. Chern. Soc. (A) 1970,1392-1401.
HYDROGENA TION AND HYDROFORMYLA TlON REACTIONS 237

77. C. K. Brown and G. Wilkinson, I. Chem. Soc. (A) 1970,2753-2764.


78. D. E. Morris and H. B. Tinker, Chem. Tech. 2, 554-559 (1972).
79. R. F. Heck and D. S. Breslow, I. Amer. Chem. Soc. 83, 4023-4027 (1961).
80. M. van Boven, N. H. Alemdaroglu, and J. M. L. Penninger, Ind. Eng. Chem., Prod. Res.
Dev. 14, 259-264 (1975).
81. N. H. Alemdaroglu, J. L. M. Penninger, and E. Oltay, Monatsh. Chem. 107,1153-1165
(1976).
82. R. Whyman, I. Organometal. Chem. 66, C23-C25 (1974).
83. J. P. Collman and L. S. Hegedus, Principles and Applications of Organctransition Metal
Chemistry (University Science Books, Mill Valley, CA, 1980), pp. 424-426.
84. J. Halpern, Pure Appl. Chem. 51, 2171-2182 (1979).
85. D. J. Thornhill and A. R. Manning, I. C. 5., Dalton Trans. 1973,2086-2090.
86. J. P. Collman and L. S. Hegedus, Principles and Applications of Organotransition Metal
Chemistry (University Science Books, Mill Valley, CA, 1980), pp. 423, 429.
87. E. L. Muetterties, Science 196,839-848 (1977).
88. A. K. Smith and J. M. Basset, I. Molecular Catal. 2,229-241 (1977).
89. E. L. Muetterties, T. N. Rhodin, E. Band, C. F. Brucker, and W. R. Pretzer, Chem. Rev.
79,91-137 (1979).
90. a) H. Storch, N. Golumbic, and R. Anderson, The Fischer- Tropsch and Related Syntheses
(John Wiley and Sons, New York, 1951); b) C. N. Satterfield, Heterogeneous Catalysis
in Practice (McGraw-Hill, New York, 1980).
91. S. O. Grim, P. H. Smith, I. J. Colquhoun, and W. McFarlane, Inorg. Chem.19, 3195-3198
(1980).
92. J. P. Farr, M. M. Olmstead, C. H. Hunt, and A. L. Balch, Inorg. Chem. 20, 1182-1187
(1981).
93. R. G. Nuzzo, D. Feitier, and G. M. Whitesides, I. Amer. Chem. Soc. 101, 3683-3685
(1979).
94. X. Guo and A. R. Sanger, unpublished results.
95. J. T. Mague, unpublished results.
96. J. D. Unruh and J. R. Christenson, I. Molecular Catal. 14,19-34 (1982).
97. M. J. Nappa, R. Santi, S. P. Diefenbach, and J. Halpern, I. Amer. Chem. Soc. 104,
619-621 (1982).
7
Functionalized Tertiary
Phosphines and Related
Ligands in Organometallic
Coordination Chemistry
and Catalysis
Thomas B. Rauchfuss

1. INTRODUCTION

The prospect of predictably manipulating the chemo- and stereoselectivity


of metal catalysts is both aesthetically and economically attractive. The
most straightforward way to affect the chemical behavior of a given metal
ion is through changes in its ligands. It is well recognized that changes
within the first coordination sphere of a metal ion can have a relatively
dramatic impact on the properties of the entire complex. On the other
hand, modifications of the ligand superstructure and substituents have
subtler and somewhat more predictable effects. Such ligand modifications
as applied to the coordination chemistry of tertiary phosphines have begun
to receive increasing attention because of the applicability of this technology
to catalysis. In this chapter, I discuss several studies which have concerned
the coordination chemistry of tertiary phosphine ligands which bear chemi-

Dr. Thomas B. Rauchfuss • School of Chemical Sciences, University of Illinois, Urbana-


Champaign, Illinois, 61801.

239
240 THOMAS B. RAUCHFUSS

cally active substituents. This review is not comprehensive, rather the ligand
systems discussed were selected because of the novel properties they impart
to their complexes, the usefulness of their complexes in mechanistic or
catalytic studies, or their synthetic versatility. Areas related to the theme
of this review which were intentionally de-emphasized are di- and polyphos-
phines, chiral phosphines, and polymer-bound phosphines. These topics
are covered more fully in other chapters of this book.

2. ETHER PHOSPHINES

The first reported application of a functionalized phosphine resulted


from work done at Monsanto on the use of chiral methoxyaryl phosphines
for asymmetric hydrogenation. It was found that the rhodium(I) complexes
of the resolved forms of cyclohexylanisylmethylphosphine, P(C6 H ll )-
(0- C 6 H 4 0CH 3 )CH 3 , catalyze the hydrogenation of acylaminoacrylic esters
with good enantioselectivity.(1,2) Subsequent refinements by the same group
culminated in the application of a system based on 1,2-R,R-his(phenyl-
anisylphosphino )ethane whose synthesis (Figure 7.1) is itself a significant
achievement. (3) In the cases of both of these chiral phosphine ligands, the
presence of the methoxy group is crucial to the performance of the catalyst
systems since its replacement by sterically comparable substituents
results in substantially decreased enantioselectivity. (2) There exist literature
precedents which demonstrate that an ether group in similar ligands can
either chelate to the metal center or interact with the polar catalytic
substrates.

OH
1 3 Steps
Ph-P(OICH 3 ~

(R, R-isomerl

Figure 7.1. The synthesis of 1, 2-R, R-bis(anisylphenylphosphino)ethane.


FUNCTIONAL/ZED TERTIARY PHOSPHINES 241

, ,
©t
:' :'
p
-
OCH 3
rr
©r:rr ~-
CH3
Figure 7.2. Proposed interaction of methoxy group with iridium in the reaction of methyl
iodide with trans-IrCl(CO)(o-Me2PC6H40CH3h·

Shaw and co-workers have described the influence of 0- anisylphos-


phines vis-a-vis p- anisylphosphines on the reactivity of compounds of the
type trans- IrCI(CO)(PMe2C6H40CH3h. (4) It was found that these
isomeric complexes oxidatively add methyl iodide at very different rates,
the o-methoxy derivative reacting ca. 500 times faster than the p-isomer,
to give the iridium(III) compounds IrCI(CH3)I(CO)(PMe2C6H40CH3h.
Since the donor strengths (at phosphorus) for these isomeric ligands are
quite similar, it follows that the equilibrium basicities of two iridium(I)
complexes are also similar. These findings led these workers to invoke a
neighboring-group effect, i.e., anchimeric assistance, to explain the unusual
rate enhancement observed for the 0- anisylphosphine complex. The
methoxy group is proposed to function as a donor ligand which enhances
the nucleophilicity of the iridium (I) center and hence the rate of the
oxidative addition (Figure 7.2). The methoxy group is in effect a kind of
internal catalyst since its influence is almost purely kinetic. This mechanistic
rationale would appear to receive support from recent studies which indicate
that the rates of oxidative addition of polar substrates to square planar
iridium(I) are enhanced by the binding of anions to the 16e - metal center. (5)
The 0- anisylphosphines can simulate this situation by coordination of the
oxygen donor.
X-ray crystallography has established that the ether oxygens are coor-
dinated in the complex RuCh(POMeh,(6) where POMe is o-diphenylphos-
phinoanisole (Figure 7.3). This remarkable compound is easily prepared
in high yield and can be isolated as red, air-stable crystals. The structural
study revealed very long Ru··· 0 distances (2.28 ± 0.03 A), while the Ru··· P
distances were found to be short and similar to those found in five-coordinate
ruthenium(II) phosphine complexes. Consistent with the structural results,
the ether donors are easily displaced by good 1T'- acceptor ligands.
RuCh(POMeh represents a very unusual example of a complex which
reacts readily with carbon monoxide, but is completely resistant to oxygen.
This observation is taken to be significant since it demonstrates that this
unsymmetrical chelate ligand protects the metal from oxidation while
permitting softer substrates to coordinate. Because of the lability associated
with one of its two donors, o-diphenylphosphinoanisole has been described
as a hemilabile chelating agent. Such ligands are of interest since they
242 THOMAS B. RAUCHFUSS

Figure 7.3. ORTEP plot of RuCl2(o-Ph2PC6H40CH3h. (Reprinted with permission from


Inorganic Chemistry.)

represent a means of stabilizing incipiently coordinatively unsaturated


complexes. In this regard, it is important to note that although
RuClz(POMeh is not itself catalytically active, it reacts quickly with basic
ethanol to afford a yellow, highly active, olefin hydrogenation-isomerization
catalyst. (7)
1,5-Bis(diphenylphosphino)-3-oxopentane (POP) is a potentially
tridentate p ... 0··· P chelating ligand. This coordination mode has
been established crystallographically in [Rh(POP)CO]PF6 , although no
reactivity studies have been reported for the complex. (8) Also described
was 1, 11-his (diphenylphosphino )-3,6,9-trioxoundecane(POOOP) and
[Rh(POOOP·H 2 0)CO]PF6 • In this complex, the polyetherdiphosphine not
only chelates to the rhodium (I) moiety but binds water through hydrogen
bonding to the oxygen donors. (8) It is intriguing that the structures of the
two rhodium complexes are very similar insofar as both contain trans
phosphines adjacent to carbonyl and oxygen donors, the latter being derived
from an ether or water (Figure 7.4). It is conceivable that compounds of
FUNCTIONAL/ZED TERTIARY PHOSPHINES 243

~5

Figure 7.4. Molecular structures of [Rh(Ph2PCH2CH20CH2CH2PPh2l-(COlt and


[Rh(Ph2PCH2CH20CH2CH20CH2CH20CH2CH2PPh2·H20)-(Colt. (Reprinted with
permission from I. C. S., Dalton Trans.)

this type could be useful for reactions which involve hydration or are
influenced by inter-ligand hydrogen-bonding effects.
A recent report has described how the ether component of polyether-
phosphinite ligands can be used to enhance the reactivity of coordinated
carbon monoxide. The polyether diphosphinite complex shown in Figure
7.5 resembles a crown ether, which in turn activates a molybdenum-bound
carbonyl towards nucleophilic attack by organolithium compounds. (9) The
acyl-lithium interaction of the product is stabilized by complexation of the
lithium ion by the polyether chelating agent. The profound influence of
Lewis acid-acyl interactions on the facility of the migratory insertion
reaction has been previously well-established.(lO) In the case of the molyb-
denum diphosphonite, the effect is clearly a result of the ligand substituents,
244 THOMAS B. RAUCHFUSS

2 M=Cr.Mo.W
R = Me. Ph. tsu. EtlN

Figure 7.5. Ligand assisted activation of CO in a polyetherdiphosphinite carbonyl complex.


(Reprinted with permission from I. Am. Chern. Soc.)

since the control compound, cis-Mo(PPh 2 0CH3 h(CO)4, does not react
with the same organolithium reagents under mild conditions. These studies
establish how remote, but chemically active substituents, can favorably
influence the reactivity of the central metal ionYl)

3. AMINOPHOSPHINES

In view of the extensive coordination chemistry of polyamine chelating


agents, it is not too surprising that the aminophosphine ligands have become
relatively common. Of particular interest are those chelating agents that
contain primary or secondary amines whose presence permits further
functionalization or promotes metalation.
Bis (2- diphenylphosphinoethyl)amine can be prepared on a large scale
from sodium diphenylphosphide and the nitrogen mustard, bis (2-
chloroethyi)amine. (12) The secondary amine function can be acetylated
readily, thus providing a reliable means for attaching the diphosphine
moiety to a number of molecular and polymeric substrates. These deriva-
tives can then be combined with labile rhodium(I) precursors to afford
modified hydrogenation catalysts. One innovative application of this
methodology involved the preparation of a protein-bound hydrogenation
catalyst. In particular, the P-N-P unit was linked to biotin via amide
formation (Figure 7.6). (13) Biotin itself is known to be very tightly bound
to the globular protein avidin, and it was found that the rhodium(I) complex
of the biotin-derivatized phosphine was also bound by the same protein.
The rhodium-protein complex functions efficiently as an asymmetric hydro-
genation catalyst, the chirality being transmitted from the chiral protein
environment to the coordination sphere of the rhodium ion. While the
actual enantioselectivities observed for this particular catalyst are modest
FUNCTIONAL/ZED TERTIARY PHOSPHINES 245

ft
.r-;-PPh2 +
R-CN + Rh ( diene 1 52 - - -
~PPh2

o Ph2 Ph2
II r - P,- + r--P +
RCN~ ~ ,Rh(dienel + AVIDIN _ AVIDIN---BIOTIN-N ~h(dienel
....---p ~P
Ph ~ 1),-CH2 - Ph2
(R=O=< 1
~ A
Figure 7.6. Synthetic sequence for the preparation of a protein bound rhodium hydrogenation
catalyst.

by 1982 standards, this work demonstrates that appropriately function-


alized phosphines can effect the marriage between organometallic and
biological homogeneous catalysis.
The aminophosphines depicted in Figure 7.7 have been described. (14.15)
Both undergo facile deprotonation in the presence of metal ions to yield
anionic P-N or P-N-P chelating agents. Preliminary results are that these
hybrid ligands may induce unusual reactivity at the bound metal ion.(14)
Phosphoranes have attracted attention as ligands for reactive metal
complexes.(16) Riess and coworkers have focused on the distinctive coordi-
nation chemistry of the bicyclophosphorane Ph(H)P(OCH 2 CH 2 hN(phoran)
which can, in principle, exist in different tautomeric forms (Figure 7.8).(17)
The novelty associated with this ligand system concerns its ability to

Figure 7.7. 1, 3-Bis [(diphenylphosphino )methyl]tetramethyldisilazane and 0- diphenylphos-


phinoaniline.
246 THOMAS B. RAUCHFUSS

9~
Ph- P. NH
0--.1

Figure 7.8. The phosphorane-phosphonite tautomerism in phoran.

generate coordinative unsaturation via this tautomerization within the


coordination sphere of a metal complex. In this way, phoran represents a
particularly innovative version of a hemilabile chelating agent. In the
complexes Rh(phoranhCI and RhCI(CO)(phoran), the rhodium ions li~ in
the cradle of the cyclic phosphine-amine chelating agent.(1S) These com-
plexes are simply prepared from the phosphorane and Rh 2ClzL4 (L = C2H 4 ,
CO) via reactions which illustrate the facility of the ring-chain tautomeri-
zation. The lability of the nitrogen donor in the carbonyl complex may
explain its reactivity as a hydroformylation catalyst. (19) The enhanced
reactivity of other metal complexes of P-N chelates has been previously
observed by Roundhill. (20)
The cationic complex CpMo(COh (phorant can be reversibly deprot-
onated with MeLi to generate the 7] 2 - P-N and 7] 2- P-O phosphoranide
linkage isomers.(21) Different chemistry has been observed for CpFe(CO)'
(phoranr where deprotonation of the amine leads to migration of the
phenyl group from P to Fe concomitant with the formation of a caged-type
phosphite(22) (Figure 7.9). The phenyl migration from P to Fe is all the

Figure 7.9. Rearrangement pathways observed for organometallic derivatives of phoran.


FUNCTIONAL/ZED TERTIARY PHOSPHINES 247

more remarkable since it is reversible: treatment of the Fe-phenyl deriva-


tive with acid promotes the migration (migratory insertion?) of the phenyl
group back to phosphorus concomitant with reformation of the bidentate
P-N chelate. These interesting interconversions are beautiful illustrations
of the novel chemical behavior peculiar to 1,5-diheterocyclooctanes, a
theme of recurring interest to chemists. (23)

4. CARBONYLPHOSPHINES

The synthetic versatility of the carbonyl functional group virtually


guarantees that the carbonylphosphines will become a heavily exploited
class of ligands. This view is strengthened by the recognition of the key
role that organic carbonyls play in classical coordination chemistry, par-
ticularly in the formation of imino- and enolate-type ligands. (24)
The carbonylphosphine particularly worthy of attention is 0-
diphenylphosphinobenzoic acid. (25-27) This compound is conveniently pre-
pared from sodium diphenylphosphide and the salts of o-chlorobenzoic
acid. Workers at Shell have shown that the combination of this ligand with
bis (1,5 - cyclooctadiene )nickel(O) affords a very active catalyst for ethylene
oligomerization. (26.28) This technology is related to the highly successful
Shell higher olefin process (S.H.O.P.) which is currently used for preparation
of detergent-range a- olefins. (29) Understandably, detailed discussions of
this catalytic process have not been published by the Shell workers,
although insight is provided from other sources.(311) In view of their applica-
bility and their ease of synthesis, it is surprising that the 0- diorganophos-
phinobenzoates have not received greater attention.
Keirn and co-workers have found that the treatment of Ni(CODh with
both triphenylphosphine and the phosphorane, Ph 3 PCHC(O)Ph, affords a
nickel(II) chelate complex formally derived from the enolate of
Ph 2 PCH 2 C(O)Ph (Figure 7.10).(30) This crystalline compound, which can
be conveniently prepared on a large scale, has been characterized by single
crystal X-ray diffraction. Much like 0- diphenylphosphinobenzoate, the
novel enolate ligand functions as an anionic P-O donor. What is particularly
intriguing is that its nickel complex also efficiently catalyzes the formation
of linear a- ole fins from ethylene.

o Ph Ph2
+ - II
Ni (CODI 2 + PPh 3 + Ph3 P - CHC Ph -2 COD 'frH
• N( l
p~pl O' Ph

Figure 7.10. Synthesis of a phenylnickel(II) phosphinoenolate from nickel(O) and a phos-


phorane.
248 THOMAS B. RAUCHFUSS

L..... P1H
Ph2

"Pd~:: l£.. o
L ....Pd-
+ C02 ~ L...... , .'.,'..
Ph2 OR

H
L 0 OR o 0""
Figure 7.11. Reversible CO 2 fixation by a palladium(II) phosphinoenolate complex.

A thorough study of palladium(II) complexes derived from the enolate


of ethyl-2-diphenylphosphinoacetate, Ph2PCH2C02C2H5, has been pub-
lished.(31) In many of these complexes, the P-O chelates playa relatively
passive role not unlike that found for 0- diphenylphosphinophenoxides. (32)
The derivatives shown in Figure 7.11, however, distinguish themselves by
reversibly binding CO 2. The CO 2 is captured via addition to the enol ate
carbon and, once attached, its binding is stabilized by intramolecular
hydrogen bonding from the pendant functional groups. This finding is
important because of the topical interest in using CO 2 as a possible carbon
feedstock.
Reaction of methyl 0- diphenylphosphinobenzoate, 0- Ph 2PC6 H 4-
C0 2CH 3, with the potassium enolate of pinacolone yields the phosphine-
diketone ligand shown in Figure 7.12. This chelating agent, called HacacP,
is representative of what is known as a compartmentalized ligand, i.e., one
which features two or more geometrically and often chemically distinct
metal-binding sites.(33) In the case of HacacP, these sites are provided by
the 0 .. ·0 or "acac" chelate and the phosphine moiety. A novel
homobimetallic coordination compound that can be made from HacacP
and copper(I) is [Cu(acacP)h.(34) This complex possesses an unusual struc-
ture wherein the two coordination spheres are planar and mutually cofacial.
This species is of particular interest, since not only are the two metal ions
adjacent but they are also coordinatively unsaturated.
The use of HacacP for the assembly of heterobimetallic complexes
takes full advantage of the compartmentalization of this novel chelating
agent. Representative of the synthetic potential of HacacP is the three-step
synthesis of [Pt{Cu(acacP)}](BF4h (Figure 7.13). The X-ray structure of
the intermediate, PtCIz{Cu(acacPh} (Figure 7.14), reveals that the
platinum and copper entities are square planar and bound to the portion
of acacP- which best suits their soft or hard acid character. (35) The presence
of an adjacent, paramagnetic, Lewis acidic Cu(acach moiety could lead
H
Ph P 0·-· ....0

~~I
Figure 7.12. Synthesis of o-diphenylphosphinobenzoylpinacolone (HacacP), 0-
Ph2PC6H4C(O)CH2(OlC(CH3l3'
FUNCTIONAL/ZED TERTIARY PHOSPHINES 249

+2 HococP
-COD
..

-2AgCI

Figure 7.13. Preparation of [Pt{Cu(acacPh}](BF4 h.

Figure 7.14. The molecular structure of PtCI 2{Cu(acacPh}.


250 THOMAS B. RAUCHFUSS

®
o CHO
Br
-0 ®:? Br
---0 @t.? ®CHO
-0
n ER 3 _n n ER3 _n

E=P, R=Ph, n=I-3


E=P, R=Me,n=1
E=As, R=Ph,n=1

Figure 7.15. The synthetic sequence for the preparation of 0- diphenylphosphinobenzal-


dehyde.

to some interesting perturbations on the chemistry of the platinum center.


Another interesting heterobimetallic derivative of HacacP is trans-
IrCI(CO){Cu(acacPh}, wherein a Cu(acac)z moiety is strapped across
the face of a 16e - iridium(I) center. That species is particularly novel since
the iridium(I) component possesses the ability to undergo facile oxidative
addition of small electrophilic substrates. (35)
0- Diphenylphosphinobenzaldehyde, or PCHO, can be prepared from
0- bromobenzaldehyde (Figure 7.15) and exists as yellow, air-stable crys-
tals. (27,36) Like diphenylphosphinostyrene, PCHO has proven to be a useful
mechanistic probe in organotransition metal chemistry. (37-39) The synthetic
versatility characteristic of the formyl group suggests that PCHO should
represent a useful precursor to a range of new ligand systems. For instance,
PCHO condenses easily with a wide variety of aryl and alkyl, mono- and
diamines to afford a number of iminophosphine chelating agents (Figure
7.16). (40-42) By varying the steric and electronic characteristics of the

o ~
C§X
CHO
+RNH 2 -OC§XPPh2
CH=NR

monoamines diamines

C~O@NH2 ~-@-N~
~NH2
~N0NHZ

~ NHZ
C~ ./"..,/NHZ \.J
®,HZ
HzU NHz

~HZ HzN NH2


.~ )--I
Figure 7.16. Iminophosphines prepared from o-diphenylphosphinobenzaldehyde.
FUNCTIONAL/ZED TERTIARY PHOSPHINES 251

amines, it is possible to modify the properties of the iminophosphine


complexes. The resultant iminophosphines form stable, 6-membered P-N
chelate rings. The ligand derived from R,R-1,2-diaminocyclohexane and
PCHO coordinates stereospecifically to form a chiral tetrahedral copper(I)
compound which is geometrically incapable of interconverting between the
.l and A forms.(41) The iminophosphines can be easily reduced with NaBH4
to give the corresponding secondary amine ligands. Upon coordination the
amine group becomes configurationally stable, and its chirality could, in
principle, be exploited for asymmetric synthesis. A number of formyl- and
ketophosphine ligands like PCHO have been described although few have
received any study as ligands. (37)

5. ALKENYLPHOSPHINES

Alkenylphosphine ligands, particularly diphenylphosphinostyrene and


the butenylphosphines, have received considerable attention as
organometallic chelating agents. Studies involving the reactions of the
olefinic component of such ligands have provided a wealth of information
relating to metal-olefin reactivity patterns.(43) Those studies have focused
on how metal ions can affect the reactivity of the olefinic group. In the
present case, we are more concerned with the reverse situation: how the
presence of unsaturation in a phosphine ligand can affect the chemistry of
the metal ion.
Fenske and co-workers have described a new class of 1,2-diphos-
phinoalkenes which are unique for their radical anion-forming abilities.
This property has led to some very novel chemistry and spectroscopy. The
ligands of interest are prepared from the reaction of 2,3-dichloromaleic
anhydride (and related derivatives) with trimethylsilylated phosphines(44)
(Figure 7.17). The free ligands exist as air-stable, crystalline solids. With
classical, electrophilic metal salts, they form "normal" complexes, for
instance of the type NiX 2 (chelate). (44) With electron-rich metal ions,
however, these chelating agents react via a curious version of the oxidative

o
:~CI +2 M~Si PPh2 - - - X~I
PPh2
+ 2 Me3SiCI
~CI PPh2
o o
x =0, NCH 3

Figure 7.17. The synthesis of 1, 2-bis(diphenylphosphino)-maleic anhydride and -N-methyl-


maleimide.
252 THOMAS B. RAUCHFUSS

o
PPh2 ~~h2
,. ?
X~I - LnM,P I-~. .
+,
~
+ CO
PPh 2
o Ph2 6

Figure 7.18. The cleavage of metal-metal bonds with 1, 2-bis(diphenylphosphino)-N-methyl-


maleimide.

addition reaction. For example, two equivalents of 2,3-bis(diphenylphos-


phino)-N-methylmaleimide displace all four PPh 3 ligands from yellow
Pd(PPh 3 )4 to form the intensely green, paramagnetic complex, Pd(che-
lateh.(4S) Electron spin resonance studies indicate the presence of two
radical anion ligands attached to a palladium(II) center. A single crystal
X-ray diffraction study supports this assignment by revealing a centrosym-
metric, square planar palladium complex. Other spectroscopic measure-
ments, especially JR, reinforce the view that these ligands function as
one-electron oxidizing agents.
A reaction which is characteristic of Fenske's alkenyldiphosphines is
the oxidative cleavage of binuclear metal carbonyls. (46.47) The monometallic
products are best described as zwitterions containing radical anionic diphos-
phine chelating ligands coordinated to cationic (18e -) metal centers (Figure
7.18). These compounds can be subsequently converted to diamagnetic
cations via ligand-based oxidations. The ability of ligands to undergo facile
reduction/oxidation is consistent with the presence of vacant, low-energy
molecular orbitals. Related behavior can be found in the chemistry of
metalloporphyrins and the dithiolenes. Since many fundamental
organometallic reactions appear to proceed via odd-electron intermedi-
ates,(48) complexes which promote or stabilize the formation of paramag-
netic species may be expected to display distinctive chemistry. A demonstra-
tion of such ligand-based reactivity is provided by the ability of the
(CsHs)Mo(COh derivative of Fenske's diphosphine to abstract a hydrogen
atom from toluene. The products of this reaction are the diamagnetic
molybdenum cation, CsHsMo(COhCdiphosphineH)"'", and biphenyl(46)
(Figure 7.19).

q PIIz 0 Ph2
~~.J\ + .~, + I
CH3~~iMO(COl2(C5H5l + PhCH3-CH:fI~~~MO( COl2 (CsH5l + 2" PhcHzCHzPh
o Ph2 0 Ph2

Figure 7.19. Hydrogen atom abstraction from toluene by a 1, 2-bis-(diphenylphosphino)-N-


methyl maleimide complex.
FUNCTIONAL/ZED TERTIARY PHOSPHINES 253

@r
,
PPh2
,
Fe
~PPh2
CH- N1CH3)2
I
CH 3
Figure 7.20. Kumada's chiral ferrocenyldiphosphine ligand.

6. CYCLOPENTADIENYLPHOSPHINES

Cyclopentadienylphosphines have received considerable attention as


they represent the fusion of two very popular ligand types. Depending on
the nature of the linkage between the cyclopentadienyl moiety and the
phosphine, such systems can either function as net 7 e - chelating agents or
as binucleating ligands. In the former capacity, it is not exactly clear what
new chemistry will ensue from these monometallic complexes, although
they are stereochemically unusual. (49)
Considerable effort has been expended on using cyclopentadienylphos-
phines as templates for the assembly of bimetallic complexes. The central
problem here is that effective strategies have yet to be devised for connecting
two different reactive metals. For this reason, one metallic component of
cyclopentadienylphosphine complexes is generally chemically dormant.
This gives rise to the all too familiar situation of structurally novel com-
pounds that are chemically inert. Clearly the challenge of the coming years
will be the preparation of reactive heterobimetallic complexes which are
constructed in such a way that both metals can interact in a concerted
manner with the substrates.
Ferrocenylphosphines have become common, and while the iron of
the ferrocene moiety is relatively inert, these ligands are stereochemically
novel. The rhodium(I) derivative of the ferrocenyldiphosphine shown in
Figure 7.20 has been shown to function as a highly selective asymmetric
hydrogenation catalyst. (50)
Perhaps the most versatile approach to substituted cyclopentadienyl
phosphines has been described in a series of reports by Kauffmann et al.
concerning the chemistry of a spiro-cycloheptadiene. (51.52) This derivative,
which apparently can be easily prepared from cyclopentadiene and ethy-
lenedibromide, reacts with a range of lithiated ligand precursors. This route
provides access to a wide range of relatively complex hybrid ligands.

O<J +8PR 2 -. ~
PR 2
254 THOMAS B. RAUCHFUSS

Figure 7.21. Schore's iron(O)-zirconium(IV) complex prepared using a phosphinocyclopen-


tadiene ligand system.

The preparation of a phosphinocycIopentadiene containing an intervening


silane bridge has been reported by Schore, (53) This wQrk has involved not
only the preparation of the usual ferrocenyl derivatives and their subsequent
complexation, but a novel zirconium(IV)-iron(O) complex (Figure 7,21).
Since cycIopentadienyl zirconium halides represent precursors to useful
synthetic reagents, this work represents a significant step towards the
assembly of heterobimetallic complexes which contain two reactive metal
ions.

7. CONCLUDING REMARKS

In the coming years one can expect to see greater use of functionalized
phosphines in synthesis and catalysis. One reason for this trend is the
increasing infusion of organic synthetic methodology into the field of
organometallic coordination chemistry. t Furthermore, the recent applica-
tions of homogeneous platinum metal catalysts justifies the modest expense
associated with the use of somewhat elaborate phosphines. In this regard,
it is important to remind ourselves that aside from considerations of
convenience, there is no reason for assuming that the perennial favorite,
triphenylphosphine, is the optimum phosphine ligand for any catalytic
application.

ACKNOWLEDGMENTS

Our own work on phosphines has been funded by the Petroleum


Research Fund and by Dow Chemical Company. I thank John Hoots, Edith
Landvatter, and Debra Wrobleski for their assistance in the preparation
of this review.

t For an impressive illustration of this trend, see Reference 54.


FUNCTIONAL/ZED TERTIARY PHOSPHINES 255

REFERENCES

1. W. S. Knowles and M. J. Sabacky, Ger. Offen. 2, 123,063 (1971).


2. W. S. Knowles, M. J. Sabacky, and B. D. Vineyard, Adv. Chem. Ser.132, 274-282 (1972).
3. B. D. Vineyard, W. S. Knowles, M. J. Sabacky, G. C. Bachman, and D. J. Weinkauff,
1. Am. Chem. Soc. 99, 5946-5952 (1977).
4. E. M. Miller and B. L. Shaw, 1. C. S., Dalton Trans., 480-485 (1974).
5. W. J. Louw, D. J. A. de Waal, and J. E. Chapman, 1. C. S., Chem. Comm., 845-846
(1977).
6. J. C. Jeffery and T. B. Rauchfuss, Inorg. Chem. 18, 2658-2666 (1979).
7. T. B. Rauchfuss, Plat. Met. Rev. 24, 95-99 (1980).
8. N. W. Alcock, J. M. Brown, and J. C. Jeffery, 1. C. S., Dalton Trans. 583-588 (1976).
9. J. Powell, A. Kuksis, C. J. May, S. C. Nyberg, and S. J. Smith, 1. Am. Chem. Soc. 103,
5941-5943 (1981).
10. J. P. Collman, R. G. Finke, J. N. Cawse, and J. I. Brauman, 1. Am. Chem. Soc. 100,
4766-4772 (1978).
11. For an example of true crown ether-phosphine, see E. M. Hyde, B. L. Shaw, and I.
Shepherd,l. C. S., Dalton Trans., 1693-1705 (1978).
12. M. E. Wilson, R. G. Nuzzo, and G. M. Whitesides, 1. Am. Chem. Soc. 100, 2269-2270
(1978).
13. M. E. Wilson and G. M. Whitesides, 1. Am. Chem. Soc. 100, 306-307 (1978).
14. M. K. Cooper and J. M. Downes, Inorg. Chem. 17,880-884 (1978); 1. C. S., Chem.
Comm., 381 (1981).
15. M. D. Fryzuk and P. A. MacNeil, 1. Am. Chem. Soc. 103,3592-3593 (1981).
16. W. Keirn, R. Appel, A. Storeck, C. Kruger, and R. Goddard, Angew. Chem. Int. Ed.
Eng. 20, 116-117 (1981).
17. D. Houalla, J. F. Brazier, J. Sanchez, and R. Wolf, Tetrahedron Lett., 2969-2970 (1972).
18. C. Pradat, J. G. Riess, D. Bondoux, B. F. Mentzen, I. Tkatchenko, and D. Houalla, 1.
Am. Chem. Soc. 101, 2234-2235 (1979).
D. Bondoux, B. F. Mentzen, and I. Tkatchenko, Inorg. Chem. 20, 839-848 (1981).
19. D. Bondoux, D. Houalla, C. Pradat, J. G. Riess, I. Tkatchenko, and R. Wolf, Funda-
mental Research in Homogeneous Catalysis, edited by M. Tsutsui (Pergamon Press, New
York, 1979), Vol. 3, pp. 969-981.
20. D. M. Roundhill, R. A. Bechtold, and S. G. Roundhill, Inorg. Chem. 19,284-289 (1980).
21. F. Jeanneaux, A. Grand, and J. G. Riess, 1. Am. Chem. Soc. 103,4272-4273 (1981).
22. P. Vierling, J. G. Riess, and A. Grand, 1. Am. Chem. Soc. 103, 2466-2467 (1981).
23. W. K. Musker, Acc. Chem. Res. 13,200-208 (1980).
24. D. SI. C. Black and A. J. Hartshorn, Coord. Chem. Rev. 9, 219-274 (1972-1973).
25. K. Issleib and H. Zimmerman, Z. anorg. aUg. Chem. 353, 197-206 (1967).
26. D. M. Singleton, P. W. Glockner, and W. Keirn, British Patent 1 364 870 (1974).
27. J. E. Hoots, T. B. Rauchfuss, and D. A. Wrobleski, Inorg. Syn. 21, 175-179 (1982).
28. For related patents, see also R. F. Mason, United States Patent 3686351 (1972) and
E. F. Lutz, United States Patent 3 825 615 (1974).
29. E. R. Freitas and C. R. Gum, Chem. Eng. Progs. 75(1), 73-75 (1979).
30. W. Keirn, F. H. Kowaldt, R. Goddard, and C. Kruger, Angew. Chem. Int. Ed. Eng. 17,
466-467 (1978); M. Peuckert and W. Keirn, Organometal. 2,594-597 (1983).
31. P. Braunstein, D. Matt, Y. Dusausoy, J. Fischer, A. Mitschler, and L. Ricard, 1. Am.
Chem. Soc. 103, 5115-5125 (1981).
32. T. B. Rauchfuss, Inorg. Chem. 16,2966-2968 (1977).
33. O. Casellato, P. A. Vigato, D. E. Fenton, and M. Vidali, Chem. Soc. Rev. 8, 199-220
(1979).
256 THOMAS B. RAUCHFUSS

34. T. B. Rauchfuss, S. R. Wilson, and D. A. Wrobleski, I. Am. Chem. Soc. 103, 6769-6770
(1981).
35. D. A. Wrobleski and T. B. Rauchfuss, I. Am. Chem. Soc. 104,2314-2315 (1982).
36. G. P. Shiemenz and H. Kaack, Liebigs Ann. Chem., 1480-1495 (1973).
37. T. B. Rauchfuss, I. Am. Chem. Soc. 101, 1045-1047 (1979).
E. F. Landvatter and T. B. Rauchfuss, Organometallics 1,506 (1982).
38. T. B. Rauchfuss, Fundamental Research in Homogeneous Catalysis, edited by M. Tsutsui,
(Pergamon Press, New York, 1979), Vol. 3, pp. 1021-1032.
39. G. D. Vaughn and J. A. Gladysz, I. Am. Chem. Soc. 103, 5608-5609 (1981).
40. T. B. Rauchfuss, I. Organometal. Chem. 162, C19-C21 (1978).
41. J. E. Hoots, J. C. Jeffery, T. B. Rauchfuss, S. P. Schmidt, and P. A. Tucker, Adv. in
Chem. Ser. 196,303-311 (1982).
42. J. C. Jeffery, T. B. Rauchfuss, and P. A. Tucker, Inorg. Chem. 19, 3306 (1980).
43. M. A. Bennett, R. N. Johnson, G. B. Robertson, 1. B. Tomkins, and P. O. Whimp, I.
Am. Chem. Soc. 98, 3514-3523 (1976) and references therein.
44. D. Fenske and H. J. Becher, Chem. Ber. 108,2115-2123 (1975).
H. J. Becher, W. Bensmann, and D. Fenske, Chem. Ber. 110, 315-321 (1977).
45. W. Bensmann and D. Fenske, Angew. Chem. Int. Ed. Eng. 18,677-678 (1979).
46. D. Fenske and A. Christidis, Angew. Chem. Int. Ed. Eng. 20,129-131 (1981).
47. D. Fenske, Angew. Chem. Int. Ed. Eng. 15,381-382 (1976).
48. J. P. Collman and L. S. Hegedus, Principles and Applications of Organotransition Metal
Chemistry (University Science Books, Mill Valley, CA 1980).
49. C. Charrier and F. Mathey, I. Organometal. Chem. 170, C41-C43 (1979).
50. T. Hayashi, T. Mise, S. Mitachi, K. Yamamoto, and M. Kumada, Tetrahedron Lett.,
1133-1137 (1976).
51. Th. Kauffmann, J. Ennen, H. Lhotak, A. Rensing, F. Steinseifer, and A. Woltermann,
Angew. Chem. Int. Ed. Eng. 19,328-329 (1980).
52. K. Berghus, A. Hamsen, A. Rensing, A. Woltermann, and Th. Kauffmann, Angew.
Chern. Int. Ed. Eng. 20, 117-118 (1981).
53. N. E. Schore, I. Am. Chern. Soc. 101,7410-7412 (1979).
54. E. P. Kyba, A. M. John, S. B. Brown, C. W. Hudson, M. J. McPhaul, A. Harding, K.
Larsen, S. Niedzwiecki, and R. E. Davis, I. Am. Chern. Soc. 102,139-147 (1980).

Note Added in Proof

The following papers are related to the topics discussed in Sections


7.2 and 7.3, respectively: J. Powell, M. Gregg, A. Kuksis, and P. Meindl,
J. Am. Chern. Soc. 105, 1064-1065 (1983), and M. D. Fryzuk and P. A.
MacNeil, Organometallics 2,355-356 and 682-684 (1983).
8
Polydentate Ligands and
Their Effects on Catalysis
Devon W Meek

1. INTRODUCTION

In recent years, tremendous interest and research activity has been expen-
ded on transition metal complexes of tertiary phosphine ligands. (1,2) Metal-
phosphine complexes display a tremendous range of coordination numbers,
stereochemistries, catalytic properties, and selectivities. The effects can
also be "fine-tuned" by purposeful variation of the steric and electronic
effects of the substituent groups on phosphorus and from synthesis of
polydentate phosphines, (2)
It is expected that the use of polyphosphine ligands will become
increasingly important in future studies, especially in the areas of catalysis,
asymmetric synthesis, and organometallic stereochemistry, since the special
properties of phosphine ligands can be accentuated by a chelating polyphos-
phine.(Z,3) This chapter will concentrate on the unique aspects that can be
achieved with polyphosphine ligands, since other chapters in this book will
discuss monodentate phosphine complexes. In fact, this chapter will empha-
size tridentate ligands, since Balch and Sanger are discussing binuclear and
phosphine-bridged complexes with ligands like Ph zPCH 2 PPh 2 (Chapters 5
and 6) and Brown and Chaloner are discussing chiral diphosphine com-
plexes (Chapter 4). However, selected examples of studies with diphosphine
ligands will be presented in areas other than asymmetric catalysis and

Dr. Devon W. Meek • Department of Chemistry, The Ohio State University, Columbus,
Ohio, 43210,

257
258 DEVON W. MEEK

dinuclear complexes, where the studies show particular advantages or


implications for future work with polydentate ligands.

2. ADVANTAGES OF CHELATING POL YDENTATE LIGANDS

If one wishes to control the number of phosphine groups attached to


a metal, clearly a chelating polyphosphine ligand will provide more control
than a monodentate phosphine. (4) In addition, one can better predict the
basicity (or nucleophilicity) of the metal and the stereochemistry of a
complex resulting from a polydentate ligand. Also, slow intra- and inter-
molecular exchange processes and detailed structural and bonding informa-
tion in the form of metal-phosphorus and phosphorus-phosphorus coupling
constants can be observed, particularly when the metal has a magnetic
nucIeus, e.g., 103Rh,Pt,195 or 199 Hg. (5)
Polydentate phosphine ligands may be used to change the usual
magnetic states, coordination geometries, and reactivity of complexes
by judiciously selecting parameters such as (i) sets of donor atoms,
(ii) 'chelate bite angle," and (iii) sterically demanding substituent groups. (4-6)
For example (1) Venanzi has used the rigid ligand, 1, to form "trans-
spanning" complexes of the type, 2,(7); (2) Otsuka et at.(S) found that the

;"~-
-- - -;--f
'
Y
I
I "/,

M
",
/
I

x-------P
I " I

£
(2)=Irons- [MX2(1)]:

M= N i ; X = CI • Br. I. NCS
M= Pd; X = CI • Br • I
M = PI ; X = CI • I

bulky phosphine (t- BuhP made the crowded 14-electron species


[Pd{P(t- BuhhJ unreactive toward O 2 ; and (3) Shaw(9) has used long-chain
diphosphines of the type (t- BuhP(CH2)nP(t- Buh (n = 9, 10, 12) to produce
the mono-, di-, and tri-nuclear complexes [MCh{(t-BuhP(CH2)nP(t-
Buh}]x (M = Pd, Pt) with 12- to 45-membered rings, whereas the same
type of ligand with a (CH 2 )50r6 chain produces dimeric [M2 CI4 {(t-
BuhP(CH 2 )nP(t-BuhhJ complexes, 3, and cyclometallated complexes of
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 259

=-
40
3
M Pd, PI M= PI, Pd

the type 4a. In the case of the pentamethylene chain, the cyclometallation
reaction occurs under remarkably mild conditions. Treatment of rhodium
trichloride with the pentamethylenediphosphine (t- BuhP(CH2)5P(t- Buh
gives two principal products (the binuclear, square-pyramidal rhodium(III)
hydride 4b, with a 16-atom ring and the cyclometallated rhodium(III)
hydride 4c) and a small amount of an olefinic diphosphine rhodium(I)
complex, 4d.
t t
BU2~Bu2
HP \ H
~ ....CI
Rh
CI .......
'-
Rh
c( '\ ~\ CI
BU2
t P t
BU2

4.E 4,S
-
4d

Flexible polyphosphine ligands (e.g., 5-9), which contain either


ethylene or trimethylene linkages, have been used to form mononuclear
CH 2CH 2 PPh 2
/
P -CH 2 CH 2 PPh 2
\
CH 2CH2PPh2

,§, Iri pod L, PP 3

PhP(CH 2 CH 2 PPh 2 l 2 PhP( CH 2 CH2CH2PPh2l2


~, elp 2" lip

complexes with different coordination numbers (four, five, and six) and
different structures (distorted tetrahedral, planar, square pyramidal,
trigonal bipyramidal and octahedral)yo-12) A properly designed polyphos-
phine ligand can fix the relative positions of the phosphorus atoms while
allowing the remaining ligands to be varied; thus, the stereochemical
positions and the nature of the variable ligands can now be monitored by
31 p nmr spectroscopy by observing the chemical shifts and coupling con-
stants of the phosphorus atoms in the polyphosphine ligand. (5)
260 DEVON W. MEEK

One can modify the coordination stereochemistry for a given donor


set (e.g., P 3 X 2) around a metal ion simply by changing the number of
methylene groups in the connecting chains of a polyphosphine ligand. Such
a study with the flexible ligands 8 (etp) and 9 (ttp) permitted the effect of
the chelate ring size on the preferred structure of some 5-coordinate
cobaJt(I) complexes to be determined. (13) The details of the stereochemistry
and exchange processes for the 5-coordinate complexes [Co(COh(etp)]X,
[Co(COh(ttp)]X, [Co(CO)L(etp)]X, and [Co(CO)L(ttp)]X (X = BF4 or
PF 6 ; L = P(OMeh, PPh 2H, PPh 2 Me, PEt3, or PPh 3 ) depend on both the
monodentate ligand and on the triphosphine ligand. For example, both the
mono- and di-carbonyl complexes [Co(CO)(etp)Lt and [Co(COh(etp)r
are trigonal-bipyramidal with the equatorial positions occupied by a car-
bonylligand and the two terminal phosphorus atoms (p2) of the triphosphine

He( PR 2)3 Me 3Si (PBu2)3


Jj 1,,6
R=Me,Ph

ligand; the axial positions are occupied by the central phosphorus atom
(pI) of the triphosphine and either a carbonyl, 13, or a monophosphine
ligand (P 3 ), 14. The [Co(CO)L(ttp)t complexes have square-pyramidal
structures with L at the axial position. The change from square-pyramidal
coordination geometry in [Co(CO){P(OMeh}(ttp)t, 15, to a trigonal-
bipyramidal geometry for the etp complexes can be attributed to a decrease

14 15
'"
in chelate bite angle of the etp ligand. As a consequence of the smaller
bite angle in 5-membered chelate rings, the central phosphorus atom (pI)
either would be pulled in abnormally close to the metal in a square-
pyramidal geometry or the ligand must adopt a different coordination
geometry where the central phosphorus atom is more equivalent to those
of the terminal phosphorus atoms (p2), e.g., as in the trigonal-bipyramidal
geometries 13 and 14.
If the connecting chain between phosphorus atoms is only one atom
(e.g., dppm, 10, HC(PMe2h, and HC(PPh 2h, 11, or CH 3 Si(PPh 2h, 12),
the phosphino units tend to bridge across two metals (see Chapters 4 and
5 for dppm examples). In fact, this bridging tendency has been used to
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 261

advantage by Osborni1 4 . 1S ) and Masters (16 ) in attaching the tridentate ligands


11 and 12 to a face of a metal cluster, i.e., complexes of type 16.

!§ (b)

3. METHODS FOR SYNTHESES OF POL YDENTA TE LIGANDS

The complexes cited above illustrate that the steric and electronic
properties of ligands can be varied relatively easily by changing the organic
substituents on phosphorus, provided viable synthetic schemes are avail-
able. Until 1971, most poly tertiary phosphines were prepared by reactions
of organic polyhalides with alkali metal dialkyl- or diaryl-phosphides (e.g.,
the relatively easy preparation of Ph 2 PCH 2 CH 2 PPh 2 ) by treating
ClCH 2 CH 2 Cl with excess LiPPh 2 , which is most economically prepared by
lithium cleavage of Ph 3 P).Il7)

3.1. Grignard and Alkali-Metal Reagents

(B
Most of the tetradentate "tripod-like" ligands of the type A-C

~
that were prepared by the research groups of Venanzi,(l2) Meek,1l0) and
Sacconii1 1) in the 1960s and early 1970s utilized the Grignard or alkali-
metal organophosphide methods (Equations 1-4). However, synthesis of
more complicated tri- and tetra-phosphines by these methods is severely
262 DEVON W. MEEK

limited by the unavailability of the appropriate phosphorus or organic


polyhalide intermediates.

N(CH 2CH 2Clh + 3MPPh 2 -. N(CH 2CH 2PPh 2h + 3MCl (1)


\) Mgin Et,O
Me2AsCH2CH2CH2Cl ~ P(CH2CH2CH2AsMe2h (2)
2) PCl,

11 BuLi
21 PCl, (3)
SCH 3
(Se)

CH 3C(CH 1 Clh + 3NaPR 2 -. CH 3 C(CH 2 PR 2 h + 3NaCl (4)


R = Ph, Et

Although there are limitations with synthetic routes that use the
classical Grignard and organolithium reagents, these methods are still
widely used. Some recent applications to somewhat unusual ligands are
illustrated below. Capka et alys.19) treated a series of alkoxylsilyl-
substituted alkyl halides with LiPPh z to produce alkyldiphenylphosphine
chains that could be attached covalently to silica (Equations 5-6).

(EtOhSi(CH2)nBr + LiPPh 2 -. (EtOhSi(CH2)nPPhl + LiBr (5)


n = 1-6

(EtO)m(CH3h_mSiCH2Cl + LiPPh 2 -. (EtO)m(CH3h_mSiCH2PPh2 + LiCl (6)


m = 1-3

The interesting tripod molecule MeSi(PBuzh was prepared in 70% yield


by treating MeSiCh with LiPBuz in diethyl ether at -80c c. (16.20) The
corresponding carbon ligand CH 3 C(PPh zh was synthesized similarly,(21.22)
whereas the tris(phosphino)methane ligands HC(PPh 2h and HC(PMezh
are best prepared by the route shown in Equation 7.
R PCH PP I-BuLi, hexane L"[HC(PR)
2 2 2 -BuH ~ 1 22
J~
R,PCl HC(PR)
23 (7)
R = Me,Ph R = Me, Ph

Syntheses of derivatives of tris(diphenylphosphino)methane have been


achieved by a more circuitous route, (23) which is required because of C-P
bond cleavage when the elements (e.g., O 2, Ss, Se) are combined directly
with the triphosphine. Thus, lithiation of the methylene group of a bis-
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 263

phosphorus compound followed by reaction of a phosphinous chloride


(Equation 8) or reaction of lithiomethyldiphenylphosphine sulfide (in
excess) with a phosphinous chloride (Equation 9) produces the triphosphine
derivatives in reasonable yields.
BuLi R'PCI
• R 2 P(X)CHLiP(X)R 2 2 • [R 2 P(X)hCHPR; (8)
x = 0, S, lone pair

MeLt. Ph 2 P(S)CH 2 Li Ph 2PU. Ph 2 P(S)CH 2 PPh 2 I) BuLl.


2) Ph PCI
Ph 2 P(S)CH(PPh 2 )2
2 (9)
Stille has used the alkali-metal phosphide route to prepare polymer-
attached optically-active diphosphine ligands. Either a monomer (e.g., 17)
was copolymerized and treated with an excess of PPh 2 - or the monomer-
diphosphine unit (e.g., 18) was synthesized before polymerization.(24) In
both cases, the polymeric materials were treated with a rhodium complex

~
Ht-\---H
(0) (b)
TsO OTs
18
17

/CH2CH2PPh2
HN
'CH2CH2PPh2

19

and used as supported catalysts (Schemes 1 and 2; see also Section 5.4).
In an important application of the organophosphide synthetic method,
Whitesides et al.(25) synthesized the chelating diphosphinoamine, 19, and
then attached it to several different types of organic molecules by coupling
the amine with carbonyl halides, anhydrides, active esters, and isocyanides
(Scheme 3); they even attached the NP 2 unit to a protein as an asymmetric
support. To quote from Whitesides's paper, "the usefulness of this pro-
cedure for diphosphine synthesis rests on four features. First, the formation
of amides by acylation of amines is one of the best understood and most
general coupling methods in organic chemistry. The fact that it is possible
to acyl ate the secondary amine of 19 without interference by the
diphenylphosphine groups makes it possible to utilize this reaction for the
264 DEVON W. MEEK

+ ©~_fH2~~CH2~RCH~102
+

~01 0.7 -CH,-CH-·


H

/
CH2 CH2
\
Hn" H
CH2
/
CH2
\
OTs OTs OTs OTs

17

Scheme 1 124e )

preparation of a wide variety of diphosphines. Second, since the preformed


diphosphine moiety is introduced as a unit, yields are relatively high. Third,
the coupling reaction is compatible with a range of functionalities. Fourth,
carboxylic acids and their derived acylating agents are readily available in
great variety. ,,(25c)
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 265

Me

HO~Me
H

O=C
F
0--1-, Me +
I
0,
CH 2CH 2
O~ '0

l
I

AIBN =<,=0
Me

/CH 2
CH
I
c=o
I
N

P~
PPh 2

0.05 0.85 0.10

Scheme 2 124d )

I. KOC(CH 3 13
~CI THF ,reflux,ISh
CI H2N+ + 2HPPh2 - - - - - -
~CI 2. 2M HCI

Scheme 3 (250 )
266 DEVON W MEEK

3.2. Other Synthetic Methods


During the past ten years, two valuable alternatives to the Grignard
and alkali-metal-organophosphide syntheses of tertiary phosphines have
been developed. By choosing the appropriate route and reagents, one
should now be able to "tailor-make" polydentate ligands to incorporate
just about any electronic and/or steric effects desired.

3.2.1. Base-Catalyzed P-H Addition to Vinyl Compounds

King et al.(26) developed a route to poly(tertiary phosphines) by use


of a base-catalyzed addition of a phosphorus-hydrogen bond to the carbon-
carbon double bond of various vinylphosphine derivatives; the first exten-
sive series of methylated polyphosphines were prepared this way. The
synthesis of unsymmetrical diphosphines of the type R 2 PCH 2 CH 2 P(CH 3 h
involves addition of a P-H bond to CH 2 =CHP(S)(CH 3 h via potassium
tert-butoxide catalysis, followed by desulfurization with LiAIH4 in boiling
dioxane (i.e., Equations 10, 11).

Variations of the base-catalyzed method were used to prepare the three


tritertiary phosphines R'P(CH 2 CH 2 PR 2 h (R' = CH 3 or C6 H 5 , R = C 6 H 5 ;
R' = C 6 H 5 , R = CH 3 ), the mixed aliphatic-aromatic, open-chained
tetratertiary phosphine, 20, the branched pentatertiary phosphine, 21,
C6 H5 C6H5
1 1
(CH3)2PCH2CH 2 PCH 2CH2PCH 2CH 2 P( CH 3 )2

20

21
~

and the optically active di- and tri-phosphines, 22 and 23. Compound 23
is an example of a chiral triphosphine, a rarity. (26c)
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 267

3.2.2. Radical-Catalyzed P-H Addition to Vinyl and Allyl Compounds

Although the base-catalyzed process is valuable, it is limited to


R 2 PCH 2 CH 2 PR 2 units and to available vinyl phosphines. A more general
method for the preparation of polyphosphines, "mixed" phosphorus-sulfur,
"mixed" phosphorus-arsenic, and "mixed" phosphorus-nitrogen com-
pounds is the radical-catalyzed addition of phosphorus-hydrogen, arsenic-
hydrogen, or sulfur-hydrogen bonds across carbon-carbon double bonds
in either vinyl or allyl compounds. (27) Even in those cases where reagents
are available for the base-catalyzed method, the radical method appears
to give higher yields and purer compounds, especially if UV radiation is
used. (27c) The reactions are fast, high-yield, and general for vinyl compounds
and selected allyl (i.e., electronegative substituents such as Cl, OH, and
NH 2 ; but not Br, I, SR, or PPh 2 ); they are summarized by Equations 12-15.
R 2 P-H + CH 2 =CH-PR 2 --. R 2 PCH 2 CH 2 PR 2 (12)

R 2 As-H + CH z=CH-PR 2 --. R 2 AsCH 2 CH 2 PR 2 (13)

RS-H + CH 2 =CH-PR z --. RSCH 2 CH 2 PR 2 (14)

R 2 P-H + CH 2 =CH-C(O)NH 2 --. R 2 PCH 2 CH 2 C(O)NH 2 (15)


The radical-catalyzed method for preparation of "mixed" PP, PAs, PS,
and PN combinations in polydentate ligands is noteworthy, as it represents
268 DEVON W. MEEK

an easy route to many potentially useful combinations of donor units in


R
ligands. Compounds of the type "'-PCH=CH 2 may be exploited for the
/
i-PrO
syntheses of more complicated "mixed" ligands. For example, reduction
of the R2t-OR group to the secondary phosphine, R 2 PH, will then make
o
additional reactions possible, via either the base-catalyzed or the
radical-catalyzed method. One can envisage using several different types of
vinyl compounds for addition of R 2 P-H, RS-H, or R 2 As-H groups. For
example, treatment of diphenylphosphine with vinyltriethoxysilane
produced Ph 2 PCH 2 CH 2 Si(OEth in 90% yield (Equation 16).(27c)

This phosphine can easily be attached to a solid support (e.g., silica gel)
by a reaction of the triethoxysilane end of the ligand.
Mono- or di-substitution on a primary phosphine (e.g., PhPH 2 ) can
be controlled by the stoichiometric ratio of the reagents. For example,
Uriarte(27c) prepared PhP(CH 2CH 2CH 2NH 2h in 85% yield using excess
Ph
allylamine, whereas the monosubstituted product "'- PCH 2 CH2 CH 2 NH 2
/
H
is the predominant product when a 1: 1 ratio of PhPH 2 and allylamine is
used, particularly when both UV and AIBN are used. The molecule
Ph
"'-PCH 2 CH 2 CH 2 NH 2 is a valuable intermediate for additional condensation
/
H
reactions that would produce "mixed" tridentate ligands (e.g.,
P~PNH2 or As'--P~NH2 combinations).
The radical-chain addition reactions appear generally applicable for
vinyl compounds; in the case of allyl derivatives, the clean, high-yield
reactions appear to be limited to allylamine, allylalcohol, and allylchloride.
However, the alcohol function is easily converted to a chloride, bromide,
or tosylate derivative, which can be used to add another ligand donor group
(e.g., PR 2 , AsR 2 , or SR). The principal advantages of the radical-catalyzed
method are (i) the flexibility of designing many different related polydentate
ligands that contain a variety of donor groups with either -CH 2CH 2- or
-CH 2CH 2CH 2-connecting units by simply choosing the appropriate R 2 P-H,
R 2 As-H, RS-H, and vinyl derivatives; (ii) the experimental simplicity of
the homogeneous solutions and the one-pot reaction; (iii) the faster reaction
POL YDENTA TE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 269

times compared to the base-catalyzed addition to vinyl phosphines; (iv)


the ease with which the reaction impurities or by-products are removed in
vacuo, leaving a crude material that is most often sufficiently pure to use
as a ligand; and (v) the selective monosubstitution of a RPH2 molecule,
especially when the radical is generated by UV at room temperature or
below.(27)

3.2.3. Phosphonium Route to Unsymmetrical Bisphosphines

Although symmetrical bisphosphine ligands of the type Ph 2P(CH 2)n-


PPh 2 (n = 1-6) can be prepared relatively easily, attempts to synthesize
unsymmetrical bisphosphines, in which the two phosphorus atoms have
different substituents, have met with little success. However, recently Briggs
and Dyer (28 ) have developed a relatively rapid and efficient synthesis of
the unsymmetrical alkane-based bisphosphines RPhP(CH2)nPPh2 (n = 3,
R = ethyl; n = 4, R = methyl or ethyl; n = 6, R = methyl, ethyl or cyclo-
hexyl) by the general method shown in Scheme 4, which involves new
unsymmetrical biphosphonium salts and bisphosphine dioxides as inter-
mediates.

(n ~ 3; Ph= phenyl,
R=alkyll
Ph 3 P RPh2P
Br(CH 2l n Br • [BdCH2lnPPh3]Br • [RPh2P(CH2lnPPh31Br2

2NaOH

Scheme 4

The synthesis route (Scheme 4) fails for n = 1 or 2, due to diffi-


culties at the hydrolysis stage, but it is expected to be general for un-
symmetrical bisphosphines of types RPhP(CH2)nPPh2, R;P(CH2)nPPh2,
RPhP(CH2)nPPhR, and other permutations, where R = alkyl, R' = alkyl
or aryl, and n ~ 3. (28 ) This method should be particularly attractive for
synthesis of chiral bisphosphines of the type RPhP*(CH2)nPPh2 (R = Me,
Et, C 6 H ll , t-C4H9, etc.), all of which contain a chiral center at P*.

3.2.4. High-Dilution Syntheses of Phosphorus Macrocycles

During the past four years, Kyba has achieved an important break-
through in the syntheses of phosphorus-containing macrocycles. (29 ) He has
270 DEVON W. MEEK

All P example
high
dilution

x= PPh; M = H ,Li Y = PPh,5,O, x= Y = PPh 3 isomers


X= 5; M=H,Li NMe, NPh, X = PPh,
CH2 Y = 0,5 , N Me, NPh , CH 2
L = CI, Br X = 5,0 ,NMe ; Y= PPh
X= Y = 5

(l
L"l
(l p/ Ph Ph

©t
"P

©r: XLi
Xli
+
:)QJ o X YJQJ
Y X
V
~P pJQJ
Ph V'Ph
L--./

X = 5, PPh 1. Y X 1. 5 isomers
CI 5, PPh PPh PPh ,5
CI NMe,OH PPh NMe ,0
OH,OMe 0

Scheme 5

designed a high-dilution apparatus that facilitates condensation of phos-


phorus or sulfur nucleophiles with halide or tosylate ligand units to produce
II-membered P 3 and mixed P 2 S, PS 2 , and P 2 N cycles, as well as 14-
membered P4 , P 2 S2 , P 2 N 2 , and P 2 0 2 macro cyclic ligands (Scheme 5) in 18
to 45 % yields. (29d) Owing to the tetrahedral stereochemistry of the phos-
phorus atoms, a large number of isomers is produced with this synthetic
method, however; the five possible isomers of a P 4 macrocycle are illustrated
below.

Although no catalysis results have been reported with these macrocycles


to date, the P3 macrocycles may provide interesting catalytic effects, as
they are constrained to occupy the face of a polyhedron. Open-chain P3
ligands, which can chelate around an edge of a polyhedron, are proving to
be useful for homogeneous catalysis (see Sections 4 and 5),
Ciampolini et al. very recently reported the synthesis of the first
crown-ether-type phosphorus-containing macrocycle 4,7,13, 16-tetra-
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 271

phenyl-1, 1O-dioxa -4, 7,13,16-tetraphosphacyclooctadecane ([ 18]ane P 402)


in 18% yield in one step by treating the dilithioderivative of
Ph Ph
"- PCH 2 CH 2 P / with bis(2-chloroethyl)ether (Equation 17).(30)
H
/ "- H

LiPh
(17)
rHF,2()OC

/ 1) O(CH 2 CH 2 Clh in THF, -20°C


2) warmed to RT, and then refluxed 0.5 hr.

24

Although the synthesis of 24 was not done under high dilution conditions,
the 18% yield could probably be increased significantly by using Kyba's
apparatus.(29d) Synthesis of other complicated crown-ether-type polyphos-
phine molecules can be expected in the near future.
All of the five possible diastereoisomers of 24 were isolated and
characterized. Ciampolini et at. found that any given diastereoisomer racem-
izes at 140°C through inversion at the phosphorus atoms, thus giving an
isomeric mixture whose composition is close to the statistically expected
one. The two most easily available isomers were used to form nickel(II)
and cobalt(II) complexes of general formula [M([18]aneP 40 2)]X2, (X =
BF4 -, BPh 4-). The nickel(II) and cobalt(II) complexes of 24 are low-spin,
with the hexa- and penta-coordinate geometries of the cobalt(II)-
([18]aneP40 2) complexes being due to the configurations of the two ligand
isomers. The four phosphorus atoms occupy the equatorial positions of
both the octahedral and square-pyramidal Co(II) structures; for the a-
isomer only one oxygen atom is located in a proper stereochemical position
for coordination to the metal. (30)

3.2.5. Template Synthesis of Polyphosphine Ligands

Whereas many tri- and tetra-dentate combinations of nitrogen, oxygen,


and sulfur donors have been accomplished via the "template effect" for
two decades, syntheses of polyphosphine ligands via a template synthesis
272 DEVON W. MEEK

have been accomplished only recently. (31-33) As extensions of the base-


catalyzed and radical-catalyzed additions of P-H bonds to vinyl groups
(discussed in Sections 3.2.1 and 3.2.2), diphosphine(31,32) and triphos-
phine(33) ligands have been obtained in high yields by performing the
condensations on coordinated phosphines (Schemes 6 and 7). This method

Ph2PH
(Ph 2 PCH 2 CH 2 PPh 2) W(CO)5
AI BN
81%

t- BuOK
or
AI BN cis, 72,82 %

base

cis

Bu Li ,
THF -
cis - M(CO)4(PPh 2H)2 - - - - - " cis- [M(CO)4(PPh 2HHPPh2)]
-78°C

(1) RC=CR'
o Ph2
C P
(-78°C) OC, I / 'CHR
M I
/1\ /CHR'
OC C P
o Ph2
M = Cr', 25- 64 %
M=Moj 22-56%

(M= Cr, Mo ; RC = CR' = MeOOCC == CCOOMe , PhC=CCOOEt ,PhC=CH)

Scheme 6131.32)
POL YDENTA TE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 273

Addition Reactions to Coordinated Secondary Phosphines

PtCl z(PPH) +
-EtOH
[PtCI (eptpl]CI

EtOH
""""'PPh Z - [PtCI(CyZP~ ~""""PPhZ)]CI
Ph

PtC1Z(CyPPHl + CI/".../'-...PCyZ ~ [PtCI(CyZP~ PVVPCYZ)]CI


THF I
Ph

NEt3
+ CI/".../'-... PPh Z TH'F [PtCI(PhZP~ ~v'VPPhZl]c1
Ph

Scheme 7(33)

appears extremely attractive for making unsymmetrical ligands like


Ph Ph
I I , (33)
Ph2PCH2CH2PCH2CH2PR2 and R2PCH2CH2PCH2CH2CH2PR2. In addition,
the "template control" may make certain isomers of P 3 and P4 ligands much
more favored, as compared to the high-dilution method.
Base- or radical-catalyzed additions of P-H bonds (also As-H and
S-H) to vinyl groups should be advantageous for building certain desired
coordination geometries or mixed polymetallic complexes. Keiter has
used this method to prepare all six mixed-metal complexes of the
type (OC)sMP~PM'(CO)s (where M = Cr, Mo, Wand PP =
Ph 2 PCH 2 CH 2 PPh 2 ) by the condensation of the appropriate
(OC) sM(PPh 2 CH=CH 2 ) and (OC) sM'(PPh 2 H) complexes (Equation
18).(34)
KO(t-Bui
(OC)sM(PPh 2CH=CH 2) + (OC)sM'(PPh 2 H)

M = Cr,Mo, WandM' = Cr,Mo, W


274 DEVON W. MEEK

Keiter's data suggest that the best results are obtained by building the
polyphosphine from a vinylphosphine ligand bonded to the transition metal
complex of interest. This approach should be limited only by the lability
of the complex and the susceptibility of other attached ligands to reaction
under the experimental conditions employed. It should also lead to syntheses
of new cyclic and acyclic tri- and tetra-phosphines coordinated to the metal
(or cluster) used for the template.

4. POL YPHOSPHINE HOMOGENEOUS CA TAL YSTS

As illustrated in the first three sections, chelating di-, tri-, and tetra-
dentate ligands (particularly those incorporating more than one phosphorus
donor) have become an important class of ligands in inorganic and
organometallic chemistry. The author strongly believes that polyphosphine
ligands will become increasingly important, particularly in the areas of
asymmetric synthesis, supported catalysts, and clusters. The much lower
tendency for M-P bond dissociation in such ligands allows the chemist
unique possibilities for "fine tuning" electronic and/or structural para-
meters, while maintaining a fixed stoichiometry and stereochemistry in the
complex.
For metal-phosphine catalysts of the types RhCI(PPh 3 h, RuHCI(PPh 3 h,
and RhH(CO)(PPh 3 h, it has become accepted that dissociation of
triphenylphosphine is a necessary condition for hydrogenation or hydrofor-
mylation catalysis by such complexes. In the sections to follow, examples
have been selected to show that three phosphino groups are bonded to the
metal centers of the "active species;" yet rapid hydrogenation and selective
hydroformylation of terminal olefins still occur.
For a metal catalyst that oscillates between a sequence of 16- and
18-electron complexes, the catalyst cycle does not require dissociation of
one of the donor groups of a tridentate ligand. The primary requirement
for hydrogenation and hydroformylation catalysis is that three sites be
available on the metal for bonding two hydrogen atoms and the substrate
molecule (olefin or CO, respectively). Thus, a tightly bound tridentate
ligand can occupy the other three sites of an octahedral metal. Even a
4-coordinate dB metal [e.g., Co(I), Rh(I), Ir(!), Pd(II), Pt(II)] complex of
a chelating tridentate ligand can function as an effective catalyst, if the
fourth position is a participating ligand in the catalysis cycle (e.g., hydride
or CO for hydrogenation and hydroformylation reactions, respec-
tively). Thus, a complex such as HCo(ttp) or HRh(ttp)
(ttp = PhP(CH 2 CH 2 CH 2 PPh 2 h) should undergo the accepted sequence of
steps, i.e., for hydrogenation: coordination of an olefin, hydride transfer
to produce the metal-alkyl, oxidative-addition of H 2 , a second hydride
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 275

transfer, elimination of the alkane, and regeneration of the HM(ttp)


catalyst.
Subtle metal-ligand and metal-substrate interactions are critical for
understanding catalytic mechanisms involving organometallic and coordi-
nation compounds. Phosphorus-31 chemical shifts and coupling constants
are uniquely sensitive to these interactions, since a phosphorus atom is
bonded directly to the metal center in most cases. Consequently, phos-
phorus-31 nmr is rapidly becoming the analytical method of choice for
rapid characterization of organometallic and coordination compounds that
contain phosphorus ligands. It is expected that future 31 p studies will provide
definitive identification of "intermediates" in reaction mechanisms and in
catalytic systems, since 31 p nmr spectra provide direct characterization of
the phosphorus-containing species in solution.

4. 1. Hydrogenation Catalysis with Chelating Triphosphine Ligands


DuBois and Meek(36) found that the metal-hydride complexes
H 3Co(ttp), HCo(ttp), and HRh(ttp) (where ttp = the chelating triphos-
phine ligand PhP(CH 2CH2CH 2PPh 2h) rapidly catalyzed the hydrogenation
of terminal olefins at room temperature and 1 atm of H2 pressure. Also,
RhCI(ttp), which can be handled in air as a solid, is an effective
homogeneous catalyst in ethanol in the presence of NaBH4' which generates
RhH(ttp) in situ. The most rapid hydrogenation rates were with "HCo(etp)"
(etp = PhP(CH 2CH 2PPh 2h), which was generated in situ, and with an
isolated sample of HRh(ttp). Under ambient conditions, these two hydrides
convert l-octene quantitatively to n- octane at rates comparable to the
Wilkinson catalyst. However, the relative stabilities of the HRh(ttp) and
HCo(etp) catalysts differ markedly in solution. The HRh(ttp) complex
retains high catalytic activity through several batch operations in benzene
solution, whereas the activity of "HCo(etp)" ceases when the olefin is
depleted. DuBois also observed that initially H3CO(ttp) rapidly hydro-
genated l-octene to n -octane;(36) however, the rate decreases rapidly due
to isomerization of l-octene to cis and trans-2-octene in the presence of
the cobalt catalyst.
These HRh(ttp), RhCl(ttp), HCo(ttp), and H3CO(ttp) complexes differ
significantly from Wilkinson's catalyst, RhCI(PPh 3h, in that the triphos-
phine donor groups remain bonded through the catalyst cycle, whereas
some PPh 3 dissociation occurs when RhCI(PPh 3h is used. (37) The most
important distinction to be gained from the catalysis experiments using
chelating triphosphine ligands is that dissociation from the metal of one of
three phosphino groups of the triphosphine ligand is not required for
catalytic hydrogenation rates comparable to (or faster than) those obtained
with RhCl(PPh 3 h. (36)
276 DEVON W. MEEK

Subsequently, Niewahner demonstrated that the complex RhCI(ttp)


in the presence of either triethylaluminum or diethylaluminum chloride is
an effective homogeneous catalyst for hydrogenation of terminal ole fins
and l-octyne. (38) Proton and phosphorus-31 nmr spectra were used to
identify several different chemical species [including RhH(ttp)] in the
catalytically active RhCI(ttp)-AIEt 2X solutions. The observed rate of
hydrogenation of l-octene to n -octane at 20 ± 0.3°C and under a constant
H2 pressure of 750 torr is 6.4 x 10 4 M- 1 min-\ i.e., 25 times more rapid
than the Wilkinson catalyst, RhCI(PPh 3h, under comparable conditions.
In the presence of excess olefin and >2: 1 AI: Rh ratio, the observed rate
law for the catalytic hydrogenation of l-octene is -d[H 2 ]/ dt =
kobsd[Rhh[H2].
Proton and 31 p {lH} NMR spectra of toluene solutions of RhCI(ttp)
and AlEt3 (or AIEt 2Cl) indicate that RhH(ttp) is formed rapidly and cleanly
in the solutions and that addition of terminal ole fins results in very rapid
conversion of RhH(ttp) to RhR(ttp). The nmr data also indicate that one
ethylaluminum species is bonded directly to the rhodium and another is
bridged via the Rh-H linkage.(38) The exact nature of the ethylaluminum
species is not known because of exchange between coordinated and free
ethylaluminum units and the likelihood that AIEt 2 CI is the product from
the first step of the reaction between RhCI(ttp) and AlEt 3. Hence, the
ethyl aluminum species are indicated by EtAI in Equations 19 and 20.
EtAI:::
I
RhCI(ttp) + xsAlEt3 ¢ (ttp)Rh-H-AIEt + C 2 H 4 + EtAI::: (19)
EtAI::: EtAI:::
I \! I
(ttp)Rh-H-AIEt + olefin (ttp)Rh-R + EtAI::: (20)
Formation of the hydride complex RhH(ttp) in toluene results from (3-
hydride elimination of ethylene from RhEt(ttp), which is formed by an
initial alkylation of RhCI(ttp) by either AIEt3 or AlEt 2Cl.
On the basis of the 31 p eH} and the IH nmr spectra, the kinetic data,
and the fact that solutions of RhCI(ttp) and either AlEt3 or AIEt 2CI rapidly
catalyze the hydrogenation of olefins at ambient conditions, the following
mechanism involving RhR(ttp), RhH(ttp) associated with an ethyl-
aluminum species, and H2 was proposed. (38) (Association of the ethyl-
aluminum species with the rhodium complexes is omitted for clarity.)
KJ
RhCI(ttp) + AIEt 2 Cl ~ RhEt(ttp) + AIEtCI 2
K2
RhEt(ttp) ~ RhH(ttp) + C 2 H 4
K3
RhH(ttp) + alkene ~ RhR(ttp)

RhR(ttp) + H2 ~---.. RhH(ttp) + alkane


25
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 277

The analogous system containing Et 2AICI and the rhodium complex of


PhP(CH 2CH 2CH 2PCY2h, where Cy = the cyclohexyl group, also catalyzes
l-octene at a rate comparable to that of RhCI(ttp) + Et 2AICI; however, in
this case, the rate depends on the concentration of olefin.
In another recent study with the triphosphine ligand ttp, Mazanec,
Letts, and Meek discovered that excess NaBH4 in refluxing THF converts
[RuClz(ttp )]2 into the mixed hydride-borohydride complex [RuH( 1/ 2_
BH4)(ttp)], 26.(39) Compound 26 is unique in that: (1) the 1/2-BH4 group
is static on the nmr time scale at room temperature; and (2) the BH4-
protons undergo a distinct two-step exchange process above room tem-
perature, with the Ha protons being scrambled before the Hb protons
(Figure 8.1). Owing to the presence of both Ru-H and doubly-bridged
Ru-BH4 linkages in 26, addition of an acid (e.g., HBF4) in the presence

of neutral ligands L produces the cationic complexes [Ru(H)(L)z(ttp)t


(L = CO, CH 3CN, and P(OMeh), whereas addition of a base produces a
different series of complexes, i.e., the neutral complexes RuH 2 (L )(ttp)
(L = CO, PPh 3, and P(OMeh).(39)
Several of the [RuH(L)z(ttp)t and RuH 2(L)(ttp) complexes may
prove to be effective hydrogenation and hydroformylation catalysts. For
example, preliminary studies on the hydrogenation of l-octene showed
that the parent compound RuH(1/2_ BH4)(ttp), 26, is inactive as a catalyst
at 25°C and 1 atm of H2 pressure. However, after addition of one equivalent
of HBF4'Et 2 0, catalytic hydrogenation commenced immediately at a rate
~0.75 that of RhCI(PPh 3h in THF; the hydrogenation gave only n-octane
as a product. (39) In a separate set of hydrogenations with NEt3 as a
cocatalyst, an induction period of about 10 minutes was needed before the
maximum rate was obtained; the slow step appears to be cleavage of
the RU:::::::::B bridge. From the spectral data, it may be inferred that
RuH 2(ttp) is the active catalyst in the presence of base (Figure 8.2). In the
presence of acid, the catalytic species is probably cationic, e.g., [RuH(sol-
r.
vent)x (ttp) The isolated, stable, 18-electron cationic complexes
[RuH(CH 3CNh(ttp)]BF4 and [RuH(P(OMehh(ttp)]BF4 are not effective
catalysts; this difference in the catalysis rates may simply reflect the need
for one of the nonphosphine ligands to dissociate to generate a vacant
278 DEVON W. MEEK

378 0 K
j

~33~ (J:P
,A 1\ M
Itl~ ~ "( " P I"H,
rt-318~
"'" JlA JL
Ho-'B,

I., H,

~230~_
He ~Hb H
2 ,-5 ,10 ,-158
Figure 8.1. Proton (11 B decoupled) nmr spectra of RuH( 1/ 2 - BH 4 )(ttp), 26, in dg-toluene for
the temperature range 230-378°K. Note that the resonance for Ha begins to collapse in the
temperature range 338-348°K, whereas the resonance of Hb remains. Both Ha and Hb
resonances collapse as the temperature is raised to 378°K. Reprinted with permission from
I. Am. Chern. Soc. 104,3898-3905 (1982). Copyright (1982) American Chemical Society.
SP1 = 16.4 ppm
8P2 =38.0
SP3 = 153.3

JP1- P2 = 36 Hz
JP1- P3 = 17
JP2-P3 =33
JP1-Pi'(fJP1-P3

Pl ---
STD
I Ho __ I
152 ppm Oppm
Figure 8.2. 31p{lH} nmr spectrum of cis-[RuH 2 (P(OMeh)(ttp)] in C 6 D 6 , note that all of the
J p . p couplings are relatively small, in contrast to the 2lp.p values of 300-500 Hz observed for
trans 2J p.p couplings in similar complexes. Reprinted with permission from J. Am. Chern. Soc.
104,3898-3905 (1982). Copyright (1982) American Chemical Society.
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 279

coordination site before the ruthenium-ttp complex can function as a


hydrogenation catalyst under mild conditions.(39)
In both the RhHCttp)- and RuHC1/2- BH4)(ttp)-catalyzed hydrogena-
tion studies discussed above, the 31 p {lH} and IH NMR spectra show that
the ttp ligand does not dissociate from the metal over the temperature
range studied, i.e., 203-353°. Thus, rapid hydrogenation catalysis occurs
at ambient conditions, even when three phosphino groups are bound to
the metal center. In the future, we should see particularly selective catalytic
results with chelating tridentate ligands, as a result of the steric and
electronic control that can be incorporated into such a metal complex.

4.2. Hydroformylation Catalysis with Chelating


Oiphosphine Ligands

Slaugh and Mullineaux of Shell(40) and Eisenman(41) of Diamond Alkali


were the first investigators to report that rhodium-phosphine Cor -phosphite)
complexes are efficient catalysts for hydroformylation of l-alkenes
(Equation 21), using I-hexene for illustration. Their work was followed

~+H2+CO--+ ~CHO+~
CHO (21)

+~+~

closely by extensive Rh-PPhrcatalyzed studies by Wilkinson and co-


workers between 1965 and 1970.(42-46) They found that at 1 atm of 1: 1
H 2: CO, 25°C, and 50 mM HRh(CO)(PPh 3h, a 20: 1 linear-to-branched
aldehyde (l: b) ratio was obtained by hydroformylation of 1-pentene.(44)
However, at the higher temperatures and pressures and low-rhodium
concentrations anticipated for commercial operation, I: b ratios of only 2-3
were reported.(47) This poor selectivity was improved substantially by
increasing the PPh 3 concentration (46,48) or by decreasing the partial pressure
of carbon monoxide.(47b) This requirement for excess PPh3 ligand was
attributed to a need to maintain at least two phosphine ligands bonded to
rhodium during the selectivity step of the reaction. (46,48)
Oliver and Booth(49) and Pruett and Smith(48) showed qualitatively
that both rate and selectivity of hydroformylation increase as the phosphine
ligands are changed from the more basic alkyl phosphines to the less basic
aryl phosphines. These substituent effects, as well as the need for a large
concentration of monodentate phosphines, suggest that increased I : b selec-
tivity could be achieved by chelating diphosphine or polyphosphine ligands,
since dissociation of a chelated ligand from the metal should be more
difficult than dissociation of a monodentate ligand. In fact, Unruh, Hughes,
280 DEVON W. MEEK

Ph 2 P(CH2'n PPh 2
~PAr2
Fe
'\§;;rPAr2
n = 2 dppe
n=3 dppp Ar = CSH5 fd pp
n = 4 dppb Ar = p- CF3- CSH4 fd pp - CF 3
Ar=m-F-C SH4 fdpp-F
Ar=p-CI-CS H4 fdpp-CI

~PPh2 ~PPh2
{~PPh2 PPh2
PPh2
LZ::.PPh 2
PPh2 PPh2
diop t- bdcb c - bdc b t -bdch

27 28 29 3,9

Figure 8.3. Examples of bidentate diphosphine ligands used in the Celanese hydroformylation
studies (27-30).

and Christenson at Celanese(SO) recently reported hydroformylation studies


with the chelating ligands shown in Figure 8.3.
The Celanese investigators(SO) found that t-bdcb, mop, and ferrocene-
type ligands led to highest selectivities and when the ligand: rhodium mole
ratio was :;;,;1.5: 1. When substituted phenyl groups were attached to phos-
phorus, good correlations were obtained between the Hammett (T para-
meters and the I: b aldehyde ratios, as well as the rate of formation of
linear aldehyde (Figure 8.4). Electron-withdrawing substituents on the
benzene rings (less basic ligands) improved the hydroformylation rates and
selectivities to linear aldehyde. The 3: 2 diphosphine-rhodium complex
that has been proposed(Soa.b) to account for these observations is shown in
Figure 8.5; it has one bulky diphosphine ligand (e.g., t-bdcb, mop, or the
ferrocene-based P 2 ligands) chelated to each rhodium atom and a third
diphosphine ligand bridging the two rhodium atoms. The 3 : 2 diphosphine-
rhodium ratio is based on the catalysis results, and the distorted trigonal-
bipyramidal structure is based on the 31 p {lH} nmr spectral data. (SOc)
Consistent with the 3 : 2 diphosphine: Rh stoichiometry, a mixed ligand
effect was observed, wherein 1: 1 molar combinations of a diphosphine
and HRh(CO)(PPh3h require a monophosphine of optimum size (i.e.,
phosphine ligands approximately the size of Ph 2PEt and Ph2PCHz-, where
(J == 140°) to produce linear aldehyde selectivities comparable with
stoichiometric (i.e., 1.5: 1 ratio) use of diphosphines alone.
In the studies on the Rh-fdpp-CF3 catalyst system,150dl the effects of
H2 and CO partial pressures and total reaction pressure (1: 1 H 2/CO)
indicate that CO, rather than the bridging diphosphine ligand, dissociates
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 281

20.0
0
I-
<[
a: 10.0
.0
8.0
6.0
eo(

4.0

20.0

10.0
u
G>
II>
8.0
" 6.0
q
0
4.0
><
a.
G>
s: 2.0
.x
H p-CI m-F p- CF 3

-0.4 0,0 0.4 0.8


HAMMETT (J

Figure 8.4. Hammett up plot for fdpp and derivatives. khep, = (Effhep,)(kobs)/100 [Rhl. where
[Rh] = catalyst concentration in millimoles/liter. Conditions: 790 kPa. 110°C. 1: 1 H 2 : CO.

Ar ~
Ar'~-f0
OC'~h_P,
/1 dr Ar
~pH
~ l'Ar
I
Ar
Fe
Ar ~r I
Ar H/'P~
Ar,1 I
P-Rh

@
o
~ "CO
\'Ar
Ar

Figure 8.5. Proposed structure of the 3: 2 diphosphine: rhodium selective catalyst species.
282 DEVON W. MEEK

Table 1. 1-Hexene Hydroformylation 8 with Rigid Diphosphines

Percent selectivity to

Added Ratio 2-MethyI-


P-P P -P:Rh 1: b Hexane 2-Hexene hexanaI Heptanal

( + )-diopb 1.0 2.1 1.0 53.0 15 31


1.5 5.5 0.1 1.2 15 83
2.0 5.5 0.8 0.2 15 84
t-bdcb C 1.0 3.2 0.4 23.0 18 58
1.5 7.2 0.5 1.4 12 86
2.0 7.2 0.5 1.2 12 87
3.0 7.2 0.5 1.3 12 86
10.0 7.9 0.5 1.0 11 87
c-bdcb d 1.5 3.2 0.4 24.0 18 58
2.0 3.4 0.4 12.0 20 68
5.0 3.7 0.6 0.6 21 78
55.0 3.7 0.7 0.7 21 78
t-bdch e 1.5 2.4 0.0 29.0 21 50
5.0 1.1 0.0 1.0 46 52

a AJI runs with 790 kPa of 1 : 1 H,: CO.


h 106 ± 3°e.
c 106 ± 2°C.
d 100 ± l°e.
e 103 ± l°e.

in order to provide a 16-electron complex capable of binding the olefin.


In contrast to the results with t- bdcb, mop, and the ferrocene diphosphines,
the "nonrigid" diphosphines, dppe, dppp, and dppb did not lead to high
selectivities for linear aldehyde. Also, they did not maximize total aldehyde
efficiencies at a ligand-to-rhodium ratio of 1.5: 1, as did the more rigid
ligands. The t- bdch ligand (Table 1) behaved very much like the nonrigid
ligands, while c-bdcb (Table 1) was intermediate between the high selec-
tivities of the ferrocene-based ligands and the nonrigid ligands.
The mechanism suggested recently by the Celanese investigators to
account for the fdpp, t-bdcb, and mop results is shown in Figure 8.6. The
scheme does not specify the geometry of the 3 : 2 complex, and it assumes
that the two rhodium atoms within the complex act independently. The
Celanese mechanism depicted in Figure 8.6 differs in two important aspects
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 283

o ~ CO
II I
RCRhL2L k" RRhL2L

~I
RCRh(H),L2L
~o
RRh(HbL 2L ~<===)::;k~B' RRhL L

RCH~'" :H~ R' ' ••/{'


HRhCOL2L
)
.c"
C,~
) k:
I.
HRhL2L < ko
Rb
\
b
H~hL2L
Ik"
R)-CHol \~ •••.

R'
o
II
RH~ R' L Bb R'

:>--CR\ >-Rh(H)U . ~:~_,fc:L'L

-~H~
R' II
.:>- CRhL2L
0
k_"
'
R' I
CO

>- RhL2L
r 2b

k3b

Figure 8.6. Proposed Celanese mechanism for the rhodium-phosphine catalyzed hydroformy-
lation. L2 = chelating ligand; L = 1/2 of bridging ligand; R = CH 2 CH 2 R'.

from the Wilkinson mechanism, (44--46) which has been the mechanism
generally accepted for rhodium-phosphine catalyzed hydroformylation
reactions (51.52):
1. Wilkinson suggested that only two phosphine ligands were bound to
rhodium in the "active-catalyst," i.e., dissociation of one PPh3 ligand
from HRh(CO)(PPh 3h is a prerequisite for formation of an active
catalyst. According to mechanistic studies that are summarized by Figure
8.7, HRh(COh(PR 3h can be formed readily from HRh(CO)(PR 3h
and CO, HRh(COh(PR 3h being the active species capable of reacting
with the olefin. In contrast, the Celanese catalysis and 31p{1H} results
suggest that under hydroformylation conditions, three phosphorus
atoms must be bound to rhodium at the instant that straight-chain
selectivity is determined. (50d) The high I: b selectivity obtained with rigid
284 DEVON W. MEEK

PPh 3 CO
HRhCO(PPh 3 b ::;::=./'===~> HRhCO(PPh 3 ), :.;;:::=\,.====:=::=: HRh(CO),(PPh 3 )'

RCHO ~ 1~ 1-alkene
o 1-alkene
II I
RCRh(H)'CO(PPh 3 ), HRh(CO),(PPh 3 ),

H21l
o
II
0
II
RCRh(CO),(PPh 3 )' ::;::=~'\==::: RCRhCO(PPh 3 ), :::;:::::===::::
1l
RRh(CO),(PPh 3 ),
co
(a)
PPh 3 1-alkene 1-alkene
.2
HR:~TPPh'), ::::==~=== HRh(CO),PPh 3 ~=====::::::::
\
HRhlT'PPh'
I

HRhCO(PPh 3 )' RRh(CO),PPh 3

:HO~ 0 ~PPh'
II II
RCRh(H 2 )CO(PPh 3 ), ~=~=::: RCRhCO(PPh 3 )' ::;::===== RRh(CO),(PPh 3 ),
~
'\
H2
(b)
Figure 8.7. Wilkinson mechanism for rhodium-catalyzed hydroformylation: (a) associative
path; (b) dissociative path. R = R'CH 2 CH r or R'CH(CH 3 )-.

diphosphine ligands indicates that complexes of the type HRh(CO)L 2L


(where L2 = a chelating phosphine, L = a monophosphine or a diphos-
phine acting as a monodentate to that rhodium atom) influence the
important step in the catalytic cycle at which Rh-H adds to coordinated
olefin.
2. Variations of the H 2 : CO ratio show that the rate-controlling steps for
linear and branched' aldehydes are different, and that CO may not be
bound to rhodium when the olefin is first coordinated to Rh. These
results suggest: (a) either the CO ligand dissociates to form the 16-
electron complex HRh(L 2 )(L), which then coordinates an olefin
molecule to form HRh(L 2 )(L)(olefin); or (b) L dissociates to allow
formation of HRh(CO)(L 2 )(olefin), but it quickly reassociates, driving
the Rh-H addition step preferentially to the linear alkyl intermediate
(n-alkyI)Rh(CO)(Lh(L). In either case, the stereochemistry imposed
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 285

by the three phosphine groups provides the influence for linear aldehyde
selectivity.
The important differences between the Celanese and Wilkinson
mechanisms (i.e., composition and coordination geometry of tris-phosphine
vs. bis-phosphine active species) merit further careful study. This may
prove to be an area where chelating specially designed triphosphine ligands
can provide definitive choices between the two mechanisms.

5. SOLID-SUPPORTED POL YPHOSPHINE CA TAL YSTS

Applications of di-, tri-, and tetra-dentate ligands in the areas of


asymmetric synthesis and catalysis by polymer-bound metals have begun
to require incorporation of polyphosphine units into complicated organic
structures. Introduction of multiple phosphine units onto a solid support
can be accomplished (1) either by attachment of the entire polydentate
ligand via functionality on one of the substituent chains, or (2) by introduc-
tion of multiple individual phosphine units into a precursor, ordinarily by
nucleophilic displacement of halides or tosylates by a R 2 P- ion (most
commonly the anion has been Ph 2 P-) or by addition of R 2 PH molecules
to activated olefins. (53) The latter routes suffer from unpredictable yields
and incompatibility with many functional groups. Three studies, which
illustrate the special advantages that can be gained by covalent attachment
of chelating polydentate ligands, are presented below.

5.1. Coordination of Multiple Phosphino Groups from


Supported Monophosphines

In contrast to organic polymers such as polystyrene, rigid nonswellable


inorganic supports such as silica or alumina should increase the efficiency
of anchored catalysts by minimizing undesirable deactivation reactions,
such as intermolecular aggregation of active species or multiple-ligand
chelation of the metal center. The latter factor is a problem with most of
the supported monophosphine catalysts that have been studied. For
example, spacing of the -PR 2 groups uniformly on the support surface is
very difficult to control during the synthesis. Consequently, one metal may
be bonded to only one -PR 2 group, whereas another metal may be bonded
to two -PR 2 groups (Figure 8.8). Thus, the electronic and stereochemical
properties of the two metals will differ significantly, causing the catalysis
specificity and rate to differ at the two metal sites.
286 DEVON W MEEK

solid ~
SUPPO'J

Figure 8.8. Illustration of two types of metal-phosphine bonding that can occur with supported
monophosphine ligands.

In an interesting demonstration of the variable coordination environ-


ments created by anchored monodentate diphenylphosphino groups, Capka
et at. compared the kinetics of liquid-phase hydrogenation of alkenes
catalyzed by homogeneous rhodium(I) complexes prepared in situ from
11-, 11- ,_ dichloro-bis- [di(alkene) rhodium] and tertiary phosphines of the type
RPPh 2 (R = -(CH2)nSi(OEth; n = 1-6 and R = -CH2SiMe3-m(OEt)m;
m = 1-3) and by their heterogenized analogues anchored to silica.(ls.19)
The hydrogenation with both types of catalysts at 1.1 atm hydrogen pressure
and 37-67°C were first order in the alkenes. With one exception, the rate
constants of the homogeneous catalysts do not depend on the length of
the phosphine-alkylene chain, -(CH2)nPPh2. Except for the (CH 2h chain,
all the rates for the homogeneous catalysts are in the range of 4.5-
5.5 liter mol- 1 min -1. The length of the -CH2CH2CHZ- chain apparently
allows the ethoxysilyl group to interact-either sterically or electronically-
with the rhodium dimer, and this rate is one-half the other values.
The ligands (EtOhSi(CH2)nPPh2 and (EtOhMeSi(CH2)nPPh2 were
anchored onto silica and treated with [RhCI(ethylenehJ2 to produce the
supported catalysts. The resulting heterogenized complexes were highly
active for both hydrogenation and hydrosilylation of alkenes under mild
conditions. The observed variation in the catalytic activity of 3--
(CH2)nPPhz-RhCILn complexes clearly shows that two different active sites
are formed on the silica surface. The activity is significantly higher (a factor
of 10 for hydrosilylation and a factor of 3-7 for hydrogenation) when the
rhodium atom is linked through a short chain, i.e., the -CH 2PPh 2 group,
compared to linkage via alkylene chains containing 2-6 methylene groups.
Variations in the activity were attributed to differences in the nature of
the catalytic species formed on the surface. From calculations of the surface
area and the phosphine concentrations, it was concluded that the short
-CH 2PPh 2 chain forms monophosphine, monorhodium complexes (i.e.,
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 287

31), whereas the longer chains lead to dimeric structures (32) for 2-4
carbons and (33) for 5, 6 carbons, respectively; such types of dimeric
complexes are known to be less active hydrogenation and hydrosilylation
catalysts for olefins. (35)
}-CH2 P" )C
Rh
C( "J(
31
~2 n=2-4

33 n=6,5

5.2. A Triphosphine Ligand Attached to Glass Beads

The structure and stoichiometry problems that are inherent with sup-
ported monophosphines, which were illustrated above with Capka's studies,
can be avoided by choosing a polydentate ligand to contain the requisite
number of phosphine donors. Uriarte and Meek recently demonstrated
the utility of this approach by: (1) attaching the first examples of tridentate
ligands onto a controlled-pore glass (c.P.G.) solid support; (2) characterizing
the CoCl z complex of the attached triphosphine ligand; and (3) demonstrat-
ing efficient hydrogenation catalysis of terminal olefins at 25°C and 1 atm
Hz pressure by this bound CoCJz·triphosphine complex in the presence of
a hydride source, OMH-l. The attachment sequence of reactions is presen-
ted below. (27d)

-
KI03
1- 0
II
OCH2C-H

t H
I
OCH2C=N~P
r-/PPh2
--
' - \ No BH4
PPh 2
t H H
II
OCHt:-N"-"'P
I
H
r/PPh2
"--"\
PPh2
288 DEVON W. MEEK

5.3. A NP2 Ligand Attached to Organic Compounds

Whitesides et al. introduced a chelating diphosphine unit into several


different organic molecules by taking advantage of the selective reactivity
of the amino group compared to the phosphorus atoms of
HN(CH 2CH 2PPh 2h toward carbonyl halides, anhydrides, active esters, and
isocyanates (see Section 3.1 for the attachment reactions). (25) They demon-
strated that the attached amino-diphosphine ligand (NP2) leads to com-
plexes of transition metals having a wide range of structures and physical
properties. Particularly, the water-soluble complexes offer tremendous
potential for homogeneous catalysis in aqueous solutions. For ease of
synthesis and isolation, ligand 34 provides the most practical rhodium
catalyst for effective homogeneous hydrogenation in aqueous solutions. (25d)
o
II
HOOC ~C'-..N/CH2CH2PPh2
HOOC~ 'CH 2CH 2PPh2
34

The NP 2 unit and the resultant achiral [Rh(NP 2)(NBD)t moiety can
also be attached easily at a specific site in a protein. The protein structure
then provides the chirality required for enantioselective hydrogenation.
Thus, hydrogenation of a- acetamidoacrylic acid to N -acetylalanine
catalyzed by [Rh(NP2)(NBD)t bound to avidin at RT and 1.5 atm of H2
showed -40% S enantiomeric excess. Although these hydrogenation
results with avidin are modest, it does demonstrate that asymmetric syn-
thesis is accomplished by the bis -phosphine rhodium catalyst attached
covalently to a protein. (25b)

5.4. Chelating Diphosphines Attached to Organic Polvmers

Stille et al. have prepared a series of rhodium-bound catalysts contain-


ing optically-active diphosphine ligands by two different routes. In some
cases, they first prepared monomers containing the diphosphine unit and
then performed the polymerization. (24) In other cases, the polymer was
treated with excess diphenylphosphide to replace tosylate groups (Schemes
1 and 2). For example, the optically-active monomer 2-p-styryl-4,5-
bis[(tosyloxy)methyIJ-1,3-dioxolane was polymerized with a variety of
co-monomers to provide polymer-attached optically active ligands for
rhodium-catalyzed hydroformylation (24e) and hydrogenation (24a,b) reactions
that produced optically active aldehydes and amino acids with the same
enantioselectivity as their homogeneous counterparts. More recently, two
optically active bis(phosphino)pyrrolidine monomers were prepared and
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 289

copolymerized with hydrophilic co-monomers and a divinyl monomer to


provide a cross-linked insoluble polymer that will swell in polar solvents.
Exchange of rhodium(I) onto the polymer gave catalysts that were active
for the asymmetric hydrogenation of N -acyl a -amino acids in high optical
yields, where the phosphine that was derived from the enantiomer of the
naturally occurring 4-hydroxy-proline gave (s)-amino acids. The catalysts
could be recycled by simple filtration with no loss in selectivity. (24c,d)

6. CONCLUDING REMARKS

It is anticipated that polydentate ligands will become increasingly


important in studies of catalysis, particularly for definition of the
stoichiometry and stereochemical specificity of "active species." Increasing
use of polyphosphorus ligands can also be expected, since 31 p nmr spectra
are now readily available on FT nmr instruments, and the 31 p spectra
provide direct information on the phosphorus-containing species in
solution.(S) As illustrations of the special situations one can control with
polydentate ligands, a few recent examples are cited below.
In a particularly interesting application of the radical- and base-
catalyzed synthesis of polyphosphine ligands (Section 3.2.2), Keiter et at. (54)
have prepared all five possible nonchelated triphosphine complexes of
W(CO)s (35-39).

35 36

P
(OC)sWPPh 2~ I ~PPh2W(CO)S
Ph
37

W(CO)s
I
P
(OClsWPPh 2 ~ I ~PPh2
Ph

38 39

The synthesis strategy involved building each specific complex from judi-
ciously selected coordinated fragments (Equations 22-24), similar to the
290 DEVON W. MEEK

method used by Waid and Meek for platinum complexes of new unsym-
metrical triphosphines (Scheme 7, p. 273).(33)
Ph
AlBN I
(OC)sWPPh(CH=CH 2h + 2PPh 2H --------+ Ph2PCH2CH2PCH2CH2PPh2 (22)
75'C I
W(COls
35

Complex 36, a structural isomer of 35, is synthetically difficult and best


prepared by a two-step reaction involving Ph 2 PCH 2 CH 2 PPh(CH=CH 2 )
(Equations 23 and 24):
PhP(CH=CH 2h + PPh 2H ~ PhP(CH=CH 2)(CH 2CH 2PPh 2) (23)
PhP(CH=CH 2)(CH 2CH 2PPh 2) + (OC)sWPPh 2H AlBN.

Ph
I
(OC)s WPPh2CH2CH2PCH2CH2PPh2 (24)
36
Complex 37 was prepared also by the free radical reaction, whereas com-
plexes 38 and 39 were prepared by the base-catalyzed route (see p. 289). (54)
Tridentate tripod ligands, such as CH 3C(CH 2 PPh 2 h, triphos,
CH3C(CH2PEt2h, etriphos, and CH3C(CH2AsPh 2h, triars, have been
used to stabilize a remarkable series of P3, As 3, and P4 fragments in triple-
decker sandwich dinuclear cations of the type [(tripod)M(t-t-(1/3-
P3)M(tripod)]2+.(55) More recently, these tripod ligands have led to
hydride- and halide-bridged dinuclear complexes, e.g.,
/H~ /H~
[(triphos)Fe-H-Fe(triphos)]PF6, [(triars)-Co-H-Co(triars)]BPh4, and
~H/ ~H/

[(etriphOs)Fig'pe(etriphos)]PF6.(56) The hydride-bridged complexes


~CI/
should be examined for potential hydrogenation catalysis.
Tripod ligands, such as HC(PPh 2h and CH 3Si(PPh 2 h, have been
suggested(14-16) as good candidates for coordinating to three different metal
centers in a cluster complex, and thus holding those centers together.
However, complications can also result. For example, treatment of
RU3(COhz with HC(PPh 2h in THF solution leads to formation of at least
eight products, each in < 10% yield. (57) One of the products is the desired
axial symmetrically capped RU3(COMHC(PPh2h], 40, analogous to the
previously reported RU3(COMCH3Si(PPh2hl (16) Spectroscopic data indi-
cate that two other products contain bridging CO ligands (as for 40);
however, at least three products contain only terminal CO ligands, as well
as two equivalent and one unique phosphine group. One such product was
characterized crystallographically as the artha -disubstituted complex
RU3(CO)9{HC(PPh2)(PhPC6H4PPh)}, 41.(57) In the latter complex, the
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 291

triphosphine ligand has been transformed so that a phenyl ring of one


phosphine moiety has undergone ortho -substitution by the phosphorus
atom of a second phosphine moiety, together with the loss of benzene.
Other perturbations sometimes arise from polydentate ligands, i.e.,
either not all the donor atoms are bound to metal atoms, or each donor
atom may bond to a different metal atom, leading to polynuclear complexes.
Examples of these two situations are 4i 58 ) and 43,(59) respectively.

H
C
Ph p / <'--PPh
21 Ph 2P 1 2
Ag
1
1
1g 1
Ag

X X
X
42 43

In the case of the potential tetradentate ligand tris (2-pyridyl)phos-


phine, P(py h, Wilkinson et al. have shown that it functions preferentially
as a monodentate phosphorus ligand. (60) Complexes can also be isolated
in which P(py h acts as a bidentate, chelating P, N ligand, e.g.,
RuHC1[P(py hh and RhCl[P(py hJ2. In the presence of excess P(py hand
at low CO + H2 (1: 1) pressures, the complex RhH(CO)[P(py hh is a
catalyst for the selective (l: b ratio = 13: 1) hydroformylation of 1-hexene
to n -heptanal.
Nickel(II) complexes of the tridentate ligand 2,6-bis (diphenylphos-
phinomethyl)pyridine, pnp, react readily with CO in H 20-EtOH at room
temperature and pressure to give a Ni(O)-carbonyl complex. (61) This Ni(II)-
Ni(O)-pnp system is an effective homogeneous catalyst for the water-gas
shift reaction in water-alcohol (ethanol, propanol, or butanol) at low tem-
peratures and pressures.
Ph
I
The triphosphine ligand Ph2PCH2CH2PCH2CH2PPh2, etp, has been
used recently to prepare the complexes 44 (see Scheme 8)(62) and 45,(63)

45
292 DEVON W. MEEK

Scheme 8

which display unique properties. Reduction of the cation [(7]s_


CsHs)Fe(etp)t with LiAID4 gives 44, which slowly undergoes H-D scram-
bling at 20°e. The mechanism for H-D scrambling is shown in Scheme 8;
this scrambling process appears to be the first example of an exo -hydrogen
sigmatropic shift around a hydrocarbon ligand bonded to a transition metal.
Complex 45 was first isolated as a co-product during the reduction of
MoCh(etp) (Equation 25).
Na/Hg; N2
MoCI 3 (etp) + PR 3 THF ~ trans-Mo(N 2 h(etp)PR 3 (25)

46

Acid treatment of 46 converts a Mo-N z group into NH3 and permits


isolation of MoBr3(etp) in >90% yield (Equation 26). (64)

Use of the triphosphine ligand permitted, for the first time, characterization
of the metal-containing product (MoBr3(etp) in this case) after ammonia
formation from a metal-dinitrogen complex. (64)
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 293

Compared to the number of papers on chelating polyphosphine ligands,


there is a paucity of papers on chelating phosphites, even though such
ligands should make interesting comparisons by exhibiting stronger 1T'-
acceptor properties than a similar polyphosphine. It is expected that
the number of studies with chelating phosphites will increase rapidly.
In fact, King reported an initial study on the bidentate
(CH 3 0hPCH 2 CH 2 P(OCH 3 h in 1978;(65) recently Fryzuk(66) showed that
this ligand and the corresponding isopropyl ligand formed the interesting
polynuclear rhodium bridged-hydride complexes

~CRh-H
p(OMeh]
~CRh-H , respectively.
p(Oiprh]
and
P(OMelz 4 P(OiPrlz 2

47 48

Complexes 47 and 48 are extremely efficient homogeneous hydrogenation


catalysts for simple olefins; however, it is not yet known whether the
integrity of the hydride-bridged cluster is maintained during the hydrogena-
tion cycle. These compounds should lead to some fascinating results in the
future.
King has recently studied the first examples of chelating tridentate
phosphine-phosphite ligands, i.e., RP[CH 2 CH 2P(OCH 3 hh (R = CH3 and
C 6H 5).(67) These ligands react with metal(II) chlorides of Fe, Co, and Ni
in methanol solution to form the [MCI(P 3 ligand)t cations, which are
planar for nickel and tetrahedral for cobalt and iron. (67)

ACKNOWLEDGMENTS

The author is grateful to the Guggenheim Foundation for a 1981-1982


Fellowship, to The Ohio State University for a Professional Leave, and to
Professor E. Muetterties and the Chemistry Department at the University
of California, Berkeley, for their hospitality during the period this chapter
was written.

REFERENCES

1. (a) 1. Chatt, Advances in Organometallic Chemistry, edited by F. G. A. Stone and R.


West (Academic Press, New York, 1974), Vol. 12, pp. 1-29; (b) G. Booth, Advances in
Inorganic Chemistry and Radiochemistry, edited by H. 1. Emeleus and A. G. Sharpe
294 DEVON W. MEEK

(Academic Press, New York, 1964), Vol. 6, pp. 1-69; (c) A. Pidcock, Transition Metal
Complexes of Phosphorus, Arsenic and Antimony Ligands, edited by C. A. McAuliffe
(Wiley and Sons, New York, 1973), pp. 1-31.
2. R. Mason and p. W. Meek, Angew. Chem. Int. Ed. Engl. 17, 183-194 (1978).
3. D. W. Meek, Strem Chemiker 5,3-11 (1977).
4. For reviews on the spectral and magnetic data of transition metal complexes of polydentate
ligands, see: (a) c. Furlani, Coord. Chem. Revs. 3, 141-167 (1968); (b) M. Ciampolini,
Structure and Bonding (Berlin) 6, 52-93 (1969); (c) J. S. Wood, Progress in Inorganic
Chemistry, ed~.ed by S. J. Lippard (Wiley-Interscience of John Wiley and Sons, New
York, 1972), 01. 16, pp. 227-486.
5. D. W. Meek a d T. J. Mazanec, Acct. Chem. Res. 14,266-274 (1981).
6. L. Sacconi, Tr nsition Metal Chemistry, edited by R. L. Carlin (Marcel Dekker, New
York, 1968), Vol. 4, pp. 199-298.
7. N. J. DeStefano, D. K. Johnson, and L. M. Venanzi, Angew. Chem. Int. Ed. Engl. 13,
133-134 (1974).
8. M. Matsumoto, H. Yoshiska, K. Nakatsu, T. Yoshida, and J. Otsuka, 1. Am. Chem. Soc.
96,3322-3324 (1974).
9. (a) N. A. AI-s1em, W. S. McDonald, R. Markham, M. C. Norton, and B. L. Shaw, 1. C.
S. Dalton, 59- 3 (1980); (b) N. A. AI-Salem, H. D. Empsall, R. Markham, B. L. Shaw,
and B. Weeks, 1. C. S. Dalton, 1972-1982 (1979); (c) B. L. Shaw, Catalytic Aspects of
Metal Phosphi e Complexes, edited by E. C. Alyea and D. W. Meek, Advances in
Chemistry Series (Am. Chern. Soc., Washington, 1982), Vol. 196, pp. 101-115.
10. (a) G. S. Benner, W. E. Hatfield, and D. W. Meek, Inorg. Chem. 3, 1544-1549 (1964);
(b) G. Dyer and D. W. Meek, Inorg. Chem. 4, 1398-1402 (1965); (c) C. A. McAuliffe
and D. W. Meek, Inorg. Chim. Acta 5,270-272 (1971).
11. (a) L. sacconi,[oord. Chem. Revs. 8, 351-367 (1972); (b) L. Sacconi, 1. Chem. Soc. A.
1970, 248-25 ; (c) R. Morrassi, 1. Bertini, and L. Sacconi, Coord. Chem. Revs. 11,
343-402 (1973 .
12. L. M. Venanzi Angew. Chem. 76, 621-628 (1964); Angew. Chem. Int. Ed. Engl. 3,
453-460 (1964 .
13. D. L. DuBois and D. W. Meek, Inorg. Chem. 15, 3076-3083 (1976).
14. A. A. Arduini, A. A. Bahsoun, J. A. Osborn, and C. Voelker, Angew. Chem. Int. Ed.
Engl. 19, 1024-1025 (1980).
15. J. A. Osborn and G. S. Stanley, ibid., 19, 1025-1026 (1980).
16. J. J. deBoer, J~A. van Doorn, and C. Masters, 1. C. 5., Chem. Comm. 1978, 1005-
1006.
17. (a) J. Chat! an F. A. Hart, 1. Chem. Soc. 1960, 1378-1389; (b) W. Hewertson and
H. R. Watson, . Chem. Soc. 1962, 1490-1494.
18. M. Czakova and M. Capka, 1. Mol. Catal. 11,313-322 (1981).
19. Z. M. Michalska, M. Capka, and J. Stoch, 1. Mol. Catal. 11, 323-330 (1981).
20. G. Fritz, G. Becher, and D. Kummer, Z. anorg. aI/gem. Chem. 372, 171-179 (1970).
21. K. Issleib and H. P. Abicht, 1. Prakt. Chem. 312, 456 (1970).
22. H. H. KarSCh'[. Schubert, and D. Neugelbauer, Angew. Chem. Int. Ed. Engl. 18,484
(1979).
23. S. O. Grim an E. D. Walton, Phosphorus and Sulfur 9,123-126 (1980).
24. (a) N. Takaishi H. Imai, C. A. Bertelo, and J. K. Stille, 1. Am. Chem. Soc. 100,264-268
(1978); (b) T. asuda and J. K. Stille, 1. Am. Chem. Soc. 100, 268-272 (1978); (c) G. L.
Baker, S. J. Fritschel, J. R. Stille, and J. K. Stille, 1. Org. Chem. 46, 2954-2960 (1981);
(d) G. L. Baker, S. J. Fritschel, and J. K. Stille, 1. Org. Chem. 46, 2960-2965 (1981);
(e) S. J. Fritschel, J. J. H. Ackerman, T. Keyser, and J. K. Stille, 1. Org. Chem. 44,
3152-3157 (1979).
POL YDENTATE LIGANDS AND THEIR EFFECTS ON CATAL YSIS 295

25. (a) R. G. Nuzzo, S. L. Haynie, M. E. Wilson, and G. M. Whitesides, I. Org. Chern. 46,
2861-2867 (1981); (b) M. E. Wilson and G. M. Whitesides, I. Arn. Chern. Soc. 100,
306-307 (1978); (c) M. E. Wilson, R. G. Nuzzo, and G. M. Whitesides, I. Arn. Chern.
Soc. 100, 2269-2270 (1978); (d) R. G. Nuzzo, D. Feitler, and G. M. Whitesides, I. Arn.
Chern. Soc. 101,3683-3685 (1979).
26. (a) R. B. King, Acct. Chern. Res. 5, 177-185 (1972); (b) R. B. King and J. C. Cloyd,
Jr., I. Arn. Chern. Soc. 97, 53-60 (1975); (c) R. B. King, J. Bakos, C. D. Hoff, and L.
Marko, I. Org. Chern. 44, 3095-3100 (1979).
27. (a) D. L. DuBois, W. H. Myers, and D. W. Meek, I. C. S. Dalton, 1011-1015 (1975);
(b) R. Uriarte, T. J. Mazanec, K. D. Tau, and D. W. Meek, Inorg. Chern. 19, 79-85
(1980); (c) R. Uriarte, Ph.D. Dissertation, The Ohio State University, Columbus, Ohio,
1978; (d) R. J. Uriarte and D. W. Meek, Inorg. Chirn. Acta 44, L283-L284 (1980).
28. J. C. Briggs and G. Dyer, Chern. Ind., 163-165 (1982).
29. (a) E. P. Kyba, R. E. Davis, C. W. Hudson, A. M. John, S. B. Brown, M. J. McPhaul,
L.-K. Liu, and A. C. Glover,!. Arn. Chern. Soc. 103, 3868-3875 (1981); R. E. Davis, E. P.
Kyba, A. M. John, and J. M. Yep, Inorg. Chern. 19,2540-2544 (1980); (c) E. P. Kyba and S.
B. Brown, Inorg. Chern. 19, 2159-2162 (1980); (d) E. P. Kyba and S.-S. P. Chou, I. Org.
Chern. 46, 860-863 (1981).
30. M. Ciampolini, P. Dapporto, A. Dei, N. Nardi, and F. Zanobini, Inorg. Chern. 21,489-495
(1982).
31. (a) R. L. Keiter, R. D. Borger, J. J. Hamerski, S. 1. Garbis, and G. S. Leotsakis, I. Arn.
Chern. Soc. 99,5224-5225 (1977); (b) R. L. Keiter, Y. Y. Sun, J. W. Brodack, and L. W.
Cary, I. Arn. Chern. Soc. 101,2638-2641 (1979).
32. P. M. Treichel and W. K. Wong, 1. Organornet. Chern. 157, C5-C9 (1978).
33. R. L. Waid and D. W. Meek, unpublished data (1981); R. L. Waid, Ph.D. Dissertation,
The Ohio State University, Columbus, Ohio, June, 1982.
34. R. L. Keiter, S. L. Kaiser, N. P. Hansen, J. W. Brodack, and L. W. Cary, Inorg. Chern.
20,283-284 (1981).
35. (a) B. R. James, Hornogeneous Hydrogenation (John Wiley and Sons, New York, 1973);
(b) E. L. Muetterties, ed., Transition Metal Hydrides (Marcel Dekker, New York, 1971).
36. D. L. DuBois and D. W. Meek, Inorg. Chirn. Acta 19, L29-L30 (1976).
37. (a) 1. A. Osborn, F. H. Jardine, J. F. Young, and G. Wilkinson, I. Chern. Soc. A. 1966,
1711-1732; (b) c. A. Tolman, P. Meakin, D. L. Lindner, and J. P. Jesson, I. Arn. Chern.
Soc. 96, 2762-2774 (1974).
38. (a) J. Niewahner and D. W. Meek, Inorg. Chirn. Acta 64, L123-L125 (1982); (b) J.
Niewahner and D. W. Meek, A. C. S. Advances in Chernistry Series 196,257-272 (1982).
39. (a) T. J. Mazanec, J. B. Letts, and D. W. Meek, I. Chern. Soc., Chern. Cornrnun. 356
(1982); (b)J. B. Letts, T. J. Mazanec, and D. W. Meek, I. Arn. Chern. Soc. 104, 3898-3905
(1982).
40. L. H. Slaugh and R. D. Mullineaux, United States Patent No.3 239 566 (1966).
41. 1. L. Eiseman, United States Patent No.3 290 379 (1966).
42. J. A. Osborn, G. Wilkinson, and J. F. Young, I. Chern. Soc., Chern. Cornrn. 1965,17-18.
43. D. Evans, G. Yagupsky, and G. Wilkinson, I. Chern. Soc. A. 1968,2660-2665.
44. D. Evans, J. A. Osborn, and G. Wilkinson, I. Chern. Soc. A. 1968, 3133-3142.
45. M. C. Baird, J. T. Mague, J. A. Osborne, and G. Wilkinson, I. Chern. Soc. 1967,
1347-1360.
46. C. K. Brown and G. Wilkinson, I. Chern. Soc. A. 1970,2753-2764.
47. (a) F. B. Booth, United States Patent NO.3 511 880 (1970); (b) K. L. Oliver and F. B.
Booth, Hydrocarbon Processing 49 (4),112-114 (1970); (c) A. Hershman, K. K. Robinson,
1. H. Craddock, and 1. F. Roth, Ind. Eng. Chern. Prod. Res. Dev. 8, 372-375 (1969);
(d) J. H. Craddock, A. Hershman, F. E. Paulik, and J. F. Roth, ibid. 8, 291-297 (1969).
296 DEVON W. MEEK

48. R. L. Pruett and 1. A. Smith, f. Org. Chern. 34,327-330 (1969).


49. K. L. Oliver and F. B. Booth, Am. Chern. Soc. Pet. Div. Prepr., Gen. Pap. 14 (3), A 7 (1969).
50. (a) 1. D. Unruh and 1. R. Christenson, f. Mol. Caral. 14, 19-34 (1982); (b) O. R. Hughes
and J. D. Unruh, f. Mol. Caral. 12,71-83 (1981); (c) J. D. Unruh, O. R. Hughes, J. R.
Christenson, and D. Young, in Proceedings of Eighth Conference on Catalysis in Organic
Synthesis, edited by W. R. Moser, in press; (d) J. D. Unruh and J. R. Christenson,
Organometallic Reactions and Syntheses, edited by E. I. Becker and M. T. Tsutsui, Vol.
7 (1982); (e) O. R. Hughes and D. A. Young, f. Am. Chern. Soc. 103, 6636-6642 (1982).
51. B. Comils, "Hydroformylation, Oxo Synthesis, Roelen Reaction," New Syntheses with
Carbon Monoxide, edited by J. Falbe (Springer-Verlag, Berlin, 1980), pp. 1-225.
52. R. L. Pruett, "Hydroformylation," Advances in Organometallic Chemistry, edited by
F. G. A. Stone and R. West (Academic Press, New York, 1979), pp. 1-60.
53. G. M. Kosolapoff and L. Maier, eds., Organic Phosphorus Compounds (Wiley-Interscience
of John Wiley and Sons, New York, 1972), Vol. 1; (b) B. W. Bangerter, R. P. Beatty,
J. K. Kouba, and S. S. Wreford, f. Org. Chern. 42, 3247-3251 (1977), and references
therein.
54. R. L. Keiter, J. W. Brodack, R. D. Borger, and L. W. Cary, [norg. Chern. 21,1256-1259
(1982).
55. (a) C. A. Ghilardi, S. Midollini, A. Orlandini, and L. Sacconi, [norg. Chern. 19, 301-306
(1980); (b) M. Di Vaira, S. Midollini, and L. Sacconi, f. Am. Chern. Soc. 101, 1757-1763
(1979); (c) M. Di Vaira, C. A. Ghilardi, S. Midollini, and L. Sacconi, f. Am. Chern. Soc.
100,2550-2551 (1978).
56. (a) P. Dapporto, S. Midollini, and L. Sacconi, [norg. Chern. 14, 1643-1650 (1975); (b)
C. Bianchini, P. Dapporto, C. Mealli, and A. Meli, [norg. Chern. 21, 612-615 (1982).
57. M. M. Harding, B. S. Nicholls, and A. K. Smith, f. Organornet. Chern. 226, C17-C20
(1982).
58. (a) K. D. Tau and D. W. Meek, f. Organornet. Chern. 139, C83-C86 (1977); (b) K. D.
Tau, R. Uriarte, T. 1. Mazanec, and D. W. Meek, f. Am. Chern. Soc. 101,6614-6619
(1979).
59. A. F. M. 1. Van Der Ploeg and G. Van Koten, [norg. Chirn. Acta 51, 225-239 (1981).
60. K. Kurtev, D. Ribola, R. A. Jones, D. J. Cole-Hamilton, and G. Wilkinson, f. Chern.
Soc., Dalton 55-58 (1980).
61. P. Giannoccaro, G. Vasapollo, and A. Sacco,!. C. 5., Chern. Cornrnun. 1136-1137 (1980).
62. S. G. Davies, H. Felkin, and O. Watts, f. C. 5., Chern. Cornrnun. 159-160 (1980).
63. M. C. Davies and T. A. George, f. Organornet. Chern. 224, C25-C27 (1982).
64. J. A. Baumann and T. A. George, f. Am. Chern. Soc. 102,6153-6154 (1980).
65. R. B. King and W. M. Rhee, [norg. Chern. 17, 2961-2963 (1978).
66. M. D. Fryzuk, Organornetallics 1,408--409 (1982).
67. R. B. King and 1. W. Bibber, [norg. Chirn. Acta 59,197-201 (1982).
9
Cationic Rhodium and
Iridium Complexes in
Catalysis
Robert H. Crabtree

1. INTRODUCTION

Among the factors that affect the reactivity of transition metal com-
pounds, the net charge carried by the complex has up to now tended to
be neglected. In this review, we will first survey the effects of net charget
on the reactivity of metal complexes in general, before going on to the
specific case of cationic complexes of rhodium and iridium in catalysis. We
will not discuss work bearing on asymmetric hydrogenation, since this is
covered in Chapter 3.
While we will emphasize the effect of overall charge, particularly of
positive charge, we cannot ignore such well-recognized effects as the oxida-
tion state of the metal or the nature of the coligands. Overall charge is
just one factor among these. We will see, however, that it can in some
cases be the major factor determining reactivity.

j- By net charge, we refer to the charge carried by the complex as a whole. RhClL, has no
net charge, RhL4 + a net charge of + 1.

Dr. Robert H. Crabtree • Sterling Chemistry Laboratory, Yale University, 225 Prospect
Street, New Haven, Ct. 06511.

297
298 ROBERT H. CRABTREE

2. THE EFFECTS OF NET IONIC CHARGE

2.1. Effects on the Reactivitv of Bound Ligands

The charge on a complex cation or anion tends to be delocalized on to


the ligands in accordance with Pauling's electroneutrality principleY)
Hydrogens in the immediate coordination sphere of a cation, for example,
tend to become positively charged. This is one reason for the striking ability
of H 2 0 or NH3 to stabilize the simple cations of the transition metals (e.g.,
Cu(H 2 0)l+), compared to the far poorer abilities of even relatively unhin-
dered alcohols, ethers, and primary, secondary, and tertiary amines. One
of the roles a metal can play in catalysis is to change the normal reactivity
pattern of the substrate on going from the free state to the bound state.
Clearly one would expect electrophilic behavior of a substrate to be
encouraged by binding to a cation, and nucleophilic behavior by binding
to an anion.
The activation of the nitrile group, RCN, by metals is a good example.
Free nitriles are very resistant to nucleophilic attack by both BH4 - and
H 2 0. Coordination to ruthenium in a complex with an overall charge of
2+ has a dramatic effect. The nitrile can now be very rapidly reduced by
NaBH 4/EtOH as follows:

In this case, the BH4 - attacks the nitrile, transferring two H- groups to
carbon. The nitrogen atom is then protonated in the ethanolic medium,
releasing the free amine.(2) Catalysis of RCN reduction by NaBH 4/EtOH
has been observed for a variety of simple metal ions. The same mechanism
probably applies. (3a) In the catalyzed hydrolysis of nitriles the metal ion
plays the same role, but the nucleophile is now H 2 0. (3b)
The same type of behavior can be found for nucleophilic attack of
coordinated olefins and arenes. The C=C group of (cod)PdCI 2 is attacked
only by basic MeOH, while [(cod)Pd(PR 3 )Clt is rapidly attacked even in
neutral MeOH. (4) Hydride addition to coordinated pi-ligands tends to be
controlled by the charge on the metal, for example(S):

Re(C6H6h + ~ Re(C 6 H 6 )(C 6 H 7) ~

Fe(C 6 H 6 h 2 + ~ Fe(C 6 H 6)(C 6 H 7t ~ Fe(C 6 H7h

Overall positive charge also tends to increase the acid dissociation


constant of a bound ligand. The best known example is that of water. Many
CA T10NIC RHODIUM AND IRIDIUM COMPLEXES 299

metal ions are so polarizing that the H 2 0 ligand immediately loses protons
on binding.

Transition metal ions having fewer positive charges are far less polarizing,
but the charge effect can still be seen. (6)

not isolated

A single positive charge seems to be insufficient to cause ionization of the


coordinated water. (7)

2.2. Effects on the Tvpes of Ligand Bound

An overall positive charge on a complex most directly affects the


binding of charged ligands. In particular, anionic ligands, or ligands which
can dissociate a proton to give anions, are bound very strongly. This is one
reason why the tetra-anion of ethylenediaminetetracetic acid is so effective
at sequestering metal ions of high-positive charge.
The same effect can make BPh 4 - a coordinating anion, to the extent
that it can even displace tertiary phosphine(7):

Its use as a counter ion in catalytic studies can lead to catalyst deactivation(S)
for this reason.
Strong effects are also seen in the binding of neutral ligands. As
mentioned above, ligands with several H-substituents are very effective at
de localizing positive charge. This feature increases the binding constant of
H 2 0, for example, to the extent of giving relatively stable aquo-complexes
even with soft metal ions, which on hard-soft considerations would not be
expected to bind well. Some examples of hard-ligand complexes of soft
metals follow:

[(C 6 H 6 )Os(H 20hf+(O', Pddpe (thflz 2+( "", Rh(CsMes)(Me2COh 2+(6(,

Pd(MeCN)4 2+(\ ", IrH 2(PPh 3 h(H 20lz +(7(, Rhdpe(MeOHlz +( (2(.

These solvent ligands can be quite labile, and, for this reason, the solvento-
complexes themselves are often useful catalyst precursors, as discussed in
a recent review by Davies and Hartley. (10)
300 ROBERT H. CRABTREE

Among complexes of harder ions, such as those in the first row, the
absolute tendency to bind hard ligands such as H 20 is greater, yet the
same trends mentioned above seem to apply. The vast majority of isolable
aquo complexes, for example, bear a net positive charge.
The binding of dmso to metals by sulfur or oxygen is often thought
to be dictated by the hardness or softness of the metal. Yet Rh(I), Pd(II),
and Pt(II), normally soft, often bind via O. This often occurs in complexes
of net positive charge, e.g., [Rh(cod)(0-Me2S0hr, [Pd(0-Me2S0h(S-
Me2S0hf+, [Pd(dpe)(0-Me2S0hr or [Pd(0-R2SO)4]2+ (R = isoamyl),
while in cationic or neutral complexes S -binding often occurs, e.g.,

[Rh(S- Me 2 S0 h(CO)CI].(13)

The arguments we have presented suggest that one reason for 0-


binding in the cationic systems may be the ability to de localize positive
charge in the product as illustrated below. Other solvents can de localize
positive charge; for example, dmf probably binds via an oxygen lone pair,
in which case charge can be delocalized on to nitrogen.

M+ ....... O=SR 2 - M -O-SR2 +


M+ ....... O=CHNR 2 - M -O-CH=NR2 +

Clearly, a continuous range of ligand-binding ability exists. This will


change with the metal and coligands. So one might ask when a solvent is
coordinating, and when it is not. We will refer to a solvent as noncoordinat-
ing when the solvent does not appear to occupy a vacant site in the inner
coordination sphere, as weakly coordinating when solvento-complexes can
be detected or isolated but the substrate(s) for the catalytic reaction can
displace the solvent, and as strongly bound when no detectable solvent
dissociation in favor of the substrate takes place.
Anomalous displacement effects can sometimes be seen among
classical ligands in ionic complexs. For example, H2 would not normally
be expected to displace CO, but rather the reverse. In a cationic system,
however, the positive charge on the metal seems to weaken the M -CO
bond more than the MH2 system so that H2 does become able to displace
CO in a series of cationic iridium complexesY4l

Probably, CO is a better overall acceptor than is H 2 • Indeed we have argued


that H2 addition to some cationic iridium species can result in net electron
transfer to the metal (metal reduction) rather than away from it (oxidation),
CA TfONIC RHODIUM AND IRIDIUM COMPLEXES 301

as implied by the classical nomenclature "oxidative addition." This is


consistent with the idea that part of the net positive charge can be delocal-
ized over the H-ligands, which become 8+ rather than 8- in character.
In the case of alkene binding, ethylene normally binds better than,
say, butene, so that an ethylene dimerization catalyst would not usually
give C6 or higher products because CZ H4 would exclude butenes from the
metal site. Sen(!5) has decribed interesting catalytic reactions of
[Pd(MeCN)4f+ in which the reverse seems to be true. The electrophilic
character of the metal encourages the binding of substituted alkenes rather
than C 2H 4, and so C6 , Cs, and even C lO , species are formed directly from
C2 H 4 .
One might imagine that a charged mononuclear complex might be
incapable of agglomeration to give polynuclear species, because of coulom-
bic repulsion. This does not seem to be always the case. Cluster formation
may be slowed but is not prevented in 1+ species. This process can lead
to the deactivation of cationic catalysts. (16) The presence of potentially
bridging groups (RS, CI, H) can encourage oligomerization, especially of
systems also containing weakly bound ligands, which can be displaced by
the bridging groups.

2.3. Effects on Redox Properties

The main point of interest for our purposes is the greater resistance
to oxidation shown by cationic complexes. They can be less oxidation- and
air-sensitive than their neutral analogues. As extreme examples,
[CpIrHL 2]BF4 (L = PPh 3) can be refluxed in CCI4/CHCh for 3 hr in a
vessel open to the air without change, and [1r(cod)L 2 ]BF4 (L = PMe2Ph)
seems to be unique among hydrogenation catalysts in being catalytically
fully active even in H Z/0 2 mixtures.(7) The iso-electronic CpIrL2 and
Ir(cod)LCI, in contrast, are air-sensitive.

2.4. Effects on Solubility and Catalyst Separation

Since all the complexes under discussion are ionic, it is not surprising
to find that most are soluble only in polar solvents. This can be very useful
in separating the desired product from the spent catalyst. For example,
Professor Bill Suggs has shown that our catalyst, [Ir(cod)PCY3(py)]PF6 , is
useful in certain steroid reductions. Here, the reaction solvent is CHzCh.
To isolate the steroidal product, the solvent is removed and the residue
extracted with hexane. The extract contains only the steroid and not the
insoluble catalyst. (17)
An ion exchange resin has been used to immobilize a cationic
catalyst. (1S) Although in this case the cationic center was formed by
302 ROBERT H. CRABTREE

protonation of a remote nitrogen group in a P-bound P(C6H4NMe2h


ligand, the same effect could probably be obtained with complex having
an overall change.
Anionic catalysts should have the same advantages mentioned above.
Organometallic species are not normally water-soluble, but net charge
encourages solubility in H 20, particularly if some hydrogen-bonding func-
tionality is present. Water-soluble catalysts may well prove to be most
useful in two-phase systems, e.g., hexane/water, because catalyst separation
then becomes much easier than in classical systems. Examples of water-
soluble systems are [RhL 4]BF4 (L = I-phospha-3,5,7-tri-azaadaman-
tane)(19) and [Rh(cod)(PPh2C 6H 4S0 3h], which is a zwitterionic monomer
in MeCN, but a P,O-bridged dimer in the solid state.(20)

2.5. Counter Ion Effects

Some anions are more coordinating than others. The rough order of
increasing binding ability Rh(I) and Ir(1) 1+ systems seems to be SbF 6 , PF6,
BF4, CF3S0 3, CI04, CF3C0 2, CH3C0 2, BPh4 • Some transition metal
systems show substantial anion effects even between such apparently nonco-
ordinating anions as BF4 and PF6. In our alkane activation experiments(21)
(see Section 4.3), for example, BF4 salts were found to be three times
more effective than PF6 salts. Curiously, in alkene hydrogenation with a
closely related system, the PF6 salts are slightly superior to the BF4 salts. (7)
Such effects are relatively rare, however, and are to be expected only where
the substrate is a poor ligand.

2.6. Mechanistic Effects

The overall charge on a complex ion can dictate the mechanism of an


ionic reaction. For example, oxidative addition of HCI normally occurs by
protonation, followed by anation, as exemplified by(22)

In some cationic iridium complexes, the reaction occurs by the inverse


mechanism (23):

t
[Ir(cod)L 2 ...9..::.. IrCl(cod)L2 t
~ [IrHCl(cod)L 2

More than just charge effects are involved, since some related neutral
complexes also add HCl in this way, but in the neutral cases the intermedi-
ates are not stable enough to be isolated, in contrast to the cationic system
shown above.
CA TIONIC RHODIUM AND IRIDIUM COMPLEXES 303

By increasing the affinity of the catalyst for substrate functional groups


such as C=O, -NH2' and -OH, a positive charge may assist catalytic
reactions such as asymmetric hydrogenation where such chelate-formation
has been shown to be important. (24) Solvolysis of M -Cl bonds to give
cationic species may account for the increase in optical yields some-
times encountered with RhCIL n systems on going to more ionizing
solvents. (25)
Normally, 20e species are so unstable that they are not found. In one
case where a transition state of this type has been postulated, axial-
equatorial site-exchange in [NiBrL3JBr (L = PMe3), attack of the Br- ion
on the organometallic cation takes place. (26) This attack may well be favored
by the neutralization of charge in the transition state.

2.7. Conclusion

Just as one can vary the chemical behavior of a metal by changing the
coligands, one can also do so by changing the overall charge. Increasing
positive charge makes the metal more polarizing, oxidizing, and harder,
and makes the complexes much less soluble in nonpolar solvents. This is
a variable that has perhaps received less study than it merits.

3. SYNTHETIC METHODS

3.1. Halide Abstraction

Chloride abstraction is perhaps the most generally useful method of


forming cationic organometallic complexes. Where good ligands (e.g., PR 3,
CO, C2F 4) are to replace the halide in a labile system, often all that is
needed is the ligand, a polar solvent, and a simple salt of the counter-ion(27)
(e.g., NH 4PF6 or NaBF4).
When a poor ligand (e.g., H 20, Me 2CO, or C 2H 4) is to replace the
halide, or the metal is substitution inert, this method is less likely to work.
AgPF 6 in a polar solvent is often useful(28) but suffers from the disadvantage
that Ag + is a good oxidizing agent. The formation of metallic Ag indicates
that an alternative method must be used. TIPF6 is suitable,(29) since it is
far less oxidizing. [Me30JPF6 has also been used;(30) this has the advantage
that the products, MeCl and Me 20, can be pumped away. This is simpler
than the removal of the fine precipitates of AgCI and TICI that are obtained
in the above methods.
Chloride abstraction by such reagents as SbCl 5 and AlCh is relatively
unusual.(31)
304 ROBERT H. CRABTREE

3.2. Protonation

Protonation is an excellent method of producing cationic complexes.


Quite a large number of complexes MXxLy will proton ate, especially if a
lone pair is available (i.e., where x + y < 9 and x < max valency of metal).
In the particular case of cationic rhodium and iridium systems,
examples(7) are:
IrHsLz ~+ ~ IrHzSzL z + (S = coordinating solvent)

RhHL4 J£... RhL4 + + Hz


Sometimes an excess of acid can be used, but where an equivalent is
required, it has been found(32) that hydrolysis of an equivalent of [Ph 3C]PF6
is a convenient acid source, since the organic salt can be weighed out
precisely. Where water must be avoided, the etherate HPF 6 .Et 20 can be
used. Methylation with Me 30+ should be completely analogous, although
it has not been widely used.

3.3. Electrophilic Attack at a Ligand

Electrophilic attack at a coordinated ligand is also useful. Some


examples(33.34) follow:
Ru(COzMehL z + HA ~ [RuL 2 S x f+ + MeCOOH
(S = solvent; A = PF6 , BF4; L = PPh 3 )
Rh(cod)(acac) Ph3C+Lor H+ ~ [Rh(cod)Lzt

3.4. Redox

Oxidation of organometallic species electrochemically, or by chemical


oxidants such as FeCh, Br2, or Ph 3 C+, is still rare but will probably grow
more important in the future. No catalytic applications have yet been
described.
Oxidation of the bulk metal itself by a noncoordinating oxidant has
been used(11) but, among the platinum metals, is probably limited to Pd.
NOPF6
Pd ~ [Pd(MeCN)4l(PF6 h + 2NO

4. CA TIONIC RHODIUM AND IRIDIUM COMPLEXES IN


CATALYSIS

4.1. Hvdrogenation of C=C Groups

During the development of RhCI(PPh 3h as the first efficient alkene


hydrogenation catalyst by Wilkinson et al.,(35) it was realized that only two
CA T/ONIC RHODIUM AND IRIDIUM COMPLEXES 305

of the three PPh 3 groups remained permanently bound to the metal during
the catalytic cycle. An early inference from this idea was that if the P: Rh
ratio (n) could be adjusted to 2, rather than 3 as in RhCl(PPh 3h itself, a
more efficient catalyst might be obtained. The effect of a variable P: Rh
ratio was studied using mixtures of PR 3 and [Rh(cod)Clh as catalyst
precursors. Not only was n = 2.0 found to be the most effective ratio, but
those phosphines, such as PEt 3, which were totally inactive at n = 3, became
active at n = 2. (36)
Osborn, (37) who had contributed to this work at Imperial College,
subsequently took up the problem at Harvard. He made the important
discovery of the existence and catalytic activity of the first examples of the
cationic rhodium and iridium complexes which have been used the most
extensively in catalytic studies. For example, the readily available
[Rh(diene)Clh reacts with excess PPh 3 in EtOH to give [Rh(diene)-
(PPh 3ht, which can be isolated as the perchlorate, tetrafluoroborate, or
hexafluorophosphate. These are air-stable, crystalline complexes, and very
convenient for use as catalyst precursors. Under H 2, the diene is hydroge-
nated, generating the reactive RhL2 + fragment.
As expected, this is an excellent catalyst and it has several unusual
properties that depend in part on its ionic character. For example, inter-
mediates can be isolated relatively easily from the coordinating solvents
that Osborn used. These complexes [RhH 25 2L 2]A (5 = solvent (Me2CO,
EtOH); A = noncoordinating anion (Cl0 4 , PF6, BF4 )) have two solvent
ligands bound through their oxygen lone pairs in the immediate coordina-
tion sphere of the complex. This relatively firm binding of hard ligands by
a soft organometallic center seems to be particularly favored by a positive
charge on the complex, as mentioned above, and by the presence trans to
5 of the strong trans-influence hydride ligands (the antisymbiotic
effect). 10
Interesting selectivity effects were found for various dienes and alkynes,
the reduction of both of which could be stopped at the monoene stage.
Almost exclusive cis-addition of H2 to the alkyne was observed. Ketones
can be reduced, although this is not a result of the cationic character of
the system, since the catalysts can be deprotonated with NEt3 to give
neutral species that also effect this reduction equally well. This illustrates
a possible ambiguity in studying cationic catalysts if dissociation of a positive
species, usually H+, can occur. Fortunately, in this particular case, each
system can be separately prepared and studied.(37·38)
As might be expected in a system that binds solvents well, rates of
reduction of various substrates are very solvent dependent, (37) the substrate
being in competition with the solvent for the metal.
In contrast to the relatively high solvent dependence, the Osborn
system shows a relatively small dependence on L. Wilkinson's catalyst
requires PPh 3 or ligand of very similar cone angle, because the steric
306 ROBERT H. CRABTREE

compression of these relatively bulky phosphines encourages ligand dissoci-


ation, which is needed to generate a site at the metal. The smaller PEt 3,
as mentioned above, is inactive at n = 3 but active for n = 2. It seems
likely that, as in the Wilkinson system, two Ls remain bound to the metal
through the catalytic cycle in the Osborn system, too. No dissociation of
L is required, so the steric requirement is relaxed and L can be varied
rather widely. Simple amines and nitriles can be used in certain circum-
stances, for example. (38)
Even [Rh(diene)L 3 ]A salts were shown to be active,(37) probably
because alkene hydrogenation requires no more than three sites at the
metal, two for hydrogen and one for the substrate. Induction periods are
sometimes observed in these cases, no doubt because L(or cod) does not
have to dissociate before H2 can add. Compared to the Wilkinson system,
the side-reaction leading to isomerization of the alkene (e.g., I-butene
2-butene) is a more serious competitor with hydrogenation itself. In applica-
tions requiring specific deuteration or where the isomerization product is
nonhydrogenable, this can be a problem. (7)
[RhlnbdllPPh 3 )2r
3-cyclohexenal ---=--a-ce-:-to-ne~'------.. l-cyclohexenal
RhCllPPh,h
3-cyclohexenal ethanol - - • cyclohexanal

Osborn et at.(37) showed that the active isomerization catalysts were


t
indeed the dihydrides [RhH 2S 2L 2 and not [RhS 2L 2 which were pre- t,
pared by an independent route.

However, the addition of acid to the catalyst system (70% aq HCI0 4 )


partially suppresses isomerization, while the addition of base (NEt3)
encourages it. This suggests that neutral monohydrides are formed from
RhH 2S2L 2+ with base and that these are the more active isomerization
catalysts. It also suggests that the untreated normal Osborn catalyst is really
a mixed system containing solvated RhH2L2 + and RhHL 2. Possibly the
presence of the extra equivalent of dissociated PPh 3 in the Wilkinson system
helps prevent reversal of the first hydrogen transfer (leading to isomeriz-
ation) by coordination of L to the metal (see Figure 9.1). In the Osborn
system, step (a) can be followed by step (b), leading to isomerization. In
the Wilkinson system, the extra equivalent of L, which dissociates during
the catalytic cycle, may reassociate [step (c)]. This may help prevent {3-
elimination in the alkylrhodium intermediate. The Osborn system was
found to catalyze HID exchange between D2 and H 20, suggesting that
the same deprotonation process in the RhH 2S2L 2+ cations is responsible.
CA T10NIC RHODIUM AND IRIDIUM COMPLEXES 307

t->
H H
I
M-H (a) • ~ ~
I L
L
-~IL

H
I
M-H
I
I \
Figure 9.1. A proposed scheme for olefin isomerization.

Exchange between O 2 and the ortho- hydrogens of P(OPhh was also obser-
ved, suggesting that reversible ortho- metallation occurs.
Oeprotonation tended to be suppressed in systems containing the more
basic phosphines, such as [Rh(diene)(PMe2PhhJA. Using a combination of
acidic conditions and a basic phosphine, isomerization was substantially
reduced (e.g., for 1-hexene k isom./k hydrog. = 0.03 for L = PMe2Ph,
HCI0 4 : 1.4 mol-equiv, compared to kJkh = 1 for L = PMePh 2, no HCI0 4
added).
Our own work(16) took these results as a starting point. We were
particularly interested to discover what would be the effect of using a
solvent that would be less coordinating than those used by Osborn, and,
in particular, less coordinating than the substrate olefins. After many
failures, George Morris, then a graduate student in Paris, at last found the
successful solvent: CH 2Ch. It seems that the solvent must be polar so that
it can dissolve the ionic catalyst and may be slightly coordinating (see
below), though much less so than the substrate. For the rhodium system,
the results were encouraging but not dramatically different on changing
solvent. The reduction rates increased by six- to eightfold on going from
acetone to CH 2Ch. Oeprotonation could be effected with NEt 3 , as in
coordinating solvents, in which case reaction rates tended to fall threefold
or so, and isomerization rates tended to rise.
It was only when we extended our studies to the analogous iridium
complexes [Ir(cod)L2JA that exceptional rate effects were observed. Osborn
had found these to be poor hydrogenation catalysts in his coordinating
solvents and only briefly studied them, concentrating instead on the rhodium
system.
We found that in CH 2Ch, a rate acceleration of 10 3 _10 6 was observed
over the rates in acetone. Not only were monosubstituted alkenes reduced
308 ROBERT H. CRABTREE

very rapidly, but the much more hindered tri- and tetra-substituted alkenes
were reduced at similar rates. (39)
The ability to rapidly reduce hindered substrates makes the iridium
system unique among homogeneous catalysts. The rate acceleration ob-
served on going from Me2CO to CH 2Ch seems to be the result of a
changeover of active species (or more accurately, principal species present),
from IrH2(Me2COhL2 + in acetone to IrH 2(0lefinhL 2+ in CH 2Ch (olefin =
substrate). In acetone, the olefin has to displace Me2CO to be reduced, so
only unhindered ole fins that are good ligands are reduced. In CH 2Ch, that
restriction is lifted.
An interesting feature of the iridium system is that these olefin dihy-
dride intermediates can be observed directly by nmr at -80°. Two isomers
of (cod)lrH2L2 +, for example, have been isolated. Only the isomer contain-
ing a coplanar M(C=C)H grouping is active in hydrogenation; in the other,
the M(C=C) groups are orthogonal to the cis-M-H bonds, and insertion
is not possible. In this system, hydrogenation is observed, but no alkyl
hydride intermediate can be trapped. In contrast, no olefin dihydride
complexes have yet been observed for the analogous rhodium system, but
instead Halpern et al. (40) have successfully trapped the first alkyl hydride
intermediate.
A disadvantage of the iridium systems is that they deactivate under
certain conditions to give the unusual complexes: [L 2Hlr(/.L -HhlrHL2]A
and [(lrHLL'h(/.LrH)]A2 (L = PR 3 ; L = amine, or RCN). Certain solvents
(e.g., PhCl) seem to prevent deactivation without affecting the rate of
reaction very much, at least for [lr(cod)(PMePh 2h]PF 6 .(41) Presumably
these solvents bind less well than the olefin but better than CH 2Ch. We
suggest that even CH 2Ch itself can act as a chelate to iridium via the'
halogens. We have recently shown that the softer iodine can bind much
more strongly to these iridium systems, for example, in the novel complex
[lrH 2(o-diiodobenzene)L 2]A,(42a) in which a crystallographic study shows
that the aromatic ligand is indeed chela ted via the halogens. Such species
may be important intermediates in oxidative addition of RX to metals.
The catalysts have the activity of a heterogeneous catalyst in their
ability to reduce hindered olefins, but they differ sharply from these in
retaining the selectivity of homogeneous catalysts. Heterogeneous catalysts
tend to reduce 1 to 3 (cis and trans); classical homogeneous catalysts tend
to give 2 and only very slowly 3. Only the trans-isomer of 3 is formed in
the latter case. [lr(cod)(PCY3)Py]PF 6 catalyzes the reduction of 1 to 2 by
one equivalent of H 2, and of 2 to trans -3 by an excess. (17) These substrates
bind to the iridium to form complexes of the type [lrH 2(substratehL 2 r,
in which they appear to be 0- not C=C bound. The order of binding is
1 > 2 > 3. This helps account for the selectivity and suggests that unsatur-
ation promotes Ir-O binding. Indeed PhCH=CHC0 2Et binds so strongly
CA TlONIC RHODIUM AND IRIDIUM COMPLEXES 309

that the C=C function cannot be reduced, at least by [Ir(cod)-


(PMePh zh]PF 6 •
The catalyst is also insensitive to O 2 and oxidizing or other sensitive
functionality. For example, 4 is reduced cleanly to 5, while other types of
catalyst tend to cleave C-Br or cyclopropane C-C bonds, or both.

o o o

J 0
1 2 3

¢~¢
Br Br Br Br

4 5

[IrH 2 (thfhL 2 ]A systems, prepared in situ, are effective in the isomeriz-


ation of allylic alcohols to carbonyl compounds,(43) and isomerization is
always a side reaction in hydrogenation reactions with the iridium
catalyst. (42b)
Halpern(44) has made some significant observations in the chemistry
of the cationic rhodium catalysts. Because much of this relates to asymmetric
induction, it is treated in Chapter 4. It is appropriate to mention here that
the catalyst formed from [Rh(cod)(dpe )]A (dpe = 1,2-diphenylphosphino-
ethane) has been shown to dimerize reversibly in the absence of substrate.
An aromatic ring from one of the dpe ligands of the first metal binds in
an 1J 6 -fashion to the second metal and vice versa. (44) The system is labile,
in contrast to the nonlability of simple arene complexes of Ir(1).
Other cationic systems have been studied as hydrogenation catalysts.
For example, [Rh(NH 3)sH]2+ was found to reduce a variety of aromatic
acids in aqueous solution. (4S) Since metal precipitated in the absence of
substrate, it remains a possibility, however, that the organic substrates
stabilized the formation of colloidal Rh, itself an excellent catalyst.
For Maitlis's (46) interesting hydrogenation catalysts based on
[(C sMes)MX2 h (M = Rh or Ir), no evidence exists for the involvement of
cationic species, although this is by no means out of the question in
isopropanol, the ionizing, coordinating reaction solvent used.
310 ROBERT H. CRABTREE

Interestingly, [((CsMes)Mh HXz][HXz] catalyzes the air-oxidation


of isopropanol to acetone over 3d. at 20° (M = Rh, Ir).(46b)
Since supported homogeneous catalysts are fully described in Chapter
14, we only mention two particularly imaginative examples here, in both
of which the positive charge of the catalyst system has a role to play.
Whitesides et al. (47) have bound a RhLz + species to avidin, a globular
protein. Biotin, which binds strongly to avidin, was covalently attached to
a diphosphine fragment and allowed to bind to the protein. A [(nbd)RhSzt
fragment was then introduced; this presumably binds to the diphosphine.
The resulting catalyst gave ca 45% e.e., in the reduction of ex -acetamido-
acrylic acid. The success of this system requires the rhodium fragment to
be compatible in the aqueous environment; the cationic character of the
metal probably encourages this.
Catalysts of the RhLz + type have also been incorporated into the layer
silica mineral hectorite. Once again, the positive charge on the metal
strongly encourages binding. The resulting catalyst has unusual selectivity
patterns, no doubt due to the requirement that the substrate has to interca-
late before it can be reduced. (48)

4.2. Hydrogenation of C=O Groups

The reduction of the C=O function is relatively unusual among


homogeneous hydrogenation catalysts. [Rh(cod)(PPh3)]A-based systems
reduce various ketones, both in the native cationic form, where HzO is a
cocatalyst, (Z8,37) and in the deprotonated form. (38) This does not necessarily
occur via the enol, since PhzCO can be reduced by the neutral system in
this way.
So far, esters have only been reduced by anionic systems, such as
K[RuHz(PPh3h(PPhzo-C6H4)], where the hydrogen ligands presumably
take on a more strongly hydridic character by delocalization of the net
negative charge. (49) The role of the K+ counter-ion in polarizing the C=O
group by O-binding is probably also important. This interesting problem
has had less attention than it merits.
Ketones can be very efficiently reduced by H-transfer from i-PrOH
in presence of [Ir(cod)(phen)]Cl(SO) (phen = 1,10-phenanthroline). The
positive charge on the metal probably assists coordination of i -PrOH, the
first step of the reduction.
M+ + lPrOH -H'. MOWr _ MH + Me2eO

4.3. Alkane Activation

It is well known that C-H bonds in coordinated ligands can be cleaved


by metals. This cyclometalation reaction has been studied for many
CA TfONIC RHODIUM AND IRIDIUM COMPLEXES 311

years;(51) a classic example(52) is:

(Np)HRu{(MezPCHzhh -NpH I [(MezPCHzhRuH(MezPCHzCHzPMeCHz)]z


(Np = 2-napthyJ)

in which a relatively unactivated SP3 CH bond is cleaved by the metal.


A question that has not been answered until very recently is how this
reaction could be effected intermolecularly to cleave CH bonds in alkanes.
An early report by Shilov(53) et at. that PtC1 42- in CH 3COOD
homogeneously effected HID exchange attracted much attention. The
reaction may go via electrophilic attack of the metal on the alkane, but
the details are not yet clear. (54)
In a recent example(21) of an alkane dehydrogenation by a metal
complex, [IrH 2 S2 L 2 ]BF4 was shown to react with cyclopentane to give
[CpIrHL 2 ]BF4. Yields were increased by the addition of But CH=CH 2,
which accepts the hydrogens stripped off the metal. This olefin is unique
among those tried in taking up H2 from a metal, but in itself binding only
weakly to it. C2 H 4, in contrast, is bound to the metal so strongly as to exclude
the alkane almost completely, and only a trace of alkane dehydrogenation was
observed. Olefins containing allylic hydrogens are themselves dehydroge-
nated in preference to the alkane, and various styrene derivatives give the
unusual deactivation reaction shown below instead of alkane activation. (55)

Cycloheptane and cyclooctane gave "7 5_cycloheptadienyl and cyclooc-


tadiene complexes, respectively, but other alkanes, e.g., methy1cyclopen-
tane, adamantane, or bicyciooctane, give as yet uncharacterized products
which probably contain the dehydrogenated alkane.
Mechanistic experiments tend to exclude heterogeneous, radical, or
carbonium ion mechanisms. For example, dynamic light scattering
measurements fail to detect particulates in the reaction mixture, although
a 1 nanomolar Ir colloid can be easily detected. The reaction mixtures also
fail to catalyze the reduction of PhN0 2 to PhNH 2 by H 2 , a reaction that
does occur with authentic Ir colloids. Gas chromatography experiments
show that neither the carbonium ion rearrangements expected for t-butyl-
methy1carbonium ion nor the chlorine atom abstraction reaction expected
for cyciopentyl radical occur.
Cyclometalation might be expected to have occurred in preference to
alkane activation in these PPh 3 complexes. Recent work on other systems
has shown, however, that alkane activation can occur even in PPh 3 com-
plexes. (56)
312 ROBERT H. CRABTREE

It is to be expected that some ligands will prove to be more metalation-


resistant than others. Possibly, for example, the order of increasing metala-
tion-resistance for the following ligands might be:

This is supported by some important results recently obtained by Bergman


in which a variety of alkanes have been activated photochemically. Berg-
man(S7) has shown that [(CsMes)IrLH2] on photolysis in cyclohexane gives
[(CsMes)IrL(Cy)H], but that the cyclohexyl hydride formation is competi-
tive with cyclometalation for PPh 3 , but the major product for PMeJ.
It is possible that the positive charge on the metal in our cationic
iridium system gives it an electrophilic character, which helps C-H activa-
tion to occur. The Bergman system does not, at first sight, look electrophilic,
but this could still be the case if a photoexcited state were the active species.
That photoexcited states can be active in C-H chemistry is shown by Ozin's
discovery that Cu atoms photoexcited in a CH 4 matrix give CH 3 CuH as
the first-formed product even at 12 K.(S8)
Ion cyclotron resonance experiments(s<)) have also shown that positively
charged ions are more active than neutral atoms, e.g.,

Fe + alkane ~

Fe+ + alkane -. products

An electrophilic character is also found for low-coordination number


metal atoms at kinks and steps in metal surfaces.(60) These sites are also
known to be much more active for alkane reactions than the flat metal
surfaces. (61)
These results suggest that, at least in electrophilic complexes, electron
withdrawal from the C-H bond of the alkane may be more important than
electron donation into the C-HO'* orbital. The most likely transition state
in which this can be achieved would be T -shaped:

H
H : H
"'c';
I--- M
H

Slightly bent CH··· M systems have been observed in several


studies(62): intermediates of this type may lie on the route of approach to
the transition state.
There is currently much interest in cluster complexes as models for
surface chemistry. These studies have shown the unusual bonding modes
CA T10NIC RHODIUM AND IRIDIUM COMPLEXES 313

that are possible on metal arrays. Catalysis, however, requires the availabil-
ity of sites. The types of clusters usually studied, metal carbonyls, have
tended not to be catalytically active. One important aspect of the metal
surface, its lack of masking ligands, has not been modeled in the vast
majority of cluster studies. In contrast, ligand-deficient clusters containing
open sites or very labile ligands may prove to have useful catalytic activity.
Bifunctional catalysts in which two chemically different metals are close
together but not directly bonded may have more interesting properties.
Another feature of the metal surface that may be important for certain
reactions is its e1ectrophilic character. The interesting catalytic behavior of
the [M(diene)L 2 ]A systems, mentioned in this review, may in part depend
on the ligand-deficiency and electrophilicity of these systems. In these two
properties, they may parallel the properties of a metal surface. It remains
to be seen whether those cooperative effects that may occur are incidental
or essential to catalytic action in heterogeneous, as well as homogeneous,
systems.

4.4. Hydroformylation

A number of reports show that [Rh(cod)L 2]A complexes can be


precursors for alkene hydroformylation. (63) Both the cationic and deproton-
ated neutral systems are active. The deprotonated system appears to be
indistinguishable from that derived from RhH(CO)(PPh 3 h; the cationic
system is distinct, though broadly comparable.
A study of the rates and product ratios in the neutral system suggested
that dinuclear species may playa role in the catalytic cycle.

4.5. Oecarbonylation

Pignolet has shown that the trinuclear iridium compounds


[(IrL2H2h(lLrH)]A2 are catalysts for aldehyde decarbonylation.(64) The
use of cationic complexes in this connection is interesting. Decarbonylation
with RhCl(PPh 3 h (6) gives RhCl(CO)(PPh 3 h, and subsequent reaction is
very slow. CO loss is probably the slow step that limits the activity of the
usual catalysts. Indeed, 6 is usually used as a stoichiometric decarbonylation
reagent for this reason. Since an overall position charge should weaken
M-CO bonding, cationic complexes may prove to be catalytically active
under milder conditions than are the classical ones.

4.6. Water Gas Shift (WGS)

In the WGS reaction, CO is activated to nucleophilic attack by water.


Clearly cationic complexes are well suited to this task, and it is not surprising
314 ROBERT H. CRABTREE

to find that [lr(cod)L 2 ]A is an effective catalyst. (65)

OH
M-CO+ ~ M-C/ + H+ ...... M+ + H2 + CO 2
'\,
o
It is not yet clear if the system remains cationic throughout the catalytic
cycle; added base does not suppress the reaction.

4.7. Polvmerization
[Rh(NO)(MeCN)4]A 2 (66) is an interesting alkene oligomerization
catalyst, apparently involving allylic intermediates; it may resemble Sen'sOS)
[Pd(MeCN)4]A 2 •

5. CONCLUDING REMARKS

The effects of net charge on catalysis have not been given the emphasis
due them. In this review, I have tried to outline some of these effects, and
hope the ideas presented will provoke further studies in the area.

ACKNOWLEDGMENTS

I thank Gregory Hlatky for discussions, and Yale University for a


Junior Faculty Fellowship, which allowed me to write this review.

REFERENCES

1. L. Pauling, I. Chern. Soc. 150, 1461 (1948).


2. R. H. Crabree and A. J. Pearman, I. Organornetal. Chern. 157,335 (1978).
3. a) T. Satoh, S. Suzuki, Y. Suzuki, Y. Miyaji, and Z. Imai, Tet. Lett. 1969, 4555. b) S.
E. Diamond, B. Grant, G. M. Tom, and H. Taube, Tet. Lett. 1774,4025.
4. 1. Chatt, L. M. Vallarino, and L. M. Venanzi, I. Chern. Soc. 1957, 2496 and 3413; C.
Eaborn, N. Farrell, and A. Pidcock, J. C. S., Dalton 289 (1976).
5. S. G. Davies, M. L. H. Green, and D. M. P. Mingos, Tetrahedron 34,3047 (1977).
6. C. White, S. 1. Thompson, and P. M. Maitlis, I. Chern. Soc. 1977, 1654.
7. R. H. Crabtree and 1. M. Mihelcic, recent observations (1981).
8. L. M. Haines, Inorg. Chern. 19, 1685 (1971); G. Mestroni, G. Zassinovich, and A. Camus,
J. Organornetal Chern. 140,63 (1977).
CA TlONIC RHODIUM AND IRIDIUM COMPLEXES 315

9. Y. Hung, W-J. King, and H. Taube, Inorg. Chern. 20, 157 (1981).
10. J. A. Davies, F. R. Hartley, and S. G. Murray, f. C. S., Dalton 1980,2246.
11. R. F. Schramm and B. Wayland, Chern. Cornrn. 1968, 898.
12. J. M. Brown, P. A. Chaloner, and P. N. Nicholson, Chern. Cornrn. 1978, 646.
13. J. A. Davies, Adv. Inorg. Radiochern. 24, 116 (1981).
14. M. J. Church, M. J. Mays, R. N. F. Simpson, and F. P. Stefanini, 1. Chern. Soc., (A)
1970, 2909 and 3000.
15. A. Sen and T.-W. Lai, J. Arner. Chern. Soc. 103,4627 (1981).
16. R. H. Crabtree, Accounts Chern. Res. 12, 331 (1979).
17. J. W. Suggs, S. D. Cox, R. H. Crabtree, and J. M. Quirk, Tet. Lett. 22,303 (1981).
18. S. C. Tang, T. E. Paxson, and L. Kim, J. Mol. Cat. 9, 313 (1980).
19. R. Uriarte and R. H. Crabtree, recent observations (1981).
20. A. F. Bozowski, D. J. Cole-Hamilton, and G. Wilkinson, Nouveau Journal de Chirnie 2,
137 (1978).
21. R. H. Crabtree, J. M. Mihelcic, and J. M. Quirk, f. Arner. Chern. Soc. 101, 7738; (1979);
R. H. Crabtree, M. F. Mellea, J. M. Mihelcic, and J. M. Quirk, J. Arner. Chern. Soc.
104,107 (1982).
22. U. Belluco, M. Giustiniani, and M. Graziani, f. Arner. Chern. Soc. 89, 6494 (1967).
23. W. J. Louw, D. J. A. de Waal, and J. E. Chapman, Chern. Cornrn. 1977, 845; R. H.
Crabtree, J. M. Quirk, T. Khan-Fillebeen, and G. E. Morris, J. Organornetal. Chern.
157, C13 (1978).
24. A. S. C. Chan, J. J. Pluth, and J. Halpern, J. Arner. Chern. Soc. 102,5952 (1980).
25. T. Hayashi, T. Mise, S. Mitachi, K. Yamamoto, and M. Kumada, Tet. Lett. 1976,49.
26. P. Meier, A. E. Merback, M. Dartiguenave, and Y. Dartiguenave, Chern. Cornrn. 1979,
49.
27. J. R. Shapley, R. R. Schrock, and J. A. Osborn, f. Arner. Chern. Soc. 91, 2816 (1969).
28. R. R. Schrock and J. A. Osborn, 1. Arner. Chern. Soc. 98, 2134 (1976).
29. F. W. S. Benfield, M. L. H. Green, and B. R. Francis, f. Organornetal. Chern. 44, C13
(1972).
30. C. Eaborn, N. Farrell, and A. Pidcock, J. C. S., Dalton 1976,289.
31. N. G. Gaylord and H. F. Mark, Adv. Chern. Ser. 34, 127 (1962).
32. O. W. Howarth, C. H. McAteer, P. Moore, G. E. Morris, J. Chern. Soc., Dalton 1981,
1481.
33. R. W. Mitchell, A. Spencer, and G. Wilkinson, J. Chern. Soc., Dalton 1973,846.
34. M. Green, T. A. Kuc, and S. H. Taylor, Chern. Cornrn. 1970, 1553.
35. J. A. Osborn, F. H. Jardine, J. F. Young, and G. Wilkinson, J. Chern. Soc., (A) 1964,
1711; F. H. Jardine, J. A. Osborn, and G. Wilkinson, ibid 1967, 1574.
36. S. Montelatici, A. van der Ent, J. A. Osborn, and G. Wilkinson, J. Chern. Soc., (A) 1968,
1054.
37. R. R. Schrock and J. A. Osborn, Chern. Cornrn. 1970,567, and f. Arner. Chern. Soc. 13,
2397 and 3089 (1971); J. R. Shapley, R. R. Schrock, and J. A. Osborn, J. Arner. Chern.
Soc. 91, 2816 (1969); R. R. Schrock and J. A. Osborn, J. Arner. Chern. Soc. 98, 2134
and 2134 (1976).
38. R. H. Crabtree, A. Gautier, G. Giordano, and T. Khan, f. Organornetal Chern. 141, 113
(1977).
39. R. H. Crabtree, H. Felkin, and G. E. Morris, f. Organornetal Chern. 141,205 (1977).
40. A. S. C. Chan and J. Halpern, J. A mer. Chern. Soc. 102,838 (1980).
41. E. Martin and K. B. Wiberg, unpublished results (1980)
42. a) R. H. Crabtree, J. W. Faller, and M. F. Mellea, manuscript in preparation, (1981); b)
R. H. Crabtree, H. Felkin, T. Khan and G. E. Morris, J. Organornet. Chern. 168, 183
(1979).
316 ROBERT H. CRABTREE

43. D. Baudry, M. Ephritikine, and H. Felkin, Nouv. J. Chirn. 2, 355 (1978).


44. J. Halpern, D. P. Riley, A. S. C. Chan, andJ. J. Pluth,!. Arner. Chern. Soc. 99,8055 (1977).
45. A. R. Powell, Platinurn Metal. Revs 11, 58; (1967); G. C. Bond, British Patent no. 197
723 (Chern. Abs. 73,98401 (1970)).
46. a) P. M. Maitlis, Accts. Chern. Res. 11,301; b) C. White, A. J. Oliver, and P. M. Maitlis,
J. Chern. Soc., Dalton 1973, 1901.
47. M. E. Wilson and G. M. Whitesides, J. Arner. Chern. Soc. 100,306 (1978).
48. T. J. Pinnavia, R. Raythatha, J. G-S. Lee, L. J. Halloran, J. F. Hoffman, J. Arner. Chern.
Soc. 101, 6891 (1979).
49. G. Pez, Chern. Cornrn. 1980,783.
50. A. Camus, J. Mol. Cat. 6, 231 (1979).
51. M. 1. Bruce, Ang. Chern. (Int. Ed.) 16,73 (1977).
52. J. Chat! and J. M. Davidson, J. Chern. Soc. 1965, 843.
53. N. F. Gol'dschleger, M. D. Tyabin, A. E. Shilov, and A. A. S'hteinman, Zh. Fiz. Khirn.
43,2174 (1969).
54. A. E. Shilov, Sov. Sci. Revs B 4, 71 (1982).
55. R. H. Crabtree, M. F. Mellea, and J. M. Quirk, Chern. Cornrn. 1981, 1217.
56. D. Baudry, M. Ephritikine, and H. Felkin, Chern. Cornrn. 1980, 1243; M. A. Green, J.
C. Huffman, and K. G. Caulton, J. Organornetal. Chern. 218, C39 (1981).
57. R. G. Bergman, personal communication (1981).
58. G. A. Ozin, D. F. McIntosh, and S. A. Mitchell, J. Arner. Chern. Soc. 103, 1574 (1981).
59. J. Allison, R. B. Feas and D. R. Ridge, J. Arner. Chern. Soc. 101, 1332,4998 (1979).
60. T. H. Upton, W. A. Goddard, and C. F. Melius, J. Vac. Sci. Technol. 16, 531 (1979);
B. Krahl-Urban, E. A. Niekisch, and H. Wagner, Surf. Sci. 64, 52 (1977); J. Holzl and
F. K. Schulte, Springer Tracts Mod. Phys. 85, 1 (1979).
61. H. Wagner, Springer, Tracts Mod. Phys. 85, 151 (1979).
62. R. K. Brown, T. M. Williams, A. J. Schultz, G. D. Stucky, S. D. Ittel, and R. L. Harlow,
1. Arner. Chern. Soc. 102,981 (1980).
63. R. H. Crabtree and H. Felkin, J. Mol. Catal. 5, 75 (1979).
64. H. H. Wang and L. H. Pignolet, Inorg. Chern. 19, 1470 (1980).
65. J. Kaspar, R. Spogliarich, G. Mestroni, and M. Graziani, J. Organornetal. Chern. 208,
C15 (1981).
66. N. G. Connelly, P. T. Dragget!, and M. Green, J. Organornetal Chern. 140, C10 (1977).
10
Hydrogenation Reactions of
CO and CN Functions Using
Rhodium Complexes
Balint Heil, Laszl6 Mark6, and Szilard Toros

1. INTRODUCTION

Stimulating results achieved in hydrogenation of C=C double bonds with


phosphinerhodium complexes led to the application of these compounds
for the catalytic reduction of C=O and C=N double bonds. The present
review summarizes the results in this area reported up until now. To give
a more complete picture of the subject, some reports on iridium and
ruthenium phosphine complex catalysts have been included as well.
The catalytic reduction of C=O and C=N double bonds of a hydrogen
acceptor (A) molecule may be carried out by "direct" addition of hydrogen
(1), by transfer hydrogenation (2), or through hydrosilylation (3). In the
case of transfer hydrogenation, the hydrogen necessary for reduction is
abstracted from donor (D) organic compounds like secondary alcohols.
Hydrosilylation with mono- or dihydrosilanes leads to alkoxi- or alkyl-
aminosilanes, which subsequently have to be hydrolized.

(1)
(2)

Drs. Balint Heil and Laszl6 Mark6 • University of Chemical Engineering, Institute of
Organic Chemistry, Veszprem, Schonherz Zoltan u. 12, Hungary.
Dr. Szilard T6ros • Research Group for Petrochemistry, Hungarian Academy of Sciences,
Veszprem, Schonherz Zoltan u. 12, Hungary.

317
318 BALINT HElL. LASZLO MARKO. AND SZILARD T6ROS

(3)

Catalytic reduction of a carbon-oxygen double bond is generally more


difficult than that of an olefinic one. This experience is probably related
to the following two factors, among others:
a. the low stability-and accordingly small number-of complexes
between H2 activating metals and organic carbonyl compounds (as
compared to the vast number of metal olefin complexes), and
b. the ability of alcohols, the products of the reaction, to act as ligands,
in contrast to saturated hydrocarbons, which are inert against the
catalytically active complexes.
The efficiency of ketone reduction may be diminished also by transfer
hydrogenation, the secondary alcohol serving as a hydrogen source and,
consequently, the hydrogenation ending up in a reversible reaction deter-
mined by thermodynamic parameters. This is supported by papers describ-
ing dehydrogenation reactions with catalysts successfully used in ketone
reduction at lower temperatures. For example, Ohkubo(l) dehydrogenated
racemic 1-phenylethanol with a RhCI(NMDPPh catalyst prepared in situ,
and according to Fragale, (2) 1-phenylethanol and other secondary alcohols
are partly transformed to the corresponding ketones in the presence of
RhCI(C g H 12 )PPh3 + NaOH under nitrogen at 80-150°.

2. HYDROGENA TION OF ALDEHYDES

Wilkinson-type rhodium-phosphine catalysts are reported to be not


suitable for aldehyde reduction(3) because of decarbonylation of the sub-
strate under the reaction conditions and formation of RhCI(CO)(PPh 3)z,
which is catalytically inactive. In contrast to this, Fujitsu(4) achieved the
hydrogenation of phenylacetaldehyde with [Rh(NBD)(PR 3 )ntCI04 - (n =
2 or 3) catalysts under mild conditions. The catalytic activities
decreased in the order of PEt3 > PPh 3 - PMe3 »DPE complexes. The
total turnover number of the PEt3 complex was greater than 77.5 with
this aldehyde.
A very active aldehyde hydrogenation catalyst is formed by treating
IrH 3(PPh 3h with acetic acid. (4,5) While Coffe/ 5) hydrogenated n- butyral-
dehyde under mild reaction conditions, Strohmeier(6) obtained turnover
numbers up to 8000 in the reduction of several saturated and unsaturated
aldehydes with the same catalyst at 80-11 0° and 1 bar working without
solvent. The acetato complex IrH 2 (OOCCH 3 )(PPh 3h was suggested as
the active intermediate.
HYDROGENA TfON REACTIONS OF CO AND CN FUNCTIONS 319

Ruthenium catalysts have also been successfully used(7-10) for the


hydrogenation of aliphatic and aromatic aldehydes. Although decarbonyla-
tion of the substrate may not be avoided using RuCh(PPh3h or
RuHCl(PPh 3h, the carbonyl complex RuHCl(CO)(PPh 3h formed is, in
contrast to the corresponding rhodium derivative, also active in hydrogena-
tion.(9) Strohmeier reported turnover numbers up to 95,000 with a
RuCh(COh(PPh 3h catalyst. (10)
In hydrogen transfer from different organic compounds to aldehydes,
excellent catalytic activity of RuH 2(PPh 3)4 was observed by Japanese scien-
tists(1l) to override that of RhH(PPh 3k
Selective hydrogenation of a C=O bond in the presence of a C=C
bond was studied by Meguro(12) using cinnamaldehyde and croton aldehyde
as model compounds. With RhCh·3H 20 or Rh 2Ci 2(CO)4 and tertiary
amines in benzene, unsaturated alcohols were produced with high (up to
85 %) selectivity. Addition of PPh 3, however, retarded the reduction of the
formyl group, and saturated aldehydes were the products.

3. HYDROGENA TION OF KETONES NOT CONTAINING OTHER


FUNCTIONAL GROUPS

3.1. Catalytic Activity and Stereoselectivity


Homogeneous hydrogenation of ketones with rhodiumphosphine
catalysts was first carried out by Schrock and Osborn(13) in 1970. The
authors reduced acetone, acetophenone, cyclohexanone, and ethyl methyl
ketone with a [RhH 2P 2S2t A-type ionic complex (P = basic phosphine,
S = solvent) at 25° and normal pressure. They also found that the initial
rate of the reaction is markedly enhanced (up to 30 times) by adding 1%
of water, while reduction of ole fins is inhibited in aqueous solution.
The reaction mechanism suggested by the authors is presented below (see
Figure 10.1). Crabtree reported on the catalytic activity of [Rh(COD)-
(PPh 3htPF6- in benzene adding Et 3 NY4,15)
Preparing the active catalytic system in situ from [Rh(diene)Clh and
the corresponding tertiary phosphines,(16) mainly covalent rhodium-phos-
phine complexes are formed. Their activity is highly influenced by the
quality of the ligand used. Best results have been achieved with phosphines
of high basicity and low steric requirements. (17) RhCl(PPh 3h is completely
inactive under such conditions.
Gargano reduced ketones with RhCl(C sH 12 )PPh 3 and Rh 2H 2Ch-
(CsHu)(PPh 3h in presence of strong alkali. Pretreatment of these precur-
sors with NaBH4 increased the catalytic activity of the system further. (1S)
l
320 BALINT HElL, LAsZLO MARKO, AND SZILARD T6R6s

>t<j" r"~/(
R2CO

r
5
"p

l,,/ .
H2
1l II

l
-alcohol
+

r"(~"O/
5 P /R
'-.,.1/ /
Rh CH ~ Rh CH
5 p
P
/ "0/
"R
" '\
H

S=solvent
P=tertiary phosphine

Figure 10.1. Mechanism of ketone hydrogenation as proposed by Schrock and Osborn.(13)

Rhodium hydroxo complexes have been considered as active intermediates


(see Figure 10.2).
Addition of small quantities of Et3N transformed the inactive
RhCl(PPh 3h into an active catalyst for ketone reductionY9) Catalytic
systems of similar activity were prepared in situ from rhodium diene
derivatives and also other aryl phosphines in presence of bases like Et3N.

Figure 10.2. Mechanism of ketone reduction in strongly basic media as proposed by


Gargano.(18)
HYDROGENA TION REACTIONS OF CO AND CN FUNCTIONS 321

x IRh(NBD)CI1 2 • PPh3. Et3N


• IRh(NBD)CI12. PBu~
° IRh(NBD)C1l2· BMPP
Figure 10.3. Effect of the P: Rh ratio on rate of ketone hydrogenation with different phos-
phinerhodium catalysts.

These latter transformed the catalyst into a hydridorhodium(I)phosphine


complex and enabled the catalytic reduction of different dialkyl-, alkyl-
aryl-, diaryl-, and cyclic ketones with acceptable reaction rates under mild
conditions (50°, 1 bar).
It was shown that by changing the phosphine/rhodium ratio in catalytic
systems prepared in situ several active species are formed, (20) but while
increasing the P/Rh ratio in systems prepared from [Rh(NBD)Clh + PR 3
(PR 3 = BMPP, PBu~, etc.), the rate of hydrogenation decreases, this
tendency is reversed in the case of the [Rh(NBD)Clh + PPh 3 + Et3N
catalysts (see Figure 10.3).(21)
Polymer-bound heterogenized homogeneous rhodium catalysts have
also been successfully used in ketone reduction. Italian scientists studied
ionic rhodium complexes(22) supported on a Merrifield resin (4), but the
Wilkinson-type analog proved to be active only in the presence of Et 3N.(23)

r~ ~
T
t- PII -
R"
t-P-Rlp+-l~
I I
Rh
I
+-1 (4)
R' R"-P-R' R' PEt,
~
R' = Ph; R" = Me, menthyl
322 BALINT HElL, LASZLO MARKO, AND SZILARD TOROS

A few authors also reported on the catalytic activity of ruthenium


complexes in hydrogenation of ketones. (8,24,25)
Transfer hydrogenation is another method widely used in ketone
reduction. As a source of hydrogen, secondary alcohols like iPrOH are
preferred, and the reaction is carried out in presence of KOH at the boiling
point of the solvent. Reactions catalyzed with ruthenium complexes working
at higher reaction temperatures were often performed in cyclohexanol or
benzylalcohol, (26) but, as a consequence of hydrogen transfer from the
secondary alcohols produced, the reaction becomes reversibleY 1)
Spogliarich reported on hydrogen transfer studies with [Rh(diene)P 2t
(P = PPh 3, PMePh 2, PMe2Ph, PBzPh 2, DPM, DPE, DPP, DPB, DPET,
DIOP; diene = COD, NBD) and Wilkinson-type complexes.(27,28) Reduc-
ing cyclohexanone, the best result (90% conversion in 15 min; cyclo-
hexanone/cat. mole ratio = 1900) was achieved with the [Rh(COD)DPEt
catalyst. (28) The activity of the complexes depends on the nature of the
phosphine used and follows the order DPE > DPP = PPh3 > DPB =
DIOP = DPET > PMe2Ph > PBzPh 2 > PMePh2. This means that, in gen-
eral, higher activities were obtained with bidentate ligands. The catalytic
activity is also dependent on the number of carbon atoms (n) in
Ph2P(CH2)nPPh2, decreasing with increasing n. Similar ionic complexes
were used by Uson to reduce acetophenone. (29)
Apart from the phosphine complexes described, the hydrogen transfer
activity of several other derivatives has also been investigated. Among
them rhodium, (301 and iridium (31) bipyridyl, and phenanthroline compounds
displayed the highest catalytic activity with turnover numbers up to 900
cycles/min. Reducing unsaturated ketones, first the C=C double bond is
saturated, but the successive reduction of the C=O group is comparable
in rate with that of the olefinic bond.
From many ketones, stereoisomeric cis and trans alcohols may be
formed. The stereoselectivity of the catalysts was mainly studied with
4-t- butylcyclohexanone as a model compound. Some of the representative
results are summarized in Table 1. The stereoselectivity with catalysts
prepared in situ is determined by the structure of the phosphine ligand.
The cis alcohol is preferred, using phosphines like PPh 3 and NMDPP in
presence of Et3N where probably HRhP 3 is formed as the active intermedi-
ate. This suggestion is supported by the fact that the reaction rate is
enhanced by increasing the P /Rh ratio from 2.2 to 3. Using basic phosphines
like PBu~, BMPP, IMPP, etc., the key intermediate formed is apparently
H 2RhP 2 (highest rates are achieved at P: Rh = 2.2: 1, and the trans isomer
is preferred). Addition of Et3N is ineffective in this case or slightly decreases
both the reaction rate and the stereoselectivity. (21)
The reduction of 4-methyl- and 3-ethylcyclohexanone by hydrogen
transfer with PPh 3 and PPh 2NR 2 type aminophosphine-rhodium(I) com-
J::
~
lJ
a
~
~
::!
a
~
Table 1. Stereoselective Hydrogenation of 4-t-Bu-cyclohexanone lJ
~
C')
Trans Hydrogen
::!
Catalyst Alcohol, % donor Solvent Temp. Reference a
~
[RhHz(PPhMezhSztCl04 - 86 Hz 25° 13
~
[Rh(COD)(PPh 3htpF6 r + Et3N 63 Hz C6 H 6 20° 15
[Rh(NBD)(DPE)tCl0 4- + KOH 67 tPrOH tPrOH 82° 28
8
».
[ Rh(COD)(DPE)tCl04 - + KOH 85 tPrOH ,PrOH 82° 28 ~
[Rh(4, 7-Me zPhen)zCl z]+Cl- + KOH 20 ,PrOH ,PrOH 82° 30 ~
[Ir(3, 4, 7, 8-Me4 Phen)(COD)tCl- + KOH 86 tPrOH iPrOH 82° 31 ~
[Rh(NBD)Clh + PPh 3 + Et3N 27 Hz C6H6/MeOH 50° 21 ~
::!
a
[Rh(NBD)Clh + PBu~ + Et3N 88 Hz C6H6/MeOH 50° 21
~
[Rh(NBD)Clh + PBu~ 90 Hz C6~/MeOH 50° 21

~
(.,.j
324 BALINT HElL, LAsZLO MARKO, AND SZILARD TOROS

plexes prepared in situ has been investigated by Svoboda.(32) By hydro-


genating the 4-methyl-derivative with a PPh 3 containing catalyst, mainly
the cis isomer was formed, while employing aminophosphines, the
thermodynamically more stable trans isomer was the major product.
Japanese scientists succeeded in stereoselective hydrosilylation of sub-
stituted cyciohexanones catalyzed by RhCI(PPh 3h and a silica-linked
I
rhodium(I) complex ( Si-O-SiCH 2CH 2PPh 2hRhCI. They found that the
I
reduction of 2- substituted derivatives leads with each catalyst preferentially
to the thermodynamically less stable alcohol isomer, but the influence of
the bulkiness of silanes was also significant: increasing steric hindrance
favored the more stable stereoisomer. (33)

3.2. Enantioselective Hydrogenation of Ketones

Homogeneous enantioselective hydrogenation of ketones has been


first carried out with the ionic complexes of Schrock and Osborn using
chiral phosphines like BMPP, EMPP, DIOP, etc., but the optical yields
were rather low. (34-36) By preparing the catalyst with the same chiral
phosphines in situ, the optical selectivity (which is determined in this case
by the covalent and ionic complexes present simultaneously) was improved
considerablyY6) RhCIP 2 and RhP 3+ type intermediates are regarded to be
responsible for this effect. (20)
Many of the chiral bidentate phosphines synthesized in the last years
have also been tested for enantioselective ketone reduction. Some of the
results achieved are compiled in Table 2. The influence of phosphine
structure on optical selectivity and catalytic activity is considerable, but a
reliable correlation could not yet be found. It seems that chiral bidentate
bis(diphenyl)phosphines, like prophos forming 5-membered chelate rings
with the rhodium atom and used with great success for the hydrogenation
of dehydroaminoacids, are not suitable for ketone reduction because of
very low reaction rates.
Apart from the structure of the chiral phosphine, optical selectivity of
homogeneous asymmetric hydrogenation of ketones is strongly influenced
by other reaction parameters, such as temperature, pressure, the P jRh
mole ratio with in situ prepared systems, (20,42) the quality and quantity of
other additives, and the solvent, etc. While the effect of the latter is shown
in Table 3, the influence of Et3N on enantioselectivity in different sol-
vents(19,43) is demonstrated in Figure 10.4. By increasing the Et3NjRh mole
ratio in methanol, there is a maximum in enantioseiectivity (ascribed to a
HRh(P-P) type catalyst), but further amounts of the base decrease the
optical purity of the secondary alcohol considerably. This may be explained
~
~
J:I
Table 2. Enantioselective Hydrogenation of Acetophenone 0
G)
n,
Temp. Pressure, Reaction Chemical Optical ~
Catalyst Solvent °C bar time, h yield, % yield, % Reference :::!
0
<:
[Rh(NBD)(R)-BMPPhtCI0 4- EtOH 20 24-80 20-40 a 8.6 34 J:I
[Rh(NBD)«R)-EMPPhtPF 6 - THF 50 72 77 0.24 35 ~
C)
:::!
[Rh(NBD)Cl]2 + IMPP EtOH 60 3.1 43 ? 11.4 37 0
[Rh(NBD)(-)-DIOptCI04- IPrOH 30 ? 8.1 36 ~
0
[Rh(HD)Clh + (+)-DIOP MeOH/C6 H6 30 173 10 51 16 "ll

[Rh(COD)(BPPFOHJrCI0 4- MeOH 0 50 8 96 43 39 8
):.
[Rh(HD)Clh + BDPCP ? 80 200 ? 22 40 <:
0
[Rh(COD)(PNNP)tCI04- MeOH/C6 H 6 20 12 36 72 16.7 41
~
[Rh(C2~hClh + prophos MeOH 50 100 ~ 44.2 42
~
[Rh(NBD)Clh + (+)-DIOP + Et3N C6 H6 50 70 6 64 80 43 <:
C)
:::!
[Rh(NBD)(cycphosJrPF6 - + Et3N MeOH 25 100 140 71 6 44 0
RhCl-( - )-DIOP-PSt b C6 H6 25 4 27 11 7.7 45 ~
[Rh(COD)Cl]2 + DPPEA + Et3N C6 H 6/ HZO 50 ? 22 46
[~RU4(COls[( - )-DIOPh 130 100 5 40 8.1 38

a moles / mole Rh.


o PSt = Styrene--<livinylbenzene copolymer.

~
Ol
326 BALINT HElL, LASZLO MARKO, AND SZILARD TOROS

Table 3. Enantioselective Hydrogenation of Ketones in


Different Solvents

Optical
Ketone Solvent yield, %

EtOH (s) 1.6


THF (s) 0.3
EtOAc (R) 0.9
2-0ctanone a
DMF (R) 5.1
(CH 3hCHCOOH (R) 11.2

CH 3COOH (R) 12.0

C6H6 (s) 19
C6H6 + 2% MeOH (s) 24
C6H6+ 2% H 2 O (s) 27
Acetophenone b
MeOH (s) 37
MeOH + 2% CH 3COOH (s) 43
CH 3COOH (s) 56

a Catalyst: [Rh(COD)(ACMPhrBF. -('7)


b Catalyst: 1/2[Rh(NBD)Clj2 + (S)-BMPP; reaction conditions: 50',
1 barH2"11

Optical
yield,·I.

(R)

70

50

30

2 3

o in methanol
x in benzene
(5)

Figure 10.4. Effect of Et3N on enantioselectivity of acetophenone hydrogenation with a


[Rh(NBD)Clh + (-)-DIOP catalyst.
HYDROGENA TION REACTIONS OF CO AND CN FUNCTIONS 327

by the effect of excess OMe - ions which transform the catalyst into a
[Rh 3(diphosh(OMeht type complex described by Halpern.(48) In benzene,
this latter effect is small, and the enantioselectivity of the system is mainly
determined by the HRh(P-P) derivative.
Optical selectivity in these reactions is obviously strongly depending
also on the structure of the substrate. It has been generally observed that
enantioselectivity is higher in the case of alkyl-aryl ketones than with
dialkyl ketones. (43)
First results with polymer-supported rhodium (I) chiral diphosphine
complexes were published by Ohkubo, (45) but the optical selectivity repor-
ted was rather moderate (7.7% e.e.). Chiral ruthenium catalysts have also
been successfully used, and the results were compiled in a review published
in 1981.(49)
Only little information is available on asymmetric ketone reduction
using hydrogen transfer. With secondary alcohols or indoline as hydrogen
donors, optical yields up to 9.9% were obtained in the presence of
H4Ru4(CO)8[( - )DIOPh as a catalyst. (5 0) Iridium compounds like
[Ir(COD)(PPEl)tCI0 4 - proved to be more enantioselective in the pres-
ence of KOH. 1511 Reducing propiophenone 30% e.e. was observed at 50%
conversion, but it decreased with increasing conversion. The activity of this
iridium catalyst was rather high: 98% conversion was obtained in 4 hours
with high substrate/catalyst ratios (1000-2000) under mild conditions.
As mentioned earlier, hydrosilylation can be used as an alternative
route of ketone reduction. By performing the addition of hydrosilanes in
the presence of suitable chiral transition metal catalysts, prochiral ketones
are transformed in the subsequent hydrolysis step into optically active
alcohols (5):

(5)

This method of ketone reduction has been widely used all the more,
since there are relatively few catalysts suitable for the homogeneous direct
hydrogenation of ketones and the enantioselectivity of hydrosilylation may
be rather high. Because of a comprehensive review published in 1977,(52)
only a few general remarks will be made and some of the latest results
presented (Table 4). Both optical purity and configuration of the secondary
alcohols formed are strongly dependent on the structure of the hydrosilane.
Rather low enantioselectivities were observed with monohydrosilanes using
a rhodium-DIOP catalytic system, but activity and enantioselectivity are
enhanced if dihydrosilanes are used. Best results have been achieved with
a- NpPhSiH 2. Reasonable selectivities could be obtained also with other
chiral diphosphine derivatives using dihdyrosilanes, but in the presence of
328 BALINT HElL, LASZL() MARKO, AND SZILARD T6R6s

Table 4. Asymmetric Hydrosilylation of Ketones

Optical
Ketone Silane Phosphine yield, % Reference

PhCOMe a-NpPhSiH 2 (+)-DlOP (s) 58 53


PhCOMe Ph 2 SiH 2 MPFA (R) 49 52
PhCOMe a-NpPhSiH 2 naphos (s) 18 54
PhCOMe Ph 2SiH 2 glup (s) 47 55
PhCOMe a-NpPhSiH 2 PPPM (R) 60 56
Ph COMe a-NpPhSiH 2 glucophinite - 65 57
PhCOCH 2 CI a-NpPhSiH 2 (+)-DlOP (s) 63 58
PhCOEt a-NpPhSiH 2 (+)-DlOP (s) 56 59
PhCOEt Ph 2SiH 2 cycDIOP (R) 52 60
PhCOEt a-NpPhSiH 2 camphinite - 52 57
PhCO(IPr) PhMe2SiH (s)-BMPP (R) 56 61
PhCO(tBu) EtMe2SiH (R)-BMPP (R) 56 61
PhCO(tBu) PhMe2SiH (R)-BMPP (s) 61.8 62
PhCO(cHex) PhMe2SiH (R)-BMPP (s) 58 63


CH 3 (CH 2)sCOCH 3 Ph(cHex)SiH 2 (-)-DlOP (R) 44 58

PhZPO~
b
PhzPO 0
o CHzOPPhz
PhCHzO 0 CHzOPPh z

glucophinite camphinite
the monodendate BMPP ligand, application of monohydrosilanes was more
favorable. In contrast to the hydrogenation of ketones, no characteristic
difference could be detected between ionic and covalent rhodium complexes
as catalysts of hydrosilylation.
Benes prepared some new chiral phosphines suitable as ligands for
this reaction.(60) Capka anchored (C 2 H sOhSi(CH 2 hP(Ph)Men to silica (6).
Complexed with rhodium, a catalyst for the enantioselective hydrosilylation

E
of ketones like acetophenone and propiophenone was obtained. (64)

o Ph
O~Si-(CHzh-p/ (6)
/ .J,,,
/ 0 Rh menthyl

Chiral phosphinites were also employed as ligands in hydrosilylation.(55)


HYDROGENA TlON REACTIONS OF CO AND CN FUNCTIONS 329

Based on kinetic and spectroscopic studies, it was suggested(65,66) that in


hydrosilylation of t-butylphenylketone the interaction of the silyl hydrido-
rhodium phosphine species with the ketone is rate determining.
If prochiral ketones are hydrosilylated with dihydrosilanes carrying
different alkyl groups (R I ¥- R 2) and the siloxane is treated with a Grignard
reagent, optically active silanes are produced (7):

R 3 MgX
H 20
(7)
" *
OCHRR' *
RCHOHR'
In this way, a- NpPhMeSiH of 82% e.e. was formed by hydrosilylation
of (-)-menthone with a RhCI( + )-DIOP catalyst,(67) a reaction in which
two chiral centers were acting in a diastereoselective procedure.

4. HYDROGENA TlON OF KETONES CONTAINING OTHER


FUNCTIONAL GROUPS

4.1. Selective Hvdrogenation of Unsaturated Ketones

By hydrogenating ketones containing C=C and C=O double bonds


with rhodium phosphine catalysts, the olefinic bond is saturated first.
RhCI(PPh 3 h is inactive in hydrogenation of the keto group (except activated
groups in ketoesters, as discussed in Section 4.2.). This fact made the
selective reduction of mesityl oxide to methyl-isobutylketone possible.(68)
By adding H 2 0 2 to the catalyst it becomes heterogeneous, but its activity
is enhanced enormously. Using it for more than 500 hours, turnover
numbers above 100,000 were measured.(69) The Wilkinson compound
hydrogenates the olefinic bond in A4 -a,~- unsaturated 3-keto steroids
stereoselectively, and the Sa-form is produced. A very strong influence of
reaction parameters was detected. (70)
Solodar studied the reduction of piperitenone (8), a compound contain-
ing two different olefinic bonds and a prochiral carbon-oxygen double
bond.(7I) He found that mainly the sterically less hindered C=C bond was
favored in the hydrogenation reaction, and thus pule gone was the major
product, but also other compounds were formed in minor amounts.

?o + H, r,"",,",>',]'",··?o (8)

Lz = mop, MDPP, PCH3(2-RIO-C6H4)Rz


(R I = Me, iPr, Bz; R2 = Ph, cHex, iPr)
330 BALINT HElL, LASZUJ MARKO, AND SZILARD TOROS

Reducing a,{3- unsaturated ketones with catalysts prepared from


[Rh(diene)Cl]2 and PR 3 in situ again the C=C bond is saturated preferen-
tially, but performing the reaction in primary alcohols hydrogen transfer
as a side reaction may take place, and the aldehyde thus formed is decar-
bonylated and the catalyst deactivated. (72)
No difference in selectivity was detected reducing unsaturated ketones
in a hydrogen-transfer reaction. Beaupere and co-workers reported experi-
ments in a-phenylethanol with HRh(PPh 3)4 as catalyst(73) and found that
electron-attracting groups (X) on substrates like XC 6 H 5 CH=CHCOR
increase the reaction rate. In a similar reaction, ruthenium complexes were
the catalysts, and hydrogen was transferred from isopropylidene- or cyclo-
hexylidene-1,2-a- D-glucofuranose to benzalacetophenone, forming the
saturated ketone with 100% yield. (74)
In contrast to the results discussed above, hydrosilylation of unsatur-
ated ketones with rhodiumphosphine catalysts makes the selective
saturation of either the C=C or the C=O bond possible. Using monohy-
drosilanes, l,4-addition takes place and saturated ketones are formed,
while dihydrosilanes prefer the 1,2-addition and unsaturated alcohols are
the products (9):
Rl Rl
R,SiH
"
/
CHCH=CR 3
1
H2O
----.
R2
"CHCH CR'
/ 211
0
R' OSiR 3
1,4-addition
Rl
"C=CHCR' [Rhl
(9)
R/ ~
Rl Rl
R 2 SiH,

/
"c=CHCHR 3
1
H2 O
~
"C=CHCHR
/ 1
3

R2 OH
R' OSiR,H
1,2-addition

Accordingly a-ionone(a) and citral(d) (a,{3-unsaturated carbonyl com-


pounds containing an additional isolated double bond) could be transformed
with 96-97% selectivity(74) into dihydroionone(c) and citronellal(f) in the
presence of RhCl(PPh 3h and Et 3SiH (10, 11):

~O ~OSiE"~~O
a b c

(10)
HYDROGENA T/ON REACTIONS OF CO AND CN FUNCTIONS 331

~HO ---+
(11)

d e

Similar observations were made reducing {3-ionone and pUlegone:


using PhMe2SiH mainly, saturation of the a,{3-olefinic bond was achieved,
while in the presence of Et 2SiH 2 or Ph 2SiH 2, the unsaturated alcohol was
formed exclusively from both substrates.
Later, Ojima reported on the hydrosilylation of a,{3-unsaturated
ketones with chiral rhodium catalysts (Rh-BMPP and Rh-DIOP) and
a-NpPhSiH 2. As models, {3-ionone and 2-methylcyclohexenone were
reduced to the corresponding alcohols with 33.5 and 43% optical selec-
· ·t y. (76)
t IVI

4.2. Hvdrogenation of Keto Acids and Keto Esters

In contrast to the hydrogenation of simple ketones, RhCI(PPh 3h was


found to be an active catalyst(77) for the reduction of keto esters. In addition,
several ionic or covalent rhodium complexes containing other types of
phosphines have been used successfully.
Some characteristic results in asymmetric reduction of keto acids and
keto esters are summarized in Table 5. It should be mentioned here that
there are some remarkable differences between the hydrogenation of simple
ketones and keto esters. For example, the optical selectivities are generally
much higher with keto esters. This is probably due to the bidentate ligand
character of these compounds.
Recent results suggest that small amounts of water decrease the reac-
tion rate of hydrogenation of propyl-pyruvate.(77) It should be mentioned,
however, that Schrock(13) and Solodar (47 ) made their earlier opposite
observations while working with monophosphine ligands. Ojima obtained
the highest optical selectivities in benzene and THF with neutral rhodium
complexes, while methanol decreased both optical selectivity and the reac-
tion rate. (77) The fact that neutral and ionic rhodium-DIOP complexes
produce enantiomers of opposite configuration, and further, the finding
that enantioselectivity of keto ester reduction depends on the covalent or
ionic character of the catalytic system, is in accordance with earlier observa-
tions in asymmetric hydrogenation of acetophenone. 2o
While these and other reactions discussed earlier were carried out in
different organic solutions, several water-soluble transition-metalphos-
phine complexes have also been synthesized and used as catalysts in the
~
'"
Table 5. Asymmetric Reduction of Keto Acids and Keto Esters

Reducing Optical
Substrate agent Phosphine Solvent yield, % Reference

CH 3COCH 2COOMe H2 ICMP EtOH 71 48


CH 3COCOOH H2 BPPFOH MeOH 83 39
~
f-
CH3COCOOEt H2 CPPM c-C6 H12 67.3 78 ~
-I
CH3COCOOEt H2 BPPM C6 H 6 65.3 78 :x:
!:!:!
CH3COCOOPr n H2 BPPM C6 H 6 75.8 77 r--
CH3COCOOBu n H2 CPPM c-C 6 H12 62.8 78 ~
CH3COCOOPrn a- NpPhSiH2 + H 2O (-)DIOP C6 H 6 85.4 79 fG
r-
0-
PhCOCOOMen a-NpPhSiH2 + H 2O (+)DIOP C6 H 6 77 79
PhCOCOOMen
~
~
a-NpPhSiH 2 + H 2O PPh3 C6 H 6 17 79
CH3COCOOMen a- NpPhSiH2 + H 2O (+)DIOP C6 H 6 85.6 79 ~
):,.
CH3COCH2CH2COOBui a a- NpPhSiH 2 + H 2O (+)DIOP C6 H 6 84.4 79
~
CH3COCOOPr n H2 (-)DIOP THF 41.9 77 VI
!:::!
f-
):..
a The product: 4-methyl-y- butyrolactone. ~
0
Cj,
~
Q,
VI
HYDROGENA TlON REACTIONS OF CO AND CN FUNCTIONS 333

last years. Joo prepared ruthenium complexes containing m-sulphophenyl-


diphenylphosphine as ligand and found that among them HRuCl(mSPPh 2h
and HRu(OAc)(mSPPh 2h are active in hydrogenation of keto-acids.(80)

4.3. Hydrogenation of Aminoketones and Other Biologically Active


Derivatives

Increasing interest has been shown in recent years for the synthesis
of biologically active secondary alcohols by homogeneous catalytic
stereoselective and enantioselective hydrogenation reactions.
The first result in this field was the reduction of 2-(6-carbomethoxy-
hexyl)cyclopentane-l,3,4-trione with [Rh(COD)(ACMPhtBF4 - as
catalyst to 2-( 6-carbomethoxy-hexyl)-4-(R)-hydroxycyciopentane-l ,3-
dione, \HI) an intermediate of the EJ prostaglandine synthesis (12):

Q
o 0
(CH2)6COOCH3 * Q(CH2)6COOCH3
+H2~ (12)
o 0 HO" ~0

Japanese authors reported(82) on the synthesis of (R)( - )-pantolactone


(a precursor of pantothenic acid, which is an important constituent of
coenzyme A) by hydrogenation of a- keto-t3,t3- dimethyl-/,- butyrolactone
with rhodium catalysts containing different chiral pyrrolidinediphosphines
(13):

[Rh(diene)CI], + l

(13)

R = (BuO, (Bu, Ph, Me, H

They found that the enantioselectivity of the reaction is highly influen-


ced by the steric bulkiness of the N- substituents in the phosphine, while
electronic factors play only a minor role. Working with the Rh-BPPM
catalytic system (R = (BuO), 86.7% optical yield was achieved under
optimum reaction conditions, (83) enabling the preparation of pure D( -)-
pantoyl lactone in a single recrystallization step.
~
""-
Table 6. Enantioselective Reduction of Aminoketones

Reducing Reaction Conversion, Optical


Substrate agent Catalyst time, h % yield, % Reference

PhCOCH2NHCH2Ph H2 [Rh(COD)(ACMPhtBF4 - _1.7 oa 37


CH3 COCH(COOH)NHCOCH3 b [Rh(NBD)(MDPPhtBF4 - 6 80 95" 84
H2
3, 4-(MeOhC6 H 3 COCH2NH2·HCl H2 [Rh(HD)Cllz + BPPFOH 48-96 100 89 85
~
r-
3,4-(HOhC6 H 3 COCH2NHMe'HO H2 [Rh(NBD)(BPPFOHJrCl04 - 168 100 95 85
~
PhCOCH2NEt 2 H2 [Rh(NBD)Cllz + (+)DlOP 20 70 93 86 ""i

2-NpCOCH2NEt2 H2 [Rh(NBD)Cllz + (+)DlOP 20 93 95 86 ~


r=:
PhCOCH2NBu~ H2 [Rh(NBD)Clh + (+)DlOP 20 52 90 86 r-
l:>,
C/)
4-Et-C6 H 4 COCH2NEt2 H2 [Rh(NBD)Clh + (+)DlOP 20 39 90 86 ~
0.
MeCOCONHCH(CH 2Ph)COOMe H2 [Rh(COD)Clh + (+)DlOP 20 100 26" 87
MeCOCONHCH(CH2Ph)COOMe H2 [Rh(COD)Cl]2 + BPPM 64 100 28" 87 ~
:0
MeCOCONHCH(CH 2Ph)COOMe H2 RhCl(PPh 3 h 64 100 20" 87 ~
)::.
PhCOCONHCH(CH 2Ph)COOMe a-NpPhSiH 2 [Rh(COD)Clh + (+)DlOP 24 79 82" 87
~
PhCOCONHCH(CH2Ph)COOMe a-NpPhSiH2 RhCl(PPh 3 h 24 62 56" 87 C/)
~
r-
o [an' value. l:>,
b 90% d, 10% I. :0
C)
, diastereomeric excess.
(j.
:0
0,
C/)
HYDROGENA TION REACTIONS OF CO AND CN FUNCTIONS 335

Biological and therapeutic effects of 2-amino-1-arylketones and their


derivatives are well known. Some results in asymmetric hydrogenation of
these compounds are summarized in Table 6. High optical yields were
observed also in these cases, apparently again pointing to the simultaneous
coordination of the carbonyl group and the nitrogen atom of the substrate
to the central rhodium atom in the active intermediate.
It has been shown by Ojima(87) in the synthesis of depsipeptides (see
the last five lines in Table 6) that optical selectivities are much higher in
the hydrosilylation reactions, where asymmetric induction by the chiral
catalyst predominates over that of the chiral center already present in the
N- (a- ketoacyl)-a- aminoester substrates.
Special biological effects are often displayed only by given
stereoisomers of secondary alcohols. This is the case for example with
tropines, constituents of some well-known biologically active derivatives
like cocaine and scopolamine. In the preparation of tropine stereo isomers,
rhodiumphosphine catalysts prepared in situ could be used with success.(21)
The structure of alcohols is mainly determined by the structure of phos-
phines in the catalyst: both a- and {3- hydroxy derivatives could be synthe-
sized as major products. With PAr3 type phosphines, the reaction rate is
higher at P /Rh = 3 than at P /Rh = 2.2, and a- hydroxy tropines are formed
with up to 98% yield. Using basic alkyl phosphines as ligands, the catalytic
system shows highest activity at P /Rh = 2.2, and the thermodynamically
more stable {3- isomers can be prepared in yields as high as 89%. Selectivity
of the reduction is not depending on the quality of the R group in nort-
ropinones. Substantial similarities can be observed between the main
characteristics of the hydrogenation of these derivatives and those of
4-tBu-cyclohexanone (see Table 1, Section 3.1.).

5. HOMOGENEOUS HYDROGENA TION OF CARBON-


NITROGEN DOUBLE BONDS

Although both [Rh(diene)P 2t A-type complexes,(36) RhCI(PPh 3h,(88)


and catalysts prepared in situ(88) from [Rh(diene)Clh + PR 3 are active in
hydrogenation of the C=N bond, only a few papers have been published
in this field. In contrast to ketone reduction, catalytic activity increases in
the order PBu~ < PEt 2Ph < PEtPh 2 < PPh 3 and has its optimum at a
P /Rh = 2.2 mole ratio. While increasing amounts of ligand inhibit the
catalyst, (88) the system remains homogeneous and active even at lower mole
ratios. Phosphites are also suitable as ligands in this system. (89)
Reduction of Schiff-bases has been performed also with hydrogen
transfer reactions(90) catalyzed by RhCI(PPh 3h, RhCI(CO)(PPh 3 h, and
336 BALINT HElL, LASZU] MARKO, AND SZILARD T6R6s

RuClz(PPh 3h complexes using i-propyl alcohol as hydrogen donor. It was


found that Me or MeO groups on the ring of RC6H4RIC=NC6H4R2 type
substrates increase the reaction rate. KOH and minor amounts of water
promote the hydrogenation of PhCH=NPh with RhCI(PPh 3h as
catalyst. (91)
Aldimines have also been hydrogenated by hydrogen transfer from
iPrOH catalyzed by RhCI(PPh 3h, RuCI 2(PPh 3)4, HOsCI(CO)(PPh 3h, and
RhCI(CO)(PPh 3h in the presence of sodium carbonate.(92) The formation
of hydrido-metal complexes as active species was suggested. To prove this
assumption, HRh(PPh3)4 was prepared and used for reduction of
PhCH=NPh with a high reaction rate. Catalysts prepared in situ from
RhCh, PPh3, and sodium carbonate were less active. RuCh, IrCh, or OsCh
gave systems which were essentially inactive.
Ojima reported on the reduction of R 1R 2C=NR 3 type Schiff bases
by hydrosilylation(93) with several catalysts. The reaction was also affected
by the quality of hydrosilane. Using dihydrosilanes, which were found
to be more active than monohydrosilanes and trihydrosilanes, the activity
of the catalysts used decreased in the order RhCl(PPh 3h » Rh(CO)CI-
(PPh 3h > PY2RhCI(dmf)BH4 > [Rh(HD)CI]2 > [Rh(COD)CI]2 > PdClz >
PdClz(PPh 3h.
The reduction of C=N bonds and immonium ions was suggested to
be involved in the N- alkylation of amines by CO + H 20 with different
rhodium catalysts like RhCh, [Rh(COD)CI)z, RhCl(PPh 3h, and
RhCI(CO)(PPh 3h(94) (14, 15):
::::NH + OHCR ---. ::::N+ =CHR ---. ::::N -CH2R (14)
-NH2 + OHCR ---. -N=CHR ---. -NHCH 2R (15)
Asymmetric reduction of prochiral imines, ketoximes, and several
dihydroisoquinoline derivatives was successfully performed with chiral
rhodium and ruthenium catalysts (Table 7).
Italian scientists reported(36) on asymmetric hydrogenation of N- (a-
methyl benzylidene )benzylamine to the corresponding N- benzyl-1-phenyl-
ethylamine with [Rh(NBD)(-)DIOptCI04- as catalyst and achieved 22%
e.e. The same substrate was hydrogenated also with a catalyst prepared in
situ from [Rh(NBD)CI]2 and DIOP, but the optical induction was
much lower «3.3 % e.e.). (88) By preparing the catalyst with (R)( +)-
Ph2PCH(IPr)CH 2PPh 2 as the chiralligand,(96) however, optical selectivities
up to 65% have been achieved.(97)
Moderate optical inductions were found in the hydrogenation of
ketoximes with H4Ru(CO)g[(-)-DIOPh as catalyst. By reducing t-
butylphenylketoxime, the optical yield was 14.5%.(38)
Kagan reported on asymmetric hydrosilylation of prochiral compounds
with carbon-nitrogen double bonds. (95) N- (a- methylbenzylidene)benzyl-
HYDROGENATION REACTIONS OF CO AND CN FUNCTIONS 337

Table 7. Enantioselective Reduction of Compounds Containing C =N


Double Bond

Reducing Optical
Substrate Catalyst agent yield, % Reference

PhMeC=NCHzPh [Rh(NBD)(-)DlOptCl 4- H2 22 36
PhMeC=NCHzPh [Rh(NBD)Cl]z + (+)DlOP Hz 3.3 88
PhMeC=NCHzPh [Rh(NBD)Cl]z + (s)BMPP Hz 1.5 88
PhMeC=NCHzPh [Rh(NBD)Clh + valphos Hz 65 97
MeEtC=NOH H4Ru4(CO)8[( - )DlOP]z Hz 2.6 38
MePhC=NOH H4Ru4(CO)8[( - )DlOP]z Hz 4 38
Ph(tBu)C=NOH H4Ru4(CO)8[( - )DlOPh H2 14.5 38
PhMeC=NCHzPh RhCl( + )DlOP PhzSiH z 65 95
DHIQA RhCl(+)DlOP PhzSiH z 22.5 95
DHlQB RhCl( + )DlOP Ph 2 SiH z 5.7 95
DHIQC RhCl(+)DlOP Ph 2 SiH 2 38.7 95
MePhC=NPh RhCl(+)DlOP PhzSiH z 47 95
Me(CH zPh)C=NCH 2 Ph RhCl( + )DlOP Ph zSiH 2 11.5 95

DHIQ A B C
=H

)§Q
RI RI = OCH3 R' = OCH 3
RZ = CHzPh RZ = CH 3 OCH 3

R' =CH,-@-OCH 3
R'

amine was reduced with Ph 2 SiH 2 in the presence of RhCl(DIOP) as catalyst


with 65% optical selectivity. In a similar reaction, 1,2,3,4-tetrahy-
dropapaverine was synthesized with 38% e.e. from the corresponding
dihydroisoquinoline.

ABBREVIA TIONS FOR LIGANDS

Abbreviation Name of Ligand


ACMP 0- Anisylcyclohexylmethylphosphine
ICMP (2-Isopropyloxyphenyl)cyclohexylmethylphosphine
BMPP (R)-( + )-Benzylmethylphenylphosphine
EMPP (R)-( - )-Ethylmethylphenylphosphine
IMPP (+ )-Isopropylmethylphenylphosphine
338 BALINT HElL, LASZL(] MARKO, AND SZILARD TOROS

BPPM (2s,4s)- N- (t- ButoxycarbonyJ) -4-diphenylphosphino-2-


diphenylphosphinomethylpyrrolidine
CPPM (2s,4s)-N- (ChoJesteryJoxycarbonyl)-4-diphenyJphosphino-2-
diphenyJphosphinomethyJpyrroJidine
PPPM (2s,4s)- N- (t- B utylcarbonyl)-4-diphen yJphosphino-2-
diphenyJphosphinomethyJpyrroJidine
BPPFOH (R)-O'- [(s)-1 ,2-Bis (diphenyJphosphino )ferrocenyJ]ethyJalcohoJ
MPFA (R)-O'- [(s)-2-Dimethylphosphinoferrocenyl]ethyJdimethyJamine
DPM Bis (diphenyJphosphino )methane
DPE Bis (1,2- diphenyJphosphino )ethane
DPP Bis (1,3 - diphenyJphosphino )propane
DPB Bis (1,4- diphenyJphosphino )butane
DPET cis-Bis (1,2- diphenyJphosphino )ethyJene
DlOP 2,3-0- Isopropylidene-2,3-dihydroxy-l ,4-
bis (diphenylphosphino )butane
prophos (R)-( + )-Bis (1 ,2-diphenyJphosphino )propane
vaJphos (R)-( +)-1 ,2-Bis (diphenyJphosphino )-3-methyJbutane
cycphos (R)-1 ,2- Bis (diphen yJphosphino )cyclohexyJethane
cycDIOP (2R,3 R)-2,3 -0- cyclohexylidene-2,3 -dih ydroxy-l ,4-
bis (cyclohexyJphenyJphosphino )butane
naphos (sH - )-2,2' -Bis(diphenyJphosphinomethyJ)-1,1' -binaphtyl
BDPCP d-trans- 1,2 -Bis (diphen ylphosphinoxy )cyclopentane
gJup MethyJ-4,6-0- benzyJidene-2,3-bis-0- (diphenylphosphino )-0'- D-
gJucopyranoside
DPPEA Diphen yJphosphino- N- (0'- phen yJeth yl)-acetamide
PNNP N,N' -Bis (diphenyJphosphino )-N,N' -bis ((S)-( - )-0'- methyl-benzyl)-
ethyJenediamine
MDPP (- )-Menthyldiphenylphosphine
NMDPP (+ )-NeomenthyldiphenyJphosphine
rnSPPh 2 rn- SuJphophenyJ -diphen ylphosphine
HD 1,5-Hexadiene
COD 1,5 -Cyclooctadiene
NBD Bicyclo[2.2.1 ]hepta-2,5 -diene
Phen 1,10-Phenantroline
PPEI 2-PyridinaJ -0'- phenylethylimine

REFERENCES

1. K. Ohkubo, T. Ohgushi, and K. Yoshinaga, Chern. Lett. 1976,775-778.


2. C. FragaJe, M. Gargano, and M. Rossi, I. Mol. Catal. 5,65-73 (1979).
3. D. H. Doughty and L. H. PignoJet, I. Arn. Chern. Soc. 100, 7083-7085 (1978).
4. H. Fujitsu, S. Shirahama, E. Matsumara, K. Takeshita, and I. Mochida, I. Org. Chern.
46,2287-2290 (1981).
5. R. S. Coffey, Chern. Cornrnun. 1967,923.
6. W. Strohmeier and H. SteigerwaJd, I. Organornet. Chern. 129, C43-C46 (1977).
7. J. Tsuji and H. Suzuki, Chern. Lett. 1977, 1085-1086.
8. R. A. Sanchez-DeJgado and O. L. de Ochoa, I. Mol. Catal. 6,303-305 (1979).
HYDROGENA TlON REACTIONS OF CO AND CN FUNCTIONS 339

9. R. A. Sanchez-Delgado, A. Andriollo, O. L. de Ochoa, T. Suarez, and N. Valencia, J.


Organornet. Chern. 209, 77-83 (1981).
10. W. Strohmeier and L. Weigelt, J. Organornet. Chern. 145, 189-194 (1978).
11. H. Imai, T. Nishiguchi, and K. Fukuzumi, J. Org. Chern. 41, 665-671 (1976).
12. S. Meguro, T. Mizoroki, and A. Ozaki, Chern. Lett. 1975,943-946.
13. R. R. Schrock and J. A. Osborn, J. Chern. Soc. D. 1970, 567-568.
14. R. H. Crabtree, J. Chern. Soc., Chern. Cornrnun. 1975,647-648.
15. R. H. Crabtree, A. Gautier, G. Giordano, and T. Khan, J. Organornet. Chern. 141,
113-121 (1977).
16. B. Heil, S. Toros, S. Vas tag, and L. Marko, J. Organornet. Chern. 94, C47-C48 (1975).
17. S. Vastag, B. Heil, and L. Marko, J. Mol. Catal. 5,189-195 (1979).
18. M. Gargano, P. Giannoccaro, and M. Rossi, J. Organornet. Chern. 129,239-242 (1977).
19. B. Heil, S. Taros, J. Bakos and L. Marko, J. Organornet. Chern. 175,229-232 (1979).
20. S. Toros, B. Heil, and L. Marko, J. Organornet. Chern. 159,401-407 (1978).
21. B. Heil, L. Kollar, L. Marko, and S. Toriis, J. Organornet. Chern., in press.
22. G. Strukul, M. Bonivento, M. Graziani, E. Cernia, N. Palladino, Inorg. Chirn. Acta 12,
15-21 (1975).
23. B. Heil, S. Toros, H. J. Kreuzfeld, and H. Pracejus, React. Kinet. Catal. Lett., in press.
24. P. Frediani, U. Matteoli, M. Bianchi, F. Piacenti, and G. Menchi, J. Organornet. Chern.
150,273-278 (1978).
25. W. Strohmeier and L. Weigelt, J. Organornet. Chern. 171, 121-129 (1979).
26. B. Graser and H. Steigerwald, J. Organornet. Chern. 193, C67-C70 (1980).
27. R. Spogliarich, G. Zassinovich, G. Mestroni, and M. Graziani, J. Organornet. Chern. 179,
C45-C47 (1979).
28. - - , J. Organornet. Chern. 198,81-86 (1980).
29. R. Uson, L. A. Oro, R. Sari ego, and M. A. Esteruelas, J. Organornet. Chern. 214, 399-404
(1981).
30. G. Zassinovich, G. Mestroni and A. Camus, J. Organornet. Chern. 168, C37-C38 (1979).
31. A. Camus, G. Mestroni, and G. Zassinovich, J. Mol. Catal. 6, 231-233 (1979).
32. P. Svoboda and J. Hetftejs, Collect. Czech. Chern. Cornrnun. 42, 2177-2181 (1977).
33. J. Ishiyama, Y. Senda, 1. Shinoda, and S. Imaizumi, Bull. Chern. Soc. Jpn. 52,2353-2355
(1979).
34. P. Bonvicini, A. Levi, G. Modena, and G. Scorrano, J. Chern. Soc., Chern. Cornrnun.
1972,1188-1189.
35. M. Tanaka, Y. Watanabe, T. Mitsudo, H. Iwane and Y. Takegami, Chern. Lett. 1973,
239-240.
36. A. Levi, G. Modena, and G. Scorrano, J. Chern. Soc., Chern. Cornrnun. 1975,6-7.
37. J. Solodar, Ger. Offen. 2 306 222 (1973); Chern. Abstr. 79, 146179t (1973).
38. C. Botteghi, M. Bianchi, E. Benedetti, and U. Matteoli, Chirnia 29,256-258 (1975).
39. T. Hayashi, T. Mise, and M. Kumada, Tetrahedron Lett. 1976, 4351-4354.
40. T. Hayashi, M. Tanaka, and 1. Ogata, Tetrahedron Lett. 1977,295-296.
41. M. Fiorini, F. Marcati, and G. M. Giongo, J. Mol. Catal. 3, 385-387 (1977).
42. K. Ohkubo, M. Setoguchi, and K. Yoshinaga, Inorg. Nucl. Chern. Lett. 15, 235-238 (1979).
43. S. Toriis, B. Heil, L. Kollar, and L. Marko, J. Organornet. Chern. 197,85-86 (1980).
44. D. P. Riley and R. E. Shumate, 1. Org. Chern. 45,5187-5193 (1980).
45. K. Ohkubo, M. Haga, K. Yoshinaga, and Y. Motozato, Inorg. Nucl. Chern. Lett. 16,
155-158 (1980).
46. F. Joo and E. Trocsanyi, J. Organornet. Chern. 231, 63-70 (1982).
47. J. Solodar, CHEMTECH 1975, 421-423.
48. J. Halpern, D. P. Riley, A. S. C. Chan, and J. J. Pluth, 1. Arn. Chern. Soc. 99, 8055-8057
(1977).
340 BALINT HElL, LASZLO MARKO, AND SZILARD T6R6s

49. U. Matteoli, P. Frediani, M. Bianchi, C. B"tteghi, and S. Gladiali, I. Mol. Catal. 12,
265-319 (1981).
50. M. Bianchi, U. Matteoli, G. Menchi, P. Frediani, S. Pratesi, F. Piacenti, and C. Botteghi,
I. Organomet. Chem. 198, 73-80 (1980).
51. G. Zassinovich, A. Camus, and G. Mestroni, I. Mol. Catal. 9, 345-347 (1980).
52. 1. Ojima, K. Yamamoto, and M. Kumada, Aspects of Homogeneous Catalysis, edited by
R. Ugo, (D. Reidel Publishing Co., Dordrecht, 1977), Vol. 3, pp. 185-228.
53. W. Dumont, J. C. Poulin, T. P. Dang, and H. B. Kagan, I. Am. Chem. Soc. 95, 8295-8299
(1973).
54. K. Tamao, H. Yamamoto, H. Matsumoto, N. Miyake, T. Hayashi, and M. Kumada,
Tetrahedron Lett. 1977,1389-1392.
55. M. Capka, J. Hetflejs, and R. Selke, React. Kinet. Catal. Lett. 10,225-228 (1979).
56. K. Achiwa, Fundamental Research in Homogeneous Catalysis 3, 549-564 (1979).
57. T. H. Johnson, K. C. Klein, and S. Thomen, I. Mol. Catal. 12,37-40 (1981).
58. H. B. Kagan, J. F. Peyronel, and T. Yamagishi, Adv. Chem. Ser. 173, 50-66 (1979).
59. R. J. P. Corriu and J. J. E. Moreau, I. Organomet. Chem. 85, 19-33 (1975).
60. J. Benes and J. Hetflejs, Collect. Czech. Chem. Comrnun. 41, 2264-2272 (1976).
61. 1. Ojima and Y. Nagai, Chem. Lett. 1974,223-228.
62. T. Hayashi, K. Yamamoto, K. Kasuga, H. Omizu, and M. Kumada, I. Organomet. Chern.
113,127-137 (1976).
63. 1. Ojima, T. Kogure, M. Kumagai, S. Horiuchi, and T. Sato, I. Organomet. Chem. 122,
83-97 (1976).
64. M. Capka, Collect. Czech. Chem. Commun. 42, 3410-3416 (1977).
65. 1. Kolb and J. Hetflejs, Collect. Czech. Chem. Commun. 45, 2224-2239 (1980).
66. - - , Collect. Czech. Chem. Commun. 45,2808-2816 (1980).
67. R. J. P. Corriu and J. J. E. Moreau, I. Organomet. Chem. 91, C27-C30 (1975).
68. W. Strohmeier and E. Hitzel, I. Organomet. Chem. 91,373-377 (1975).
69. - - , I. Organomet. Chem. 102, C37-C41 (1975).
70. W. Voelter and C. Djerassi, Chem. Ber. 101,58-68 (1968).
71. J. Solodar, 1. Org. Chem. 43,1787-1789 (1978).
72. L. Kollar, S. Taros, B. Heil, and L. Marko, I. Organornet. Chern. 192,253-256 (1980).
73. D. Beaupere, P. Bauer, and R. Zan, Can. I. Chem. 57,218-221 (1979).
74. G. Descotes and D. Sinou, Tetrahedron Lett. 1976, 4083-4086.
75. 1. Ojima, T. Kogure, and Y. Nagai, Tetrahedron Lett. 1972,5035-5038.
76. - - , Chem. Lett. 1975, 985-988.
77. 1. Ojima, T. Kogure, and K. Achiwa, I. Chern. Soc., Chem. Cornmun. 1977,428-430.
78. K. Achiwa, Tetrahedron Lett. 1977, 3735-3738.
79. 1. Ojima, T. Kogure, and M. Kumagai, I. Org. Chem. 42,1671-1679 (1977).
80. Z. Toth, F. Joo, and M. Beck, Magy. Kem. Foly. 86,173-177 (1980).
81. C. J. Sih, J. Heather, G. P. Peruzzotti, P. Price, R. Sood, and L. H. Lee, I. Am. Chem.
Soc. 95, 1676-1677 (1973).
82. K. Achiwa, T. Kogure, and 1. Ojima, Chem. Lett. 1978, 297-298.
83. 1. Ojima, T. Kogure, T. Terasaki, and K. Achiwa, I. Org. Chem. 43, 3444-3446 (1978).
84. J. Solodar, Ger. Offen. 2312924 (1973); Chem. Abstr. 80, 3672h (1974).
85. T. Hayashi, A. Katsumura, M. Konishi, and M. Kumada, Tetrahedron Lett. 1979, 425-428.
86. S. Taros, L. Kollar, B. Heil, and L. Marko, I. Organomet. Chem. 232, C17-C18 (1982).
87. 1. Ojima, T. Tanaka, and T. Kogure, Chern. Lett. 1981, 823-826.
88. S. Vastag, B. Heil, S. Taros, and L. Marko, Transition Met. Chern. 2,58-59 (1977).
89. Z. Nagy-Magos, S. Vastag, B. Heil, and L. Marko, Transition Met. Chem. 3, 123-124
(1978).
90. J. S. Shekoyan, G. V. Varnakova, U. N. Krutii, K. 1. Karpeiskaya, and V. Z. Sharf, fzv.
Akad. Nauk SSSR, Ser. Khim. 1975, 2811-2813.
HYDROGENA TION REACTIONS OF CO AND CN FUNCTIONS 341

91. V. Z. Sharf, L. H. Freidlin, 1. S. Portjakova, U. N. Krutii, Izv. Akad. Nauk SSSR, Ser.
Khirn. 1979, 1414-1415.
92. R. Grigg, T. R. Mitchell and N. Tongpenyai, Synthesis, 1981,442-444.
93. 1.-0jima, T. Kogure, and Y. Nagai, Tetrahedron Lett. 1973, 2475-2478.
94. Y. Watanabe, M. Yamamoto, T. Mitsudo, and Y. Takegami, Tetrahedron Lett. 1978,
1289-1290.
95. H. B. Kagan, N. Langlois, and T. P. Dang, I. Organornet. Chern. 90, 353-365 (1975).
96. W. Bergstein, A. Kleemann, and 1. Martens, Synthesis 1981,76-78.
97. S. Vastag, 1. Bakos, S. Toros, N. E. Takach, R. B. King, B. Heil, and L. Mark6, I. Mol.
Catal., in press.
11
Decarbonylation Reactions
Using Transition Metal
Complexes
Daniel H. Doughty and Louis H. Pignolet

1. INTRODUCTION

The decarbonylation of aldehydes, acyl halides, aroyl halides, alcohols,


or ketones is a useful and important reaction in organic synthesis. 0,2,3)
Although several methods not utilizing transition metals are known(4,5)
(including various deformylation reactions and thermal and photochemical
decarbonylations), they are not general and not usually applicable under
mild conditions where undesirable side reactions are minimized.o)
Several transition metal complexes have been reported to function as
stoichiometric homogeneous decarbonylation reagents under mild condi-
tions. The following discussion will show that the organic products observed
upon de carbonyl at ion of a given substrate depend strongly on the catalyst
and type of substrate to be decarbonylated. Completely different reaction
products are observed, for example, when different types of acid chlorides
are used with the same catalyst. This fact must be considered when trying
to formulate a general reaction mechanism for these decarbonylation reac-
tions. The earliest reported metal-promoted decarbonylation reactions are

Daniel H, Doughty, • Exploratory Chemistry, Div. 8315, Sandia National Laboratories,


Livermore, California 94550.
Dr. Louis H. Pignolet • Department of Chemistry, University of Minnesota, Minneapolis,
Minnesota 55455.

343
344 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

the formation of ruthenium carbonyl compounds upon refluxing ruthenium


trichloride hydriate in alcoholic solvents containing tertiary phosphines and
base. Using th,s method, complexes such as RhHCI(CO)(PPh 3h were
prepared by Ch~tt and Shaw(6.7) and Vaska(8,9) in 1960 and 1961. In 1964,
Chatt, Shaw, amd Field proposed a mechanism(1O) that involved a metal-
promoted oxidation of the alcohol to the corresponding aldehyde and
subsequent decarbonylation of the aldehyde. In support of this mechanism,
they reported that methane and acetaldehyde were observed when ethanol
was decarbonylated.
Another early report of this class of reactions was by Vaska, (11) showing
that iridium halides, in the presence of triphenylphosphine, abstract CO
from alcohols to give IrCI(CO)(PPh3h. Dimethylformamide is also de car-
bonylated by this reaction scheme. (12) Similarly, OsBr2(PPh3h reacts with
alcohols to yield OsHBr(CO)(PPh 3h and the corresponding alkane. (13)
In 1966, Prince and Raspin reported the decarbonylation of saturated
aldehydes by R\il2Ch(PEt2Ph)6 +Cl-. (14,15) The metal product formed in the
absence of ox~~en was RuCh(CO)(PEt 2Phh.(6) These decarbonylation
reactions were slow, generally requiring approximately 80 hours to undergo
several transformations upon heating in solution (16,18) and so the actual
reactive species is not known. The organic products depended upon the
aldehyde that was used. With acetaldehyde, the product was methane.
With propanal and butanal, however, a majority of the hydrocarbon product
was the alkene. for example, with butanal at 70-80°C, the ratio of propene
to propane was. 9 : 1. The overall stoichiometry of the reaction was such
that one mole of ruthenium produced one mole of hydrocarbon and one
mole of RuCh(CO)(PEt 2Phh.
The hydrogen that is liberated when the ole fins are formed is consumed
by the reduction of butyraldehyde to butyl alcohol, but this cannot account
for all the hydrogen formed. Labeling studies using deuterated aldehyde
C 2H 5CDO produced a mixture of products: C 2H 6 (40-45%), C 2H5D
(5-10%), and C2H 4 (50%). When another deuterated aldehyde,
CD 3CH 2CHO, was used, many different products were observed. Inter-
molecular hydrogen transfer was postulated to account for the lack of
product selectivity.
Vaska's complex, IrCI(CO)(PPh3h. decarbonylates acyl halides
catalytically at 78°C to give mixtures of olefins. (19) Aroyl halides are not
decarbonylated, even at high temperatures, although oxidative addition is
observed. The difference in behavior between the two kinds of acid halides
is explained in terms of the greater stability of the metal-carbon bond in
the aryl complex and the absence of a ~-hydrogen in the aryl complex,
making elimination of olefins impossible. A mechanism for the catalytic
decarbonylation of acyl halides has been proposed(19) and is given in
Scheme 1.
DECARBONYLA TION REACTIONS 345

IrCI(CO)(PPh 3h

."H,CH,COX j +RCH,CH,COX

-co -olefin

PPh 3
-co I /CH2CH~ -olefin
OC-Ir"
I Cl
X

\'----_ _~-HX==__________'/ +RCH,CH,COX

Scheme 1

The main reaction path is believed to be the second route in which


PPh 3 has been eliminated. This accounts for the relatively long induction
times. Acyl- and alkyl-iridium intermediates have been isolated, including
some containing only one phosphine ligand. (19)
The best studied decarbonylation reagent is chlorotris(triphenylphos-
phine)rhodium(I), RhCI(PPh 3 h, which decarbonylates aldehydes and acyl
and aroyl halides under mild conditions (i.e., thermally in solution at
temperatures below lOO°C). The reactions are stoichiometric and are sum-
marized by Equations 1-4. (20-24)

o
II
RCH + RhCJ(PPh 3 h ~ RH + RhCl(CO)(PPh 3 h + PPh 3 (1)
o
II
R 2 CHCR 2 CH + RhCl(PPh 3 h ~ R 2 C=CR 2 + H2 + RhCl(CO)(PPh 3 h + PPh 3 (2)

o
II
RCCI + RhCI(PPh 3 h ~ RCI + RhCI(CO)(PPh 3 h + PPh 3 (3)
346 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

The olefin producing reactions, Equations 2 and 4, occur when a P-H is


present, but th~y are not equally important. While reaction 2 is of minor
importance (e.g., decarbonylation of heptanal using RhCI(PPh 3h yields
86% hexane and 14% I-hexene),(22) acid chlorides will be completely
converted to olefin if a p-hydrogen is present (Equation 4). Hence,
RhCI(PPh 3h is useful for the stoichiometric conversion of aldehydes into
alkanes and ole fins (if P-H is present) and of acid halides into alkyl or aryl
halides and olefins (if P-H is present). Stoichiometric decarbonylation of
general ketones using transition metal complexes has not been reported,
although a brief mention of this possibility has appeared, (25) and several
activated ketones have presumably been decarbonylated
stoichiometrically. (26)
Decarbonylation of transition-metal carbonyl complexes is also
affected by RhCI(PPh 3h. For example, CO is abstracted from
MO(CO)6, (27.28) W(CO)6, (27) and Fe(CO)5 (28) according to Equation S:

Alexander and Wojcicki(29.30) have used 13CO to show that when an acyl
transition metal carbonyl complex is decarbonylated, abstraction of the
carbonyl ligand occurs according to Equation 6:

RhCI(PPh 3h + 2( 11 5 -C 5H 5)Fe(COh 13 COR --+ RhCI(CO)(PPh 3h


+ (11 5-C 5 H 5 )Fe(CO)C 3 CO)R + (11 5-C5H5)Fe(PPh3)(CO)13COR (6)

The product distribution depends on the ability of the R group to migrate.


A recent review(31) discusses the many other catalytic and stoichiometric
reactions of RhCI(PPh 3h.
Catalytic decarbonylation of aldehydes using rhodium complexes of
chelating diphosphine ligands has been studied (vide infra). These com-
plexes give good catalytic activities (over 300 turnovers/hr for heptanal at
IS0°C) and are highly selective. The catalysts are stable for days, and
turnovers in excess of 10 5 have been achieved. Experiments on a variety
of aldehydes have established the general usefulness of this reaction in
organic synthesis (see Chapter 11, Section 3).
Even the observation of catalytic decarbonylation under ambient con-
ditions has been reported. (32) The catalyst precursor, Ru(PPh 3htpp (tpp =
dianion of tetraphenylporphyrin), is reported to decarbonylate phenyl-
acetaldehyde at very rapid rates, giving in excess of S x 10 4 turnovers/hr
at SO°C. Further research may yield even more active catalysts.
DECARBONYLA TlON REACTIONS 347

2. DISCUSSION OF DECARBONYLA TlON MECHANISM WITH


RhCI(PPh 3 J3

A considerable amount of work has been performed over the last 15


years to determine the mechanism of acid chloride decarbonylation with
RhCI(PPh 3h. (19,22-24,31,33,34) Although the discovery of aldehyde decar-
bonylation preceded that of acid chlorides, (35,36) much more time has been
spent on the acid chloride system because it is more easily studied. Many
intermediates have been isolated and characterized (see Table 1). Even
though the mechanism of the catalytic reaction is not well understood, the
mechanism for the stoichiometric decarbonylation of acid chlorides has
been proposed. However, the generally accepted mechanism has recently
been challenged. (37,38) In this section, we will first review the stoichiometric
decarbonylation mechanism for acid chlorides, followed by the
stoichiometric decarbonylation of aldehydes. Finally, the mechanism of
catalytic decarbonlyation of acid chlorides and aldehydes will be discussed.

2.1. Stoichiometric Decarbonvlation of Acid Chlorides

The general mechanism proposed for acid chloride stoichiometric


· (33 '34 "39 40) un d er ml'ld con d'Itions
d ecar bonyIation . . . . E quatlOns
IS gIven 10
.
7a-7e.
K« 1
RhCl(PPh 3 h ~ RhCl(PPh 3 h + PPh 3 (7a)
RhCl(PPh 3 h + RCOX ~ cis-RCO(X)RhCl(PPh 3 h (7b)

cis-RCO(X)RhCl(PPh 3 )2 ~ trans-RCO(X)RhCl(PPh 3 h (7c)


trans-RCO(X)RhCl(PPh 3 lz ~ R (X)RhCl(CO)(PPh 3 h (7 d)

R(X)RhCI(CO)(PPh 3 lz ~ RX + RhCl(CO)(PPh 3 h (7e)

When l3-hydrogens are present, i.e., R = R'CH 2CH 2, olefins and HCI are
predominantly formed, as shown in Equation 7f:

The only known exceptions (vide infra) to the above mechanism, Equations
7 a-7f, occur when R is an aryl group (37,38) or when R is an a, l3-unsaturated
species. (41) In the former case, decarbonylation is not observed under mild
conditions, and in the latter case, phosphonium salts are produced. Recent
work(38,39) has, however, substantiated the proposed mechanism for
aliphatic and substituted aliphatic acid chlorides,
Table 1. Isolated or Observed Intermediates during the Stoichiometric Decarbonylation of Acid Chlorides Using RhCI(PPh 3J3

Acid chloride Aroyl (or acyl) complex Aryl (or alkyl) complex Reference ~
til

MeCOCI RhCI 2(COMe)(PPh 3lz 20b


CH 2ClCOCI RhCh(COCH 2Cl)(PPh 3lz 45
EtCOCI RhCI 2(COEt)(PPh 3lz 20b
PrCOCI RhCI 2(COPr)(PPh 3h 45
Me(CH 2)4COCI RhCI2[CO(CH2)4Me](PPh3lz
PhCOCI RhCI 2(COPh)(PPh 3h RhCI 2(CO)(C 6H s )(PPh 3h 20b,24a
p-ClC 6H 4COCI RhCh(p-COC 6H 4CI)(PPh 3lz RhCI 2(CO)(C 6H 4Cl)(PPh 3lz 24a
p-N0 2C6H 4COCI RhCI2(p-COC6H4N02)(PPh3h RhCI2(CO)(C6H4N02)(PPh3h 24a
~
p-MeOC6H 4COCI RhCh(p-COC 6H 4OMe)(PPh 3lz RhCI2(CO)(C6H40CH3)(PPh3lz 24a <:
~
PhCH 2COCI RhCI 2(COCH 2Ph)(PPh 3lz RhCI2(CO)(CH2C6Hs)(PPh3h 24a
;t:
PhCH 2CH 2COCI RhCI 2(COCH 2CH 2Ph)(PPh 3lz RhCl 2(CO)(CH 2CH 2Ph)(PPh 3lz 46 t::J
C
threo-PhCHDCHDCOCI RhCI 2 (threo-COCHDCHDPh)(PPh 3 )2 46 c::
G')
=l:
C 6D sCD 2CH 2COCI RhCI2(COCH2CD2C6Ds)(PPh3lz 24c
~
p-ClC 6H 4CH 2COCI RhCI2(p-COCH2C6H4Cl)(PPh3lz RhCI2(CO)(CH2C6H4Cl)(PPh3) 24a ):.

p-N0 2C 6H 4CH 2COCI RhCI2(p-COCH2C6H4N02)(PPh3)2 RhCI2(CO)(CH2C6H4N02)(PPh3h 24a ~


r-
p-MeOC 6H 4CH 2COCI RhCI2(p-COCH2C6H40Me)(PPh3lz RhCh(CO)(CH2C6H40CH3)(PPh3lz 24a C
S;
0)
C S H 17 COCI RhCI2(COCsH17)(PPh3lz 22
;t:
Me(CH 2)n COCl a RhCI2[CO(CH2)nMe](PPh3h 1,22
l!
G')
erythro-MeCHPhCHPhCOCI RhCL 2(erythro-COCHPhCHPhMe )(PPh 3lz 24c <:
cr-
threo-MeCHPhCHPhCOCI RhCl 2(threo-COCHPhCHPhMe )(PPh 3lz 24c III
-t
(s)-( - )-a-CF3CHPhCOCI (s)-RhCh(CO-a-CF 3CHPh)(PPh 3lz (s)-RhCI 2(CO)(a-CF 3CHPhj(PEt 2Phlz 24b

an = 4, 5, 14. 16.
DECARBONYLA TfON REACTIONS 349

The stoichiometric de carbonyl at ion reaction begins with the oxidative


addition of acid chloride to RhCI(PPh 3h (Equation 7b), which is presumably
a solvent-stabilized, very reactive intermediate. (31) Tolman (42) and
Halpern (43) have presented kinetic evidence for the importance of
RhCI(PPh 3h in the catalytic hydrogenation of ole fins by RhCI(PPh 3h. In
addition, the solvated species, RhCI(S)(PPh 3h (where S = DMF,
acetonitrile), was observed in the stoichiometric decarbonylation of
aldehydes(44) (vide infra).
The 5-coordinate acyl complex has been the subject of many studies
since it is isolable from the reaction mixture (see Table 1). Proton and
phosphorus nmr studies show that the acyl complex isomerizes from a cis-
to a trans-phosphine stereochemistry which have different chemical shifts
and coupling constants(38,45) as shown in Table 2. In addition, the infrared
spectra of the isomers are different in the Rh-Cl and carbonyl regions (e.g.,
when R = CH2Ph, lieD = 1720 cm -1 for the cis isomer and lieD = 1770 cm- 1
for the trans isomer). (38)
The conversion of the cis isomer to the trans isomer has been clearly
demonstrated by monitoring the 1H and 31 p nmr as a function of time. (34,38)
However, without more detailed kinetic information, it is not certain that
the migration of the carbonyl group to form the alkyl complex, as shown
in Equation 7d, occurs exclusively from the trans isomer.
The detailed structure of these isomers remains somewhat clouded,
however. Preliminary results from single crystal X-ray structural determina-
tions of the intermediates are not in agreement. RhClz(COCH 2CH 2Ph)-
(PPh 3h is reported to have square pyramidal geometry, (34) whereas
RhClz(COCH 2Ph)(PPh 3h is reported to have a trigonal bipyramidal
structure. (33) The similarity in 31 p nmr spectra of these two compounds
makes it unlikely that the differences in reported solid-state structure are
maintained in solution. In addition, both reports of structural determination
were made in the preliminary stages of refinement and as such are subject

Table 2. Hand
1 31 P nmr data on AcVI-Rhodium Complexes, RhClz{CORJ-
(PPh 3 J2

lHnmr 31 p nmr

R Isomer 8,ppm 8, ppm J(Rh-P, Hz) Reference

CH 3 cis 3.37 (s) 29.8 (d) 145 34,45


CH 3 trans 2.49 (s) 23.6 (d) 108 34,45
CHzPh cis 5.29 (s) 27.3 (d) 145 38,45
CHzPh trans 4.26 (s) 23.4 (d) 108 38,45
350 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

to significant uncertainty. Since no further reports of these structures have


appeared from either group, this area merits further investigation.
Solution molecular weight determinations indicate that the acyl com-
plex is monomeric, (24) not dimeric with chI oro bridges. These 5-coordinate
complexes must have an open coordination site for the next step in the
reaction, alkyl migration, to occur, as given in Equation 7d.
Generally, migration occurs with retention of configuration at the
a-carbon. 0.24.46.47) For example, decarbonylation of an optically active acid
chloride, (S)-a- 2H-phenylacetyl chloride, yields S-benzyl-a-2H-chloride
with net retention of configuration. (47)
The elimination reaction, given in Equation 7e, is accompanied by a
formal reduction in the oxidation state of the rhodium from (III) to (1).
The stereochemistry of this reaction has also attracted substantial interest.
\
If (3-hydrogens are absent, the alkyl halide is eliminated with retention of
configuration at the carbon. (47) If (3-hydrogens are present, (3-hydride
elimination occurs as shown in Equation 7f, giving an alkene and hydrogen
h alI'd e. (24334647)
. " I nvestIgatlOns
.. .
mto t he (3 -h yd'd
nee I"ImmatlOn
. have sown
h
that Saytzeff (rather than Hoffman) elimination is preferred. (24)
In order to further study the stereochemistry of this elimination, acid
chlorides having asymmetric carbons at the a- and (3-position were reacted
with RhCl(PPh 3h. The alkyl complexes formed from erythro- and threo-
2,3-diphenylbutanoyl chloride are given in Equation 8. (24c)

Me Me
H*Ph Ph*H
(8)
Ph H Ph H
Rh Rh

erythro threo

Product analysis shows that the E isomer is produced from erythro-2,3-


diphenylbutanoyl chloride, and the Z isomer is produced from the threo-
acid chloride. Both these results are consistent with cis (3-hydride
elimination. (24,33.46) Further studies on the elimination reaction have shown
a large kinetic isotope effect (kH/ kD "'" 7) for decarbonylation of
PhCH 2CH 2COCl and PhCD 2CH2COCl. (24) This result suggests that (3-
hydride elimination is the rate-determining step and that C-H bond scission
is involved. (47) Failure to observe the alkyl complex led the authors to
postulate a concerted elimination from the acyl complex. (24) However, the
complex RhCh(CO)(CH2CH2C6H5)(PPh3h has been observed by
another group(46) when a toluene solution of RhC}z(COCH 2CH 2C6H 5)
(PPh3h was analyzed at 80°C by infrared spectroscopy (v co = 1996 cm -1).
Thus, it seems that cis (3-hydride elimination as given in Equation 7e is
most consistent with the available data.
DECARBDNYLA TION REACTIONS 351

The proposed mechanism (Equation 7) has been challenged


recently(37,38) on two points. First, benzoyl chloride did not produce
chlorobenzene under reaction conditions that should lead to stoichiometric
decarbonylation, in contradiction of earlier work. (24a,48) Under mild condi-
tions (30-80°C) benzoyl chloride did react with RhCI(PPh 3h, but the major
product was an aryl-rhodium phosphine complex, Rh(C 6 H s)Ch(PPh 3h (see
Equation 9), not RhCI(CO)(PPh 3 h and C6 H sCI, as expected in Equation
7e:
(9)
The previous workers have confirmed this result. (49)
The earlier kinetic investigation (24a) had used infrared spectroscopy
and UV -VIS electronic absorption spectroscopy to monitor the concentra-
tion of metal carbonyl compounds. They had often observed the production
of alkyl chlorides(24) or olefins plus HCI(24c,40) from alkyl-rhodium carbonyl
complexes at low temperatures and had relied (49) on an earlier incorrect
report of chlorobenzene production from benzoyl chloride. (48) Thus, they
misinterpreted the products formed on disappearance of the aryl-rhodium
carbonyl complex. (49)
This seemingly anomalous result can be understood by the increased
stability of the metal-carbon bond in aryl complexes. (50,51) This added
stability may give the aryl-rhodium carbonyl complex, Rh(C 6 H s)Ch(CO)
(PPh 3 h, another reaction path that is unfavorable for alkyl-rhodium car-
bonyl complexes. In fact, the stability of the metal-carbon bond was invoked
to explain the difference in reactivity between alkyl- and aryl-acid chlorides
with IrCI(CO)(PPh 3 h. (19) Only the former were decarbonylated, even at
elevated temperature. In addition, aryl-rhodium phosphine complexes of
the type observed in the above report have been postulated(21) as intermedi-
ates in the catalytic decarbonylation of para-substituted benzoyl chlorides.
For example, when p-chlorobenzoyl chloride was heated briefly with
RhCI(PPh 3 h, yellow rhodium complexes were obtained (but not isolated
in analytical purity) that did not show any carbonyl absorptions in the
infrared (21) and were postulated to be the aryl-rhodium complex
Rh(C 6 H 4CI)Ch(PPh 3 h.
The other exception to the general mechanism occurs when an a, {3-
unsaturated acid chloride is used as the substrate. The presence of a
{3-hydrogen would allow the possibility of {3-hydride elimination to occur,
but the product would have a terminal carbon-carbon triple bond, which
has not been observed previously. Apparently, reductive elimination of
R-CI is also too slow, and as a result phosphonium salts are the major
product (Equation 10):
R-CH=CH r RhCI 2 (CO)(PPh 3 h + PPh 3
~ RhCl(CO)(PPh 3 h + R-CH=CH r PPh 3 Cl- (10)
352 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

Both of these results are important in that they point out that the
reductive elimination step is poorly characterized for these complexes. It
is clear that the reaction will be sensitive to the relative bond strengths of
reactants and products. The strength of the metal-carbon and metal-X
bond must be compared to the strength of bonds formed during the
elimination reaction. Very little thermodynamic data exist for organometal-
lic complexes, and so these comparisons are difficult to make. The general
trend is that reductive elimination becomes less favored as one progresses
in the series of products: olefin + HX > R-CI > phosphonium salts> ArCl.
The study of reductive elimination reactions from these types of complexes
merits further investigation.

2.2. Stoichiometric Decarbonylation of Aldehydes

The mechanism of aldehyde decarbonylation is thought to follow the


established mechanism for acyl halide decarbonylation discussed in the
previous section (Equation 7, where X = H). Several observations support
this idea, even though intermediates are much more labile than those of
the acid chloride system.
The only Rh(III) intermediate that has been isolated from the reaction
sequence is given below. (52) This compound was prepared by reacting an
aldehyde, 8-quinoline carboxaldehyde, with RhCI(PPh 3 h. The ability of
the aldehyde to form a chelate after oxidative addition has occurred (termed
"chelate trapping" by the author) imparted sufficient stability to the com-
pound to allow isolation and characterization. Prolonged heating in
refluxing xylene yields the expected decarbonylation products. Other
examples of oxidative addition of aldehydes to Rh(I) complexes are presen-
ted in Chapter 7, Section 4.

&:\I~~l
I
Ih cl
Rh
"H
II PPh 3
o
The stereochemistry of aldehyde decarbonylation has received much
attention. Walborski and Allen (53) have shown that the decarbonylation of
optically active aldehydes proceeds with 93% retention of configuration,as
shown in Equation 11:

(11)
DECARBONYLA TION REACTIONS 353

Retention of configuation at the a-carbon was also observed with acid


chlorides (vide supra), but aldehydes give a product of higher optical
purity. (47,53)
Decarbonylation of deuteroaldehydes has been used for specific
deuteration of compounds. The incorporation of deuterium into the prod-
ucts as given in reaction 12 occurs with high yield. (53,54)

Ph, 1\ ,CDO Ph, 1\ P + RhCI(CO)(PPh 3 h + PPh 3


~ + RhCI(PPh 3 h ~ ~
Ph Me Ph Me

(12)

The decarbonylation of other de ute rated aldehydes provides deuter-


ated products according to Equations 13 and 14. (15)

RhCI(PPh 3 h + EtCDO ~ EtD + RhCI(CO)(PPh 3 h + PPh 3 (13)

RhCI(PPh 3 h + CD 3 CH 2 CHO ~ CD 3 CH 3 + RhCI(CO)(PPh 3 h + PPh 3 (14)

RhCI(PPh 3h is known to isomerize double bonds during


decarbonylation. (22) For example, a-methyl cinnamaldehyde is converted
to cis- and trans-f3-methylstyrene, as shown in Equation 15:

Ph
140°C , , /CH 3 Ph
,,/
H
C=C + C=C
H
/ "H H
/ "-CH 3
(91 %) cis (9%) trans
+ RhCI(CO)(PPh 3 h + PPh 3 (15)

Isomerization does not occur. (54)


When f3-hydrogens are present in the substrate aldehyde, olefins and
H2 are produced, but in substantially smaller quantities than observed with
analogous acid chlorides. For example, decarbonylation of heptanal yields
14 % hexene, (22) whereas exclusive formation of alkene is observed from
similar acyl chlorides. This difference is due to the ease with which hydride
transfer (as opposed to chloride transfer) occurs in the respective alkyl
intermediates. (30)
The results discussed above lead to the conclusion that the mechanism
of decarbonylation of aldehydes is very similar to that postulated for the
decarbonylation of acid chlorides. However, kinetic studies of the reaction
show that a different rate -limiting step is operative with aldehydes. With
acid chlorides, the rate-limiting step is thought to be migration or reductive
elimination, depending on the R-group. (24) A detailed kinetic study on the
stoichiometric decarbonylation of aldehydes with RhCI(PPh 3h (44) has
354 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

shown a different kinetic behavior. This result is consistent with the observa-
tion that Rh(III) intermediates are insolable in the acid chloride system
but not with aldehydes.
The kinetic study(44) was performed on RhCI(PPh 3 h and the chlorine-
bridged dimer, Rh 2 Clz(PPh3 )4, with a variety of aldehydes. The progress
of the reaction was monitored by UV -VIS electronic absorption spectros-
copy at 25°C, and decarbonylation products RhCI(CO)(PPh 3 h and alkane
were analyzed.
The study shows that the decarbonylation reaction is second-order,
and a rapid pre-equilibrium step is observed when RCHO forms an associ-
ation complex with RhCI(PPh 3 h, as given in Equation 16:
Kj
RCHO + RhCl(PPh 3 h ~ RhCl(RCHO)(PPh 3 lz + PPh 3 (16)
RhCl(RCHO)(PPh 3 lz ~ RhCl(CO)(PPh 3 lz + RH

For the reaction of propanal with RhC1(PPh 3 h at 25°C in benzene,


Kl = 3.7 ± 0.6 x 10-3 and k2 = 8.7 ± 2 x 10-4 sec-l. Phosphine inhibition
was measured and the observed kinetics agree with the rate law given in
Equation 17:

d[RhCl(CO)(PPh 3 h] = [ k 2 K,[RCHO] ] [Rh]


(17)
dt Kj[RCHO] + [PPh 3 ] total

Two isobestic points were observed during the course of the reaction.
In order to study the reaction without excess phosphine, the rhodium
dimer, Rh 2 Clz(PPh 3 )4, was used as the reagent. Coordinating solvents (DMF
or CH 2 CN = S) were added to stabilize the formation of the solvated
species, RhCI(S)(PPh 3 h. Decarbonylation of propanal by this solution
yielded the expected products, ethane and RhCI(CO)(PPh 3 h. Again, the
solution was monitored by UV-VIS spectroscopy, and isobestic points were
observed that corresponded to mixtures of RhCI(S)(PPh 3 h and
RhCI(CO)(PPh 3 h with no indication of intermediates. (44) The kinetic
behavior was consistent with either a rapid pre-equilibrium step or the use
of a steady-state approximation on RhCl(RCHO)(S)(PPh 3 h.
The postulated mechanism is shown in Equations 18 and 19.
kj
RhCl(S)(PPh 3 lz + RCHO ~ RhCl(RCHO)(S)(PPh 3 lz (18)
L j

RhCl(RCHO)(S)(PPh 3 lz --S RhCl(CO)(PPh 3 lz + S + RH (19)

Since the rate laws for the two mechanisms (rapid pre-equilibrium or
steady-state approximation) are indistinguishable, the mechanism could
not clearly be established.
DECARBONYLA TION REACTIONS 355

The work(44) also provided activation parameters: for propanal,


Ml'" = 13.2 ± 0.3 kcal/mol, ilS'" = -18 ± 1 e.u.; for n-butanal, ilH'" =
12.9 ± 1 kcal/mol, ilS'" = -19 ± 3 e.u. In addition, the kinetic isotope
effect was determined for propanal: kH/kD = 1.55. These activation para-
meters are very similar to values measured for the oxidative addition of
H2 and O 2 to IrCI(CO)(PPh 3 h. (55-57)
The conclusions, therefore, are that oxidative addition is the rate-
limiting step and that C-H(D) bond breaking (or making) is involved to a
small extent. This conclusion is further supported by the fact that electronic
effects measured by analysis of a Hammett plot for para-substituted benzal-
dehydes gave p = 1.5, nearly identical to the value (p = 1.4) obtained for
oxidative addition of para-substituted benzoylchlorides to IrCl(CO)-
(PPh 3 h. (56)
Recent work by Kampmeier,3'J) has excluded the possibility of free-
radical intermediates taking part in the decarbonylation of aldehydes.
Decarbonylation of exo- and endo-5-norbornene-2-carboxaldehyde gives
norbornene and nortricyclane, respectively. Reaction of citronella with
RhCl(PPh 3 h gives only 2,6-dimethyl-2-heptene. Thus, the individual steps
in the decarbonylation mechanism do not exhibit free-radical character
and must be intramolecular. This work also reported a kinetic isotope
effect, kH/kD = 1.86, for the decarbonylation of C6H 5CH2CH(D)0 by
RhCI(PPh 3 h. These results imply that the oxidative addition step is a
concerted insertion of the metal into the C-H bond. These results are also
consistent with the previous work(44) which suggests that oxidative addition
is the rate-limiting step.

2.3. Catalytic Decarbonylation of Acid Chlorides and Aldehydes

At elevated temperature, RhCI(PPh 3 h becomes a catalyst (or catalyst


precursor) for the decarbonylation reaction. Acid chlorides are reported
to be decarbonylated catalytically at 170°C. (21,58.59) A recent report has
lowered the temperature where catalytic activity is observed to 111°C, (38)
but this appears to be more a result of lower detection limits than an
observation of different reactivity. Aldehydes are reported to be catalyti-
cally decarbonylated above 200°C(21,59) by RhCl(PPh 3h or RhCl(CO)-
(PPh 3 h.
The essential question to be addressed is whether the rhodium complex
re-enters the decarbonylation sequence previously discussed (Equation 7)
or if a different pathway is available at elevated temperature. Loss of CO
from RhCI(CO)(PPh 3 h is not observed even at high temperature (300°C)
or under UV irradiation. (60) Therefore, mechanisms that have been pro-
posed involve either oxidative addition to this complex(24,31,59) or loss of
CO from the alkyl-rhodium carbonyl complex. (21) These two mechanisms
356 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

PPh3 PPh 3
+RCOX 1 X OC 1 X
RhCl(PPh3h • R CO-Rh/ - - - - - + . "Rh/
4-- I'"'-CI - - R/I "Cl

-'~'H
PPh3 PPh 2

HRX

Scheme 2(24.31.591

are illustrated in Schemes 2 and 3, respectively, Scheme 2 seems more


consistent with experimental observations than does Scheme 3, Evidence
supporting Scheme 2 includes:
a, Rh(RCO)ChCO(PPh 3 h is observed upon reaction of
Rh(RCO)C}z(PPh 3 h with CO at room temperature for 90
minutes, (38) Replacing the CO atmosphere with nitrogen causes
dissociation of CO from the complex followed by reductive elimina-
tion of the product, RCI (R = C6 H sCH 2 ), to produce RhCI(CO)-
(PPh 3 h.

PPh3
1 X
R-Rh/
1 "Cl ~ ~ (major pa1h)
PPh 3 ~ -co

RhCl(PPh3h

- PPh3 1I ":::::::::'-co:"". . . . . . . . . . . . .
RhCl(PPh 3h ' :: ~ ~Ph3
OC-Rh-Cl
1

PPh3

Scheme 3(211
DECARBONYLA TfON REACTIONS 357

b. RhCI(CO)(PPh 3h is a catalyst for the carbonylation of certain


alkyl halides, (59) as shown in Equation 20.

PhCH 2 Ci 100 atm, ~~'c ~ PhCH 2 COCl (20)


RhCI(CO)(PPh 3 h

c. Oxidative addition of acid chlorides (e.g., acetyl chloride(61)) occurs,


albeit slowly, at room temperature to RhCI(CO)(PPh 3h.
Scheme 3 is supported by the isolation from the reaction mixture of
a yellow rhodium complex that does not exhibit carbonyl stretching frequen-
cies in the infrared. (21) This compound, originally postulated to be RhClz(Ar)
(PPh3h, has been observed in recent work by Kampmeier, where Ar =
C6H5 and p-CH3C6H 4 • (38,39) It is possible that both mechanisms are correct
for certain specific aldehydes or acid chlorides. Additional work on phos-
phine inhibition of the catalytic reaction, however, leads to the conclusion
that another mechanism may be important at elevated temperature, (38)
Phosphine inhibition is observed in two types of experiments. Catalytic
activities were measured for several catalyst precursors having different
PPh3/Rh ratios, (38) Compounds studied were: RhCI(PPh 3h, [PPh 3/Rh] =
3; RhCl(CO)(PPh 3h and Rh 2Clz(PPh 3)4, [PPh3/Rh] = 2; and
Rh 2Clz(COh(PPh 3h. [PPh 3/Rh] = 1. The second method was to simply
add excess PPh 3. Results(38) for decarbonylation of phenylacetyl chloride
at 111°C are listed in Table 3. The results show a consistent trend of
increasing catalytic activity as the PPh3/Rh ratio decreases to 1.0.
This inhibition by PPh 3 can be accommodated better by Scheme 3
than by Scheme 2. Phosphine inhibition only occurs in the original

Table 3. The Catalytic Activity of Decarbonylation of Phenylacetyl Chloride


at 111°C B

Catalytic
activity
Catalyst PPh 3 b (eq.) PPh 3 /Rh Time (h) PhCH 2 Ci c (turnovers/hr)

RhCl(PPh 3 b 3 5.5 25 4,5


RhCl(CO)(PPh 3 h 3 6 24 4.0
RhCl(CO)(PPh 3 h 2 2 41 20.5
Rh 2 CI 2 (PPh 3 )4 2 2 37 18.5
Rh 2 Ci 2 (COh(PPh 3 h 1.25 44 35.2

a Taken from Reference 38, [Catalyst]o = 0.06 M; mol [PhCH 2 COCl]/mol [Catalyst]o = 45; nitrogen
sweep.
b Equivalents of PPh 3 added per mol [Catalyst]o'
, Mol [PhCH 2 Cl]/mol [Catalyst]o.
358 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

dissociation of PPh 3 from RhCI(PPh 3 h, not in the ensuing catalytic sequence


given in Scheme 2. Scheme 3, on the other hand, postulated the catalytically
active intermediate RhCI(PPh 3 h, and the concentration of this species is
related to the concentrations of RhCI(PPh 3 h and PPh3 through the dissoci-
ation equilibrium constant. Increasing the concentration of PPh3 would
diminish the concentration of RhCI(PPh 3 h. thus lowing the reaction rate.
However, since RhCI(PPh 3 h is not observed upon quenching the reaction,
this explanation seems tenuous.
These results are best explained if dissociation of PPh 3 from one of
the intermediates in the above mentioned schemes led to another more
reactive intermediate. This type of mechanism was proposed for
decarbonylation of RCH 2 CH 2 COX with IrCI(CO)(PPh 3 h (19) and is pre-
sented in the lower pathway of Scheme 1. In order to explain phosphine
inhibition observed with the rhodium system, a similar mechanism has been
tentatively proposed. (39)
It seems reasonable that some alternate pathway may exist for the
following reasons:
a. Migration reactions and reductive elimination are not inhibited by
excess phosphine. (24.38)
b. The observed phosphine inhibiition is too large to be attributed to
reversible association of PPh 3 to RhCI(CO)(PPh 3 h, (38) which
would have the result of diminishing the reaction rate by lowering
the concentration of the metal complex required for oxidative
addition.
c. While it is possible that excess PPh 3 may retard CO dissociation
from RhC}z(RCO)(PPh 3 h, there is no precedent to show that this
reaction should occur.
A plausible, but speculative, reaction scheme for the catalytic decar-
bonylation of acid chlorides (and aldehydes) that involves phosphine dis-
sociation from one of the intermediates has been postulated. (24c.38) It is
clear, though, that there are presently not enough facts to substantiate this
hypothesis. Future work on the mechanism of catalytic decarbonylation
using RhCI(CO)(PPh 3 h and other catalysts which investigates phosphine
inhibition could be very informative.

3. CA TAL YTIC DECARBONYLA TION OF ALDEHYDES WITH


CA TlONIC DIPHOSPHINE COMPLEXES OF Rh(l)

As discussed in the previous sections, the stoichiometric decarbonyla-


tion of aldehydes is effectively carried out by using RhCI(PPh 3 h. It would
DECARBONYLA TfON REACTIONS 359

be extremely advantageous if this reaction made catalytic at mild tem-


peratures (i.e., <: 100°C), especially when the high cost of rhodium is
considered. Catalytic decarbonylation of aldehydes using trans-RhCI(CO)-
(PPh 3 h requires very high temperatures, and at 178°C a catalytic activity
of only 10 turnovers per hour is observed. (62) This temperature is too high
to be of practical synthetic use. Early attempts at making this reaction
catalytic at mild temperatures involved methods that were designed to
labilize CO from trans-RhCI(CO)(PPh 3 h, thereby regenerating the reac-
tive catalyst "RhCI(PPh 3 h" (see Equation 7). For example, photolysis of
trans-RhCI(CO)(PPh 3 h was examined, but unfortunately the CO ligand
was not photolabilized. (60) Other attempts that met with some success
involved the use of solvent-stabilized cationic complexes of the types
[Rh(PPh 3 ht and [Rh(PPh 3 h(COW. (65) The latter complex reversibly binds
CO. The increased lability of CO compared with that of the neutral complex
trans-RhCI(CO)(PPh 3 h is presumably due in part to the lowered effective
basicity of rhodium. [Rh(PPh 3 ht was shown to decarbonylate benzal-
dehyde at 100°C catalytically and with a rate considerably faster than that
obtained using trans-RhCI(CO)(PPh 3 h. (66) Unfortunately, this catalyst
system decomposed readily at 100°C and therefore the reaction was not
synthetically useful. (66)
Cationic complexes of Rh(I) with chelating diphosphine ligands have
been known for some time but only recently have their catalytic properties
been examined (see Chapter 4). (62.64,66-70) Complexes of the types [Rh(P-
Ph]x and [Rh(P-P)]X, where X = CI or BF4 and P-P = Ph2P(CH2)nPPh2
(named dppm, dppe, dppp, and dppb for n = 1,2,3, and 4, respectively),
are known to bind CO weakly and reversibly, and in the case of [Rh(dppeht
the reaction with CO does not even give a detectable adduct. The fact that
these complexes require cis phosphine stereochemistry is important and
results in major reactivity differences compared to their triphenylphosphine
analogues. (71) In the case of aldehyde decarbonylation, this difference is
quite remarkable. For example, the catalytic decarbonylation of benzal-
dehyde at lS0°C using [Rh(dppph]BF4 gives a catalytic activity for benzene
production of 1.1 x 10 2 turnovers per hour, whereas an activity of only
0.60 turnovers per hour was obtained using trans-RhCI(CO)(PPh 3 )z. (64,66)
Data for the catalytic decarbonylation of benzaldehyde using diphosphine
and triphenylphosphine complexes of Rh(I) is presented in Table 4.
The major points concerning the catalytic decarbonylation of benzal-
dehyde using diphosphine complexes are: (i) the activities are significantly
larger than with RhCI(PPh 3 h; (ii) the activities show a marked dependence
on chelate ring size with the order in activity dppp > dppe > dppb > dppm
observed at all temperatures; (iii) the activities using [Rh(P-Pht are
approximately two times larger than with [Rh(P-P) t; (iv) the activiites are
independent of the counter ion type (Cl-, BF4 -, or PF6-); (v) the activities
360 DANIEL H. DOUGHTY AND LOUIS H. PIG NOLET

Table 4. Catalytic Decarbonylation of Benzaldehyde Using Rhodium


Catalysts B

Catalytic activity
(moles benzene/moles Temperature
Catalyst catalyst/hour)b °C

RhCl(PPh 3 h Stoichiometric 115


[Rh(dppmlzr 0.40 115
[Rh(dppelzr 3.6 115
[Rh(dppphr 11 115
[Rh(dppblzr 1.2 115
RhCl(PPh 3 h 0.60 150
[Rh(dppplzr 1.1 x 102 150
Rh(dppp)Cl 30 145
[Rh(dppp)]BF4 40 145
[Rh(dppplzr 1.1 X 103 178
[Rh(dppelzr 2.1 X 10 2 178
[Rh(dppmlzr 7.4 X 10 1 178
RhCl(PPh 3 h c 1.0 X 10 1 178

a Catalyst concentrations between 1 x 10-4 and 1 x 10- 3 M and neat benzaldehyde as the solvent. (See
References 62 and 66 for details.)
b Values for benzene production are averaged over a 30-40 hr period and first-order dependence on
[catalyst] is assumed.
, Same results using trans-RhCl(CO)(PPh 3 )2'

for the [Rh(p-Phr complexes remain approximately constant for at least


several days, whereas the [Rh(P-P)t catalyst systems are less stable; and
(vi) the [Rh(dpppht system shows catalytic behavior at temperatures as
low as 100°C (activity = 3 turnovers hour-i). Since the [Rh(dpppht com-
plex is the best diphosphine catalyst, the remaining discussion will concen-
trate on this system. A wide variety of aldehydes have been decarbonylated
using this catalyst. Results are presented in Table 5.
In order to achieve the activities tabulated, it is important to run the
reactions under a continuous purge of dinitrogen. If the concentration of
CO is allowed to build up, the reactions become significantly inhibited.
This observation has important mechanistic implications (vide infra) and
is of great practical importance. Also, oxygen irreversibly destroys the
catalyst, so the gas purge must utilize nitrogen that has been purified
(passage through an activated BASF catalyst column works well). The data
DECARBONYLA TION REACTIONS 361

Table 5. Catalytic Decarbonylation of Aldehyde Using {Rh(dpppMBF/66!

Catalytic
Observed volatile Temperature, activity
Aldehyde product °C (turnovers/hr)

Heptanal Hexane 150 3.3 x 10 2


2-Ethylbutanal Pentane 115 5.0
2-Phenyl-2-methylbutanal 2-Phenylbutane 180 1.3

2-Phenylpropanal Ethylbenzene 150 6.7 x 10 1


Cinnamaldehyde Styrene 150 1.4 x 10 2
a-Methylcinnamaldehyde cis-/3-Methylstyrene 150 1.6 x 10 1
Benzaldehyde Benzene 150 9.0 x 10 1
para- Tolylbenzaldehyde Toluene 150 3.4 x 10 1
para-Methoxybenzaldehyde Anisole 150 1.0 x 10 1
para-Chlorobenzaldehyde Chlorobenzene 150 1.0

in Table 5 illustrate the general usefulness of this catalyst. Important points


are that the reaction shows a high degree of selectivity and that the activity
is much lower for sterically hindered tertiary aldehydes. The selectivity is
best illustrated by noting that heptanal and 2-phenylpropanal yield only
saturated products. Decarbonylation of these aldehydes using RhCI(PPh 3 h
at similar temperatures produces significant amounts of hexene and styrene,
respectively. Importantly, decarbonylation of heptanal using the mono-
diphosphine complex [Rh(dppp)]BF4 is not selective. In this case hexene
and hexane are produced in a 1: 2 mole ratio at 150°C. The catalytic
decarbonylation of the optically active aldehyde (-)-(R)-2-phenyl-2-
methylbutanal using [Rh(dppp)]BF4 at 165°C produced (+)-(s)-2-phenyl-
butane with 100% retention of absolute configuration. (72)
For most aldehydes studied, the decarbonylation reaction occurs with
high product yields (yields based on aldehyde). Typical data are shown in
Table 6 using [Rh(dpppht as the catalyst. However, the saturated
aldehydes, heptanal and 2-ethylbutanal, are decarbonylated to hexane and
pentane, respectively, in rather low yield (but with good catalytic activities,
see Table 5). Further experiments revealed that the low yields observed
for these aldehydes are due to thermal decomposition, which also occurs
in the absence of catalyst. Even lower yields (based on aldehyde) are
obtained for these aldehydes using RhCI(PPh 3 h at the same temperature,
since with this complex decarbonylation activities are much smaller.
362 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

Table 6. Yield and Conversion for Various Aldehydes·

Temp., Yield, Conversion,


Aldehyde (mmol) Time °C Product %b %b

Benzaldehyde (2.45) 2 days 140 Benzene 79 100


Cinnamaldehyde (4.00) 2 days 118 Styrene 95 100
2-Phenylpropanal (64.0) 1.5 days 175 Ethylbenzene 83 93
Heptanal (1.85) 2 days 118 Hexane 39 46
2-Ethylbutanal (60.8) 1.5 days 110 Pentane 15 43

a 0.02 to 0.05 mmol catalyst, Rh(dppp), + in solvent m-xylene or a-methyl-naphthalene.


b % yield = (moles decarbonylation product/moles aldehyde added) x 100; % conversion = (moles de car-
bonylation product/moles aldehyde consumed) x 100.

3.1. Discussion of Mechanism with [Rh(P-PM+ Complexes

In order to gain insight into the mechanism of this catalytic reaction


a variety of experiments have been carried out. Monitoring of the reaction
during actual catalytic conditions by 31 p nmr and UV -VIS absorption
spectroscopy reveals that only the complex [Rh(P-Pht is present in any
significant concentration (only cases where P-P = dppp and dppe and
aldehyde = benzaldehyde have been examined in detail). Upon quenching
the reaction by the rapid precipitation of rhodium species (via addition of
cold pentane), the complexes [Rh(dpppht and [Rh(dppphcot are
observed, and in the case of the dppe system only [Rh(dppeht
is observed. The original catalytic activity can be restored by reusing the
complexes isolated from the reaction in a new experiment. During at least
a two-day period of running the catalytic reaction (benzaldehyde----.
benzene), there are no signs of Rh(III) intermediates and catalyst decompo-
sition is minimal. Identical activities are obtained using [Rh(dppphcot
or [Rh(dpppht as the catalyst.
Deuterium-labeling experiments have been carried out. Deuterobenz-
aldehyde, C6 H sCDO, and p-tolualdehyde, CH 3C6 H 4 CHO, were mixed in
equimolar amounts and decarbonylated at 140°C using [Rh(dpppht.
Toluene and benzene were produced simultaneously, separated by GLC
and analyzed by mass spectroscopy (MS). The isotopic purities of C6 H sCH3
and C6 HsD were determined to be 100 ± 1 % and 100 ± 4%, respectively,
and therefore all possible processes that are intermolecular in aldehyde
can be excluded. This experiment, along with the one that used an optically
active aldehyde (vide supra), argues against a free-radical chain mechanism.
The deuterium isotope effect (kH/kD) for C6HSC(H, D)O was 1.6 ± 0.1
using [Rh(dpppht at 140°C. A deuterium isotope effect of 1.55 has been
DECARBONYLA TlON REACTIONS 363

observed for the stoichiometric decarbonylation of propionaldehyde at


25°C with RhCI(PPh 3h(DMF), and this result was used to support the
theory that oxidative addition of the aldehyde is the rate-determining
step. (44)
The above results are consistent with a mechanism that involves
oxidative addition of the aldehyde, R-group migration (deinsertion), and
reductive elimination of RH. The active catalyst is then regenerated by
rapid CO loss from [Rh(dppphcot. Loss of CO from this complex is
known to be a facile process. (66) Also, since Rh(III) intermediates are not
observed, the migration and reductive elimination steps cannot be rate-
limiting. There are several problems with this simple mechanistic picture.
First, the para-substituted benzaldehyde substituent effect on activity does
not follow a simple trend as would be expected if oxidative addition was
rate-limiting. With [Rh(dpppht the trend in kobs. where rate =
kobs[complex], is para-H > -CH3 > -OCH3 > -Cl. In cases where oxidative
addition is known to be rate-limiting, the observed trend in rate constant
is para-CI > -H > -CH3 > -OCH3. (44,73) Also, the decarbonylation of
benzaldehyde is inhibited by the presence of added reagents such as CO,
benzonitrile, PEtPh 2 , and dppp. This is not unexpected for [Rh(dpppht,
where the 5-coordinate adducts [Rh(dppp)zLt are formed (vide injra),(64)
thereby diminishing the active catalyst concentration. However, inhibition
also occurs using [Rh(dppe)zt as catalyst, although this complex does not
form 5-coordinate adducts with these ligands. Finally, the absence of a
large kinetic isotope effect suggests that C-H bond breaking or making
may not be important in the transition state.
The rate of benzene production from neat benzaldehyde solution was
measured as a function of [Rh(dppp)zt concentration. The reaction is
first-order such that

rate = d[benzene]/dt = kobs[Rh(dppph +].

The rate of benzene production was also measured as a function of benzal-


dehyde concentration while keeping the concentration of the catalyst con-
stant. A plot of the rate constant, kobs. vs. benzaldehyde concentration is
shown in Figure 11.1. At low aldehyde concentrations the rate is first-order
in (benzaldehyde), but as the aldehyde concentration is increased a satura-
tion effect is observed. This behavior can be interpreted in terms of a rapid
pre-equilibrium followed by a slow rate-limiting step as shown by Equation
21. (64)

+ K +
Rh(dppph + PhCHO ~ (PhCHO)Rh(dppph (21)
fast
+ k +
(PhCHO)Rh(dppph ~ (PhCO)(H)Rh(dppph
364 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

35

30
25

,-
L
20 •

15
.cL
"'
D
0

10

5 .
00 2 4 6 8 10
[PhCHOJ ,M

Figure 11.1 Plot of kobs for benzene production vs. benzaldehyde concentration at 150°C
using [Rh(dppph]BF4 as catalyst.

The rate expression for this mechanism is


k _ kK[PhCHO]
obs - K[PhCHO] + 1
which requires
1 1 1( 1 )
kobs = k + kK [PhCHO]

A plot of l/kobs vs. 1/[PhCHO] is linear and is shown in Figure 11.2. An


analysis of the data in terms of this rate law gives k = 1.1 X 10 2 hr- 1 and
K = 0.12 M- 1 at 150°C. Additional support for the mechanism shown by
Equation 21 is the direct observation of the aldehyde adduct
(PhCHO)Rh(dppp)z + by low-temperature 31 p nmr. (64,74) At temperatures
above 25°C, this adduct is not observable by 31 p nmr due to its nonrigid
nature; however, its presence may be inferred.
Although the direct observation of [(PhCHO)Rh(dppp)zt lends sup-
port to Equation 21, the possibility exists that this species is unimportant
kinetically. For example, it is possible that the kinetically important inter-
mediate contains a monodentate dppp ligand such as [(PhCHO)Rh(dppp)-
(dppp*)t, where dppp* is monodentate. The rate law for this case would
be identical, and since the establishment of the pre-equilibrium is fast such
an intermediate is likely to be unobservable by 31 p nmr. The actual isolation
or observation of "intermediates" in catalytic reactions is known to often
lead to erroneous mechanistic predictions (see Chapter 4). (75)
The mechanism of Equation 21 is plausible and may indeed be opera-
tive for [Rh(dpppht; however, several pieces of data are a cause for
DECARBONYLA TfON REACTIONS 365

0.7

0.6

0.5
L
.c
0.4

,. 0
</l
.0 0.3
.Y
0.2

0.1
...
0
0 1.0 3.0 5.0 9.0
-1
[PhCHO] ,

Figure 11.2. Plot of k;;-~s vs. [PhCHOr 1 for benzaldehyde decarbonylation using
[Rh(dppphlBF4 as catalyst at 150°C.

concern. First, the lack of a large kinetic isotope effect may be inconsistent
with C-H bond breaking being involved in the rate-limiting step as required
by Equation 21. Second, if the reaction mechanism is the same for
[Rh(dpppht and [Rh(dppeht, the rate inhibition by CO (see Table 7) is
inconsistent with Equation 21 for the dppe system. [Rh(dpppht forms a

Table 7. Inhibition in the Rate of Catalytic Decarbonylation of Benzaldehyde a

Catalytic
Temperature, activity
Catalyst °C Inhibitor (turnovers/hour)

[Rh(dppplzt 145 None 60


[Rh(dppplzt 145 Benzonitrile (5.0 M) 3.3
[Rh(dpppht 145 CO (saturated) 0.1
[Rh(dpppht 145 PEtPh 2 (0.2 M) 0.2
[Rh(dpppht 115 None 11
[Rh(dpppht 115 PPh 3 (0.05 M) 7
[Rh(dppplzt 115 dppp (0.03 M) 0.1
[Rh(dppelzt 145 None 20
[Rh(dppeht 145 CO (saturated) 1.5

a For conversion of benzaldehyde into benzene, using [catalyst] = 1 mM and benzaldehyde as solvent, see
Reference 66 for details.
366 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

carbonyl adduct, and the inhibition observed in the rate of benzaldehyde


decarbonylation can be explained by the competitive equilibria shown in
Equation 22:
+ KI +
Rh(dppph + PhCHO +===! (PhCHO)Rh(dppph (22)
+ K2 +
Rh(dppp)z + CO +===! (CO)Rh(dppph

K2 is larger than Kh since the carbonyl stretch in the IR is very intense


when recorded using benzaldehyde saturated with CO as solvent. The
problem is that in the case of [Rh(dppeht, a CO adduct is not observed
by IR. Also, the low-temperature 31 p nmr of [Rh(dppeht recorded using
benzaldehyde and acetone as solvents shows no evidence for adduct forma-
tion. Therefore, the mechanism shown by Equation 21 is inconsistent with
the facts for [Rh(dppeht. An alternate mechanism, which is consistent
with all of the data for both catalysts, is shown in Equation 23.

Rh(p-ph + ~ Rh(p-p)(p-p*)+ (23)


* + + PhCHO
Rh(p-p)(p-p) r,-
K2
(PhCHO)Rh(p-p)(p-p*) +

(PhCHO)Rh(p-p)(p-p*)+ I::,. Rh(PhCO)(H)(p-p)(p-p*t


where p-p* = monodentate diphosphine (dppp or dppe)

The rate expression for this mechanism, assuming a steady-state concentra-


tion of the undetectable intermediate [Rh(dppp)(dppp*)t and assuming
that L2 is negligibly small, is

which requires

This rate expression is experimentally indistinguishable from the one


derived from Equation 21, since a plot of 1/kobs vs. 1/[PhCHO] is linear
for both mechanisms. From the data shown in Figure 11.2 and assuming
the mechanism in Equation 23, the values of kl and Ldk2 are 1.1 x
10 2 hr- 1 and 8.5 M, respectively.
The major difference between the mechanisms of Equations 21 and
23 is that in the former the rate-determining step is oxidative addition and
the aldehyde adduct concentration builds up, while in the latter Rh-P bond
rupture becomes the slow step as [aldehyde] increases. The absence of a
DECARBONYLA TfON REACTIONS 367

large kinetic effect for the [Rh(dppp)zt system favors the latter mechanism;
however, the magnitude of such effects is not well understood in
organometallic chemistry. Also, the inhibition by CO in the [Rh(dppeht
system is explained by the latter mechanism as a competition between CO
and aldehyde for the undetectable intermediate [Rh(dppe)(dppe*)t.
Therefore, the failure to observe a CO adduct in this system is explained,
since the concentration of [(CO)Rh(dppe)(dppe*)t needed for effective
inhibition would be very low and unobservable. It should be realized that
Equation 23 incorporates the features of Equation 21. For example, if
k3« k1' k- b kz, k-z, we have the essential features of Equation 21, except
that a monodentate dppp ligand is implicated in the aldehyde adduct. The
relative magnitudes of these rate constants may depend significantly on
the ligand and aldehyde type, and it is possible that k3 « kz, k-z, so that
the plot of Figure 11.2 will give values of k 1 and Ld k 3 • In fact, this latter
case is probably the most reasonable by analogy to results on the
stoichiometric decarbonylation of aldehydes using RhCI(PPh 3 h. (44) In this
case the initial PPh3 dissociation and aldehyde adduct formation occur
much faster than oxidative addition (rapid pre-equilibrium and slow rate-
limiting step).
In principle, the relative magnitudes of the specific rate constants can
be determined by complete kinetic analyses for the dppe and dppp systems
using several para-substituted benzaldehydes. Providing that all of these
systems exhibit the same mechanism, this series of experiments should
permit determinations of the specific rate and/or equilibrium constants of
Equation 23. Although the observed para-substituent effect on kobs for the
decarbonylation using [Rh(dppp)zt (Table 5) can be rationalized by either
mechanism, a detailed comparison of the specific rate constants and equili-
brium constants should permit a mechanistic distinction. These experiments
have not yet been carried out.
In light of the above arguments, an overall mechanistic scheme may
be postulated. Such a scheme is shown in Figure 11.3. Note that in this
scheme the steps of both Equations 21 and 23 are incorporated. For
Equation 21, k3 is rate-limiting whereas for Equation 23, k1 and kz are
the slow steps. The 5-coordinate CO and aldehyde adducts are only
observed in the dppp system, while all species with monodentate diphos-
phine ligands are unobservable. However, there are many examples in the
literature where metal complexes having monodentate diphosphine ligands
have been isolated and characterized. (76-79) This scheme provides an
hypothesis for future experiments.
There is good precedence for the oxidative addition of aldehydes to
meta I compIexes. (5Z "76 80) Th e migration
. . . e I"ImlOatlOn
an d re d uctIve . steps
are fast and are consistent with the high selectivity observed with the dppp
system (vide supra) and by analogy to known organometallic reactions. (81)
368 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

Figure 11.3. Possible mechanism for the catalytic decarbonylation of aldehydes using
[Rh(P-Pht as catalyst where (P-Pj = dppp or dppe.

For example, the decarbonylation of heptanal using [Rh(dpppht gives


hexane as the only volatile product, whereas both hexane and 1-hexene
are observed using [Rh(dppp)t or RhCI(PPh 3 h. In order for hexene to
be produced, a /3-hydride abstraction must occur from an intermediate that
contains Rh-R. As shown in Figure 11.3, the species that contains the
Rh-R grouping [Rh(R)(H)(CO)(dppp)(dppp*)t is coordinatively sat-
urated, and a /3-hydride abstraction is highly unlikely. Even if CO dissoci-
ation occurs or another Rh-P bond ruptures in this intermediate, the
dangling P atom would quickly resaturate the coordination. For
monodiphosphine complexes or for PPh 3 complexes, the analogous reaction
intermediates are coordinatively unsaturated and /3-H abstraction may
easily occur, leading to olefin and H2 production as shown in Equation 24.

S = solvent

The marked rate dependence on n, the number of methylene groups


in the chelate backbone (Table 5), warrants some discussion. The rate
DECARBONYLA TlON REACTIONS 369

increases as a function of n up to n = 3 but falls off significantly for n = 4.


A similar trend has been noted in other catalytic reactions using diphosphine
ligands, (70,82,83) and an understanding of this effect will obviously involve
both steric and electronic factors. For the catalytic decarbonylation of
aldehydes the rate of Rh-P bond rupture may be important (Figure 11.3).
The available evidence suggests that 5-membered chelate rings are more
inert to bond rupture than 6-membered chelate rings. (84) This could explain
the increase in rate ongoing from dppe to dppp. Additional data which
supports this is that [Rh(dppeht does not react with H2 (1 atm pressure),
while [Rh(dpppht reacts with H2 readily, giving a dihydride. (70) It is
expected that the 4- and 7 -membered chelate ring cases will be even more
labile. In the case of dppb, bimetallic complexes with a dppb ligand bridging
both metal centers are easily formed. (85) Indeed, the true form of
[Rh(dppbht in aldehyde or acetone solution is complex, and high con-
centrations of bimetallic species have been implicated. (74) The single crystal
X-ray structure of [Rh(dppbh]BF4 shows a highly distorted coordination
geometry (twisted ca,44% towards tetrahedral) with one unusually long
Rh-P bond. (74) Also, the reaction of this complex with CO gives the
bimetallic dppb-bridged complex [Rh 2 (COl4(dppbhf+. (85) Clearly, the
slow rate of decarbonylation using [Rh(dppbht could be due to the
presence of bimetallic species. Although it is expected that [Rh(dppmht
should be less stable than its dppe and dppp analogues, it is possible that
a monodentate dppm ligand lacks the flexibility to swing away from the
metal, thus enabling the transformations shown in Figure 11.3. It is likely
that a delicate balance is needed between Rh-P bond lability, chelate ring
flexibility, and thermodynamic stability of less reactive species.

4. CATAL YTIC DECARBONYLA TION OF ALDEHYDES USING


CA TlONIC DIPHOSPHINE COMPLEXES OF Ir(l)

Catalytic decarbonylation of benzaldehyde using several iridium com-


plexes has also been examined. (72) Results of these experiments are shown
in Table 8. The main points to be made here are: (i) [Ir(P-Pht catalysts
have activities that are ca. twenty times lower than their Rh analogs; (ii)
the iridium mono-diphosphine catalysts are better than the bis-diphosphine
Ir catalysts (opposite trend noted using rhodium, see Table 5); and (iii)
IrCl(CO)(PPh 3 h is a much better catalyst than RhCl(CO)(PPh 3 h and is
also better than most of the iridium diphosphine catalysts. The results for
the [M(P-Pht catalysts may be explained in terms of the proposed
mechanistic scheme in Figure 11.3. Since Ir-P bonds should be stronger
than Rh-P bonds, the value of kJ will be smaller for the Ir catalysts, thus
370 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

Table 8. Catalytic Decarbonylation of Benzaldehyde Using Iridium Catalysts·

Catalytic activity
(moles benzene/moles Temperature
Catalyst catalyst/hr) roC)

[Ir(dppeht 2.5 178


[Ir(dpppht 4.4 178
[Ir(dppbht 4.2 178
[Ir(dppphCOt 4.6 178
[Ir(dppp )(cod)]BF4 b 55 178
[Ir(dppp)(cod)]BF/ 27 150
[Ir(dppp)]BF/ 1.3 x 10 2 178
[Ir(dppp)]BF/ 39 150
IrCl(CO)(PPh 3 h 66 178
RhCl(CO)(PPh 3 h 10 178

a Catalyst concentration is 10-15 mg in 25 ml of neat benzaldehyde as solvent (see Reference 72 for details).
b Initial values observed during first several hours of reaction.
C Made in situ by first purging the system with H2 to hydrogenate the cod (1,5 -cycJooctadiene) to cycJooctane.

accounting for their lower activities. Also, since k3 should be larger for
the Ir system, (86) kl is likely to be the rate-limiting step. Evidence which
supports faster oxidative addition reactions for Ir over Rh is that benzal-
dehyde is decarbonylated much faster by IrCI(CO)(PPh 3 h than by its Rh
analogue (Table 8). In these complexes, oxidative addition is expected to
be rate-limiting. (87) An alternate explanation for the slow rates with [Ir(P-
Pht is that the rate of CO loss from [Ir(P-PhCOt is rate-limiting. Further
experiments are needed in order to determine the slow step.
The rates of catalytic de carbonyl at ion of benzaldehyde using
mono-diphosphine complexes of Rh and Ir provide an interesting com-
parison. First of all, the mono-diphosphine complexes of Rh and Ir are
not robust under the conditions of the catalytic reaction and therefore are
of little practical use. However, they do provide useful data for mechanistic
arguments. With Rh, the bis-diphosphine catalysts [Rh(P-Pht are always
more active than their mono-diphosphine analog [Rh(p-p)r when neat
aldehyde is used as solvent. (66) Although Rh-P bond rupture is not
necessary with the coordinatively unsaturated mono-diphosphine com-
plexes, the rhodium may not be "electron-rich" enough to promote facile
oxidative addition. In support of this argument, the presence of the diolefin
cod in the coordination core, [Rh(cod)(dppp)t, increased the activity of
decarbonylation by a factor of 6 compared with [Rh(dppp)t.(66) With Ir
DECARBONYLA TION REACTIONS 371

there is no problem with oxidative addition to [Ir(dppp)t, since Ir com-


plexes are well known to undergo oxidative addition much more readily
than their Rh analogs. Therefore, we see that the decarbonylation activity
of [Ir(dppp)t is ca. 30 times greater than that of [Ir(dpppht, and that
the presence of cod slows down the activity by a factor of 2. These
comparisons show the delicate balancing of effects that is required in order
to tune a catalyst system. They also provide supporting evidence for the
importance of M-P bond rupture as shown in Figure 11.3.

5. ADDITIONAL STUDIES WITH BIS-CHELA TE COMPLEXES


OF Rh(l)

A good example of the difficulties that can be encountered in attempts


to improve the performance of a catalyst by ligand changes is the use of
chelating P-N ligands. Since the rate of Rh-P bond breaking appears to
be an important factor in determining the activity of catalytic decarbonyla-
tion (Figure 11.3), a complex that contains a potentially more labile ligand
was synthesized-[Rh(P-NhJBF4' where

has been characterized by single crystal X-ray diffraction and has a distorted
square-planar cis geometry. (88) This ligand is expected to undergo facile
Rh-N bond rupture by analogy to other P-N complexes. (84) However, this
complex is not effective as a decarbonylation catalyst and exhibits an activity
of only 10 turnovers/hour at 178°C for benzene production from
benzaldehyde. (88) The reason for the inactivity of this complex is unknown;
however, since Rh(I1I) species are not observed during the reaction and
[Rh(P-NhcOt readily loses CO, the inactivity must result from slow
oxidative addition or initial bond dissociation. Examination of crystal
structure of complexes with diphosphine ligands reveals that interligand
steric interactions play a major role in determining geometric
distortions. (74,89) For example, in [Rh(dppbhJBF4 the distortion toward
tetrahedral geometry and the unusual lengthening of one Rh-P bond is
due to phenyl ... phenyl interaction. (74) This complex readily forms dppb-
bridged bimetallic species in solution. (85) It is likely that the desired bond
lability in bis-chelate complexes of Rh is affected more by steric forces
than by electronic factors. More experiments on various P-N com-
plexes are needed in order to determine the relative importance of these
factors.
372 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

6. DECARBONYLA TION OF BENZOYLCHLORIDE WITH


(Rh(dppp)2r

Benzoylchloride cannot be decarbonylated catalytically or


stoichiometrically by [Rh(dpppht or [Rh(dppp)t under the same condi-
tions used for aldehyde decarbonylation. In fact, the addition of ben-
zoylchloride to active aldehyde decarbonylation systems completely and
irreversibly stops the aldehyde decarbonylation. (64) It has been found that
PhCOCI quickly reacts with [Rh(dppp)]CI at 2SoC or with [Rh(dppph]CI
at IS0°C to produce the oxidative addition product Rh(Clh(PhCO)-
(dppp). (90) This Rh(III) complex has a square pyramidal geometry, with
the benzoyl ligand occupying the axial position. Surprisingly, the benzoyl
phenyl group in this complex will not migrate even at 180°C and, therefore,
chlorobenzene cannot be produced. A similar observation has been made
by Baird. (45) It is interesting that the bis-triphenylphosphine analog
Rh(CI)z(PhCO)(PPh 3 h, which has trans phosphine ligands, exhibits facile
Ph group migration but not reductive elimination. (37) Obviously, there are
subtle effects at work here. In light of recent results by Kampmeier (vide
supra), (37) it would be interesting to examine the decarbonylation of alkyl
acid chlorides using diphosphine catalysts. An important observation with
respect to the decarbonylation of benzoylchloride using [Rh(dpppht is
that both chlorobenzene and benzene are catalytically produced using H2
as the purging gas. (91) Catalytic activities at I9SoC for chlorobenzene and
benzene production are 38 and 27 turnovers per hour, respectively, using
neat benzoylchloride as solvent and a catalyst concentration of ca. 1 mM.
The mechanism of this reaction has not been elucidated; however, radical
intermediates are suspected. (91)

ACKNOWLEDGEMENTS

The research on catalytic decarbonylation using diphosphine com-


plexes was supported by the National Science Foundation. The Johnson
Matthey Company is acknowledged for generous loans of rhodium and
iridium trichloride.

REFERENCES

1. J. Tsuji and K. Ohno, Synthesis 1, 157 (1969).


2. C. W. Bird, Transition Metal Intermediates in Organic Synthesis (Logos Press, London,
1967), pp. 112,239.
DECARBONYLA TION REACTIONS 373

3. A. Kozikowski and H. Wetter, Synthesis 1976,561.


4. W. M. Schubert and R. R. Kinther, The Chemistry of the Carbonyl Group (Wiley-
Interscience of John Wiley and Sons, New York, 1966), p. 695.
5. J. N. Pitts and J. K. S. Wan, The Chemistry of the Carbonyl Group (Wiley-Interscience
of John Wiley and Sons, New York, 1966), p. 823.
6. J. Chatt and B. L. Shaw, Chem. and Ind. 1960, 931.
7. - - , Chem. and Ind. 1961,290.
8. L. Vaska and J. W. DiLuzio, 1. Am. Chem. Soc. 83, 1262 (1961).
9. L. Vaska, Chem. and Ind. 1961,1402.
10. J. Chatt, B. L. Shaw, and A. E. Field, 1. Chem. Soc. 1964, 3466.
11. L. Vaska and J. W. DiLuzio, 1. Am. Chem. Soc. 83,2784 (1961).
12. J. P. Collman, C. I. Sears, and M. Kubota, Inorg. Synthesis 11, 101 (1968).
13. L. Vaska, 1. Am. Chem. Soc. 86, 1943 (1964).
14. R. H. Prince and K. A. Raspin, 1. C. S., Chem. Comm. 1966, 156.
15. - - , 1 . Chem. Soc. (A) 1969, 612.
16. - - , 1 . Inorg. Nuc!. Chem. 31,695 (1969).
17. K. A. Raspin, 1. Chem. Soc. (A) 1969, 461.
18. N. W. Alcock and K. A. Raspin, 1. Chem. Soc. (A) 1968, 2108.
19. J. Blum, S. Kraus, and Y. Pickholtz, 1. Organomet. Chem. 33, 18 (1971).
20. (a) M. C. Baird, C. J. Nyman, and G. W. Wilkinson, 1. Chem. Soc. (A) 1969, 346; (b)
M. C. Baird, J. T. Mague, J. A. Osborn, and G. Wilkinson, 1. Chem. Soc. (A) 1967, 1347.
21. J. Blum, E. Oppenheimer, and E. D. Bergman, 1. Am. Chem. Soc. 89, 2338 (1967).
22. K. Ohno and J. Tsuji, 1. Am. Chem. Soc. 90,99 (1968).
23. J. K. Stille, M. T. Regan, R. W. Fries, F. Huang, and T. McCarley, Adv. Chem. Ser.
132, 181 (1974).
24. (a) J. K. Stille and M. T. Regan, 1. Am. Chem. Soc. 96, 1508 (1974); (b) J. K. Stille and
R. W. Fries, 1. Am. Chem. Soc. 96, 1514 (1974); (c) J. K. Stille, F. Huang, and M. T.
Regan,l. Am. Chem. Soc. 96, 1518 (1974).
25. A. Rusina and A. A. Vlcek, Nature 206, 295 (1965).
26. E. Muller, A. Segnitz, and E. Langer, Tetrahedron Lett. 1969,1169.
27. Y. S. Varshavskii, E. P. Shestakova, N. A. Buzina, T. G. Cherkasova, N. V. Kiseleva,
and V. A. Kormer, Koord. Khim. 2,1410 (1976); through Chem. Abstr. 86,37002 (1977).
28. Y. S. Varshavskii, E. P. Shestakova, N. V. Kiseleva, T. G. Cherkasova, N. A. Buzina,
L. S. Bresler, and V. A. Kormer, 1. Orgnomet. Chem. 170, 81 (1979).
29. J. J. Alexander and A. Wojcicki, Inorg. Chem. 12,74 (1973).
30. - - , 1 . Organomet. Chem. IS, C23 (1968).
31. F. H. Jardine, Progress in Inorganic Chemistry, edited by S. J. Lippard, Vol. 28, pp.
63-202 (Wiley-Interscience of John Wiley and Sons, New York 1981).
32. G. Gomazetis, B. Tarpey, D. Dolphin, B. R. James, 1. C. S., Chem. Commun. 1980,939.
33. K. S. Y. Lau, Y. Becker, F. Huang, N. Baenziger, and J. K. Stille, 1. Am. Chem. Soc.
99,5664 (1977).
34. D. L. Egglestone, M. C. Baird, C. J. L. Lock, and G. Turner, 1. C. S., Dalton 1977, 1576.
35. J. A. Osborn, F. H. Jardine, and G. Wilkinson, 1. Chem. Soc. (A) 1966, 1711.
36. M. C. Baird, C. J. Nyman, and G. Wilkinson, 1. Chem. Soc. (A) 1968,348.
37. J. A. Kampmeier, R. M. Rodehorst, and J. B. Philip, Jr., 1. Am. Chem. Soc. 103, 1847
(1981).
38. J. B. Philip, Jr., Ph.D. Thesis, University of Rochester, 1980.
39. J. A. Kampmeier, S. H. Harris, and D. K. Wedegaertner, 1. Org. Chem. 45,315 (1980).
40. J. K. Stille, M. T. Regan, R. W. Fries, F. Huang, and T. McCarley, Adv. Chem. Series
132, 181 (1974).
41. J. A. Kampmeier, S. H. Harris, and R. M. Rodehorst,l. Am. Chem. Soc. 103,1478 (1981).
374 DANIEL H. DOUGHTY AND LOUIS H. PIGNOLET

42. C. A. Tolman, P. Z. Meakin, D. L. Lindner, and J. P. Jesson, J. Am. Chem. Soc. 96,
2762 (1974).
43. J. Halpern and C. S. Wong, J. C. S., Chem. Commun. 1973, 629.
44. C. S. Wong, Ph.D. Thesis, University of Chicago, 1973.
45. D. A. Slack, D. L. Egglestone, and M. C. Baird, J. Organomet. Chem. 146, 71 (1978).
46. N. A. Dunham and M. C. Baird, J. C. S., Dalton 1975,774.
47. J. K. Stille, Ann. New York Academy of Science 295, 52 (1977).
48. J. Blum, Tetrahedron Lett. 1966,1605.
49. J. K. Stille, private communication.
50. P. J. Davidson, M. F. Lappert, and R. Pearce, Chem. Rev. 76, 219 (1976).
51. M. C. Baird, J. Organomet. Chem. 64,289 (1974).
52. J. W. Suggs, J. Am. Chem. Soc. 100,640 (1978).
53. H. M. Walborsky and L. E. Allen, J. Am. Chem. Soc. 93, 5465 (1971).
54. - - , Tetrahedron Lett. 1970, 823.
55. P. B. Chock and J. Halpern, J. Am. Chem. Soc. 88, 3511 (1966).
56. J. Y. Chen, Ph.D. Thesis, The University of Chicago, 1972.
57. R. Ugo, A. Pasini, A. Fusi, and S. Cenini, J. Am. Chem. Soc. 94, 7364 (1972).
58. W. Strohmeier and P. Prohler, J. Organomet. Chem. 108,393 (1976).
59. J. Tsuji and K. Ohno, Tetrahedron Lett. 1966,4713.
60. G. L. Geoffroy, D. A. Denton, M. E. Keeney, and R. R. Bucks, Inorg. Chem. 15,2382
(1976).
61. J. Chat! and B. L. Shaw, J. Chem. Soc. (A) 1966,1437.
62. D. H. Doughty and L. H. Pignolet, J. Am. Chem. Soc. 100,7083 (1978).
63. D. H. Doughty, M. F. McGuiggan, H. Wang, and L. H. Pignolet, Fundamental Research
in Homogeneous Catalysis 3, 909 (1979).
64. D. H. Doughty, M. P. Anderson, A. L. Casalnuovo, M. F. McGuiggan, C. C. Tso, H.
H. Wang, and L. H. Pignolet, Adv. Chem. Ser. 196, 65 (1982).
65. R. R. Schrock and J. A. Osborn, J. Am. Chem. Soc. 93,2397 (1971).
66. D. H. Doughty, Ph.D. Thesis, University of Minnesota, 1979.
67. A. S. C. Chan, J. J. Pluth, and J. Halpern, J. Am. Chem. Soc. 102, 5952 (1980).
68. M. D. Fryzuk and B. Bosnich, J. Am. Chem. Soc. 99, 6262 (1977).
69. W. S. Knowles, M. J. Sabacky, and B. D. Vineyard, Adv. Chem. Ser. 132, 274 (1974).
70. B. R. James and D. Mahajan, Can. J. Chem. 57, 180 (1979).
71. J. Halpern, D. P. Riley, A. C. S. Chan, and J. J. Pluth, J. Am. Chem. Soc. 99, 8055 (1977).
72. H. H. Wang, Ph.D. Thesis, University of Minnesota, 1981.
73. J. Y. Chen, Ph.D. Thesis, University of Chicago, 1972.
74. M. P. Anderson and L. H. Pignolet, Inorg. Chem. 20, 4101 (1981).
75. P. S. Chua, N. K. Roberts, B. Bosnich, S. J. Okrasinski, and J. Halper:1, J. Chem. Soc.,
Chem. Comm. 1981, 1278.
76. S. D. Ittle, C. A. Tolman, A. D. English, and J. P. Jesson, Adv. Chem. Ser. 173,67 (1979).
77. C. A. Tolman, S. D. lttle, A. D. English, and J. P. Jesson, J. Am. Chem. Soc. 100, 4080
(1978).
78. A. R. Sanger, J. C. S., Dalton 1977, 120.
79. C. L. U. Su, Adv. Organomet. Chem. 17,269 (1979).
80. T. B. Rauchfuss, J. Am. Chem. Soc. 101, 1045 (1979).
81. J. P. Collman and L. S. Hegedus, Principles and Applications of Organotransition Metal
Chemistry (University Science Books, Mill Valley, CA, 1980).
82. Y. Kawabata, T. Hayashi, and I. Ogata, J. C. S., Chem. Commun. 1975,462.
83. J.-c. Poulin, T.-P. Dang, and H. B. Kagan, J. Organomet. Chem. 84,87 (1975).
84. W. J. Knebel and R. J. Angelici, Inorg. Chem. 13, 627, 632 (1974).
85. L. H. Pignolet, D. H. Doughty, S. C. Nowicki, M. P. Anderson, and A. L. Casalnuovo,
J. Organomet. Chem. 202, 211 (1980).
DECARBONYLA TION REACTIONS 375

86. J. P. Collman and L. S. Hegedus, Principles and Applications of Organotransition Metal


Chemistry (University Science Books, Mill Valley, CA, 1980), p. 211.
87. - - - , Principles and Applications of Organotransition Metal Chemistry (University
Science Books, Mill Valley, CA, 1980), p. 205.
88. M. P. Anderson, B. J. Johnson, and L. H. Pignolet, preliminary results.
89. L. H. Pignolet, D. H. Doughty, S. C. Nowicki, and A. L. Casalnuovo, Inorg. Chern. 19,
2172 (1980).
90. M. F. McGuiggan, D. H. Doughty, and L. H. Pignolet, 1. Organomet. Chern. 185,241
(1980).
91. M. F. McGuiggan, Ph.D. Thesis, University of Minenesota, 1982.
12
Homogeneous Catalysis of
Oxidation Reactions Using
Phosphine Complexes
D. Max Roundhill

1. SIGNIFICANCE OF METAL CA TAL YZED OXIDA TION


REACTIONS

Oxidation is one of the most important reactions in chemistry and bio-


chemistry. Combustion of hydrocarbons drives much of our economy and
transportation, and biological oxidation processes are fundamental to life
and ecology. From an industrial viewpoint, oxidation reactions occupy a
pivotal role in the conversion of hydrocarbons into required products, and
indeed it has been estimated that over 50% of such processes involve
hydrocarbon oxidations. (1)
In choosing a particular oxidant one must be particularly cognizant of
the required selectivity for the desired transformation. Among the available
reagents for oxidations are metal compounds in high-oxidation states,
molecular oxygen, and peroxides or peracids.(2-1S) Both oxygen and the
peroxidic reagents can be used in conjunction with metal complexes,
whereby the oxygen containing reagent is "activated" to reactivity, or the
selectivity of the reagent is modified. Several review articles have discussed
the structural and chemical aspects of dioxygen binding to transition metal
compounds.(16-21) Furthermore, alkyl peroxides in conjunction with high-
valent metal compounds have been used to advantage for selective alkene
epoxl'd'
atlOn. (22)

Dr. D. Max Roundhill • Department of Chemistry, Tulane University, New Orleans,


Louisiana 70118.

377
378 D. MAX ROUNDHILL

1. 1. Mechanistic Features of Metal Catalyzed Oxidations


The primary focus of research using transition metal phosphine com-
plexes for oxidations is in the complexation and activation of molecular
oxygen. These oxygen complexes have been variously regarded as com-
plexes of coordinated peroxide, superoxide, or singlet oxygen, and their
reactivity with reduced substrate has been interpreted on such a basis. In
this chapter, we will focus on the chemical reactivity of these compounds
for oxygen atom transfer oxidation reactions, with a particular emphasis
on the mechanistic features of these processes.
Atom-transfer-type oxidation reactions can occur by electrophilic,
nucleophilic, or radical pathways. Electrophilic mechanisms are prevalent
with metal compounds in high-oxidation states. Common among these
reagents are oxo complexes of Mn(VII), Cr(VI), V(V), as well as organic
reagents such as peracids.(8-10.15) Recently, a proposal has been forwarded
with respect to this chemistry that metallocycle formation between M=O
and an alkene can occur, provided that a second spectator oxo group is
available that can simultaneously convert from a metal oxygen double to
a triple bond. (23) Reagents such as hydrogen peroxide and organic peroxides
have been previously considered to follow nonselective, free-radical path-
ways. Recent work by Sharpless, however, using tert- butyl hydroperoxide
as oxidant in conjunction with high-valent metal compounds has shown
that stereospecific alkene epoxidation can be effected. (24)
Catalyzed oxidations using molecular oxygen and transition metal
phosphine complexes appear to follow two pathways. Coordination of
dioxygen to the metal center results in a formal transfer of one or two
electrons from the metal to give a complexed superoxo or peroxo ligand,
which then undergoes reaction with the reduced substrate. A second path-
way that is frequently followed involves initial oxygen attack at the hydro-
carbon substrate to form an organic peroxide, which is then catalytically
converted into oxygenated. product by reaction with the phosphine metal
complex. Both of these courses are prevalent in catalytic oxidations with
transition metal phosphine complexes, and diagnostic experiments are
necessary to deduce the function of any metal phosphine complex used as
an oxidation catalyst.

2. TRANSITION-METAL PHOSPHINE OXYGEN COMPLEXES

2.1. Synthesis and Structure

Synthetic complexes with a coordinated dioxygen have been known


for many years, mainly because of their relevance to reversible oxygen
HOMOGENEOUS CATAL YSIS OF OXIDA TlON REACTIONS 379

carriersY6,17) Much of the earlier work centered around formally divalent


cobalt complexes having nitrogen or oxygen donor ligands. In 1963, Vaska
reported that the d 8 iridium(I) compound IrCI(CO)(PPh 3h would revers-
ibly add molecular oxygen, (25) and this result instigated considerable interest
in the chemistry of low-valent, electron rich, metal ions with oxygen. Since
that time many similar compounds have been obtained by addition of
oxygen to complexes that can undergo a formal two-electron oxidative
addition reaction. Examples of these compounds are ones found with Ir(I),
Rh(I), Ni(O), Pd(O), Pt(O), Ru(O). Since substituted phosphines are com-
monly used to stabilize these precursor complexes in their low-valent
oxidation state, the complexes formed by oxygen addition are pre-
dominantly ones with such phosphine ligands. Examples of these com-
pounds that are formed by oxygen addition are IrCI(CO)(PPh3h02,(25)
Pt(PPh3h02,(26,27) Ni(CNPhh02,(28,29) [Ir(dppeh02]X,(30) and
RuX(NO)(PPh 3h 0 2.(31)

(1)

(2)

Ni(CNPh)4 + O 2 -. Ni(CNPhh02 + 2CNPh (3)

[Ir(dppeh]X + O 2 -. [Ir(dppeh02]X (4)


(5)

The reversibility of the oxygenation reaction is dependent on the nature


of the ligands. Thus, whereas the compound IrCI(CO)(PPh 3h will revers-
ibly add oxygen, the analogs with more basic phosphines, such as PMePh 2
or PMe2Ph, bind oxygen more strongly, and the O 2 is not readily dissociated.
Structurally, it has been found that each of these compounds has a TJ 2_
bonded oxygen ligand(I), and, from electronic considerations, the coordin-
ated dioxygen resembles a complexed peroxide ligand. (16) This binding
mode differs significantly from that of type II found for metals such as
Fe(II) and Co(II), which undergo one-electron oxidation.

o o
/
M~I M-O
o
II

Structural and infrared (p(O-O)) studies have been published showing


that the complexed 0-0 bond lengths are in the range 1.41-1.52 A, which
compares closely with the sum of 1.46 A for the 0-0 single-bond covalent
380 D. MAX ROUNDHILL

radii. The infrared stretch v(O-O) for these compounds I is in the 800-
900 cm- 1 range.
New palladium dioxygen complexes have been prepared from reaction
of superoxide ion and [PdCI(methoxydicyclopentadienyl-1] 1)Jz. (32) This
dioxygen-bridged complex [(MeO-DiCp)PdO)]2 reacts with methanol to
yield the methoxy-bridged compound [(MeO-DiCp)PdOMeJz. Sub-
sequently, the oxygen-bridged compound Rh 202(1,5-CODh has also been
prepared. On treatment with cyclohexanone this compound gives dehy-
drogenation products cyclohexene-3-one and phenol.(33) On pyrolysis of
Rh 202(1,5-CODh, cyclooctanone is formed quantitatively in the presence
of cyclohexene. (34)
Photolysis of Pt0 2(PPh 3h leads to the formation of Pt(PPh 3h and O 2.
Irradiation in the presence of the singlet oxygen traps 2,2,6,6-tetraethyl-
piperidine, and 1,3-diphenylisobenzofurane confirms that the oxygen is
dissociated in a singlet state. Carrying out similar experiments with
[lr(dppeh02t shows that oxygen dissociates in the triplet state. The
difference is related to the observation that the lowest energy transition in
Pt0 2(PPh3h is an O 2 --+ Pt charge transfer band, whereas in [lr(dppeh02t
the lowest energy band is assigned as Ir --+ dppe charge transfer. (35)

2.2. Reactions with Electrophiles


The reactivity of these 1] 2-oxygen compounds can be rationalized on
the basis of the coordinated dioxygen having nucleophilic character. Thus,
the compound Pt(PPh 3h02 reacts with electrophiles such as S02, (36)
CO 2,(37) NO,(36) R 2 CO,<37,38) and RCHO(37) to give cyclic or acyclic
adducts, (6-10) where the electrophilic center of the adduct molecules are
bonded to the oxygen atom from the coordinated dioxygen. The reaction

50 2
(6)

CO 2
(7)

(PPh 3 hPt(ONO)z (8)


0-0 H
RCHO / \ /
(PPh3)2Pt" C, (9)
0/ 'R

R 2 CO
(10)
HOMOGENEOUS CATAL YSIS OF OXIDA TlON REACTIONS 381

with S02 has been studied using oxygen-18 labeling techniques, where it
is concluded that the reaction pathway involves an intermediate peroxo
. subsequently transf orms to the sulfate compI ex (3940)
compound III, WhICh . :
*
0* /0"
Ir/" I + S02 - - Ir 0* (11)
'0* "s/
1'~

°°III

Kinetic studies have been made on the addition of ketones to both


PtL 2 0 2 and IrX(CO)L 202 (L = substituted phosphine, X = halide).(41,42)
The reaction with PtL 2 0 2 and ketones (R 2 CO) follows a two-term rate
law (12).
(12)

These data have been interpreted in support of a two pathway mechanism,


one being independent, and the other first-order, in ketone concentration.
The proposed intermediates are shown in Figure 12.1. The faster pathway
is binuclear with the ketone binding to platinum to give a pentacoordinate
intermediate. The slower pathway, which is ketone independent, is pro-
posed to have a slow step to give an "activated" form of PtL 2 0 2 , which
may structurally resemble the 1/ 1 -form II rather than the 1/ 2- form 1. In
agreement with the concept that the coordinated dioxygen has nucleophilic
character, the reaction rate increases for ketones having electron withdraw-
ing groups.
The reactivity of IrX(CO)L 2 0 2 to ketones is less than that of PtL 2 0 2 •
A similar metallocycle is formed (Equation 13) but only with the
electrophilic ketone hexafluoroacetone. Now the oxygen complex

Figure 12.1. Proposed pathways for the reaction of PtL 2 0 2 with ketones.
382 D. MAX ROUNDHILL

IrX(CO)L 2 0 2 is coordinately saturated, and intermediate adduct formation


with ketone is unlikely. Since the reaction does not involve prior ligand
dissociation, the results provide a kinetic measurement of the direct reac-
tivity of a coordinated oxygen molecule to an external electrophile. (42)

(13)

The reaction between Pt(PCY3h02 and hexafluorobutyne-2 or


dimethyl acetylenedicarboxylate forms an adduct (14) with formal cleavage
of the 0-0 bond:
o R
(PCY3}zPt0 2 + RC == CR ..... (PCY3)2Pt/ If (14)
'o~R
The mechanism proceeds by two steps. The first step is first-order in both
platinum oxygen complex and alkyne, and the rate of the second step is
independent of alkyne concentration (R = CF3, C0 2Me). The intermediate
may be an adduct with the alkyne coordinated to platinum, or the reaction
may proceed in the unsymmetrical metallocycle as is suggested for the S02
insertion. (43)
The most definitive study for establishing the nucleophilic character
of the complexed dioxygen in form I comes from the reaction between
PtL 2 0 2 (L = PPh3, PCY3) and alkyl or aroyl halides. (44) The products from
reacting triphenylmethyl bromide and benzoyl bromide are the correspond-
ing alkyl-peroxo platinum(II) complexes L 2 PtBr(00R) (L = PPh 3; R =
CPh 3, COPh)(15).
OOR
/
L 2 Pt0 2 + RBf ..... L 2 Pt (15)
"- Bf

Both an SN 2 type (IV) and a cyclic (V) transition state are considered
plausible. Discrimination between the mechanistic types was achieved with
R-Bf

v
(S)-( - )-a -phenethyl bromide. Treatment of this alkyl halide with
(PPh3hPt02, followed by 0-0 cleavage of the peroxo complex with
LiAIH4' gives S-phenethyl alcohol having the R configuration. This result
is in agreement with an SN 2-type mechanism involving the coordinated
HOMOGENEOUS CATAL YSfS OF OXfDA TlON REACTIONS 383

dioxygen as nucleophile. Again the concept of conversion of form I to II


being involved in the reaction pathway is considered. These authors suggest
that this conversion represents the extreme of an asymmetric vibration
which can lead to a transient species II, which will undergo preferential
electrophilic attack at the distal oxygen atom. Such a species has not been
isolated for a dB or d 10 metal center, but is suggested on the basis of a
.
vanety 0 f k'mehc
. measurements. (41424445)
. . . Th e concept 0 f a nuc Ieop h'l'
I IC
peroxidic dioxygen ligand when bonded to a formal dB or d 10 metal center
appears to be well founded. Furthermore, these reactions are supported
by calculations on Pt(PH3h02, which are in agreement with the concept
of a Pt2+(PH3h022- model.(46)
The formation of metallocyclic complexes by addition of ketones to
dioxygen complexes resembles a three-step Criegee mechanism for
ozonolysis.(47,4B) In each case one envisages the formation of a dipolar 0-0
bond that leads to electrocyclic ring closure with the ketone. No proven
oxidation of aldehydes or ketones has yet been achieved with these metal-
locycles. Nevertheless, in the catalytic conversion of cyclic ketones to
lactones using hydrogen peroxide and a molybdenum complex as catalyst,
it is believed that such an intermediate metallocycle is first formed and
that it subsequently converts to lactone and molybdenyl products (16).(49)
o
MO/? +A
'0 V (16)

3. OXIDA T/oN OF ALKENES

From the initial discovery of these T/ 2-dioxygen complexes, attempts


have been made to use them for alkene oxidation. The introduction of
oxygen functionality into an alkene hydrocarbon is a particularly significant
goal, especially if the source is molecular oxygen and the reaction can be
achieved catalytically with high selectivity. Since dioxygen will readily
coordinate to dB and d 10 metal complexes, it has been speculated that one
can modify its chemical reactivity to undergo selective reaction with alkenes.
In the presence of catalytic amounts of RhCI(CO)(PPh 3 h. Pt(PPh3 h02,
IrI(CO)(PPh 3 h, IrCI(N 2)(PPh 3 h, or RhCl(PPh 3 h, cyclohexene and cyclo-
pentene undergo autoxidation:

(17)
384 D. MAX ROUNDHILL

The major product from the former is cyclohexen-3-one, along with minor
amounts of cyclohexene oxide (17). Epoxide formation has also been
identified as a minor product from cyclopentene autoxidation. The inter-
mediacy of 3-cyclohexene hydroperoxide was proposed in this report but
not verified. (50) Subsequent work on the autoxidation of cyclohexene using
RhCI(PPh 3 h verified this premise,(51) and a number of review articles have
emphasized this conclusion. (52-55) The involvement of preformed hydroper-
oxide has been verified by comparing the rate of cyclohexene oxidation
both with hydroperoxide present, and also when the cyclohexene is purified
free from peroxide. In the former case the reaction is rapid and there is
no induction period. Under conditions where the cyclohexene is peroxide
free the reaction proceeds more slowly, and there is an induction period
of close to three hours (using IrCI(CO)(PPh 3 h as catalyst) as the hydroper-
oxide intermediate is being performed (18):

(18)

The mechanistic details of the subsequent chemistry are unclear, but it is


apparent that the catalytic decomposition of hydroperoxide follows a chain
process that resembles a Haber-Weiss pathway.(56) It has also been found
that dB complexes show a greater activity for cyclohexene autoxidation
that do d 10 complexes, but no rationale has been provided to explain this
observation.(51)
This pathway involving catalytic decomposition of preformed per-
oxides from alkenes and oxygen, whereby the transition metal phosphine
is the catalyst, is a frequently encountered situation. Any oxidation study
using metal complexes as catalyst should involve a careful check for induc-
tion periods, low selectivity, inhibition by added free-radical chain traps,
partial order kinetics, and observed zeroth order in the rate law for added
oxygen. Examples of such processes are found in the transition metal
phosphine complex catalyzed autoxidation of cyclohexene, (51.57-65)
styrene, (66-68) tetramethylethylene, (69) and butadiene. (70) Among the
various complexes that have been used are Pt(PPh 3)4, RhCI(PPh 3h,
RuCh(PPh 3h, IrCI(CO)(PPh 3h, RhCl(CO)(PPh 3 h, Ru(COh(PPh 3h, and
OsBr2(PPh3), or their oxygen adducts. The radical chain mechanism has
been generally accepted for these reactions, except for a case where the
reactions are carried out at the lower temperature of 30°. In this reaction
the final product is 2-cyclohexene-1-yl hydroperoxide (18). (71) The catalyst
used is RhCl(PPh 3 h, and it is considered that a concerted coordinated
sphere mechanism can be applied to fit the data equally well.
HOMOGENEOUS CATAL YSIS OF OXIDATION REACTIONS 385

As a final generalization., it should be mentioned that these radical


chain processes are common when transition metal compounds other than
phosphine complexes are used. Catalytic oxidations have been carried out
with a wide range of metal complexes, and a comparison has been observed
between reactions catalyzed by phosphine complexes and those with acetyl-
. (5557)
acetonate or other ligands. '

4. OXIDA TlON OF OTHER SUBSTRA TES

4.1. Isocvanides

The dioxygen complexes M0 2 (CNBu-t)z (M = Ni, Pd) have been


isolated in the pure state. (28,29,72) Treating the nickel complexes with excess
tert- butyl isocyanide at 30°C results in the formation of Ni(CNBu-t)4 and
t-BuNC0(19):

NiOz(CNBu-t)z + 4t-BuNC -4 Ni(CNBu-t)4 + 2t-BuNCO (19)

When, however, triphenylphosphine is added to the nickel complex, this


compound is preferentially oxidized to triphenylphosphine oxide (20):

NiOz(CNBu-tlz + 4PPh 3 -4 Ni(CNBu-t)z(PPh 3 lz + 20PPh z (20)

The reaction of methyl or cyclohexyl isocyanide with Ni0 2 (CNBu-t)z


produces a mixture of tert-butyl and either methyl or cyclohexyl isocyanate,
and it is considered important that the formed isocyanate complex be
kinetically labile for achieving successful isocyanide oxidation. (29)

4.2. Carbon Monoxide

Although the primary method of choice for the metal-catalyzed oxida-


tion of carbon monoxide to carbon dioxide involves nucleophilic attack of
water or hydroxide on a coordinated carbonyl,(73-761 there are an increasing
number of cases where a transition metal phosphine complex will catalyze
this conversion.
Following the initial reports that the compound Rh 6(COh6 will catalyze
the oxidation, with molecular oxygen, of carbon monoxide to carbon
dioxide,m-79) it has been found that this process is more favorable when
phosphines are added.(80-81l The addition of triphenylphosphine and CO
to a mixture of Rh 6(COh6 and benzene in the presence of oxygen leads
to the formation of both CO 2 and OPPh 3 . The stoichiometric yield of CO 2
386 D. MAX ROUNDHILL

is greater than that of OPPh3, and the addition of water increases the
quantity of CO2 formed. Labeling studies have been carried out to probe
the mechanism, and further details are discussed in Section 5.3.
The addition of triphenylphosphine causes the Rh6 cluster to convert
to a lower homologue. Among the solution species identified are the
compounds Rh4(COho(PPh3h and Rh 2(COMPPh3k These species pre-
sumably react with oxygen to give adducts in a manner similar to that
found with Rh 2(COMPPh3)4 (21) when prepared from Rh 4(COh2 and
PPh 3,(82) and also with the compound Rh 4Cl4(COMPPh 3h(02h (22)(83):

[RhHCI(PPh 3 h]2 + 2RhCI(CO)(PPh 3 h + 2H 2 0 + 2C0 2 (22)

A zerovalent platinum bimetallic complex Pt2(dppmh will also catalyze


the oxidation of CO to CO2 using either O 2 or NO as oxidant.(84) Again,
no comment can be made regarding the reaction mechanism. Oxidation of
CO on platinum metal surfaces involves the reaction of chemisorbed atomic
oxygen, (85) but it is unlikely that such species are formed under the mild
conditions of temperature used in these examples. In the conversion
of the coordinated carbonyl in IrCI(CO)PN (PN =
o-(diphenylphosphino)-N,N-dimethylaniline) to CO 2 though it has been
tentatively suggested that a metallocyclic intermediate is formed (23)(86):

IrCI(CO)PN + O 2 ~ (23)

Such an intermediate can be formed because of the nucleophilic reactivity


of a dioxygen molecule when complexed to iridium(I) and because of the
known propensity of a coordinated carbonyl ligand to undergo such
nucleophilic attack.

4.3. Aldehydes and Ketones

In benzene solvent a series of d 8 and d 10 phosphine complexes cata-


lyze the oxidation of benzaldehyde to benzoic acid and perbenzoic
acid under an oxygen atmosphere. The catalytic activity follows the
order RhCI(CO)(PPh 3h > Pd0 2(PPh 3h = Pd(PPh 3)4 > RhCI(PPh 3h >
[RhCl(PPh 3hJ2> IrCI(CO)(PPh3h02 = Pt02(PPh3}z.(87) These authors
propose that a nonradical pathway is followed, but subsequent work showed
HOMOGENEOUS CATAL YSIS OF OXIDATION REACTIONS 387

the oxidation to occur by a free-radical mechanism. (88) This correction was


based on a number of new experiments. The kinetics of the oxidation using
Pd0 2(PPh 3h as catalyst resembles that found using cobalt(III) catalysts,
and also the reaction is retarded by adding the free-radical inhibitor
2,6-di-tert- butyl-p- cresol. Furthermore, the presence of free radicals is
verified using the spin traps nitroso-tert-butane and phenyl-tert- butyl
nitrone.
Subsequent to the finding that the compounds Rh 6(COh6 and
Re2(COho will catalyze the autoxidation of ketones,(77,89,90) it was also
found that the phosphine complexes IrCI(CO)(PPh 3h and Pt(PPh 3h were
effective catalysts for this oxidation. (91,92) The organic oxidation products
are carboxylic acids (Equations 24-27), but the phosphine compounds are
not recovered unchanged. Inhibition by free-radical scavengers is observed,
and a reaction pathway is followed whereby the phosphine complex acts
to accelerate the conversion of preformed hydroperoxides and peracids to
carboxylic acid. (90)

(24)

(CH 3 CH zhCO ~ CH 3 CH 2 C0 2 H + CH 3 C0 2 H (25)

6 (26)

(27)

4.4. Cumene

The compound Pd(PPh3)4 can be used as a liquid-phase autoxidation


catalyst for cumene oxidation. (93,94) It is again concluded that the role of
the transition metal compound is in its reaction with preformed cumene
hydroperoxides. (94)

4.5. Tertiary Phosphines

Tertiary phosphines are among the easiest molecules to catalytically


oxidize with molecular oxygen. The reactions can usually be effected under
ambient temperature and pressure conditions. Metal complexes of both
the platinum metal group and from the first row transition series have been
used as catalysts, although in only a few cases have detailed mechanistic
388 D. MAX ROUNDHILL

ROH+PPh 3

OPPh 3 PPh 3

Figure 12.2. Mechanism for the Pt(PPh 3 h catalyzed oxidation of triphenylphosphine.

studies been carried out. Among the most studied is the compound
Pt(PPh3 h.(95-98) Under protolytic conditions, Halpern suggests that the
compound Pt02(PPh3h forms free hydroperoxide ion, which then converts
the tertiary phosphine to oxide (Figure 12.2). The resulting compound
Pt(PPh 3 )/+, in conjunction with OH-, is then proposed to further oxidize
PPh 3 to yield a second mole of OPPh3, with concomitant reduction of
platinum(II) to the zerovalent oxidation state. The selectivity to oxidation
by Pt0 2(PPh3h is related to the nucleophilicity of the ligand to platinum,
and hence its ability to undergo substitution at platinum and generate free
hydroperoxide ions. It is suggested, but not verified, that the oxidation of
triphenylphosphine with Ru(NCS)(NO)(PPh 3h (31,99) as catalyst may fol-
Iowa similar pathway, Other platinum metal complexes that have been
used in a similar manner for triphenylphosphine oxidation are
RhCl(PPh 3 h (100-102) and IrX (CO)(PPh 3h. (66)
Complexes of cobalt have also been used as catalysts for the oxidation of
tertiary phosphines with molecular oxygen. In methanol solvent the com-
pound C02(CN)4(PMe2Ph)s02 reacts with PMe 2Ph, converting it into the
oxide, (103) Since the oxygen compound is readily formed again from the
product complex Co(CNh(PMe2Phh, a catalytic cycle can be obtained (28,
29) for the phosphine oxidation,
2Co(CNh(PMe2Phh + O 2 ~ C02(CN)4(PMe2Ph)s02 + PMe2Ph (28)
HOMOGENEOUS CATAL YSIS OF OXIDA TlON REACTIONS 389

Studies on the reactions of CoCh(PEt 3h. (104) and also on a mixture


of Co(acach and PBU3/lOS) show the reactions to form only OPEt3 and
OPBU3, respectively. The reactions show no inhibition upon addition of
free radical scavengers. For CoCh(PEt3h, the rate is first-order in both
CoCh(PEt3h and O 2 , but the kinetics of the latter reaction show first-order
dependence on Co(acach(PBu3) and only half-order on oxygen concentra-
tion. This difference between the two studies is shown in the latter inter-
mediate being a bimetallic dioxygen complex, whereas the former is pro-
posed to be a monomer CoCh(PEt3h02. Extension of the work to phos-
phinite and phosphite esters(106) of type EtnP(OEth-n shows conversion
to oxide with no change in the value of n. Autoxidation of uncomplexed
EtnP(OEth-n by a free radical pathway does cause a change in the value
of n. (107) Thus the oxidation of the coordinated phosphorus ligand must be
considerably faster and cannot involve prior dissociation, which would lead
to oxidation of free phosphine and a change in the value of n. For
triphenylphosphine oxidation, the reaction is catalytic using the compound
CoClz(PPh 3h with oxygen. (l08)
Although the oxidation of tertiary phosphines by these catalytic pro-
cesses has minimal useful application, it needs to be considered as a
problematic side reaction in homogeneous catalysis. Much effort is being
currently expended to immobilize platinum metal phosphine complexes on
heterogenized tertiary phosphine supports, and irreversible oxidation at
phosphorus on these supports effectively destroys the supported catalyst.
Recent observations that the compound Rh 6(COh6 catalyzes the oxidation
of tertiary phosphines(80) correlate with the report that phosphine oxidation
occurs with molecular oxygen on Rh 6(COh6 bound to diphenylphosphino-
functionalized poly(styrenedivinylbenzene)y09) Thus, in order to use these
phosphinated polymer-supported rhodium catalysts, one needs either to
rigorously exclude oxygen, or to find a way to inhibit the simultaneous
catalyzed phosphine oxidation.

5. CO-OX/DA T/ONS

A challenge for chemists designing homogeneously catalyzed oxidation


reactions is the frequent necessity to incorporate multiple substrates in the
reaction mixture for selective co-oxidation. If the oxygen atom transfer
reagent is molecular oxygen, it is unlikely that both oxygen atoms will be
transferred to reduced substrate in a single step. Following transfer, there-
fore, of a single oxygen atom from dioxygen to reduced substrate in the
390 D. MAX ROUNDHILL

coordination sphere of the metal, the second oxygen will remain coordin-
ated to the metal. To complete the catalytic cycle, it is therefore necessary
to incorporate a second reaction step whereby this bonded oxo ligand is
transferred to another substrate or a second mole of the same reduced
substrate.

5.1. Alkenes and Tertiary Phosphines

Read and co-workers found that the complex RhCI(PPh 3h will selec-
tively catalyze the co-oxidation of both terminal alkenes to methyl ketones,
and triphenylphosphine to triphenylphosphine oxide(30):
RhCI(PPh 3 h
RCH=CH 2 + PPh 3 + O 2 ~ R-C-CH3 + OPPh3 (30)
II
o
The oxidation of alkene is stoichiometric unless excess triphenylphosphine
. present. (110-112) I n t he reactIon
IS . mech ·
amsm, one oxygen atom from th e
dioxygen is transferred to olefin, while the triphenylphosphine acts as a
co-reducing agent for reaction with the second oxygen atom to produce
OPPh 3. The proposed reaction pathway involves metallocycle formation
(31) followed by reductive elimination to form the oxetane. The oxetane
can undergo subsequent reaction with triphenylphosphine to yield OPPh3
and methyl ketone (32):
0-0
---. Rh+ I I
H 2C-CHR (31)

RHC-CH2
I I + PPh 3 ---. R-C-CH + OPPh 3 (32)
0-0 II 3
o
The reaction is inhibited by added water. The formation of such a metal-
locyclic intermediate resembles the isolation of such compounds from the
reaction of Pt02(PPh 3h with electron-deficient alkenes such as
tetracyanoethylene [33l 113 ):
Ph 3 P
"I0_ 0
/
Pt I
'\ ..... C
(33)
Ph 3 P (C~L (CN)2

Mares has also added support to this mechanism by showing that octene-1
is oxidized to unlabeled 2-octanone using RhCI(AsPh 3h as catalyst in the
HOMOGENEOUS CATAL YSIS OF OXIOA TfON REACTIONS 391

presence of 16 02 and 180H2,(114) thereby making it unlikely that a Wacker-


type mechanism is operable.
In an elegant extension of this work, Mimoun has used the compound
[Rh0 2L 4 ]X (L = AsPh 3 , AsPhMe2; X = CI0 4, PF 6 ) to convert octene-l
into 2- octanone, and he has shown that oxygen from 18 0 2 is exclusively
incorporated into the methyl ketone. (115) Mimoun has also shown that
RhCh and Cu(CI0 4h, or even Rh(CI0 4h alone, will convert terminal
alkenes into methyl ketones, and he proposed that this mechanism involves
a peroxymetalation pathway coupled with a Wacker cycle. (116) The hydroxyl
group for the Wacker step in the cycle is produced in the protonation of
the oxorhodium species remaining at the termination of the peroxymetala-
tion cycle (Figure 12.3). A similar chemistry has been explored using Rh(I)
and Cu(II) immobilized on site-separated organosulfide complexes. As the
gel's surface sulfide concentration is decreased the catalyst activity
increases, a result interpreted to support the concept that the active catalyst
. (117)
precursor IS a monomer.

Figure 12.3. Rhodium-catalyzed conversion of alkenes into methyl ketones with molecular
oxygen involving both a peroxymetalation and a Wacker cycle.
392 D. MAX ROUNDHILL

5.2. Alkenes and Hydrogen

This co-oxidation has been used by James(52) to convert a mixture of


hydrogen, oxygen, and cyclooctene into water and cyclooctanone. The
catalytic intermediate is proposed to be [IrHCh(CsHu)]2' The reaction is
accompanied by the catalytic hydrogenation of cyclooctene to cyclooctane.
No definitive mechanistic details have been given, and no data yet produced
to confirm whether the two oxidations occur by a coupled pathway.

5.3. Isocyanides and Carbon Monoxide

A brief report on the reaction of CO with Ni0 2 (tert-BuNCh at 20°


in chlorobenzene solvent shows that both CO 2 and t-BuNCO are formed
(34).(72) The yield of isocyanate is substantial, but again no mechanistic
details are given as to whether the oxygen transfer reactions occur by
separate or integrated pathways.

NiOz(t-BuNC)z + 4CO -. Ni(CO)z(t-BuNC)z + 2CO z + "I-BuNCO" (34)

'rr
2.:)00 1800
, I

II
~

Figure 12.4. Rhodium carbonyl intermediates in the Rh 6(COh6 catalyzed co-oxidation of


carbon monoxide and triphenylphosphine.
HOMOGENEOUS CATAL YSIS OF OXIOA TfON REACTIONS 393

5.4. Triphenvlphosphine and Carbon Monoxide

The compound Rh 6(COh6, in the presence of triphenylphosphine,


carbon monoxide, and oxygen, will catalyze the oxidation of reduced
substrates to form triphenylphosphine oxide and carbon dioxide. (SO-SI) The
solution changes color variously through red and yellow during the catalyzed
oxidation, and using infrared spectroscopy in the carbonyl region,
the solution species Rh 2(CO)4(PPh 3 )4, Rh 2(COh(PPh 3 )4(solventh,
Rh 2(COh(PPh 3 )6, and Rh 4(COho(PPh 3 h have been identified (Figure
12.4). The oxidation of CO only occurs when PPh 3 is present, and after
complete conversion to OPPh 3 the formation of CO 2 ceases. Decreasing
the pressure of oxygen while maintaining constant carbon monoxide and
other concentrations results in a drop in the quantity of CO 2 produced.
There is no 1: 1 correspondence in the quantity of carbon dioxide and
triphenylphosphine oxide formed-the amounts vary with time-however,
the ratio of CO 2 : OPPh 3 always remains above unity. Addition of water
to the reaction mixture produces an increase in the quantity of CO 2 formed.
Carrying out the reaction in the presence of IS02 and 1602/S0H2 confirms
that dioxygen is the only oxygen atom source for OPPh 3 • Analysis of the
formed carbon dioxide concludes that no oxygen-18 label is incorporated
into CO 2 from Is0H2, provided the contact time between the CO 2 formed
and the Is0H2 present in the reaction mixture is short. This latter condition
is necessary because the rhodium complexes in solution will catalyze the
oxygen exchange between CO 2 and water, even under the neutral pH
conditions of the reaction (35, 36):

co + PPh 3 + O 2 ~ CO 2 + OPPh 3 (35)


Rh(O)
CO z + 18 0Hz ~ CO z + 18 0CO + 18 0C 18 0 (36)

The suggested mechanistic pathways for these observations are shown in


Figure 12.5.

6. OXYGEN ATOM TRANSFER FROM METAL PHOSPHINE


HYDROPEROXIDES AND SUPEROXIDES

The involvement of hydroperoxy metal complexes in oxidation reac-


tions has been variously postulated and discussed.(4,6,l1.l1S,119) Metal hydro-
peroxides and alkylperoxides can be potentially prepared from hydrogen
peroxide or alkylperoxides. Alternately, it is apparent that since a dioxygen
molecule coordinated to a low-valent phosphine metal complex is an
394 D. MAX ROUNDHILL

2+
,!oH-
I' C
~~ 2+
JRhJ-CO
p- JRhJ-CO
OW HOO-
~
P=O P

Figure 12.5. Proposed pathways for both the Rh 6(COlt6 catalyzed co-oxidation of carbon
monoxide and triphenylphosphine, and the catalyzed oxygen atom exchange between carbon
dioxide and water.

electron rich center, it can also be protonated or alkylated by electrophile


to produce metal hydroperoxides or alkylperoxides. Such compounds can
be anticipated to show behavior resembling peroxides, and thus allow one
to do oxygen atom transfer chemistry from peroxidic compounds where

0-C
the initial oxygen source is molecular oxygen (37):

H
+ MOOH

M~ I
o
R+ +
MOOR
(37)

6.1. Metal Hydroperoxides and Alkylperoxides

When the compounds Pt(OH)(CF3 )L 2 and PtH(CF 3 )L 2 (L 2 = dppe,


2PPh 2Me) are treated with hydrogen peroxide, the hydroperoxy
platinum(II) compound Pt(OOH)(CF 3 )L 2 is formed (38).(120,121)

(38)

The hydroperoxide compounds having a range of ligands L are isolable,


and they react with PPh 3 , CO, and NO to form the oxidized products
OPPh3 , CO 2 , and HN0 2 • No reaction is observed with added cyclohexene.
The compound Pt0 2 (PPh 3 h under acidic conditions undergoes proton-
ation at oxygen with the formation of hydrogen peroxide. (122,123) The
HOMOGENEOUS CATAL YSIS OF OXIDA TlON REACTIONS 395

reaction with HX(X = CI04, BF6 , N0 3 ) proceeds in a stepwise manner,


and the complexes [Pt2(02)(OH)(PPh 3 )4]X and [Pt2(OH)z(PPh3)4]X2
have been isolated (39,40).

The structure of the formed peroxo-bridged compound shows an 0-0


separation of 1.547(21) A, and treatment of the compound with S02 yields
a product having a bidentate sulfate ligand. Similarly, when M0 2 (PPh 3 h
(M = Pd, Pt) is treated with PhC(O)NHOH at room temperature, hydrogen
peroxide is formed along with an aroylhydroxylamido complex of platinum
(41).(124)
M0 2 (PPh 3 h + PhC(O)HOH ---. M(ONC(O)Ph)(PPh 3 h + H 2 0 2 (41)
(M = Pd, Pt)

A similar reaction occurs with the hydrazine derivative PhC(0)NHNH2.


Addition of a stoichiometric amount of MeS03H to an anhydrous
solution of Pd0 2(PPh 3 h in CH 2Ch and octene-1 gives 2-octanone and
OPPh 3 • (125) The quantity of 2-octanone formed is proportional to the
quantity of added MeS03H, until a unit ratio of MeS03H: Pd0 2(PPh 3h is
obtained, after which the yield of 2-octanone remains constant. Other
terminal alkenes such as hexene-1 behave similarly, but internal alkenes
such as norbornene are unreactive. A mechanism is proposed involving a
cationic palladium hydroperoxide intermediate (Figure 12.6). Further work

°
-R'CMe
+5 •
[L "

L/ !
Pd

Dimers
/

"5
OR

r
R=H, Me, Ph 3 C
A=BF4,Me503,CF3503
S= CH 2 CI2
L=PPh 3

Figure 12.6. Proposed mechanism for the Pd0 2 (PPh 3 h catalyzed conversion of terminal
alkenes to methyl ketones with oxygen in acid medium.
396 D. MAX ROUNDHILL

on this project by Mimoun has shown that tert-butylperoxypalladium(II)


complexes can also be used for the selective oxidation of terminal alkenes
to methyl ketones, (126.127) and he presents a case whereby pseudocyclic
peroxymetalation can lead to either epoxidation or ketonization (42). (128.129)
Strong electrophilic metals such as Mo(VI) and V(V) direct the electron
transfer toward the alkyl carbon atom, producing epoxide, while less
electrophilic Group VIII metals direct electron transfer toward the (3-
carbon atom, from which (3 -hydride migration produces methyl ketone.

e
~
o
(42)
/'
/ MOR +RCMe
R CH=CH, II
o

6.2. Metal Peracyls

When the compound Pt0 2 (PPh 3 h is reacted with PhCOCI at -78°, a


peroxy-benzoate complex PtCI(OOCOPh)(PPh 3 h is formed (43).0 30 )

(43)

This thermally unstable compound reacts with triphenylphosphine to form


the oxide, and with alkenes to form epoxides. Yields of up to 50% epoxide
have been obtained using the alkenes norbornene and cyclohexene. Otsuka
et al. have recently isolated the bromo analog compound PtBr(OOCOPh)-
(PPh 3 h in a pure state and have shown that it will oxidize norbornene to
exo- norbornene epoxide in 40% yield. (44)

7. OXYGEN A TOM TRANSFER FROM COORDINA TED NITRITE


LIGANDS

7. 1. Transfer from Metal Nitrites to Carbon Monoxide and


Triphenylphosphine

The reaction of nitrite complexes with carbon monoxide to yield the


metal nitrosyl complex and carbon dioxide has been known for a consider-
HOMOGENEOUS CATAL YSIS OF OXIDA TlON REACTIONS 397

able time. (131-134) The initial work with Ni(N02h(PEt3h and carbon
monoxide yielded the nitrosyl complex Ni(NO)(N0 2 )(PEt3h(44)Y34)

(44)
The reaction is rapid under ambient conditions, and further work has shown
that the PEt3 ligand can be replaced by dppe, PMe2Ph, Ph 2PCH =
CHPPh 2Y35) The rate law for the reaction is -d[Ni(N0 2 hL 2 ]/ dt =
k 2 [Ni(N0 2 hL 2 ][CO], and an associative mechanism (45) is proposed.

o co 0
/ k,lsloWj I /
L 2 (N0 2 )Ni-N + CO L 2 (N0 2 )Ni-N
"0 lk "
3 (fast) 0
(45)

A similar reaction with Pd(N0 2h(PMePh 2 h and CO produces CO 2 and


the cluster compound Pd 4(CO)s(PMePh 2k(136) Further work on this reac-
tion using oxygen-18-labeled nitrite has confirmed this ligand to be the
source of oxygen, and a cyclic intermediate is proposed for the oxygen
transfer step. (137)
Similarly, complexes NiL4 (L = tertiary phosphine) react with aliphatic
and aromatic nitro compounds RN0 2 to yield the complex Ni(RNO)L 2
and phosphine oxide (46).(138) The ligand RNO is kinetically labile and
readily displaced by phosphine L(47), thereby making it possible to design
a catalytic oxidation cycle. For aliphatic nitro alkanes, oxygen transfer is
retarded both
NiL4 + RN0 2 ~ Ni(RNO)L 2 + L + L=O (46)
Ni(RNO)L 2 + 2L ~ NiL4 + RNO (47)
by branching at the a carbon and by electron release from the alkyl group.
It is believed that the mechanism involves an electron-transfer process to
z
form Ni I L 3·RNO prior to oxygen-atom transfer, and that neither N-O
bond-breaking nor P-O bond-making steps are rate-limiting. A cyclic
intermediate (48) is considered to be possible, but this suggestion is not
proven.

(48)
398 D. MAX ROUNDHILL

Johnson, in a series of articles, has investigated oxygen-atom transfer


from nitrite complexes of Ni, Pd, and Pt. (139-14Z) In addition to the catalytic
oxidation of CO to CO z with molecular oxygen, using Ni(NOzhdppe as
catalyst, these workers also observe some isocyanate formation on treating
Pt(NO zh(PEt3 h with CO (49).

Pt(N0 2 lz(PEt 3 lz + 2CO ---+ Pt(NCO)(N0 2 )(PEt3 lz + CO 2 (49)

7.2. Transfer from Metal Nitrites to Alkenes

Other metal complexes that will undergo oxygen-atom transfer from


a coordinated nitrite are M(NO zh(COh(PPh 3h (M = Ru, Os), which oxi-
dizes CO ;(143) Co(N,N- Bisalicylidene-o -phenylenediamino )py . NO z, which
oxidizes PPh3;(144) Ni(NOzh(PBu3)z, which catalyzes the oxidation of t-
BuNC to t_BuNCO;(145) Co(tetraphenylpophyrin)py·NO z, which catalyzes
alkene oxidation;(146,147) and PdCI(MeCNhNO z, which also catalyzes
alkene oxidation. (148) The complex Co(tetraphenylporphyrin)py· NO z,
which reacts by transfer of an oxygen atom to the alkene, can be considered
as having a weak, oxygen-centered nucleophile on the NO ligand. The z
alkene is complexed to palladium(II), such that it is susceptible to
nucleophilic attack. Using this concept, catalytic amounts of acetaldehyde
and acetone have been obtained from ethylene and propylene, respectively.
No palladium precipitates from the reaction mixture, since the formed
Pd(O) species are being reoxidized by the cobalt nitrite complexes. Thus,
the process involves oxygen transfer from the nitro ligand of cobalt-nitro
complexes to palladium(II)-bound alkene, followed by reoxidation of the
reduced nitrosyl ligand by molecular oxygen. The palladium serves exclus-
ively as a cocatalyst. (146) In the conversion of decene-l into 2-decanone
using PdCI(MeCNhNO z, an oxygen-l8-labeling study verifies that the
ketonic oxygen atom originates from the nitro group. Spectroscopic
evidence is presented for two intermediates, an alkene complex and a
species derived from attack by the nitro group on coordinated alkene. (148)
These alkene oxidations using oxygen-atom transfer from a coordinated
nitrite ligand offer advantages of selectivity over catalytic pathways involv-
ing complexes having dioxygen directly bonded to the metal center.

ACKNOWLEDGMENTS

We thank M. A. Andrews, R. D. Feltham, W. A. Goddard III, B. R.


James, H. Mimoun, and J. Valentine for making articles available prior to
publication. Our own research in this field is supported by the Division of
HOMOGENEOUS CATAL YSIS OF OXIDA TlON REACTIONS 399

Chemical Sciences of the U.S. Department of Energy, to whom we owe


thanks.

REFERENCES

1. S. Benson and P. S. Nangia, Accts. Chem. Res. 12, 223-228 (1979).


2. O. Hayaishi, Molecular Mechanisms of Oxygen Activation (Academic Press, New York,
1974).
3. T. G. Spiro, ed., Metal Ion Activation of Dioxygen (John Wiley and Sons, New York,
1980). Vol. 2 in Metal Ions in Biology.
4. J. Lyons, Aspects of Homogeneous Catalysis, edited by R. Ugo (Reidel, Boston, 1977),
Vol 3, pp. 1-136.
5. - - , Fundamental Research in Homogeneous Catalysis, edited by M. Tsutsui and R.
Ugo (Plenum Publishing, New York, 1977), pp. 1-52.
6. G. Modena, K. B. Sharpless, G. Costa, J. Halpern, Y. Ishii, B. R. James, J. E. Lyons,
H. Minoun, P. Rossi, R. A. Sheldon, and P. Teyessie, Fundamental Research in
Homogeneous Catalysis, edited by M. Tsutsui and R. Ugo (Plenum, New York, 1977),
pp. 193-205.
7. R. Stewart, Oxidation in Organic Chemistry, Part A, edited by K. B. Wiberg (Academic
Press, New York, 1965), pp. 1-68.
8. K. B. Wiberg, Oxidation in Organic Chemistry, Part A, edited by K. B. Wiberg (Academic
Press, New York, 1965), pp. 69-184.
9. W. A. Waters and J. S. Littler, Oxidation in Organic Chemistry, Part A, edited by K.
B. Wiberg (Academic Press, New York, 1965), pp. 185-241.
10. D. G. Lee and M. van der Engh, Oxidation in Organic Chemistry, Part B, edited by
W. S. Trahanovsky (Academic Press, New York, 1973), pp. 177-227.
11. R. A. Sheldon and J. K. Kochi, Catalyzed Oxidations of Organic Compounds (Academic
Press, New York, 1981).
12. L. H. Chinn, Selection of Oxidants in Synthesis (Marcel Dekker, New York, 1971).
13. P. M. Henry, Palladium Catalyzed Oxidation of Hydrocarbons (Riedel, Boston, 1980).
14. L. Reich and S. S. Stivala, Autoxidation of Hydrocarbons and Polyolefins (Marcel Dekker,
New York, 1969).
15. B. Plesnicar, Oxidation in Organic Chemistry, Part C, edited by W. S. Trahanovsky
(Academic Press, New York, 1978), pp. 211-294.
16. L. Vaska, Accts. Chem. Res. 9,175-183 (1976).
17. L. H. Vogt, Jr., H. M. Faigenbaum, and S. E. Wiberley, Chem. Rev. 63, 269-277 (1963).
18. J. Valentine, Chem. Rev. 73, 235-245 (1973).
19. V. J. Choy and C. J. O'Connor, Coord. Chem. Rev. 9, 145-170 (1972).
20. R. S. Drago and B. B. Corden, Acc. Chem. Res. 13, 353-360 (1980).
21. F. Basolo, B. M. Hoffman, and J. A. Ibers, Acc. Chem. Res. 8, 384-392 (1975).
22. K. B. Sharpless and T. R. Verhoeven, Aldrichimica Acta. 12,63-74 (1979).
23. A. K. Rappe and W. A. Goddard III, J. Am. Chem. Soc. 104,3287-3294 (1982).
24. T. Katsuki and K. B. Sharpless, J. Am. Chem. Soc. 102, 5974-5976 (1980).
25. L. Vaska, Science 140,809-810 (1963).
26. C. D. Cook and G. S. Jauhal, Inorg. Nucl. Chem. Lett. 3,31-33 (1967).
27. R. Ugo, G. LaMonica, F. Cariati, S. Cenini, and F. Conti, Inorg. Chim.Acta 4,390-394
(1970).
28. S. Otsuka, A. Nakamura, and Y. Tatsuno, J. Chem. Soc., Chem. Comm. 836 (1967).
400 D. MAX ROUNDHILL

29. - - , l. Arn. Chern. Soc. 91, 6994-6999 (1969).


30. M. J. Nolte, E. Singleton, and M. Laing, l. Arn. Chern. Soc. 97, 6396-6400 (1975).
31. B. W. Graham, K. R. Laing, C. J. O'Connor, and W. R. Roper, l. Chern. Soc., Chern.
Cornrn. 1272 (1970).
32. H. Suzuki, K. Mizutani, Y. Mora-oka, and T. Ikawa, l. Arn. Chern. Soc. 101,748-749
(1979).
33. F. Sakurai, H. Suzuki, Y. Mora-oka, and T. Ikawa, l. Arn. Chern. Soc. 102, 1749-1751
(1980).
34. Y. Mora-oka, H. Suzuki, R. Sugimoto, F. Sakurai, and T. Ikawa, Abstr. Xth. International
Conf. Organornetallic Chernistry, Toranto, Abstr. 5D04.
35. A. Vogler and H. Kunkely, l. Arn. Chern. Soc. 103,6222-6223 (1981).
36. J. J. Levison and S. D. Robinson, l. Chern. Soc. (A) 762-767 (1971).
37. P. J. Hayward, D. M. Blake, G. Wilkinson, and C. J. Nyman, l. Arn. Chern. Soc. 92,
5873-5878 (1970).
38. R. Ugo, F. Conti, S. Cenini, R. Mason, and G. B. Robertson, l. Chern. Soc., Chern.
Cornrnun. 1498-1499 (1968).
39. R. W. Horn, E. Weissberger, and J. P. Collman, Inorg. Chern. 91, 2367-2371 (1970).
40. J. S. Valentine, D. Valentine, and J. P. Collman, Inorg. Chern. 10,219-225 (1971).
41. R. Ugo, G. M. Zanderighi, A. Fusi, and D. Carreri, l. Arn. Chern. Soc. 102, 3745-3751
(1980).
42. W. B. Beaulieu, G. D. Mercer, and D. M. Roundhill, l. Arn. Chern. Soc. 100, 1147-1152
(1978).
43. H. C. Clark, A. B. Goel, and C. S. Wong, l. Arn. Chern. Soc. 100, 6241-6243 (1978).
44. Y. Tatsuno and S. Otsuka, l. Arn. Chern. Soc. 103,5832-5839 (1981).
45. W. J. Louw, T. I. A. Gerber, and D. J. A. de Waal, l. Chern. Soc., Chern. Cornrnun.
760-761 (1980).
46. J. G. Norman, Jr. Inorg. Chern. 16, 1328-1335 (1977).
47. R. Criegee, Angew. Chernie. Int. Ed. Engl. 14,745-752 (1975).
48. G. D. Fong and R. L. Kuczkowski, l. Arn. Chern. Soc. 102,4763-4768 (1980).
49. S. E. Jacobson, R. Tang, and F. Mares, l. Chern. Soc., Chern. Cornrnun. 888-889 (1978).
50. J. P. Collman, M. Kubota, and J. W. Hosking, l. Arn. Chern. Soc. 89, 4809-4811 (1967).
51. A. Fusi, R. U go, A. Pasini, and S. Cenini, l. Organornet. Chern. 26, 417-430 (1971).
52. B. R. James, Adv. Chern. Ser. 191, 253-276 (1980).
53. R. VaAtta, J. Burstyn, and J. S. Valentine, Reactions of Coordinated Ligands, edited
by P. S. Braterman (Plenum, New York, 1983), in press.
54. H. Mimoun, Pure Appl. Chern. 53,2389-2399 (1981).
55. R. A. Sheldon and J. K. Kochi, Adv. Ca ta I. 25,272-413 (1976).
56. F. Haber and J. Weiss, Proc. Roy. Soc. 147,332-351 (1934).
57. A. Fusi, R. Ugo, and G. M. Zanderighi, l. Catal. 34,175-190 (1974).
58. S. Cenini, A. Fusi, and G. Capparella, l. Inorg. Nucl. Chern. 33,3576-3579 (1971).
59. - - , Inorg. Nucl. Chern. Lett. 8,127-131 (1972).
60. V. P. Kurkov, J. Z. Pasky, and J. B. Lavigne, l. Arn. Chern. Soc. 90, 4743-4744 (1968).
61. H. Arzoumanian, A. Blanc, U. Hartig, and J. Metzger, Tetrahedron Lett. 1011-1014
(1974).
62. H. Arzoumanian, A. A. Blanc, J. Metzger, and J. E. Vincent, l. Organornet. Chern.
82,261-270(1974).
63. A. A. Blanc, H. Arzoumanian, E. J. Vincent, and J. Metzger, Bull. Chirn. Soc. Fr.
2175-2179 (1974).
64. W. Strahmeier and E. Eder, l. Organornet. Chern. 94, C14-C19 (1975).
65. S. Cenini, A. Fusi, and F. Porta, Gazz. Chirn. Ital. 108, 109-114 (1978).
66. K. Takao, Y. Fujiwara, T. Imanaka, and S. Teranishi, Bull. Chern. Soc. lpn. 43,
3898-3900 (1970).
HOMOGENEOUS CATAL YSIS OF OXIDA TlON REACTIONS 401

67. K. Takao, M. Wayaku, Y. Fujiwara, T. Imanaka, and S. Teranishi, Bull. Chem. Soc.
fpn. 43, 3898-3900 (1970).
68. J. E. Lyons and J. O. Turner, Tetrahedron Lett. 2903-2906 (1972).
69. - - , f. Org. Chem. 37, 2881-2884 (1972).
70. F. Mares and R. Tang, f. Org. Chem. 43,4631-4632 (1978).
71. B. H. van Vugt and W. Drenth, Reel. Trav. Chim. Pays-Bas. 96, 225-229 (1977).
72. S. Otsuka, A. Nakamura, Y. Tatsuno, and M. Miki, f. Am. Chem. Soc. 94,3761-3767
(1972).
73. J. Halpern and A. C. Harkness, f. Am. Chem. Soc. 83,1258-1259 (1961).
74. J. Halpern and S. Nakamura, f. Am. Chem. Soc. 83,4102-4103 (1961).
75. B. R. James and G. L. Rempel, f. Chem. Soc. A., 78-84 (1969).
76. J. E. Bercaw, L.-Y. Goh, and J. Halpern, f. Am. Chem. Soc. 94, 6534-6536. (1972).
77. G. D. Mercer, J. S. Shu, T. B. Rauchfuss, and D. M. Roundhill, f. Am. Chem. Soc. 97,
1967-1968 (1975).
78. c. S. Chin, M. S. Sennett, and L. Vaska, f. Mol. Catal. 4, 375-378 (1978).
79. K. L. Watters, R. F. Howe, T. P. Chojnacki, C.-M. Fu, R. L. Schneider, and N. B.
Wong, f. Catal. 66, 424-440 (1980).
80. D. M. Roundhill, M. K. Dickson, N. S. Dixit, and B. P. Sudha-Dixit, Adv. Chem. Ser.
196,291-301 (1982).
81. M. K. Dickson, Ph.D. Thesis, Washington State University (1982).
82. Y. Iwashita and A. Hayata, f. Am. Chem. Soc. 91, 2525-2528 (1969).
83. W. R. Cullen, B. R. James, and G. Strukul, Canad. f. Chem. 56, 1965-1969 (1978).
84. C.-S. Chin, M. S. Sennett, P. J. Wier, and L. Vaska, Inorg. Chim. Acta. 31, L443-L444
(1978).
85. T. Engel and G. Ertle, Adv. Catal. 28,1-78 (1979).
86. D. M. Roundhill, G. H. Allen, R. A. Bechtold, and W. B. Beaulieu, Inorg. Chim. Acta.
54, L99-LlOO (1981).
87. J.-I. Hojo, S. Yuasa, N. Yamazoe, I. Mochida, and T. Seiyama, f. Cata!. 36,93-98
(1975).
88. H. Sakamoto, T. Funabiki, and K. Tarama, f. Catal. 48, 427-429 (1977).
89. M. K. Dickson, B. P. Sudha, and D. M. Roundhill, f. Organornet. Chem. 190, C43-C46
(1980).
90. D. M. Roundhill, M. K. Dickson, N. S. Dixit, and B. P. Sudha-Dixit, f. Arn. Chern.
Soc. 102, 5538-5542 (1980).
91. D. M. Roundhill, Proc. of the 2nd International Workshop on Fundarnental Research in
Homogeneous Catalysis, edited by Y. Ishii and M. Tsutsui (Plenum, New York, 1978),
pp. 11-23.
92. G. D. Mercer, W. B. Beaulieu, and D. M. Roundhill, f. Arn. Chern. Soc. 99, 6551-6554
(1977).
93. E. W. Stern, f. Chern. Soc., Chern. Cornrnun. 736 (1970).
94. R. A. Sheldon, f. Chem. Soc., Chern. Cornmun. 788-789 (1971).
95. G. Wilke, H. Schott, and P. Heimbach, Angew. Chernie Int. Ed. Eng!. 6, 92-93
(1967).
96. J. P. Birk, J. Halpern, and A. L. Pickard, f. Am. Chern. Soc. 90, 4491-4492 (1968).
97. J. Halpern and A. L. Pickard, Inorg. Chern. 9, 2798-2800 (1970).
98. A. Sen and J. Halpern, f. Am. Chern. Soc. 99,8337-8339 (1977).
99. B. W. Graham, K. R. Laing, C. 1. O'Connor, and W. R. Roper, f. Chern. Soc. Dalton
Trans. 1237-1243 (1972).
100. R. K. Pod dar and U. Agarwala, Inorg. Nuel. Chern. Lett. 9,785-789 (1973).
101. R. L. Augustine and J. Van Peppen, f. Chern. Soc., Chem. Cornrnun. 497-498 (1970).
102. B. H. Van Vugt, N. 1. Kooke, W. Drenth, and F. P. 1. Kiujpers, Reel. Trav. Chirn.
Pays-Bas. 92, 1321-1325 (1973).
402 D. MAX ROUNDHILL

103. J. Halpern, B. L. Goodall, G. P. Khare, H. S. Lim, and J. J. Pluth, 1. Am. Chem. Soc.
97,2301-2303 (1975).
104. D. D. Schmidt and J. T. Yoke,l. Am. Chem. Soc. 93, 637-640 (1971).
105. R. P. Hanzlik and D. Williamson, 1. Am. Chem. Soc. 98, 6570-6573 (1976).
106. W.-S. Hwang, I. B. Joedicke, and J. T. Yoke, Inorg. Chem. 19,3225-3229 (1980).
107. W.-S. Hwang and J. T. Yoke, 1. Org. Chem. 45, 2088-2091 (1980).
108. J. Drapier and A. 1. Hubert, 1. Organomet. Chem. 64, 385-391 (1974).
109. M. S. Jarrell, B. C. Gates, and E. D. Nicholson, 1. Am. Chem. Soc. 100, 5727-5732
(1978).
110. C. Dudley and G. Read, Tetrahedron Lett. 52, 5273-5276 (1972).
111. C. W. Dudley, G. Read, and P. J. C. Walker, 1. Chem. Soc., Dalton Trans. 1926-1931
(1974).
112. G. Read and P. J. C. Walker, 1. Chem. Soc., Dalton Trans. 883-888 (1977).
113. R. A. Sheldon and J. A. Van Doorn, 1. Organomet. Chem. 94,115-129 (1975).
114. R. Tang, F. Mares, N. Neary, and D. E. Smith, 1. Chem. Soc., Chem. Commun. 274-275
(1979).
115. F. Igersheim and H. Mimoun, Nouv. 1. de Chimie. 4,161-166 (1980).
116. H. Mimoun, M. M. Perez Machirant, and I. Seree de Roch, 1. Am. Chem. Soc. 100,
5437-5444 (1978).
117. E. D. Nyberg and R. S. Drago, 1. Am. Chem. Soc. 103,4966-4968 (1981).
118. A. G. Davies, Organic Peroxides (Butterworths, London, 1961).
119. R. A. Sheldon, Aspects of Homogeneous Catalysis, edited by R. Ugo (Reidel, Boston,
1981), Vol. 4, pp. 3-70.
120. R. A. Michelin, R. Ros, and G. Strukul, Inorg. Chim. Acta. 37, L491-L492 (1979).
121. G. Strukul, R. Ros, and R. A. Michelin, Abstr. Xth International Conference
Organometallic Chemistry, Toronto Abstract 5D06 (1981).
122. S. Bhaduri, L. Casella, R. Ugo, P. R. Raithby, C. Zuccaro, and M. B. Hursthouse, 1.
Chem. Soc., Dalton Trans. 1624-1629 (1979).
123. R. Ugo, A. Fusi, G. M. Zanderighi, and L. Casella, 1. Mol. Catal. 7, 51-57 (1980).
124. P. L. Bellon, S. Cenini, F. Demartin, M. Manassero, M. Pizzotti, and F. Porta 1. Chem.
Soc., Dalton Trans. 2060-2067 (1980).
125. F. Igersheim and H. Mimoun, Nouv. 1. de Chimie. 4, 711-713 (1980).
126. M. Roussel and H. Mimoun, 1. Org. Chem. 45, 5387-5390 (1980).
127. J.-M. Bregeault and H. Mimoun, Nouv. 1. de Chimie. 5,287-289 (1981).
128. H. Mimoun, R. Charpentier, A. Mitschler, J. Fischer, and R. Weiss, 1. Am. Chem. Soc.
102,1047-1054 (1980).
129. H. Mimoun, 1. Mol. Catal. 7,1-29 (1980).
130. M. J. Y. Chen and J. K. Kochi, 1. Chem. Soc., Chem. Commun. 204-205 (1977).
131. W. Manchot and A. Waldmuller, Chem. Ber. 59, 2363-2366 (1926).
132. W. Hieber and J. S. Anderson, Z. Anorg. Allgem. Chem. 208, 238-248 (1932).
133. W. Hieber and H. Beutner, Z. Natuforsch. B, 15, 323-324 (1960).
134. G. Booth and J. Chatt, 1. Chem. Soc. 2099-2106 (1962).
135. R. D. Feltham and J. C. Kriege, 1. Am. Chem. Soc. 101, 5064-5065 (1979).
136. J. Dubrawski, J. C. Kreige-Simondsen, and R. D. Feltham, 1. Am. Chem. Soc. 102,
2089-2091 (1980).
137. D. T. Doughty, G. Gordon, and R. P. Stewart Jr., 1. Am. Chem. Soc. 101,2645-2648
(1979).
138. R. S. Berman and J. K. Kochi, Inorg. Chem. 19, 248-254 (1980).
139. S. Bhaduri, B. F. G. Johnson, A. Khair, A. Pickard, Y. Ben-Taarit, and R. Ugo, 1.
Chem. Soc., Chem. Commun. 694-695 (1976).
140. B. F. G. Johnson, C. J. Savory, J. A. Segal, and R. H. Walter, 1. Chem. Soc., Chem.
Commun. 809-810 (1974).
HOMOGENEOUS CATAL YSIS OF OXIDATION REACTIONS 403

141. S. Bhaduri, B. F. G. Johnson, A. Khair, I. Ghatak, and D. M. P. Mingos, 1. Chem.


Soc., Dalton Trans. 1582-1576 (1980).
142. S. A. Bhaduri, I. Bratt, B. F. G. Johnson, A. Khair, J. A. Segal, R. Walters, and C.
Zuccaro, 1. Chem. Soc., Dalton Trans. 234-239 (1981).
143. K. R. Grundy, K. R. Laing, and W. R. Roper,!. Chem. Soc., Chem. Commun. 1500-1501
(1970).
144. B. S. Tovrog, S. E. Diamond, and F. Mares, 1. Am. Chem. Soc. 101, 270-272 (1979).
145. M. A. Andrews, personal communication.
146. B. S. Tovrog, F. Mares, and S. E. Diamond, 1. Am. Chem. Soc. 102, 6616-6618 (1980).
147. F. Mares, B. S. Tovrog, and S. E. Diamond, Abstr. Xth International Conference
Organometallic Chemistry, Toronto, Abstract 5D05 (1981).
148. M. A. Andrews and K. P. Kelly, 1. Am. Chem. Soc. 103, 2894-2896 (1981).
13
Catalysis of Nitrogen-Fixing
Model Studies
T. Adrian George

NOTATION

dppe: 1,2-Bisdiphenylphosphinoethane, Ph 2PCH 2CH 2PPh 2


THF: Tetrahydrofuran
dmpe: 1,2-Bisdimethylphosphinoethane, Me2PCH2CH2PMe2
Cp: Pentahaptocyclopentadienyl, TIs -CsHs
dppm: Bisdiphenylphosphinomethane, Ph 2PCH 2PPh 2
dptpe: 1,2-Bisdi-p -tolylphosphinoethane, (p- tolylhPCH 2CH 2P( totyl-p h
depe: 1,2-Bisdiethylphosphinoethane, Et2PCH2CH2PEt2
triphos: Bis(2-diphenylphosphinoethyl)phenylphosphine, PhP(CH 2CH 2PPh 2h
Cy: Cyclohexyl
DME: 1,2-Dimethoxyethane
dppp: 1,3-Bisdiphenylphosphinopropane, Ph2PCH2CH2CH2PPh2

1. INTRODUCTION

The catalytic fixation of dinitrogen is accomplished biologically and


abiologically. In nature, both symbiotic and non symbiotic nitrogen-fixing
microorganisms reduce dinitrogen from the atmosphere or soil to
ammonia. (1-4) A common denominator among these microorganisms is the

T. Adrian George • Department of Chemistry, University of Nebraska-Lincoln, Lincoln,


Nebraska 68588.

405
406 T. ADRIAN GEORGE

enzyme nitrogenase (comprising a molybdenum-iron [MoFe] protein and


an iron protein), which has been isolated from a number of diverse microor-
ganisms: anaerobic species (e.g., Clostridium pasteurianum), strict aerobes
(e.g., Azobacter vinelandii), and facultative aerobes (e.g., Klebsiella pneu-
moniae ). (5,6) The requirements for in vitro reduction of dinitrogen to
ammonia by nitrogenase are (i) a reducing agent (such as Na2S204); (ii) a
divalent cation (usually Mg +2); (iii) ATP; (iv) an anaerobic environment;
and (v) a controlled pH.(7) Of special interest to the chemist is the nature
of the active site, the site at which dinitrogen is reduced to ammonia in
the enzyme, and the mechanism of the reaction. A number of significant
steps toward the elucidation of the nitrogenase reaction have been made
recently. Among these are (i) the isolation of an iron-molybdenum cofactor
(FeMoco) with a molecular mass of less than 5000 Daltons that will
reconstitute the cofactorless MoFe protein from a number of mutant
organisms;(8) (ii) application of X-ray Absorption Spectroscopy, par-
ticularly the analysis of the Extended X-ray Absorption Fine Structure, to
determine the coordination environment of molybdenum in FeMoco and
the MoFe proteins isolated from nitrogenase of a number of organisms;(9)
(these data have established both sulfur and iron atoms as nearest neigh-
bors for the molybdenum atom in the MoFe proteins and FeMoco) and
(iii) detection of an intermediate dinitrogen hydride, which is formed on
the enzyme when it is actively fixing dinitrogen. (10) The biological fixation
of dinitrogen is best represented by Equation (1), in which dihydrogen
production is a specific part of the nitrogenase reaction (11):

(1)

Abiologically, dinitrogen is catalytically converted to ammonia in the Haber


Process. (12) Dinitrogen and dihydrogen react to form ammonia on a pro-
moted (K 20, MgO, CaO, BaO, Mo0 3 , etc.) iron-alumina catalyst at around
350-1000 atmospheres pressure and 300-400°C. While the catalyst in this
process is cheap, an incentive to discover a better catalyst is provided by
the dramatic increases in the cost of energy, which is required for the high
temperatures and pressures, in particular of natural gas, from which most
dihydrogen is obtained for this process by the reforming reaction (methane
and steam). Economic advantages would be gained by either finding a
catalyst that will enable the Haber Process to run under milder conditions
or by discovering a cheaper source of dihydrogen, or both.

1. 1. Scope and Limitations

Nitrogen fixation can be defined as any chemical reaction in which


dinitrogen is a reagent, although it is generally thought of in terms of
CA TAL YSIS OF NITROGEN-FIXING MODEL STUDIES 407

reactions in which ammonia, hydrazine, or organonitrogen compounds are


formed. Nitrogen fixation may involve oxidation or reduction, although
only examples of the latter will be encountered in this chapter since no
examples of oxidation have been reported within the limits imposed in this
chapter (see below). A logical starting point for any discussion of nitrogen
fixation is the preparation of complexes containing ligating dinitrogen that
have been formed in reactions in which dinitrogen is a reagent (e.g.,
Equation 2)(13):

(2)

This follows naturally from (i) the assumption that nitrogenase binds
dinitrogen at a metal center prior to its reduction to ammonia, and (ii) the
fact that the rate-determining step in the Haber Process is the binding of
dinitrogen at the surface of the catalyst. (14) In many cases, direct interaction
of dinitrogen with metals or highly reduced metral complexes results in a
nitriding reaction. Subsequent hydrolysis yields ammonia and/ or hydrazine.
However, reactions in which dinitrogen is generated during the reaction
and then becomes ligated do not fit within this definition of nitrogen fixation
. 3(15)) :
(e.g., E quatlOn

CHC1 3
trans-[IrCl(CO)(PPh 3 hl + RN3 ~ trans-[IrCl(N 2 )(PPh 3 hl + RNHCOOEt 3)
(

Discussion of nitrogen fixation within this chapter is limited by the key


words homogeneous catalysis and metal phosphine complexes. Currently,
there are no examples of homogeneous, catalytic nitrogen fixation involving
metal-phosphine complexes. However, there are examples of both
heterogeneous and homogeneous catalysis involving other metal complexes
that will be reviewed in Section 3. Justification for devoting a chapter to
the subject of nitrogen fixation is not difficult to find. First, a number of
vitally important, naturally occurring microorganisms exist that effect
homogeneous catalytic nitrogen fixation. This provides a stimulus to try
and emulate the natural system and, with an understanding of the mecha-
nism of nitrogenase, contributes to the general knowledge of catalysis.
Secondly, the large strides that have been made in stoichiometric
homogeneous nitrogen fixation contribute toward a unifying view of reduc-
tive nitrogen fixation. From these studies it is hoped that realistic, catalytic
nitrogen-fixing systems will evolve. Thirdly, there are economic, political,
and social pressures to supply sufficient food to feed the ever increasing
world population; these pressures provide an incentive to develop alterna-
tive nitrogen-fixing system~ regardless of whether they are oxidative,
reductive, heterogeneous, or homogeneous.
408 T. ADRIAN GEORGE

In limiting this chapter to metal-phosphine complexes, many significant


compounds and nitrogen-fixing systems will not be discussed. However,
interested readers are referred to a number of relatively recent reviews of
the chemistry of nitrogen fixation. (16-18) Similarly, adequate descriptions
of the bonding and physical properties of the metal dinitrogen complexes
have appeared elsewhere and will only be mentioned in this chapter where
relevant.
This chapter is organized in the following manner. It is assumed that
coordination of dinitrogen to a transition metal is the initial step in all
nitrogen-fixing reactions in which a transition metal is present. Therefore,
only those systems in which it has been clearly demonstrated that chemistry
beyond that point has occurred will be discussed. Well-characterized com-
plexes containing coordinated dinitrogen and phosphine ligands will be
listed in Table 1, but only if dinitrogen is the source of the ligating
dinitrogen. The preparation of these complexes will be referenced in Table
1 but not discussed. Reactions of coordinated dinitrogen will be presented
and discussed according to the metal to which it is bonded. Reactions that
have been observed and will be reviewed include the formation of nitrogen-
hydrogen and nitrogen-carbon bonds, as well as coordination of ligated
dinitrogen to a second metal. In a few cases nitrogen fixation has been
shown to occur without isolation of an intermediate complex containing
coordinated dinitrogen and phosphine ligands. These reactions will be
included because in a successful homogeneous nitrogen-fixing system it is
unlikely that any intermediate dinitrogen complex will be isolated (or
identified). The major part of this chapter will be devoted to the chemistry
of dinitrogen complexes of molybdenum, rhenium, and tungsten, together
with recent developments of tantalum. To date, most of the chemistry of
dinitrogen coordinated to other metals is limited to displacement of
dinitrogen or reaction at the metal. Reactions of ligated nitrogen-hydrides
and similar species (e.g., Equation 4(19») will be omitted from consideration
unless they bear directly upon the mechanism of nitrogen-fixing reactions.

[MoF(NNH 2 lz(dppeht + RCHO ~ [MoF(NNCHR)(dppeht + H 2 0 (4)

2. NITROGEN-FIXING REACTIONS

The majority of nitrogen-fixing reactions begin with a preformed


dinitrogen complex. These complexes show a wide range of stabilities.
Among the most stable are the molybdenum and tungsten complexes
containing the dppe ligand; trans-M(N 2 h(dppeh, (M = Mo, W). These two
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 409

complexes have been the most thoroughly studied and together with similar
complexes containing monodentate ligands have displayed the most diverse
chemistry of coordinated dinitrogen.
Briefly, the dinitrogen complexes listed in Table 1 are prepared by a
number of different methods. The most common method involves the
reduction in an ether solvent under a dinitrogen atmosphere of either a
phosphine complex of a metal halide or a mixture of metal salt and
appropriate phosphine. The only other frequently used method is ligand
displacement by dinitrogen. To illustrate these methods, a number of
high-yield preparations of trans-Mo(N 2 h(dppeh are given in Equations
5-8 (20-22):

MoCb(THFh + 2 dppe Mg
THF,N/ trans- [ Mo(N2lz(dppe)z] (5)
Mo C Is + 2 d ppe I'<a/Hg
THF,N! trans- [ Mo(N2h(dppe)z ] (6)
MoCl2(dppeh ~~F~
, 2
I trans-[Mo(N 2 h(dppeh] (7)
MoH4(dppe)z + 2N2 THF I trans-[Mo(N 2h(dppeh] + 2H2 (8)

In the following sections, nitrogen-fixing reactions will be reviewed element


by element rather than according to the type of reaction. All reactions are
reductive. There are no examples to date of oxidative nitrogen fixation
within the confines of this chapter.

2.1. Titanium, Zirconium, and Hafnium

There are no dinitrogen complexes of any of these metals that contain


a phosphine ligand. The considerable amount of nitrogen-fixing chemistry
reported for titanium, and to a lesser extent zirconium, involves alkoxide,
cyclopentadienyl, and halide complexes. (23-25) The catalytic nitriding reac-
tions of titanium will be reviewed in Section 3.
Titanium(IV) chloride forms a number of different adducts with trans-
[ReCI(N2)(PMe2Ph)4J (see Table 1) that are believed to involve dinitrogen
bridging between rhenium and titanium. Hydrolysis or treatment of these
ad ducts with ethanol results in the virtual quantitative recovery of the
original rhenium-dinitrogen complex. (26)

2.2. Vanadium, Niobium, and Tantalum

A large amount of work has been carried out, primarily by Shilov, on


the chemical fixation of dinitrogen using various vanadium-containing
systems.(24) However, no stable dinitrogen complexes of vanadium have
been reported or in situ nitrogen fixation observed with phosphines present.
410 T. ADRIAN GEORGE

Table 1. Oinitrogen Complexes Prepared from Oinitrogen

Complex Text No. Reference

[TiCI4{(N2)ReCI(PMe2Ph)4lz] l a 26
[(THF)TiCI 4{(N 2)ReCl(PMe2Ph)4}] 2a 26
[Ti2CI60(OEt2){(N2)ReCI(PMe2Ph)4}] 3a 26
[{NbCl(dmpehh(N 2)] 4b 27
[{Ta(CHCMe3)(PMe3hCllz(N2)] 5 30
[{Ta(CHCMe3)(PMe3hRlz(N2)] 6 a ,c 30,31 c
(R = Me, CH 2CMe3C)
[{TaCl(PMe3M C 2H 4)lz (N 2)] 7 30
cis- [Cr(N2h(PMe3)4] 8 33
[{Cr(dppeh}z(N 2)] 9 34
[(THFhCI3Cr{(N2)ReCl(PMe2Ph)4}l lOa 35
[MoAr(N2)(PR 3hl 11 36,37
(Ar = C6HSMe, C6H3Me3; PR 3 = PPh 3, PPh 2Me)
[{Mo( 1) 6- C 6H 6 )(PPh 3hlz (N 2)] 12 37
[{Mo(1) 6 -C6H3Me3)(dmpe)lz(N2)l 13 c 38,39 c
[Ar(PPh 3 hMo(N 2)FeCp( dmpe) lBF4 14 a 37
(Ar = C 6 HsMe)
trans- [Mo(N 2M dppe hl 15 c 20,21,40,
41 c ,42-46
trans- [Mo(N 2h(dppmh] 16 b 40

trans- [Mo(N 2h( dppp hl 17 40

trans-[Mo(N 2 h(dptpehl 18 22
trans- [Mo(N2h(Ar2PCH2CH2P Ar2hl 19 47
(Ar = p-C 6 H 4CF 3 , p-C 6H 4Cl, p-C6H4Me, p-C6H 4OMe)
trans-[Mo(N 2h(dpeh] 20 20
trans-[Mo(N 2 h(Ph 2 PCH=CHPPh 2 h] 21 48
trans -[Mo(N 2h( dppe )(PMe2Phhl 22 49,50
trans- [Mo(N 2)( CO)( dppe h] 23 c 51
trans- [Mo(N 2h (PPh 2Me )4l 24 45
cis- [Mo(N 2h (PPhMe2)4] 25 45,52
trans -[Mo(N 2h(PEt 2Ph)4l 26 52
trans-[Mo(N 2 h(triphos)(PR 3 )] 27 50,53
[PR 3 = PPh 3, P(p-tolylh, P(p-C 6H 4OMehl
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 411

Table 1. (Continued)

Complex Text No. Reference

[M(N 2 )(COh(PCY3h] 28 54
(M=Mo,W)
[CI 4(MeO )Mo{ (N z)ReCI(PMe2Ph)4}] 29 a ,c 55,56 c
[(THF}zCI 3 Mo{(N 2)ReCl(PMe zPh)4}] 30 a 57
[(PMe zPh)CI4W{(N2)ReCl(PMe2Ph)4}] 31 a 57
[CI 4 M {(N 2)ReCl(PMe2Ph)4h] 32 a ,c 57,58 c
(M = Moe, W)

trans- [W (N 2h (dppe h] 33 20
trans- [W (N 2}z(PPh 2Me )4] 34 49
cis- [W(N2h(PMe2Ph)4] 35 20
trans-[W(N 2 h(depe}z] 36 47
[ReH(N 2)(dppe}z] 37 111
[Re(NHPh)(N 2)(PMe3)4] 38 112
[{Fe( 1)S -CsHs)(dppe)}z(Nz)](BF4h 39 115
[{Fe( 1)S -C s H s )(dmpe)}z(N 2)](BF4}z 40 116
[FeH 2(N 2)(PR 3h] 41 117-119
(PR 3 = PPh 3, PBu"Ph 2, PEtPh 2, PMePh 2)
[FeH(N 2)L]BPh4 42 120
[L = N(CH zCH 2PPh 2h]
[FeH(N 2 )L]y, 43 c 120,121 c
[L = P(CH zCH 2PPh 2h, Y = Br, I, BPh 4]
[FeH(N 2)(depeh]BPh 4 44 122
[FeH(Pri)(PPh3}z(N2)Fe(Pri)(PPh3)Z] 45 123
[{(depe}zCIFe}z(N z)][BPh 4h 46 124
[Fe(N 2)(dppe}z] 47 125
[RuH 2(N 2)(PR 3h] 48 126, 127
(R = Ph, p-C6 H 4CH 3)
[OsHCI(N 2 )(PR 3h] 49 128
(PR 3 = PEt2Ph, PMe2Ph, PPh 2Et)
[CoH(N 2)(PPh 3h] SOc 129
Na[Co(N 2 )(PR 3 h] 51 130
(PR 3 = PEt 2 Ph, PPh 3)

K[ Co(N Z)(PMe3h] 52 c 132


[{(PR 3 hCo}z(N 2 )] 53 130
(PR 3 = PPh 3 , PEt 2 Ph)
412 T ADRIAN GEORGE

Table 1. (Continued)

Complex Text No. Reference

[{ (PMe3h CorN 2) hMg(THF)4] 54 c 132


[Co(N 2)(P R 3hMg(THFh] 55 a 133
(PR 3 = PPh 3, PPh 2Et)

[RhCl(N 2)(PCY3h] 56 134


[RhH(N2)(PPhBu~h] 57 c 135
[RhCl(N 2)(PPr;h] 58' 136
[RhH(N 2)(PR 3h] 59 137-139
(R = Cy,Bu')

[{(PCY3hHRhh(N2)] 60 137-139
[{(PPr;hHRhh(N 2)] 61 c 138
[{(PCY3hNih(N 2)] 62 c 144
[Ni(N 2)(PR 3h] 63 b 142
(PR 3 = PEt 3, PBu~, PEt 2 Ph)

a Prepared using a preformed metal-dinitrogen complex.


b Quoted as being impure.
, Crystal structure determined.

An impure dinitrogen complex of niobium has been reported by Leigh


et al.(27) Complex 4 (see Table 1) is formulated as a dinuclear complex
with a bridging dinitrogen. Treatment with acid yields no ammonia, but
ca. 20% of the nitrogen is converted to hydrazine, the remainder being
evolved as N 2.
In 1980, Schrock et al. reported the first stable dinitrogen complexes
of a group 5 metal. (28-30) Of the more than ten complexes synthesized, two
are prepared directly using dinitrogen (Table 1, structures 5, 7), while the
remainder are prepared by a metathesis-like reaction of an alkylidene
complex with a diimine (Equation 9):

2[Ta(CHCMe3)(THFhCI3] + PhCH=NN=CHPh --. [(TaCh(THFhh(N 2)]


+ 2PhCH=CHCMe3 (9)

and further reactions thereof. Each of the complexes contains one


dinitrogen in a bridging position and a wide variety of other ligands.
X ray structure determinations of 6 and [{TaCh(THF)(PBz3 )}z(N 2)]-
(Bz=CH2Ph)(31,32) not only confirm a linear Ta-N-N-Ta unit but also reveal
very long N-N bonds (1.298(12) and 1.282(6) A, respectively) and
CA TAL YSIS OF NITROGEN-FIXING MODEL STUDIES 413

extremely short Ta-N bonds. The N-N bonds are the longest observed in
a simple bridging dinitrogen complex. The structural data and chemical
behavior (vide infra) of the bridging dinitrogen lead the authors to propose
a 1'-- N24- formalism (sp hybridized N atoms) with the linkage best described
as [Ta=N-N=Ta]. It appears that in these complexes prepared using
dinitrogen tantalum is able to bind and reduce dinitrogen. Addition of two
more electrons into this unit would produce 1'-- N 26- which of course would
not exist but would give two nitride ligands: [Ta=NN=Ta] +
2e- ~ 2[Ta_Nr. These electron-deficient complexes appear to contain
the most activated dinitrogen to date in a simple bridging system and should
renew peoples' interest in the preparation and chemistry of bridging
dinitrogen complexes.

2.3. Chromium, Molybdenum, and Tungsten

2.3.1. Chromium

Few dinitrogen complexes of chromium are known. Complex 8 decom-


poses at 20°C with loss of both N2 and PMe3.(33) Treatment of complex 9
with acid produces small quantities of ammonia (7%) and hydrazine
(1 %). (34) Reaction of the polynuclear complex 10 with water regenerates
the starting rhenium complex trans- [ReCI(N2)(PMe2Ph)4], and with dioxy-
gen produces the cation [ReCI(N2)(PMe2Ph)4t.(35)

2.3.2. Molybdenum and Tungsten

Following the discovery of trans-Mo(N 2 h(dppe)z, 15, by Hidai in


1969,(40) a series of four landmark reactions in the area of the chemistry
of coordinated dinitrogen were reported by Chatt and co-workers over a
period of three years. These are (i) the reaction of 33 with acyl halides to
form nitrogen-carbon bonds (Equation 10);(59) (ii) protonation of 15 and
33 to form nitrogen-hydrogen bonds (Equation 11);(60) (iii) alkylation of
15 and 33 to form nitrogen-carbon bonds using alkyl halides (Equation
12);(61) and (iv) protonolysis to produce ammonia (and hydrazine)
(Equation 13)(62):

H
Hel I
[W(N2lz(dppeh] + RCOCI ~ [WCI(NNCOR)(dppeh]CI + N2 (10)
[Mo(N 2h(dppeh] ~ [MoBr(NNH 2)(dppeh]Br + N2 (11)
[Mo(Nzh(dppeh] + RBr .... [MoBr(NNR)(dppeh] + N2 (12)
414 T. ADRIAN GEORGE

This latter reaction only occurs when there is at least one monodentate
phosphine coordinated to the metal. For the purpose of discussion, the
remainder of this section will be divided into reactions that (i) form
nitrogen-carbon bonds; (ii) form nitrogen-hydrogen bonds; and (iii) reac-
tions of polynuclear complexes.
2.3.2a. Formation of Nitrogen-Carbon Bonds. A wide variety of
organic halides react directly with the bisdinitrogen complexes of molyb-
denum and tungsten. However, nitrogen-carbon bonds are only formed
when the phosphine ligands are bidentate. In all the other complexes that
contain one or more monodentate phosphines both dinitrogen ligands are
lost and phosphine-metal-halide complexes are formed.
Successful reactions of complexes 15 and 33, some occurring in very
high yield, with various organic halides (Equations 14-19) will be presented
in approximately chronological order. Acyl and aroyl chlorides react to
form hydrazido and diazenido complexes, respectively. The latter com-
plexes were treated with one mole of hydrogen chloride to form the
hydrazido complexes. (63) The hydrazido complexes can be deprotonated
reversibly to give the corresponding diazenido complexes (Equation 14):

[M(N 2h(dppehl + RCOCI + HCI ~ [MCI(NNHCOR)(dppehlCI + N2

HCI1l base
[MCI(NNCOR) (dppehl (14)

The X ray structure of [MoCI(NNCOPh) (dppe)ZJ confirms the presence


of a nitrogen-carbon bond. (64) Important bond angles (0) and lengths (A)
are shown here:
172.1 (6) 116.7(7)
Mo ______~~~____~~_______R
1.813 (7) 1.255 (10)

Simple alkyl bromides and iodides react at room temperature to give


alkyldiazenido complexes (Equation 15): (63,65-67)

[M(N 2h(dppelzl + RX ~ [MX(NNR)(dppelzl + N2 (15)


(M = Mo, W; R = Alkyl; X = Br, I)

although they are often isolated as the corresponding alkylhydrazido com-


plex and subsequently converted to the alkyldiazenido complex by
deprotonation. A number of X ray structure determinations have estab-
lished the bonding mode for both diazenido(65,68,69) and hydrazido(7o,71)
complexes of molybdenum and tungsten. The metal-nitrogen bond
becomes progressively shorter and the nitrogen-nitrogen bond longer as
CA TAL YSIS OF NITROGEN-FIXING MODEL STUDIES 415

more atoms are attached to the end nitrogen. [Mo(N 2h(dppeh],(41)


[MoI(NNOcn(dppeh],(69) and [MoI(NNHOcn(dppe)z]I(70) are com-
pared below:

R R
/ /
(N 2 )Mo-N-N IMo-N-N IMo-N-N

'" H

Mo-N(A) 2.014(5) 1.850(12) 1.801(5)


N-N(A) 1.118(8) 1.146(13) 1.259(8)
NNR(O) 128(1) 120(1)

Normally alkyl chlorides and aryl halides do not give nitrogen-containing


products, though "activated" alkyl chlorides such as CICH2COOEt(72) and
CICOOEt(73) (Equation 16) do:

[Mo](N 2 h(dppeh] + RCI ~ [MoCI(NNR)(dppeh] + N2 (16)


(R = CH 2 COOEt, COOEt)

n - Butyl chloride has been reported to react with 33 in benzene solution


under irradiation at 45°C to give the corresponding butyldiazenido com-
plex. (74) Reaction of racemic 2-bromooctane with the pure diasteriomer
trans-[Mo(N 2h{(s,s)-chiraphosh], where s,s-chiraphos is (- )-(2s, 3s)-
bis (diphenylphosphino )butane, produces the 2-octyldiazenido complex,
which is a mixture of two diasteriomers, in 78% yield.(75) The proton nmr
spectra recorded at 100, 200, and 396 MHz show an excess (ca. 10%) of
one diasteriomer over the other.
gem- Dibromoalkanes react with 33 to give a novel series of cationic
diazoalkane complexes (Equation 17)(76):

that react with hydride ion and carbanions to form the corresponding
neutral alkyldiazenido complexes. On the other hand, the reactions of 15
with gem-dibromides are more complex, and diazoalkane complexes are,
at best, minor products.(76) For example, CH2Br2 reacts with 15 to give a
complex formulated as trans-,trans-[(dppehBrMo(NNHCH 2 NHN)-
MoBr(dppe)z]Br2. When the mother liquor from this reaction is treated
with HBr, a further complex formulated as trans-[MoBr-
(NNHCH 2CH 2Br)(dppeh]Br is isolated. The X ray crystal structures
of three diazo alkane complexes have been determined:
[WBr(NN =CHCH 2CH 2 CH 2 0H)(dppe h]PF 6,(77) [WBr(NN =CMe2)-
416 T. ADRIAN GEORGE

+
[(dppehBrMNN(CH2)nNNMBr(dppeh]

Scheme 1

(dppe)z]Br,(77) and [WF(NN=CMeCH 2COMe)(dppe)z]BF4.(78) The W-


N-N units are linear, with N-N-C angles of 116-125° and N-C bond
lengths that are consistent with a bond order of two.
a, w- Dibromoalkanes, Br(CH 2)n Br, undergo a variety of reactions with
15 and 33 depending upon the value of n. (79) The case where n = 1 has
been discussed above and for n = 2, ethene is eliminated and [MBr2(dppe)z]
formed. For n = 3, complexes containing the group NN(CH 2hBr are
formed, and for n = 4 or 5, complexes with rings NN(CH 2 )n are formed.
For n = 6-12, two series of complexes containing either NN(CH2)nNN in
which the alkyl group bridges two molecules of substrate, or NN(CH2)nBr
are isolated. These reactions are summarized in Scheme 1. The dinuclear
species are isolated as either hydrazido or diazenido complexes. Difficulties
are experienced in working up the products when M = Mo and n = 6-12.
Interestingly, I-bromo-4-chlorobutane also cyclizes with 15 to produce
[MoBr(NNCHzCHzCHz{':Hz)(dppe)z]CI in which bromine is attached to
molybdenum, and chloride is the anion. (80) Recently, the secondary alkyla-
tion of alkyldiazenido complexes to form dialkylhydrazido complexes has
been reported.(47,81l However, when 20 is treated with MeBr or EtBr, no
alkyldiazenido complex is isolated, but instead a dialkylhydrazido complex
is produced (Equation 18):
[Mo(N 2h(depeh] + 2MeI -. [MoI(NNMe2)(depeh]I + N2 (18)

The alkylation of trans-[M(Nzh(dppe)z], (M = Mo or W) by primary


alkyl bromides and iodides has been studied in considerable detail and a
mechanism proposed for the reactions observed. (74,82) The mechanism is
summarized in Scheme 2. The initial rate-determining dissociation of one
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 417

M(N 2lz(dppeh ~ M(N 2)(dppeh + N2


lRX
[(RX)M(N 2)(dppehl
! homoloysis

[XM(N 2)(dppehJ + R·
THF / ,
R ~ My II) '\.nzene
' " hPO ~~
Iii)

[MX(NNR l)(dppehl [MX(NNR)(dppehl ![MX2(dppehl


+
![M(N2h(dppehl

(i) R 1 = CHOCH 2 CH/:H 2 ; (ii) R· couples or disproportionates

Scheme 2

dinitrogen ligand is followed by formation of an adduct between the


coordinatively unsaturated metal species and the alkyl halide, [M(N 2 )
(RX)(dppe)z]. Following carbon-halogen bond homolysis (inner-sphere
electron transfer), the halogen atom remains bonded to the metal and the
organic radical either (i) attacks the remaining coordinated dinitrogen to
form an alkyldiazenido complex, (ii) attacks the solvent (e.g., THF when
R = Me) to produce a solvent-derived radical ~hich then attacks the
dinitrogen ligand to form a different diazenido complex, or (iii) dimerizes
(e.g., R = PhCH 2) or disproportionates. The initial dissociation of
dinitrogen occurs in the dark for 15 at room temperature, but for 33 a
source of visible light is necessary (without which no reaction occurs at
room temperature). The presence of light accelerates the alkylation reac-
tions of 15. A number of flash photolysis studies of 15 and 33 have been
carried out(83.84) and lend support to the above mechanism.
It is assumed that all of the other nitrogen-carbon bond-forming
reactions mentioned so far proceed according to Scheme 2 with loss of one
dinitrogen being the initial step.(74.82) It can be argued that in the 5-
coordinate intermediate, [M(N 2)(dppe)z], the remaining dinitrogen will
be more tightly held due to increased 'TT'- back bonding from the metal and
therefore less likely to be lost from the coordination sphere.
Dinitrogen can also be alkylated in complexes containing only one
dinitrogen and a labile ligand. A series of complexes of the type trans-
[M(N 2)(NCR)(dppe)z], (M = Mo or W), have been prepared(8S) by the
reaction(86) of an organic nitrile with 15 or 33. Subsequent reactions with
alkyl bromides and iodides proceed with slow loss of RCN (slower than
418 T. ADRIAN GEORGE

loss of N2 from 15 or 33) and formation of an alkyldiazenido complex


(Equation 19/82 ):

[M(Nz)(NCR)(dppeh] + RX --+ [MX(NNR)(dppeh] + RCN (19)

In these reactions loss of RCN creates the 5-coordinate intermediate that


then follows the inner-sphere redox pathway depicted in Scheme 2 to form
product.
The presence of a labile ligand is not an absolute requirement for
alkylation of coordinated dinitrogen. Thus reactions of 15 with a series of
tetrabutylammonium halides and pseudohalides produce a number of
anionic monodinitrogen complexes (Equation 20):

[Mo(Nzh(dppeh] + [Bu~N]X --+ [Bu~N][MoX(Nz)(dppeh] + N z (20)


(X = SCN, CN, N3 )

some too unstable to isolate.(82) Interestingly, reaction of [Mo(SCN)(N 2)-


(dppe)zr with Bunl shows a rate which is first-order in complex concentra-
tion and first-order in Bunl concentration. In addition, the product retained
the thiocyanate ligand. These data are inconsistent with rate-controlling
ligand loss. Instead, it is proposed that the reaction proceeds by an outer-
sphere electron-transfer reaction (Equation 21)(82):
[Mo(SCN)(Nz)(dppehr+RX --+ [Mo(SCN)(Nz)(dppeh]+R'+X- (21)

The radical R· couples with the molybdenum(I) radical to form product,


[Mo(SCN)(NNR)(dppe)z].
Those reactions that give dialkylated-dinitrogen products, e.g.,
Equation (18) and Scheme 1 (n = 4 or 5) are believed to proceed in two
steps. The first is mono alkylation according to Scheme 2, and the second
is SN2 displacement at a halogen-bound carbon atom by the carbon-bound
nitrogen atom of the diazenido ligand. Typical SN2 reactions have been
demonstrated for the reactions of Mel and EtI with preformed methyl-
diazenido complexes [MBr(NNMe)(dppe)z].(47) Complex 20, which con-
tains the basic depe ligand, reacts with Mel to form the dimethylhydrazido
complex (Equation 18). It is suggested that in this case the basic depe has
the effect of (i) slowing down the rate of loss of dinitrogen, and (ii) increasing
the nudeophilicity of the NNMe group with the net result that second
alkylation occurs faster than first alkylation.
So far in this section all reactions have resulted in organonitrogen
ligands that are firmly attached to the metal. Do subsequent reactions of
these species form amines, for example? The answer is yes, but only in
reactions in which the integrity of the metal complex is destroyed. There
CA TAL YSIS OF NITROGEN-FIXING MODEL STUDIES 419

is no evidence for protonation of alkyldiazenido complexes beyond the


alkylhydrazido form. Reactions of tungsten diazoalkane complexes with
HBr cause protonolysis of the nitrogen-carbon bond to form
[WBr(NNH 2)(dppeh]Br with no evidence for organonitrogen com-
pounds. (87) The reaction of a tenfold excess of sodium borohydride with
[MoBr(NNBu")(dppe)z] in benzene-methanol solution at lOO°C for 10 h
in an autoclave produces NH3 (51 %), n- butylamine (54%), and butyl-
methylamine (9%).(66,88) Under the reaction conditions, sodium borohy-
dride is completely converted to Na[B(OMe)4] before amine formation
occurs. The molybdenum complex is completely degraded during the reac-
tion and ca. 90% dppe is recovered as free ligand. It is proposed that
butylmethylamine arises from the reaction of butylamine with formal-
dehyde that is generated from methoxide ion (or methanol) during the
reaction. When ethanol, rather than methanol, is used as co-solvent, only
small amounts of volatile base are obtained. However, MoH4(dppeh is
isolated from the reaction in 25% yield.
Amines have also been generated from dialkylhydrazido complexes
by a variety of destructive methods involving both strong acid and strong
base. (89 ) For example, treatment of [WBr(N2Me2)(dppehr with LiAlH4
in ether followed by base distillation yielded half the available nitrogen as
dimethylamine (0.95 mol), but the other one equivalent is unaccounted
for. In a few cases tungsten-containing products are isolated; e.g.,
[WH 4(dppeh] and [WBr2H2(dppe)z].
It is not immediately obvious how any of the nitrogen-carbon bond-
making reactions discussed in this section could be adapted to provide a
catalytic reaction for the formation of organonitrogen compounds. A prere-
quisite for any catalytic nitrogen-fixing process is the continual regeneration
of a dinitrogen-binding site. In fact, this has been accomplished re-
cently in a report of the electrochemical reduction of trans-
[MoBr{NNCH 2(CH 2 bCH 2}(dppehr at a platinum electrode in THF-
0.2 M [Bu4N][BF4] under dinitrogen that yields free N-aminopiperidine,
" (90)
H 2 NNCH 2 (CH 2 bCH 2 , and trans-Mo(N2h(dppeb. Under carbon
monoxide, the reaction proceeds similarly with cis- and trans-
[Mo(COh(dppeh] being formed. The starting dialkylhydrazido complex
is prepared in a one-step reaction from 15 (see Scheme 1, n = 4 or 5), and
the authors propose a nitrogen-fixing cycle based upon these initial results,
which is summarized in Scheme 3. Solvent is presumed to be the proton
source. This work provides the most encouraging results toward the
development of a catalytic process for the synthesis of organonitrogen
compounds reported to date.
2.3.2b. Formation of Nitrogen-Hydrogen Bonds. The bisdinitrogen
complexes of molybdenum and tungsten that contain two bidentate ligands
such as 15, 20, and 33 react with strong acid to produce hydrazido complexes
420 T. ADRIAN GEORGE

Scheme 3

with the loss of one mol of dinitrogen (e.g., Equation 11). (20,60,91) Ammonia
and hydrazine are not products in these reactions. By contrast, those
complexes that have at least one monodentate phosphine react with strong
acid to yield an array of products that may include ammonia and/or
hydrazine depending upon (i) the acid, (ii) the solvent, and (iii) the metal
(624992-97) D'IscusslOn
(M o or W) ." . WI'11 focus pnman
" 1 y upon t hose reactions
.
of bisdinitrogen complexes that produce ammonia and/or hydrazine.
A clear distinction between the stoichiometry of the ammonia-forming
reactions of molybdenum and tungsten was noted from the outset by Chatt
and co-workers. (62) The yield of ammonia is essentially 2 mol per tungsten
atom but only ca. 0.7 mol per molybdenum atom. Typically, complexes
.24, 25, 34, and 35 are treated with an excess of H 2 S04 in methanol at
20°C. One mol of dinitrogen is rapidly evolved. Beyond this point, the
reaction pathway differs depending upon whether the metal is tungsten or
molybdenum. The reaction is shown for complex 35 in Equation 22.

(22)
By varying the acid and particularly the solvent (e.g., THF, benzene,
N-methyl-pyrrolidone, or dichloromethane), increasing yields of hydrazine
are obtained at the expense of ammonia: 0.6 mol of N2H4 per tungsten
atom for 35/HCl/DME.(94)
Richards and co-workers have followed the reaction of excess H 2S0 4
with 35 in THF solution by 15N nmr spectroscopy. (96) Following the disap-
pearance of 35, resonances due to 15N2 and a series of six hydrazido (NNH 2)
complexes are observed in solution. Finally, after 120 min, a signal due to
15NH4 + was the only resonance remaining. An analogous experiment with
25 shows the appearance of resonances due to 15N2 and two hydrazido
complexes but no 15NH4 +. These and other data clearly indicate that the
hydrazido unit, NNH 2, is an intermediate formed during the reduction of
coordinated dinitrogen to ammonia. In the case of the tungsten complex
35, cited above, the hydrazido group remains intact through a series of
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 421

changes in the coordination environment around tungsten before ammonia


formation. No other nitrogen hydride intermediates are isolated or detected
spectroscopically.
Treatment of 34 with HCl in dichloromethane(93) affords a complex
formulated(I00) as a hydrazido-tungsten hydride, [WHCh(NNH 2)-
(PMePh 2h). Similar complexes are prepared by treating [WX 2 (NNH 2 )-
(PMe2Phh) (X = Cl or Br) with HX in DME(94) or THF.(98) Although
originally formulated as complexes containing an NHNH2 group, (93,94)
subsequent 15 N (98) and 1H (99) NMR studies strongly suggest that the
hydrazido-tungsten hydride formulation is correct. Upon further treatment
with acid, these complexes produce ammonia and hydrazine; the latter in
yields significantly greater than those obtained when the original complexes,
34 and 35, are treated with acid. (93,94) Hidai proposes that the {WH(NNH2)}
stage of reduction is on the route to hydrazine formation. (99)
Returning now to the importance of the hydrazido ligand as an inter-
mediate in dinitrogen reduction, it is noted that hydrazido complexes can
be isolated from many reactions that ultimately give ammonia if allowed
.
to continue, e.g" E '
quatlOns 23 - 25(969320)
":

[M(N) (PMe2Ph)4] (i) H2S~4/THF/30min I [M(HS04h(NNH2)(PMe2Phh]


2 2 (11) pentane
+ N2 + [HPMe2Ph]HS04 (23)

[M(N2lz(PMe2Ph)4] HX/MeOH I [MX2(NNH2)(PMe2Phh] + N2 + [HPMe2Ph]X (24)

The X ray structure of a number of hydrazido complexes of molybdenum


and tungsten have been determined: [WBr(NNH 2)(dppeh][BPh4],(100)
[MoF(NNH2)(dppe h][BF4], (91) [WHBrCI(NNH2)(PMe2Phh][BPh4], (94)
[WCh(NNH2)(PMe2Phh], (98) and [M (NNH 2 )(quin )(PMe2Phh]x,
where quin = quinolin-8-olate; M = Mo, X = Br, I; M = W, X = I.(101)
In all these complexes the metal-nitrogen bonds are very short, indicating
considerable multiple bonding. The bonding data for the NNH2 unit are
similar in all the above complexes despite the variation in coligands. The
N-N bond lengths indicate an order greater than one with varying degrees
of noncoplanarity of the atoms in the hydrazido group, NNH 2. These data
support the following bonding representation [M" ·N:..:....:NH 2) in which there
is considerable de localization within the unit.
Hydrazido complexes with two bidentate phosphine ligands can be
deprotonated reversibly to give the corresponding neutral diazenido com-
plex (Equation 26)(102):

[MX(NNH 2)(dppeh]X [MX(NNH)(dppelz] (26)


422 T. ADRIAN GEORGE

+ P +2
P
p, p,
I .. NzH z
"'k"
.. NzH J:I.:...
"M" ~

P~ I "N z P~ I "N z
P P

1l
P +2
p, " I ... NzH z
Nz + 'M'
p~1
P

1l MeOH

p +
p, .. NzH z
H++ "'k"
p"""l "OMe
p

II
P +
P" I .. N zH 2
''M''
P~ "OMe

a P = PMezPh

Scheme 4 8

Recently, the first mechanistic study of the conversion of a coordinated


dinitrogen to a hydrazido complex was reported.(103) Treatment of cis-
[M(N2h(PMe2Ph)4J with excess Hel, HBr, or H 2S0 4 in methanol yields
[M(NNH2)(OMeh(PMe2PhhJ in a reaction that would ultimately give
ammonia. The stoichiometry of the reaction is shown in Equation (27):

[M(N zh(PMe zPh)4] + H+ + 2CH 30H --+

[M(NNHz)(OCH3h(PMe2Phh] + N z + [HPMezPht (27)

The kinetics of the reaction exhibit a first-order dependence in complex


concentration, a second-order dependence in total acid concentration, and
no dependence upon the type of anion. A mechanism is proposed to fit
the data, and it is shown in Scheme 4 (P=PMe2Ph). Loss of phosphine is
suggested as the rate-determining step. It is methoxide ion rather than the
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 423

anion derived from the "strong acid" that coordinates to the metal. Note
also that methanol effectively provides both protons in the reaction and
the added acid scavenges the liberated phosphine. It has been shown that
in methanol alone either when heated to reflux or irradiated with visible
light the tungsten complexes 34 and 35 yield NH3 in high yield. (92) This
does not occur for the molybdenum complexes 24 and 25. The greater
basicity of dinitrogen attached to tungsten is shown by the greater reactivity
of 35 compared with 25; kw/kMo = 9.2 X 102. An isotope effect observed
in the reaction of 25 (kH/ ko = 0.3) is consistent with a mechanism involving
protolytic-equilibria prior to the rate-limiting step. (103)
It seems clear that in reactions of the tungsten complexes of dinitrogen
that yield ammonia, reduction is occurring at one metal center that is also
the source of the required six electrons. Hence, tungsten is oxidized from
W(O) to W(VI), although no tungsten (VI)-containing product has yet been
isolated. This overall process is shown in Scheme 5.(92) An important part
of this scheme that is omitted is the incorporation of the necessary electron-
releasing ligands such as MeO- and HS0 4- (S04 -2) into the coordination
sphere to assist tungsten in reducing coordinated dinitrogen.(92) Unless
there are labile ligands already attached to tungsten (such as N2 and PR 3 ),
the reaction stops at some intermediate step, e.g., [WBr(NNH 2)(dppeh]Br.
This latter complex can be converted to ammonia (0.40 mol) and hydrazine
(0.44 mol) under conditions (HBr/CH 2Ch/80°C/15 h) vigorous enough to
displace at least one dppe ligand. (87)
In those reactions in which hydrazine is formed, tungsten is probably
behaving as a four-electron reducing agent. Hydrazine may result if the
hydrazido intermediate is converted into a side-on bonded NH-NH2 species
in the presence of acid. Complexes of this type (e.g., [MO('1/2 -NHNMePh)-
(NNMePh)(S2CNMe2ht(104)) have been shown to yield organohydrazines
quantitatively upon treatment with acid. An alternate source of hydrazine
is from the decomposition of N 2H 2. However, mechanistic studies are
required before anything definitive can be said about the pathway to
hydrazine formation.
The identification and isolation of a molybdenum(III) complex,
MoBr3(triphos), in almost quantitative yield (>90%) by the author and

Scheme 5
424 T. ADRIAN GEORGE

co-worker from the reaction of trans-[Mo(N 2 h(triphos)(PR3 )], 27, with a


large excess of anhydrous HBr in THF (Equation 28)(95)

2[Mo(N 2 h(triphos)(PR 3)] :;~ 2MoBr3(triphos) + 2[R 3PH]Br + 3N 2 + 2[NH4 ]Br (28)

underscores the difference in behavior between molybdenum and tungsten


in these types of reactions. Whereas the tungsten complexes liberate one
mol of N2 and two mol of NH3 per metal atom, complexes of molybdenum
yielded 1.5 mol of N2 and 1 mol of NH3 per metal atom. The latter
stoichiometry has been observed for all molybdenum complexes containing
at least one monodentate phosphine; 22,(47,97) 24,(49) 25,(49) and 27.(95)
From three of these complexes, a molybdenum(III) product has been
isolated.(95,97,105) The reaction stoichiometry indisputably establishes the
formation of one NH3 per molybdenum atom, a net transfer of 3 electrons
per molybdenum atom [Mo(O) --+ Mo(III)], and loss of 1.5 mol of N2 per
molybdenum atom.
In general, the reactions of molybdenum complexes result in a poor
nitrogen balance. Although 1.5 mol of N2 are liberated routinely, the
ammonia yields have tended to be around 70%. There is a report of 88%
NH3 from an experiment involving 25 and aqueous HBr in propylene
carbonate for 48 hr followed by solvent removal at 140°C in vacuo and
base distillation. (49) The low yields of ammonia appear to be artifacts of
the experimental conditions. The molybdenum reactions are considerably
slower than their tungsten analogs. During the course of these reactions,
acid reacts with the solvent as well as the complex so that before the
ammonia-forming reaction is complete the acid concentration is effectively
zero. Recently, it has been shown that either addition of increments of
HBr during the reaction, or addition of acetic acid and LiBr after the loss
of the first mol of N2, provides conditions under which the nitrogen balance
. . . (106)
IS quantitative.
Bisdinitrogen complexes of molybdenum react rapidly with strong
acids to liberate 1.0 mol of N2 per molybdenum atom, totally analogous
to the similar tungsten complexes. The product at this stage is a hydrazido
complex. The first step appears to require a certain minimum effective
concentration of acid, otherwise protonation of the metal occurs, which
results in loss of all dinitrogen from the molybdenum complex.(20,105,106)
The presence of a hydrazido complex after the loss of 1.0 mol of dinitrogen
is clearly demonstrated by isolation and characterization of the complex
and by in situ 15N NMR spectral studies.(92,96) A recent study using 31p
nmr spectroscopy to follow the reaction of HBr with 27 in THF reveals
the presence of two isomeric hydrazido complexes, shown below. The
differences arise in the relative positions of Br and PPh 3. Isomer A isomer-
izes to B which reacts further. A low concentration of a non-PPh r
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 425

containing complex is also observed. This is believed to be


[MoBr2(NNH2)(triphos)]. Eventually, an intense resonance due to
PPh 3 is observed as the spectral lines begin to broaden after 1 hr. (106)

<:riO Zp/i
z\
N 2H 2+

"p
~c - <f~O
I I
Br PPh 3
A B

How are the hydrazido complexes of molybdenum converted to


ammonia? First, the stoichiometry of this latter reaction must be
reexamined (Equation 29), since it implies N-N bond cleavage to

(29)

generate one [NH4t and !N2. That this does not occur as written is proved
by an isotopic-labeling experiment that establishes that no nitrogen atom
scrambling occurs. (95) Specifically, the reaction of a mixture of 27 -C 4 N 2h
and 27-C 5 N2h and HBr in THF produces dinitrogen-28 and dinitrogen-30
but no dinitrogen-29. Therefore, one-half of the {Mo(NNHx )} units are
being oxidized to dinitrogen (giving 0.5 mol of N2 per Mo atom) and the
other half of the {Mo(NNHx )} units are being reduced to ammonia (to give
1 mol of NHt per Mo atom) and therefore Equation 29 should be rewritten
as shown in Equation 30.

2[MoBr(NNH 2)(triphos)(PPh 3)t ~


2MoBr3(triphos) + N2 + 2[NH 4 t + 2[HPPh 3 ]Br (30)

Secondly, since there is no nitrogen atom scrambling, there is a net require-


ment of 6 electrons per two molybdenum atoms to accomplish the reduction
of a dinitrogen to 2[NH4t. This is summarized in Equation 31.

Assuming that a mononuclear hydrazido complex is reduced to 2[NH 4 t,


then intermolecular electron transfer from another molybdenum complex
is necessary to account for all molybdenum appearing in the product as
molybdenum(III). Intermolecular electron transfer can be ruled out,
426 T. ADRIAN GEORGE

however, for two reasons: (i) Electrochemical studies of a variety of


hydrazido complexes that will produce ammonia in the presence of acid
show no reduction waves above that of the solvent (CH 2Ch) discharge;(92)
(ii) Protonolysis of a mixture of 15 and 27 produces no increase in the
yield of [NH4t; i.e., [MoBr(NNH2h(dppeht, which is formed in this
reaction and does not afford [NH4t, under these conditions could still
behave as reducing agents towards a Mo(NNHx )(triphos) intermediate
if intermolecular electron transfer is a requirement for ammonia
formation. (106)
Thirdly, the necessary requirement that one of the ligands of the
hydrazido complex be labile in order that the reaction proceeds to generate
a further 0.5 mol of N2 and 2[NH4t can be satisfied by Br- or PPh 3 in
the triphos system. The dissociation of PPh 3 from isomer B being the more
important step is supported by the following data: (i) The fluoride complex,
[MoF(NNH2)(triphos)(PPh3 )t, in which the fluorine atom is trans to
NNH 2, upon treatment with anhydrous HBr in THF for 6 days produces
no ammonia. The strong Mo-F bond precludes facile F- dissociation. (ii)
The addition of PMe2Ph to a THF solution that contained A and B showed

-8,
-PPh 3

- PPh a

Scheme 6
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 427

incorporation of PMe 2Ph into B (but not A) and no appreciable amount


of ammonia was produced. (106)
A proposed reaction pathway by which ammonia is formed is shown
in Scheme 6. After the dissociation of phosphine, attack by another
hydrazido complex (e.g., A) to form a bridged species, C, is proposed.
Hydrazido complexes such as A are known to be good nucleophiles. For
example, they react in acid catalyzed reactions with aldehydes and ketones
to form diazoalkane complexes (e.g., Equations 4 and 32(107)).

[MoF(NNH 2 )(triphos)(PPh 3 lt + [(7]s -C s Hs)Fe(7]s -CSH 4 CHO)] ~


[(7]s -C s Hs)Fe{7]s -CSH 4 CHNNMoF(triphos)(PPh 3 )}t + H 2 0 (32)

The driving force for this step must be the formation of a strong Mo-N
bond. The bridging ligand in the proposed intermediates C, D, and E may
be of the diimine (NHNH) type, similar to those isolated and characterized
by Sellmann. (108) Formation of an intermediate with a bridging [N 2H x ] unit
and a terminal [N2HX] unit is consistent with the isotopic-labeling experi-
ment. No firm evidence is available to say whether it is the terminal or
bridging [N 2Hx] unit that is oxidized to give dinitrogen; 0.5 mol per molyb-
denum atom.
The formation of bridging nitrogen-hydride ligands as intermediates
on route to the reduction of dinitrogen to ammonia and hydrazine has
been suggested by many authors.(7,S7,109) The unique feature of this molyb-
denum system is the pathway by which the bridge is constructed. Loss of
dinitrogen from an intermediate such as D followed by N-N bond homolysis
leads eventually to a molybdenum(III) amide F that should be readily
converted to ammonia in the presence of acid. In no step is intermolecular
electron transfer a requirement.
In principle, the molybdenum system could be developed into a
catalytic cycle for ammonia formation (see Scheme 7). The molyb-
denum(III) product is the normal starting point for synthesis of 27 in a
one-step, high-yield (>60%) reaction. An attractive feature of the system
is the stability introduced by the tridentate ligand that helps to retain the
structural integrity of the coordination environment of the metal. The
major drawback of the proposed cycle is the difficulty of regenerating the
N 2-binding site without reducing protons to dihydrogen.
Treatment of 35 with alcohols in the presence of KOH at 50°C produces
ammonia and hydrazine in moderate yields. (99) Ammonia is the major
product with all alcohols studied except for 2-propanol in which hydrazine
is predominant. The molybdenum analog 25 gives no ammonia or hydrazine
under these conditions. Hidai et at. have also carried out similar reactions
with metal hydrides. Hydridic hydrides such as [(71 s -CsHshZrHCI] and
428 T. ADRIAN GEORGE

-L

Scheme 7

[NaAlH 2(OCH 2CH 20CH 3h] produce ammonia in moderate to high yield
when reacted with 25 and 35 in benzene solution. (99)
2.3.2c. Reactions of Polynuclear Complexes. A number of polynuclear
complexes of molybdenum and tungsten are known in which dinitrogen
bridges two metals: e.g., [Mo(N 2)Mo], 12 and 13; [Mo(N 2)Fe], 14;
[Mo(N2)Re], 29, 30, and 32; and [W(N2)Re], 31 and 32. Of those that
have been structurally characterized by X ray diffraction studies (13, 29,
and 32 (M = Mo», none shows significant lengthening of the nitrogen-
nitrogen bond compared with that in gaseous dinitrogen and dinitrogen
coordinated in mononuclear complexes. This is to be contrasted with the
long nitrogen-nitrogen bonds observed in dinuclear tantalum complexes,
e.g., 6.
There is very little published data on the reactivity of this class of
complex. The molybdenum complex 12 does not produce any nitrogen-
containing products upon reaction with H 20, LiAIH4' or BuLi. (37) Reaction
of 13 with HBF4 causes protonation of the metal to form [{HMo(1/6-
C6H3Me3)(dmpe)}z(N2)][BF4Jz, which upon reaction with dilute HCI or
with H2S04/FeS04 produces no ammonia. (38)
Some of these complexes appear to be good candidates for chemical
investigation. For example, 29 has a very low v(N 2 ) at 1660 cm- 1 and in
the X ray photoelectric spectrum the N(ls) binding energy of 398.6 eV
(the two nitrogen atom peaks are not resolved) is similar to those for 15
(399.6 and 398.6 eV), a complex that can be both alkylated and proton-
ated, and the rhenium precursor [ReCI(N2)(PMe2Ph)4] (400.1 and
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 429

398.4 eV)Y 19 ) A short nitrogen-nitrogen bond does not mean that there
is no activation of coordinated dinitrogen; witness the chemistry of 15
(N-N = 1.118(8) A, cf. 29, N-N = 1.18(3) A and N 2, N-N = 1.0976(1) A).

2.4. Manganese, Technetium, and Rhenium


No reactions of coordinated dinitrogen have been reported for com-
plexes 37 and 38. However, mention should be made of trans-[ReCI(N 2 )-
(PMe2Ph)4] and its analogs even though they are not normally prepared
directly from dinitrogen. (113) Dinitrogen coordinates particularly strongly
to rhenium(I) as evidenced by the most extensive series of dinitrogen
complexes known for any metal to date. (18) A few examples are given
including the only stable complexes containing both coordinated dinitrogen
and sulfur: [ReCI(N 2 )(PR 3 )4](PR 3 = PMe2Ph, PPh 2Me, Me2NPF2,
PPh 2H, P(CH 20Meh), [ReCI(N2){Ph 2P(CH2)nPPh 2}] (n = 1,2,3),
[ReCI(CO)z(N 2)(PPh 3)z], [ReCI(N2)L(PMe2Phh] (L = py, 3-picoline,4-
picoline,pyrazine), and [ReL(N2)(PMe2Phh] (L = S2CNMe2, S2CNEt2,
S2COEt, S2CPPh 2).
The electron-richness of the rhenium complexes [ReCI(N2)(PR 3)4] is
manifest in their relatively low v(N 2 ) of 1920-1980 em-I, oxidation poten-
tials of 0.00 to 0.30 volts (vs. saturated calomel electrode), and ability to
behave as Lewis bases via the noncoordinated nitrogen atom toward Lewis
acids such as AIMe3, CrCh(THFh, and other metal and nonmetal
halides. (18) In fact, the basicity of this coordinated dinitrogen is similar to
that of THF towards AIMe3, when PR 3 = PMe2Ph, and greater than for
the other dinitrogen complexes studiedY 14 ) The polynuclear complexes
containing dinitrogen bridged to titanium and molybdenum have already
been discussed.
Acids, acyl, and aroyl halides all react with [ReCI(N2)(PMe2Ph)4].
However, protonation occurs at the metal to give the hydride [ReCIH(N2)-
(PMe2Ph)4t, whereas slow acylation and aroylation occur at the end
nitrogen atom. Interestingly, this latter reaction is the reverse of the
preparation of the rhenium dinitrogen complexes. (18) Alkyl halides do not
react with rhenium complexes of dinitrogen.

2.5. Iron, Ruthenium, and Osmium


Although a large number of dinitrogen complexes of these metals have
been reported (many containing ligands other than phosphines), only a few
reactions have been reported. Complexes 39 and 40 react with LiAIH4'
NaBH4, and acids to produce the corresponding hydride, e.g., [('1/ 5_
C5H5)FeH(dmpe)] with loss of N2.(115,116) Complex 45 is prepared by the
reduction of [FeCh(PPh 3)ZJ with PriMgBr in ether at -50°C under
430 T. ADRIAN GEORGE

dinitrogen. A band at 1761 cm -1 is attributed to v (N 2) and indicates a N-N


bond of low-bond order. Treatment of the complex with HCl in ether at
-50°C gave 10% hydrazine (based upon complexed N 2). (123)

2.6. Cobalt, Rhodium, and Iridium

Complex 50 is one of the most thoroughly studied dinitrogen com-


plexes. Reaction with hydrohalic acids leads to loss of dinitrogen and
dihydrogen.(140) Similar results occur with rhodium and iridium dinitrogen
complexes. Complex 55 is prepared by the reaction of 50 with MgEt2 in
THF at O°C and has v(N 2 ) at 1840 cm- 1 (cf., v(N 2 ) at 2088 cm- 1 for 50).
Complex 55 reacts with water to regenerate 50 and acids to give hydrazine
and a trace of ammonia. With concentrated sulfuric acid, 0.29 mol of
hydrazine and 0.03 mol of ammonia per mol of complex and dinitrogen
are liberated.(133) In this complex, it is not known whether dinitrogen is
bridging between cobalt and magnesium as it does in complex 54 or
nonbridging as it is in the salt 52. Complex 51 (PR 3 = PPh 3) reacts with
water to give 50(130) but no reactions of 54 have been reported. No reaction
of dinitrogen coordinated to rhodium or iridium have been reported.

2.7. Nickel, Palladium, and Platinum

No reactions of complexes 62 and 63 have been reported. Palladium


and platinum form no well-defined dinitrogen complexes. However, until
relatively recently there were few carbonyl complexes of palladium and
platinum. This has changed rapidly with the preparation of a wide variety
of compounds such as dinuclear complexes, e.g., [M2 Ch(1L -CO)-
(IL -dppmhJ where M = Pd, Pt, (143) and neutral and anionic polynuclear
complexes such as [Os2(CO)6{J.t-Pt-(CO)(PPh 3)}zJ(144) and
[ptg(COhsr 2Y45) The absence of simple, mononuclear palladium and
platinum dinitrogen complexes should not be construed as evidence that
this is a barren area for research.

2.8. Copper

There is one report of nitrogen fixation involving copper com-


plexes. (146) When dinitrogen is bubbled through either a pyridine or THF
solution of copper(I) borohydride or THF solution of [(PPh 3hCuBH 4J for
long periods of time, ammonia is formed. Ammonia is removed from the
reaction mixture either by sweeping it out with dinitrogen or by heating
to ca. 100°C and collecting in an acid trap. Reaction conditions are mild
with almost quantitative yields, and hydrolysis is not required to form
ammonia. Blank reactions all give negative results for ammonia production.
CA TAL YSIS OF NITROGEN-FIXING MODEL STUDIES 431

These results have been successfully reproduced in one other labora-


tory. (147) Unfortunately, the crucial experiment using 1s N2 and analyzing
for 1sNH3 has not been carried out yet.

3. NITRIDING REACTIONS

Many metals have the ability to undergo nitriding reactions with


dinitrogen. Titanium has been studied particularly thoroughly because of
its facile nitriding ability. Typically, metal salts are reduced with strong
reducing agents such as Grignard reagent, organolithium, Mg and LiAIH4
in ether and hydrocarbon solvent under a dinitrogen atmosphere, usually
high pressure, over an extended period of time. Subsequent hydrolysis
produces ammonia and/or hydrazine often in stoichiometric yieldsY3)
Reduction of [TiCI4(PPh3h] and [FeCh(PPh 3h] with EtMgBr in ether
solution for 10 h under dinitrogen (150 atm) produces very little ammonia
upon hydrolysis. (148) Low yields of organonitrogen compounds such as
aniline are obtained (upon hydrolysis) in reactions in which either a lithium
aryl is the reducing agent or a titanium-aryl precursor is employed, e.g.,
[( 1/ s_CsHshTiPh 2]. (23)
Catalytic nitrogen-fixing behavior is found in the nitriding reactions
of titanium. When TiCI 4, TiBr4, or Ti(OBu)4 is heated with a mixture of
Al and AlBr3 under dinitrogen, the yields of ammonia formed by hydrolysis
constantly increase with increase in the amount of Al and AlBr3 and can
be brought up to over 200 moljg-atom of Ti.(23) Aluminum is the reducing
agent; AlBr3 effectively removes the nitride ligand from titanium, and a
reduced titanium complex is the catalyst. Intriguing solvents have been
used in these reactions, such as benzene, molten AlBr3 (m.p. 97.5°C),
AlBr3/ AICh eutectic (2.17: 1; m.p. 6rC), and the ternary low-melting
mixture AlBr3/ AICh/benzene. It is difficult to see how these systems could
be developed into commercially viable catalysts unless dihydrogen could
be used as the sole reducing agent, i.e., hydrogenolyze the metal-nitride
bond and reduce the metal to its active form.

4. DINITROGEN BINDING AND REACTIVITY

Dinitrogen has been shown to adopt an end-on bonding arrangement


in those phosphine-containing complexes that have been isolated and
characterized. In most of these complexes, dinitrogen bonds in a linear
end-on arrangement either to one metal G (e.g., 15, 23, 43, 50, 52, and
432 T. ADRIAN GEORGE

57) or bridges between two metals H (e.g., 6, 13, 29, 32, 61, and 62).
However, in complex 54 the bridging geometry I is found with CoNN and
NNMg angles of 180° and 158°, respectively. The ability of a metal to bind
dinitrogen depends upon a subtle balance between the energies of the
appropriate metal and dinitrogen orbitals. The currently accepted bonding
scheme for end-on dinitrogen ligation involves donation of (/"- electron
density from dinitrogen to the metal and accepting 7T-electron density from
the metal into its 7T* orbitals. The extreme examples of no backing-donation
and significant back-donation are shown in resonance forms J and K:

G H

- + + -
M-N=.N: M=N=N:

J K

Organophosphine ligands are particularly successful in helping the metal


provide a favorable coordination site that is sufficiently electron-rich to
provide strong back-donation into the vacant dinitrogen 7T* - orbitals and
help stabilize the metal-dinitrogen bond.
The extent of the dinitrogen-to-metal interaction varies considerably
with changes in metal, oxidation number, and coligands. For example,
dinitrogen forms a relatively weak bond with d 6 Fe (II) , but with d 6 Re(I),
Mo(O), and W(O), the bond is stronger, and dinitrogen is activated towards
a variety of reagents. Ti(II) (d 2 ) is able to bind dinitrogen and, depending
upon the reducing agent and temperature, the bound dinitrogen is reduced
to either the "diazene-Ievel" (N2"2), "hydrazine-level" (N2"4), or "nitride-
level" (N- 3 ). The classification of "reduction-level" is based upon the
nitrogen hydride(s) formed upon hydrolysis, which is not necessarily unam-
biguous in the case of diazene and hydrazine. Depending upon the reaction
condition, diazene will decompose to a mixture of N2 and H 2, N2 and N2H 4,
and/or N2 and NH 3 , and hydrazine is readily decomposed by transition
metal compounds. Ta(III) (d 2 ) binds dinitrogen and in so doing reduces
it to the "hydrazine-Ievel," and Ta(l) (d 4 ) behaves similarly. In a
series of papers, Chatt and co-workers have attempted to define the dini-
trogen binding site in terms of its electron-richness (or -poorness) and
. b'l'
po I anza llty. (149)
So far, no example of a simple side-on bonded dinitrogen complex
has been reported. A common problem in working with dinitrogen com-
plexes is the facile loss of dinitrogen that is very frequently observed,
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 433

particularly in solution. It is for this reason that great difficulty is often


encountered when trying to isolate and purify this class of complex. In
other words, there are experimental as well as electronic obstacles to the
preparation of dinitrogen complexes.
The reactivity of coordinated dinitrogen toward electrophiles,
nucleophiles, and radicals is dependent upon many factors, particularly the
charge distribution within the dinitrogen-metal unit that results from the
synergistic-bonding pattern. Attempts to determine the charge distribu-
tion(110) have been made using X ray photoelectron spectroscopy
(XPS), (110,150) infrared absorption spectroscopy, (151) dipole moment
measurement,(110) and electrochemical measurements,(149,IS2) but no one
technique can uniquely measure this term,
In the only documented example of nucleophilic attack at coordinated
dinitrogen, LiPh(and LiMe) reacts with [(1/5 -C sH s)Mn(COh(N 2)], v(N 2 ) =
2165 cm - t, to form, after protonation, a neutral phenyldiazene complex
[(1/5 -C sHs)Mn(CO)z(PhNNH)]ys3) The evidence suggests that carbanion
attack occurs at the metal-bound dinitrogen atom, The particularly high
N(ls) binding energies (403.0 and 401.8 eV) and high v (N 2) imply a
relatively positive dinitrogen ligand (or electron-poor complex), with the
bonding better described by resonance form J than K. Efforts to extend
this work to other electron-poor dinitrogen-containing complexes have not
been successful so far. (149)
The Lewis base behavior of coordinated dinitrogen has been studied
and related to v(N 2 )YI4) In many cases, stable Lewis base-Lewis acid
ad ducts have been isolated (e.g., 1-3, 10, 14, 29-32). Those Lewis acids
that are both (T- and 1T"- acceptors have been shown to cause considerable
lengthening of the N-N bond (e.g., in 29 and 32) due to depopulation of
the N2 1T" bondsY S) Resonance structure K is more important than J in
describing the behavior of coordinated dinitrogen as a strong Lewis base
and in the stable adducts. In some cases, bridging dinitrogen complexes
are obtained directly from the preparation and isolation of dinitrogen
complexes (e.g., 4-7, 9, 12, 13, 39, 40, 45, 53, 54, and 60-62). The
noticeably long N-N bond observed in the tantalum complex 6 and concom-
mitant short Ta-N bonds are undoubtably the result of the ability of
tantalum (and certain other metals such as molybdenum, titanium, tungsten,
and vanadium) to behave as a (T-acceptor and 1T"-donor, as well as a
1T"- acceptor.
The susceptibility of coordinated dinitrogen to attack by electrophiles
and radicals and their relationship to the electron-richness of dinitrogen
complexes has been addressed in an important series of articles using E~i2
and v(N 2 ) data. (82,149,154) The successful attack of an electrophile or radical
upon coordinated dinitrogen depends upon there being a certain degree
of electron-richness. However, in many cases where alkylation or
434 T. ADRIAN GEORGE

protonation of dinitrogen would be predicted to occur [based upon E ~i2


and v(N2 )], instead oxidation or protonation of the metal or no reaction
occurs. This arises because of the high degree of conjugation present in
the [M -N-NJ unit and our inability to quantitatively determine the charge
distribution in the ground state and excited state(s).
How can you determine whether or not coordinated dinitrogen is
"activated"? It is far from being as simple as just adding acid and looking
for ammonia and hydrazine! For example, the protonation of coordinated
dinitrogen may occur to give a {MNNH} or {MNNH2} unit, which will be
unstable unless strong metal-nitrogen bonding can occur, i.e., the metal
can be oxidized. This readily occurs with Mo(O) and W(O) to form M(II)
and M(IV) complexes but is much less likely in the case of Fe(II), for
example. Unstable {MNNH} will decompose to give N2 and H+, giving the
erroneous impression that coordinated dinitrogen did not react. {MNNH 2}
may decompose to give N2H 2, which would be unstable and would form
N2 and H 2, N2 and N2H 4, and/or N2 and NH3, or simple N2 + 2H+. In the
former case, no apparent reaction of coordinated dinitrogen is the erroneous
conclusion again, unless N2H2 is detected chemically or {MNNH 2} is
detected spectroscopically. Stoichiometric formation of N2H 2, N2H 4, and
2NH3 requires 2, 4, and 6 electrons, respectively. Many metals can provide
two electrons per metal (e.g., Co(l) --. Co (III) + 2e -) or two per dinuclear
metal unit, but the problems associated with recognizing dinitrogen reac-
tivity in this case have been pointed out above. It is unrealistic to expect
many metals that are coordinated to dinitrogen to be able to provide four
or six electrons to produce N2H4 or 2NH3 at a single metal; only tungsten
has unambiguously been shown to do this so far. In dinuclear tantalum(I),
-(III), (29) titanium(II), (155) and zirconium(II)(156) complexes containing coor-
dinated dinitrogen, protonation produces nearly quantitative yields of
hydrazine. High yields of ammonia are formed upon protonation of molyb-
denum complexes such as 27 by virtue of a mechanism involving a bridging
intermediate (Scheme 6), whereby three electrons come from each molyb-
denum atom. It is probably unreasonable to expect that many dinitrogen
complexes will form NH3, N2H 4, or stable nitrogen hydride intermediates
upon protonation, unless the metal has a number of stable oxidation states
available to it or unless a reducing agent (other than the metal) is present.

5. FUTURE PROSPECTS

Despite the large amount of novel chemistry that has arisen from the
chemical studies of nitrogen fixation, there is nothing to suggest an alterna-
tive to the Haber Process for the large-scale (> 100 ton per day) production
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 435

of ammonia. However, there may be a real need for alternatives to the


Haber Process. A 400 million dollar plant to produce ammonia can be
built in an underdeveloped country, but how will ammonia be distributed,
even if it is converted into urea? Small scale, local production of nitrogen
fertilizer would be of direct benefit to the food producers of underdeveloped
countries. The use of local resources such as solar, water, and wind energy
coupled with a low-technology ammonia-producing plant that would be
working intermittently year round would produce an aqueous solution of
ammonia (or ammonia nitrate if some of the ammonia is oxidized) that
could either be run into the irrigation ditches or manually distributed.
Tht: conversion of dinitrogen to ammonia is shown in Equations 33
and 34 to emphasize two approaches.
(33)
(34)
On the one hand, dihydrogen is the reducing agent, and, on the other,
electrons and protons are provided separately. They could be classified as
the chemical and electrochemical methods. It is going to be very difficult
to develop a homogeneous "Haber Process" using metal-phosphine com-
plexes or any other metal complexes. Although there are a few stable
complexes known that contain both nitrogen and hydrogen or dihydrogen,
generally dihydrogen displaces dinitrogen from a metal-binding site rather
easily. One area that seems to have received no attention is dinitrogen
complexes of metal cluster. It may be possible to assemble mixed-metal
clusters that can coordinate both dihydrogen and dinitrogen and promote
dinitrogen reduction. Phosphine rather than carbonyl ligands would be
preferred to increase the electron-richness of the cluster, thus aiding
dinitrogen binding.
The electrochemical activation of coordinated dinitrogen has received
some attention. The moiety [M(N 2)(dppe)ZJ (M = Mo or W) has been
anchored to a Sn02 electrode but with no chemistry of the coordinated
dinitrogen reported. (157) In principle, it should be possible to anchor to an
electrode those complexes that do produce ammonia or hydrazine upon
treatment with acid. For the reaction to become catalytic, the dinitrogen
binding site needs to be regenerated. In practice, strong reducing agents
are used to generate these binding sites originally, strong enough to reduce
protons to dihydrogen. However, not enough is known about the minimum
potentials necessary for creation of the binding site, and this is information
uniquely available by electrochemical studies.
The direct production of organic nitrogen compounds using dinitrogen
has received little attention outside the reactions of organic halides with
coordinated dinitrogen. Alkenes and alkynes would be viable carbon-atom
436 T. ADRIAN GEORGE

sources, although like dihydrogen they often displace coordinated


dinitrogen. If coordinated dinitrogen, in complexes such as 5-7 where
dinitrogen is "activated," reacts with alkynes (RC-CH) in an "olefin-
methatesis-type" reaction, the product would be a nitrile (RC-N) or
hydrogen cyanide (HC-N).

6. CONCLUDING REMARKS

To many people, nitrogenase is a Rosetta Stone: elucidation of the


mechanism will provide the key to nitrogen fixation under mild condi-
tionsY56) In the meantime, work is proceeding to find new dinitrogen-
binding sites with the hope of discovering new chemistry. Certainly,
organophosphine ligands have helped provide an environment to study the
chemistry of coordinated dinitrogen. It is now time to investigate ligands
containing other donor atoms and explore their ability to provide support
for the binding and activation of dinitrogen. One of the new areas under
intensive investigation is the [FeMoS] and [FeWS] clusters.(157) Within
these systems lies the hope of modeling the nitrogenase cofactor, FeMoCo.
All solutions to the problems of nitrogen-fixation under mild conditions
lie in the future, but sufficient work has been done to catalyze further
vigorous investigations.

ACKNOWLEDGMENTS

Thanks are due to Professor R. R. Schrock for informing me of results


prior to publication. Published and unpublished work by the author repor-
ted in this chapter was supported by funds from the National Science
Foundation and the University of Nebraska Research Council. Acknowl-
edgment is made to the donors of the Petroleum Research Fund,
administered by the American Chemical Society, for partial support of this
work.

REFERENCES

1. E. N. Mishustin and V. K. Shilnikova, Biological Fixation of Atmospheric Nitrogen


(Plenum Publishing, London, 1971).
2. M. J. Dilworth, Ann. Rev. Plant Physiol. 25, 81-114 (1974).
CA TAL YSIS OF NITROGEN-FIXING MODEL STUDIES 437

3. R. C. Burns and R. W. F. Hardy, Molecular Biology, Biochemistry, and Biophysics


(Springer, Berlin, 1975), Vol. 12.
4. W. D. P. Stewart and J. R. Gallon, eds., Nitrogen Fixation (Academic Press, London,
1981).
5. J. R. Postgate, The Chemistry and Biochemistry of Nitrogen Fixation, edited by J. R.
Postgate (Plenum Publishing, New York, 1971), pp. 161-190.
6. R. H. Burris, The Chemistry and Biochemistry of Nitrogen Fixation, edited by J. R.
Postgate (Plenum Publishing, New York, 1971), pp. 105-160.
7. R. W. F. Hardy, R. C. Burns, and G. W. Parshall, Bioinorganic Chemistry, edited by
R. F. Gould, Adv. Chem. Ser. (ACS Publications, Washington, D.C.), Vol. 100, pp.
219-245 (1971).
8. V. K. Shah and W. J. Brill, Proc. Natl. Acad. Sci. U.S.A. 74, 3249-3253 (1977).
9. S. P. Cramer, W. O. Gillium, K. O. Hodgson, L. E. Mortenson, E. I. Stiefel, J. R.
Chisnel, W. J. Brill, and V. K. Shah, 1. Am. Chem. Soc. 100, 3814-3819 (1978).
10. R. N. F. Thornley, R. R. Eady, and D. J. Low, Nature (London) 272, 557-558 (1978).
11. K. L. Hadfield and W. A. Bullen, Biochemistry 8, 5103-5108 (1969).
12. A. V. Slack, Ammonia, edited by A. V. Slack and G. R. James (Marcel Dekker, New
York, 1973), Vol. 1, pp. 1-142.
13. D. E. Harrison and H. Taube, 1. Am. Chem. Soc. 89, 5706-5707 (1967).
14. M. I. Temkin, Adv. Catal. 28, 250-263 (1979), and references therein.
15. J. P. Collman, M. Kubota, F. D. Vastine, J.-Y. Sun, and J. W. Kang, 1. Am. Chem.
Soc. 90, 5430-5437 (1968).
16. R. W. F. Hardy, F. Bottomley, and R. C. Burns, eds., A Treatise on Dinitrogen Fixation
(John Wiley and Sons, New York, 1979).
17. J. Chatt, L. M. da Camara Pina, and R. L. Richards, eds., New Trends in the Chemistry
of Nitrogen Fixation (Academic Press, London, 1980).
18. J. Chatt, J. R. Dilworth, and R. L. Richards, Chem. Rev. 78, 589-625 (1978).
19. M. Hidai, Y. Mizobe, and Y. Uchida, 1. Am. Chem. Soc. 98,7824-7825 (1976).
20. J. Chatt, G. A. Heath, and R. L. Richards, 1. Chem. Soc., Dalton Trans. 1974,
2074-2082.
21. T. A. George and M. E. Noble, Inorg. Chem. 17,1678-1679 (1978).
22. L. J. Archer and T. A. George, Inorg. Chem. 18,2079-2082 (1979).
23. V. E. Vol'pin and V. B. Shur, New Trends in the Chemistry of Nitrogen Fixation, edited
by J. Chatt, L. M. da Camara Pin a, and R. L. Richards (Academic Press, London,
1980), pp. 67-100.
24. A. E. Shilov, New Trends in the Chemistry of Nitrogen Fixation, edited by J. Chatt,
L. M. da Camara Pina, and R. L. Richards, (Academic Press, London, 1980), pp.
121-150.
25. J. H. Teuben, New Trends in the Chemistry of Nitrogen Fixation, edited by J. Chatt,
L. M. da Camara Pina, and R. L. Richards (Academic Press, London, 1980), pp.
233-247.
26. R. Robson, Inorg. Chem. 13, 475-479 (1974).
27. R. J. Burt, G. J. Leigh, and D. L. Hughes, 1. Chem. Soc., Dalton Trans. 1981,793-799.
28. H. W. Turner, J. D. Fellmann, S. M. Rocklade, R. R. Schrock, M. R. Churchill, and
H. J. Wasserman, 1. Am. Chem. Soc. 102, 7809-7811 (1980).
29. S. M. Rocklage and R. R. Schrock, 1. Am. Chem. Soc. 104, 3077-3081 (1982).
30. S. M. Rocklage, H. W. Turner, D. Fellmann, and R. R. Schrock, Organometallics 1,
703-707 (1982).
31. M. R. Churchill and H. J. Wasserman, Inorg. Chem. 20,2899-2904 (1981).
32. --,Inorg. Chem. 21,218-222 (1982).
33. H. H. Karsch, Angew. Chem., Int. Ed. Engl. 16, 56-57 (1977).
438 T. ADRIAN GEORGE

34. P. Sobota and B. Jezowska-Trzebiatowska, 1. Organornet. Chern. 131,341-345 (1977).


35. J. Chatt, R. C. Fay, and R. L. Richards, 1. Chern. Soc., Dalton Trans. 1971,702-704.
36. M. Hidai, K. Tominari, Y. Uchida, and A. Misono, 1. Chern. Soc., Chern. Cornrnun.
1969,814.
37. M. L. H. Green and W. E. Silverthorn, 1. Chern. Soc., Dalton Trans. 1973, 301-306.
38. - - , 1 . Chern. Soc., Dalton Trans. 1974, 2164-2166.
39. R. A. Forder and K. Prout, Acta Crystallogr., Sect. B 30,2778-2780 (1974).
40. M. Hidai, K. Tominari, Y. Uchida, and A. Misono, 1. Chern. Soc., Chern. Cornrnun.
1969, 1392. M. Hidai, K. Tominari, and Y. Uchida, 1. Arn. Chern. Soc. 94, 110-114
(1972).
41. Y. Uchida, T. Uchida, M. Hidai, and T. Komada, Acta Crystallogr., Sect. B 31,
1197-1199 (1975).
42. L. K. Atkinson, A. H. Mawby, and D. C. Smith, 1. Chern. Soc., Chern. Cornrnun. 1971,
157-158.
43. A. Frigo, G. Pousi, and A. Turco, Gazz. Chirn. Ital. 101,637-638 (1971).
44. C. Miniscloux, G. Martino, and L. Sajus, Bull. Soc. Chirn. Fr. 567, 2183-2188
(1973).
45. T. A. George and C. D. Seibold, 1. Organornet. Chern. 30, C13-C14 (1971). T. A.
George and C. D. Seibold, Inorg. Chern. 12,2544-2547 (1973).
46. J. R. Dilworth and R. L. Richards, Inorg. Syn., edited by D. H. Busch (John Wiley and
Sons, New York, 1980), pp. 119-124.
47. J. Chatt, W. Hussain, G. J. Leigh, H. Neukomm, C. J. Picket!, and D. A. Rankin, 1.
Chern. Soc., Chern. Cornrnun. 1980,1024-1025.
48. M. W. Anker, J. Chat!, G. J. Leigh, and A. G. Wedd, 1. Chern. Soc., Dalton Trans.
1975, 2639-2645.
49. J. Chat!, A. J. Pearman, and R. L. Richards, 1. Chern. Soc., Dalton Trans. 1977,
1852-1860.
50. T. A. George and R. A. Kovar, Inorg. Chern. 20, 285-287 (1981).
51. M. Sato, T. Tatsumi, T. Kodama, M. Hidai, T. Uchida, and Y. Uchida, 1. Arn. Chern.
Soc. 100,4447-4452 (1978).
52. M. Aresta and A. Sacco, Gazz. Chirn. Ital. 102,755-780 (1972).
53. J. A. Baumann, M. C. Davies, and T. A. George, unpublished results.
54. G. J. Kubas, 1. Chern. Soc., Chern. Cornrnun. 1980, 61-62.
55. M. Mercer, R. H. Crabtree, and R. L. Richards, 1. Chern. Soc., Chern. Cornrnun. 1973,
808-809.
56. M. Mercer, 1. Chern. Soc., Dalton Trans. 1974, 1637-1640.
57. J. Chatt and R. L. Richards, 1. Less-Cornrnon Metals 54, 477-484 (1977).
58. P. D. Chadwick, 1. Chern. Soc., Dalton Trans. 1976, 1934-1936.
59. J. Chatt, G. A. Heath, and G. J. Leigh, 1. Chern. Soc., Chern. Cornrnun.1972, 444-445.
60. J. Chat!, G. A. Heath, and R. L. Richards, 1. Chern. Soc., Chern. Cornrnun. 1972,
1010-1011.
61. A. A. Diamantis, J. Chat!, G. J. Leigh, and G. A. Heath, 1. Organornet. Chern. 84,
C11-C12 (1975).
62. J. Chatt, A. J. Pearman, and R. L. Richards, Nature 253,39-40 (1975).
63. J. Chat!, A. A. Diamantis, G. A. Heath, N. E. Hooper, and G. J. Leigh, 1. Chern. Soc.,
Dalton Trans. 1977, 688-697.
64. M. Sato, T. Kodama, M. Hidai, and Y. Uchida, 1. Organornet. Chern. 152, 239-254
(1978).
65. V. W. Day, T. A. George, and S. D. A. Iske, Jr., 1. Arn. Chern. Soc. 97,4127-4128
(1975).
66. G. E. Bossard, D. C. Busby, M. Chang, T. A. George, and S. D. A. Iske, Jr., 1. Arn.
Chern. Soc. 102,1001-1008 (1980).
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 439

67. D. C. Busby, T. A. George, S. D. A. Iske, Jr., and S. D. Wagner, Inorg. Chern. 20,
22-27 (1981).
68. c. S. Day, V. W. Day, T. A. George, and I. Tavanaiepour, Inorg. Chirn. Acta 45, L54
(1980).
69. V. W. Day and T. A. George, unpublished results.
70. V. W. Day, T. A. George, S. D. A. Iske, Jr., and S. D. Wagner, I. Organornet. Chern.
112, C55-C58 (1976).
71. F. C. Marsh, R. Mason, and K. M. Thomas, I. Organornet. Chern. 96, C43-C45 (1975).
72. D. c. Busby and T. A. George, Inorg. Chern. 18, 3164-3167 (1979).
73. G. Butler, J. Chatt, G. J. Leigh, and D. L. Hughes, Inorg. Chirn. Acta 30, L287-L288
(1978).
74. J. Chatt, R. A. Head, G. J. Leigh, and C. J. Pickett, I. Chern. Soc., Dalton Trans. 1978,
1638-1647.
75. G. E. Bossard and T. A. George, Inorg. Chirn. Acta 54, L241-L242 (1981).
76. R. Ben-Shoshan, J. Chatt, G. J. Leigh, and W. Hussain, I. Chern. Soc., Dalton Trans.
1980,771-775.
77. R. A. Head and P. B. Hitchcock, I. Chern. Soc., Dalton Trans. 1980, 1150-1155.
78. M. Hidai, Y. Mizobe, M. Sato, T. Kodama, and Y. Uchida, I. Am. Chem. Soc. 100,
5740-5748 (1978).
79. J. Chatt, W. Hussain, G. J. Leigh, and F. P. Terreros, I. Chem. Soc., Dalton Trans.
1980,1408-1415.
80. T. A. George and R. L. Turcotte, unpublished results.
81. G. E. Bossard and T. A. George, Inorg. Chim. Acta 54, L239-L240 (1981).
82. J. Chatt, G. J. Leigh, H. Neukomm, C. J. Pickett, and D. R. Stanley, I. Chem. Soc.,
Dalton Trans. 1980,121-127.
83. R. J. W. Thomas, G. S. Laurence, and A. A. Diamantis, Inorg. Chim. Acta 30,
L353-L355 (1978).
84. A. Caruana, H. Hermann, and H. Kisch, I. Organornet. Chem. 187, 349-359 (1980).
85. T. Tatsumi, M. Hidai, and Y. Uchida, Inorg. Chern. 14, 2530-2534 (1975).
86. B. J. Carter, J. E. Bercaw, and H. B. Gray, I. Organomet. Chem. 181,105-116 (1979).
87. J. Chatt, R. A. Head, and G. J. Leigh, I. Chem. Soc., Dalton Trans. 1980,1129-1134.
88. D. C. Busby, C. D. Fendrick, and T. A. George, Chernistry and Uses of Molybdenum,
Proceedings of the Third International Conference (Ann Arbor, 1979), pp. 290-295.
89. P. C. Bevan, J. Chatt, G. J. Leigh, and E. G. Leelamani, I. Organomet. Chern. 139,
C59-C62 (1977).
90. c. J. Pickett and G. J. Leigh, I. Chem. Soc., Chem. Commun. 1981, 1033-1035.
91. M. Hidai, T. Kodama, M. Sato, M. Harakawa, and Y. Uchida, Inorg. Chern. 15,
2694-2697 (1976).
92. J. Chatt, A. J. Pearman, and R. L. Richards, I. Chem. Soc., Dalton Trans. 1978,
1766-1776.
93. - - , I. Chem. Soc., Dalton Trans. 1977,2139-2142.
94. T. Takahashi, Y. Mizobe, M. Sato, Y. Uchida, and M. Hidai, I. Am. Chern. Soc. 101,
3405-3407 (1979).
95. J. A. Baumann and T. A. George, I. Arn. Chem. Soc. 102,6153-6154 (1980).
96. S. N. Anderson, M. E. Fakley, R. L. Richards, and J. Chatt, I. Chem. Soc., Dalton
Trans. 1981, 1973-1980.
97. G. E. Bossard, T. A. George, and R. K. Lester, Inorg. Chim. Acta Letters, L241-L242
(1983).
98. J. Chatt, M. E. Fakley, P. B. Hitchcock, R. L. Richards, N. T. Luong-Thi, and D. J.
Hughes, I. Organomet. Chem. 172, C55-C58 (1979).
99. M. Hidai, Cumnt Perspectives in Nitrogen Fixation, edited by A. H. Gibson and W. E.
Newton (Australian Academy of Sciences, Canberra, 1981), pp. 26-29.
440 T. ADRIAN GEORGE

100. G. A. Heath, R. Mason, and K. M. Thomas, f. Arn. Chern. Soc. 96, 259-260 (1974).
101. I. R. Hanson and D. L. Hughes, f. Chern. Soc., Dalton Trans. 1981,390-399.
102. J. Chatt, A. J. Pearman, and R. L. Richards, f. Chern. Soc., Dalton Trans. 1976,
1520-1524.
103. R. A. Henderson, f. Organornet. Chern. 208, C51-54 (1981).
104. J. Chatt, J. R. Dilworth, P. L. Dahlstrom, and J. Zubieta, f. Chern. Soc., Chern. Cornrnun.
1980,786-787.
105. J. R. Dilworth, R. A. Henderson, and R. L. Richards, Cumnt Perspectives in Nitrogen
Fixation, edited by A. H. Gibson and W. E. Newton (Australian Academy of Sciences,
Canberra, 1981), p. 349.
106. G. E. Bossard, Ph.D. Thesis, University of Nebraska (1983).
107. M. C. Davies, T. A. George, and R. K. Lester, unpublished results.
108. D. Sellmann, R. Gerlach, and K. Jodden, f. Organornet. Chern. 178, 433-447 (1979),
and references therein.
109. C. R. Brulet and E. E. van Tamelen, f. Arn. Chern. Soc. 97, 912-913 (1975).
110. J. Chatt, C. M. Elson, N. E. Hooper, and G. J. Leigh, f. Chern. Soc., Dalton Trans.
1975, 2392-2401.
111. M. E. Tully and A. P. Ginsberg, f. Arn. Chern. Soc. 95, 2042-2043 (1973).
112. K. W. Chiu, W. K. Wong, and G. Wilkinson, f. Chern. Soc., Chern. Cornrnun. 1981,
451-452.
113. J. Chatt, J. R. Dilworth, and G. J. Leigh, f. Chern. Soc., Dalton Trans. 1973,612-618.
114. J. Chatt, R. H. Crabtree, E. A. Jeffery, and R. L. Richards, f. Chern. Soc., Dalton
Trans. 1973, 1167-1172.
115. D. Sellmann and E. Kleinschmidt, f. Organornet. Chern. 140,211-219 (1977).
116. W. E. Silverthorn,f. Chern. Soc., Chern. Cornrnun.1971, 1310-1311.
117. M. Aresta, P. Giannoccaro, M. Rossi, and A. Sacco, Inorg. Chirn. Acta 5, 203-206
(1971).
118. D. H. Gerlach, W. G. Peet, and E. L. Muetterties, J. Arn. Chern. Soc. 94, 4545-4549
(1972).
119. Yu. G. Borod'ko, M. O. Broitman, L. M. Kachapina, A. K. Shilova, and A. E. Shilov,
Zh. Struct. Khirn. 12, 545-546 (1971).
120. P. Stoppioni, F. Mani, and L. Sacconi, Inorg. Chirn. Acta 11, 227-230 (1974).
121. C. A. Ghilardi, S. MidolIini, L. Sacconi, and P. Stoppioni, f. Organornet. Chern. 205,
193-202 (1981).
122. G. M. Bancroft, M. J. Mays, B. E. Prater, and F. P. Stefanini, f. Chern. Soc., A 1970,
2146-2149.
123. Yu. G. Borod'ko, M. O. Broitman, L. M. Kachapina, A. E. Shilov, and L. Yu. Ukhin,
f. Chern. Soc., Chern. Cornrnun. 1971, 1185-1186.
124. J. M. Bellerby, M. J. Mays, and P. L. Sears, f. Chern. Soc., Dalton Trans. 1976,
1232-1236.
125. R. A. Cable, M. Green, R. E. Mackenzie, P. L. Timms, and T. W. Turney, f. Chern.
Soc., Chern. Cornrnun. 1976, 270-271.
126. W. H. Knoth, f. Arn. Chern. Soc. 94, 104-109 (1972).
127. S. Komiya and A. Yamamoto, Bull. Soc. Chern. fapan. 49, 784-787 (1976).
128. J. Chatt, D. P. Melville, and R. L. Richards, f. Chern. Soc., A 1971,895-899.
129. Y. Yamamoto, S. Kitazume, L. S. Pu, and S. Ikeda, f. Arn. Chern. Soc. 93, 371-380
(1971).
130. M. Aresta, G. F. Nobile, M. Rossi, and A. Sacco, f. Chern. Soc., Chern. Cornrnun. 1971,
781.
131. R. Hammer, H.-F. Klein, P. Friedrich, and G. Huttner, Angew. Chern., Int. Ed. Engl.
16,485-486 (1977).
132. R. Hammer, H.-F. Klein, U. Schubert, A. Frank, and G. Huttner, Angew. Chern., Int.
Ed. Eng/. 15, 612-613 (1976).
CATAL YSIS OF NITROGEN-FIXING MODEL STUDIES 441

133. Y. Miura and A. Yamamoto, Chem. Lett. 1978,937-940.


134. H. L. M. van Gaal, F. G. Moers, and J. J. Steggerda, 1. Organomet. Chem. 65, C43-C45
(1974). H. L. M. van Gaal and F. L. A. van Bekerom, 1. Organomet. Chem. 134,
237-248 (1977).
135. P. R. Hofmann, T. Yoshida, T. Okano, S. Otsuka, and J. A. Ibers, Inorg. Chem. 15,
2462-2466 (1976).
136. D. L. Thorn, T. H. Tulip, and J. A. Ibers, 1. Chem. Soc., Dalton Trans. 1979, 2022-2025.
137. T. Yoshida, T. Okano, and S. Otsuka, 1. Chem. Soc., Chem. Commun. 1978, 855-856.
138. T. Yoshida, T. Okano, D. L. Thorn, T. H. Tulip, S. Otsuka, and J. A. Ibers, 1. Organomet.
Chem. 181, 183-201 (1979).
139. T. Yoshida, T. Okano, Y. Veda, and S. Otsuka, 1. Am. Chem. Soc. 103, 3411-3422
(1981).
140. A. Sacco and M. Rossi, Inorg. Chim. Acta 2, 127-132 (1968).
141. P. W. Jolly, K. Jonas, C. Kruger, and V.-H. Tsay, 1. Organomet. Chem., 33, 109-122
(1971).
142. C. A. Tolman, D. H. Gerlach, J. P. Jesson, and R. A. Shunn, 1. Organomet. Chem. 65,
C23-C26 (1974).
143. M. P. Brown, R. J. Puddephatt, M. Rashidi, and K. R. Seddon, 1. Chem. Soc., Dalton
Trans. 1978, 1540-1544.
144. L. J. Farrugia, J. A. K. Howard, P. Mitrprachachon, J. L. Spencer, F. G. A. Stone, and
P. Woodward, 1. Chem. Soc., Chem. Commun. 1978, 260-262.
145. J. C. Calabrese, L. F. Dahl, P. Chini, G. Longoni, and S. Martinengo, 1. Am. Chem.
Soc. 96,2616-2618 (1974).
146. C. A. Koerntgen, Reduction of Nitrogen to Ammonia and Carbon Monoxide to For-
maldehyde by Copper(l) Borohydride, Ph.D. Thesis, Boston College, 1973.
147. B. D. James, personal communication.
148. M. E. Vol'pin, N. K. Chapovskaya, and V. B. Shur, Izvest. Akad. Nauk SSSR, Ser.
Khim. 1966,1083-1084.
149. G. J. Leigh, R. H. Morris, C. J. Pickett, D. R. Stanley, and J. Chatt, 1. Chem. Soc.,
Dalton Trans. 1981, 800-804, and references therein.
150. B. Folkesson, Acta Chem. Scand. 27, 287-302 (1973); 27, 1441-1443 (1973). V. I.
Nefedov, V. S. Lenenko, V. B. Shur, W. E. Vol'pin, J. E. Salyn, and M. A. Porai-Koshits,
Inorg. Chim. Acta 7, 499-502 (1973). P. Finn and W. L. Jolly, Inorg. Chem. 11,
1434-1435 (1972). H. Binder and D. Sellmann, Angew. Chem., Int. Ed. Eng/. 12,
1017-1019 (1973). P. Brant and R. D. Feltham, 1. Less-Common Metals 54, 81-87
(1977).
151. D. J. Darensbourg, Inorg. Chem. 11, 1436-1437 (1972). B. Folkesson, Acta Chem.
Scand. 26,4008-4018 (1972); 27, 276-286 (1973). M. S. Quinby and R. D. Feltham,
Inorg. Chem. 11, 2468-2476 (1972).
152. c. M. Elson, 1. Chem. Soc., Dalton Trans. 1975, 2401-2404.
153. D. Sellman and W. Weiss, Angew Chem., Int. Ed. Engl. 16, 880-881 (1977); 17,
269-270 (1978).
154. J. Chatt, C. T. Kan, G. J. Leigh, C. J. Pickett, and D. R. Stanley, 1. Chem. Soc., Dalton
Trans. 1980, 2032-2038.
155. R. D. Sanner, D. M. Duggan, T. C. McKenzie, R. E. Marsh, and J. E. Bercaw, 1. Am.
Chem. Soc. 98, 8358-8365 (1976).
156. 1. M. Manriquez, R. D. Sanner, R. E. Marsh, and J. E. Bercaw, 1. Am. Chem. Soc. 98,
3042-3044 (1976).
157. G. J. Leigh and C. J. Pickett, 1. Chem. Soc., Dalton Trans. 1977, 1797-1800.
158. W. E. Newton, Nitrogen Fixation, Kirk-Othmer: Encyclopedia of Chemical Technology,
3rd Edition (John Wiley and Sons, New York, 1981), Vol. 15, pp. 942-968.
159. For example, D. Coucouvanis, Ace. Chem. Res. 14, 201-209 (1981), and references
therein.
14
Polymer-Bound Phosphine
Catalysts
Norman L. Holy

1. INTRODUCTION AND SCOPE

While the theme of this volume is catalysis in a single phase, the principles
and dynamics of homogeneous catalysis have impacted other areas of
catalysis as well. Today, more than ever, there is an overlapping, a meshing
of concepts from homogeneous and heterogeneous catalysis. One of the
areas that interfaces both classical divisions is catalysis via polymer-bound
transition metal complexes. It is an interface area because catalysts are
prepared typically from an organic polymer such as polystyrene and then,
after attachment of a ligand, a soluble metal complex is bound. The choice
of metal complex is almost always based upon examples from homogeneous
catalysis. One of the principal motivations for attaching a soluble complex
to a polymer is that catalyst recovery becomes much easier. There are,
however, motivations beyond simple recovery considerations for examining
the reactions of polymer-bound catalysts. Though the characteristics of the
polymer-bound phosphines may be described to a first approximation as
being predictable on the basis of the properties, performances, and
mechanistic interpretations of the homogeneous analogs, the support often
has a significant influence on catalytic activity. For example, a polymer-
bound catalyst may give a product distribution quite different from that of
the soluble catalyst. Moreover, it is often observed that polymer-supported
catalysts are less oxygen sensitive than the homogeneous ones. There is

Dr. Norman L. Holy. Department of Chemistry, Western Kentucky University, Bowling


Green, Kentucky 42101.

443
444 NORMAN L. HOL Y

also the very practical matter of solubility. Many transition metal complexes
have but limited solubilities in desirable solvents, and the researcher is
forced to utilize either very dilute solutions or expensive solvents. For
potential commercial applications, the requirement of large solvent volumes
is certainly deleterious. With polymer-bound catalysts the difficulty of
establishing a high concentration of metal centers is circumvented, because
the metal complex may be attached to a polymer in far higher concentrations
than are possible in solution.
Polymer-bound catalysts share some of the limitations intrinsic in both
classical homogeneous and heterogeneous catalysis. Like soluble catalysts,
the range of reactions they facilitate is rather limited. As with heterogeneous
catalysts it is often difficult to learn the intimate details of ligand coordina-
tion, mechanism, and catalyst poisoning. Normally, such interpretations
rely heavily upon the characteristics of soluble complexes, though it is clear
the correspondence is imperfect.
Attaching homogeneous transition metal complexes became an impor-
tant discipline in the early 1970s. In addition to the use of phosphine-
derived polymers, there are published accounts using virtually every con-
ceivable type of attached ligand. (1,2)
The principal objectives of this review are to illustrate the range of
reactions known to be catalyzed by phosphine-supported catalysts, to
discuss critically performances of anchored catalysts in comparison to their
soluble analogs, to consider the effects of the support itself, and to indicate
a few areas where research activity might be fruitful. References are
intended to be illustrative.

2. PREPARA TlON OF POL YMER-BOUND PHOSPHINE CA TAL YSTS

There are two quite different approaches to catalyst preparation. The


most common route begins with a commercially available polymer to which
a ligand and metal complex are anchored. Advantages of this route, in
addition to the simplicity of using a commercially available material, are
that the characteristics (surface area, pore size, swelling properties) of the
polymer are already known. The alternative synthesis begins with a
monomer already containing a phosphine; this monomer is then polymer-
ized. A major advantage here is the greater flexibility in tailoring the
support to optimize performance.

2. 1. Modification of Preformed Polymers


Polystyrene is the most commonly used support because of its ready
availability in several forms, its chemical inertness, and the ease with which
POL YMER-80UND PHOSPHINE CA TAL YSTS 445

ligands and metals may be anchored to it. For a variety of catalysts, three
types of polystyrene are utilized. If a polymer lacks cross-linking, catalysts
derived from such materials are normally soluble; that is, reactions with
such polymeric catalysts are homogeneous. Even with a soluble polymer
one of the real advantages of polymer-bound catalysts, the facility of catalyst
recovery, is not negated. The catalyst may be recovered quite simply by
filtration through a membrane(3) or by addition of a solvent in which the
polymeric catalyst is insoluble. (4) A second type of polystyrene, indeed the
customary one, is a gel-type (microporous) having a low degree «2%) of
cross-linking. The typical cross-linking agent is p-divinylbenzene (DVB).
By proper choice of solvent, these resins swell and open all their internal
volume to solvent and reagent. Microporous resins must be swollen prior
to the anchoring of catalyst and also during their use as catalysts. The third
type of polystyrene has a high degree of cross-linking (typically 20, 40, or
60% DVB); these "macroreticular" or "macroporous" resins have high
surface areas. (5) The high cross-link density restricts diffusion into the
interior, and attached catalysts are normally found in a thin layer at the
internal surface. Before these macroporous resins are derivatized, it is
important that the emulsifiers used in their preparation be removed; this
requires a fairly lengthy washing procedure. (6)
Phosphine attachment to preformed polystyrene is accomplished typi-
cally by one of two routes. Bromination of the phenyl residues is followed
by either metallation of the aryl moiety and treatment with
chlorodiphenylphosphine, or by direct phosphination with diphenylphos-
phide. The second route to ligand attachment utilizes chloromethylation
followed by diphenylphosphide treatment. Either method of phosphine

®-Q-Br ~ ®-Q-Li
Br,/
_/FeOrTI(IIIJ
~~h' _

~ MPPh, ~
~ ~PPh2
~SnCI4

0-Q-CH CI
CH,OCH,ci"-,.

attachment gives nearly complete halogen replacement. This is important,


particularly for the chloromethylation route, because the possibility of
quaternization is thereby avoided:
446 NORMAN L. HOL Y

The bidentate ligand bis(1,2-diphenylphosphine)ethane (DIPHOS) may be


introduced in an analogous manner:

®-OBr + LiPPhCH 2 CH 2 PPh 2 -- ®-OPPhCH 2 CH 2 PPh 2

A similar approach was used to phosphenate noncross-linked poly-


vinylchloride.

For complete chlorine replacement, it was necessary to use reflux times of


up to 40 hr. Furthermore, a considerable breakdown of the polymer
occurred. (7)
The distribution of phosphine ligands within the polymers depends
upon the type of polymer. The distribution is quite uniform in gel-type
«2% cross-linking) polystyrenes but concentrated mainly at and near the
surface in the macroreticular beads. A disadvantage of the highly cross-
linked beads is that one is unsure of the precise phosphine distribution.
Even though the catalyst distribution in macroreticular beads is limited, it
is still possible to achieve high levels of catalyst loading because a high
percentage of the phenyl residues are actually surface-exposed in
macroreticular beads.
Attachment of the metal complex to the phosphinated polymer may
be accomplished thermally or photochemically (Scheme 1).
Use of polystyrene as a support is limited by the mechanical and
thermal stabilities of the support. Beads may fragment in stirred reactors.
The upper thermal limit for most applications is 1S0-170°C.
While polystyrene is easily the most common organic support, alterna-
tives are available. These include the use. of polymethycrylate allyl
chloride/DVB, (16) polyphenylene-isophthalamide, (17) soluble poly-
siloxanes, (18) and polyvinyl chloride. (7)
The use of oxide materials such as silica, alumina, silica-alumina, and
molecular sieve zeolites offers an alternative support to organic polymers.
These thermally stable supports operate without swelling, an advantage
POL YMER-BOUND PHOSPHINE CATAL YSTS 447

Fe(COl, 8
hp
Isomerization, Hydrosilylation

RhCl(PPh 3 l 3 9
®-PPh 2hRhCI
Hydrogenation

Pd(PhCN)2CI 2 10

Hydrogenation

H 4Ru4(CO)12 11
®-PPh2Ru4H4 (CO It 1
Hydrogenation
-

Ni(CO)2(PPh 3 h 12
®-PPh2hNi(COh
Oligomerization

1) CO 2(CO)8 13
2) heat
Hydroformylation

RhH(CO)(PPh 3 h 14
®-PPh2)xRhH(CO)(PPh3)y
Hydroformylation

Mo(PPh 3 ),(NO)2CI 2 15
®-PPh2Mo(NOhCh
Metathesis

Scheme 1. Illustrative Catalyst Uses

from an engineering point of view. Many metal complexes are attached


directly to these supports, but that is not the subject of this review. A
phosphine may be grafted onto a support without great difficulty, and this
is illustrated by the approach of the research group at British
Petroleum(19-24):

~
Si-OH
;
Si-OH + (EtOhSiCH 2CH 2PPh 2 --.
\
Si-OH
~

The metal is then attached. An alternative procedure involves coordination


of the metal to the phosphine prior to the anchoring of the ligand. For
448 NORMAN L. HOL Y

some applications, it has proven advantageous to remove unreacted surface


silanol groups (through addition of trimethylchlorosilane, for example)
prior to the use of these materials as catalysts. (20.24)
Other research groups have adopted similar procedures for anchoring
catalysts to silica (25-27) and to 'Y- alumina, glass, and molecular sieve
zeolites. (27)

2.2. Polymerization of Phosphine Monomers

For support materials one is not limited to preformed products. The


support can be synthesized from phosphinated monomers. A real advantage
in this approach is that the uncertainties in understanding the support
structure are reduced. Examples of phosphinated monomers include vinyl-
diphenylphosphine, allyldiphenylphosphine, and p-diphenylphosphino-
styrene; these are copolymerized with styrene (and sometimes divinylben-
zene). (28-34) By varying the amounts of monomers a wide range of ligand
loadings and cross-linking densities are possible.

-+ ®-Q-PPh 2

The flexibility of the polymer chain may be reduced by selecting a


monomer that yields an unsaturated polymer. An example of a more rigid
material is the example of the polymer formed by heating solid 1,6-bis-
(toluenesulfonate)2,4-hexadiyne at 60°C and then treating it with lithium
diphenylphosphide (35):

Stille has prepared rather intricate hydrophilic polymers for use as


· I supports In
ch Ira . reactions
. d '
eSlgne d to In
. d uce asymme t ry. (36,37) Th e fi rs t
POL YMER-BOUND PHOSPHINE CATAL YSTS 449

of these was based upon 2,3-0-isopropylidene-2,3-dihydroxy-l,4-


bis (diphenylphosphino )-butane (mop):

azobis(isobutyronitrile)
(AIBN)

A more recent chiral support incorporated a pyrrolidine bisphos-


phine(38):

+
r'---
C0 2 CH 2 CH 2 0 2 C

r'---

CH 3 CH
I I 3

CH-CH21t----C1- CH 2 - - - - t i - - C - - - - CH 2

~=O I ~H3
I'(N'J C02CH 2 CH 2 0H C0 2 CH 2 CH 2 0 2 C-C

Ph 2 P ~ ~H2
PPh 2
0.05 0.85 0.10

It is clear from extant hydrophobic and hydrophilic polymers that a wide


range of phosphine-containing monomers could be prepared.
450 NORMAN L. HOL Y

3. PHYSIOCHEMICAL CHARACTERIZA TlON OF CA TAL YSTS

One of the major difficulties with conventional heterogeneous catalysts


is their characterization; certainly the problems posed by polymer-anchored
catalysts are very similar. It would be fair to say that in some respects
polymer-bound catalysts are rather poorly characterized as heterogeneous
catalysts. Such basic determinations as total surface area, pore volume,
and the pore distribution have been but rarely measured. It is often assumed
that the mild conditions involved in catalyst preparation do not alter
significantly the support structure. This assumption is not without some
foundation. The textural properties of several styrene-DVV copolymers
having from 10 to 60% cross-linking were examined by Hetfleis et at. (39)
Surface areas for the organic polymers ranged from 25 (10% cross-linking)
to 247 m 2 g -1 (60% cross-linking). Upon chloromethylation, treatment
with dimethylamine, and attachment of [Rh(COhCI]2, the surface areas
became 58 and 287 m 2 g-t, respectively.
While the overall physical properties of the polymers have received
rather cursory treatment, more attention has been devoted to the charac-
terization of attached ligands and the coordination of the supported metal
complexes. It is clear in this respect that polymer-anchored catalysts are
more easily characterized than are conventional heterogeneous catalysts.
Interpretations of catalyst performance rely upon both analysis of the
polymer catalyst and a thorough knowledge of the homogeneous situation.
Information about the number of anchored ligands bonded to an
attached metal complex is obtained normally from bulk properties, typically
through elemental analysis. Supportive data for assignment of catalyst
structure may also be garnered from the number of phosphines displaced
during the attachment step, (40) and from molecular-weight determina-
tions. (7,41,42) Additional insights concerning coordination and matrix effects
are obtained by various spectroscopic tools.
JR. Most work is with supported carbonyl complexes. (13,42,43,44) Argu-
ments for proposed structures are based upon close agreement between
absorptions by polymer-bound catalysts and soluble models. The polymer
matrix appears to have little influence on band energies or intensities,
though it should be added that the subtle differences generally observed
have not been interpreted. It is used rather infrequently for detection of
metal halogen bands, (10) since these are normally too weak to be seen.
1H nmr. Doskocilova (45) et at. compared the spectra of a lightly cross-

linked polystyrene with that of a linear polystyrene, using both conventional


and magic angle rotation, and 13Cnmr. The magic angle spectra of the
cross-linked polymer were almost identical with those of the linear polymer
in CCl4 over a broad temperature range. Furthermore, the 13 C spin-lattice
relaxation time (T 1) and nuclear Over hauser-effect parameters were equal
POL YMER-80UND PHOSPHINE CATAL YSTS 451

for both polymers, indicating little difference in the dynamics of internal


motion.
31 P nmr. While 31 p nmr would appear superficially to be the ideal

probe into the chemistry of polymer-bound catalysts, in actuality there


would appear to be rather severe limitations on the situations in which it
can be applied usefully. There are just two major contributions in this
area. (34,46) Both examined polymers prepared through the polymerization
of phosphine monomers; this avoided the complicating effect of an uneven
distribution of phosphine sites. These studies examined both cross-linked
and noncross-linked polymers.
When noncross-linked poly(4-diphenylphosphinostyrene) was treated
with an equimolar amount of [RhCl(COD)h or [RhCI(C2H 4hJ2 two spectra
were taken: one was accumulated from 0.5-1 hr after adding the metal
complex, the other from 8-24 hr.(34) Even with a phosphine function in
each repeating unit, up to 200,000 pulses were required. The initial spec-
trum was different from the final one, apparently reflecting a 0.5-1 hr
requirement for ligand exchange to reach equilibrium. Furthermore, the
appearance of the spectrum was highly dependent on the Rh/P ratio (Figure
14.1). Spectra from Rh/P of 0.4-0.6 are most simply interpreted as the
consequence of exchange in which apparently all phosphine ligands of this
noncross-linked polymer participate. The spectral parameters of the poly-
mer-bound complexes (Rh/P = 1) agreed very closely (Table 1) with those
of monomeric phosphines, providing strong support for structural corres-
pondence.
Stille et at. found that the greatest decreases in 31 p spin-lattice relaxa-
tion times (T 1) occurred upon polymerization of a phosphinated monomer,

______________ A~ _________ ~ R;VP


---------' '--_____A 0.15- 0.30

"'""--- 0.5

- - - - - - . . . , - - - 0.85

.......- - _ . . - . - - - -_ _ _ 1.0

60 50 o -10 ppm

Figure 14.1. 31 p nmr spectra during stepwise addition of [RhCl(C 2H 4 hJ2 to ®--PPh 2 .
452 NORMAN L. HOL Y

Table 1. 31p nmr Results of Monomeric and Polymeric Phosphines


and Their Complexes

Compound lJ(Rh-P), H2

PPh 3 -5.6

0---Q-PPh 2 -5.3

+53.3d 184.4

+53.5d 184

Ph 3 P-Rh(COD)CI +31.5d 152

®-O PPh 2 -Rh(COD)CI +30.9d 147

and that there were smaller decreases as cross-linking and metal complexa-
tion were introduced. (46) The volume swept by the bound phosphine was,
therefore, smaller than for the soluble material, an observation which has
important implications for asymmetric synthesis and regioselectivity.
Electron Spectroscopy for Chemical Analysis (ESCA). ESCA is a tech-
nique permitting the determination of the oxidation state of the metal. (47-50)
For example, from the palladium 3d binding energies listed in Table 2, it
is clear that the oxidation state of palladium in the polymer catalysts is 2+.
The method is also useful in determining the distribution of ligand and
metal. Peak intensities are proportional to concentration. After making a
surface determination, the polymer can be etched (argon) and the values
determined again. An average for the whole polymer can be made after
grinding the catalyst.
There are two limitations to the method. The most important of these
is that the peaks are rather broad (often 2 eV at the! height). This limits
the usefulness of the technique to a determination of the bulk metal valence.
If the "active" catalyst has a different oxidation state from that of the bulk
metal and is present in minor amounts, it may go undetected. The second
difficulty is that the etching process itself may effect a reduction of the metal.
Extended X-ray Absorption Fine Structure (EXAFS). Using a tunable
X-ray synchrotron, the P leI and P IBr ratios were determined for
POL YMER-BOUND PHOSPHINE CA TAL YSTS 453

Table 2. Palladium 3d Binding Energies!50) (in eV)

Complex 3d 3 /2 3d s /2

®-Q- PPh 2 hPdCl 2 344.0 338.6

®-Q- PPh 2 hPdCl 2 343.5 338.2

(Ph 3 PhPdCI 2 343.6 338.3


PdCh 343.6 338.1
PdQ 340.7 335.5

(Ph 3 PhRhCI and (Ph 3 PhRhBr and polymer-bound versions of these.(S1)


From phosphine/halogen ratios, it was possible to conclude whether the
bulk coordination involved monomeric or dimeric metal species. With
low-crass-linked polystyrene (2% DVB) both halides existed as dimers. In
a polymer with 20% cross-linking, the bromide was present as the monomer
(the chloride was not tested), demonstrating that increased cross-linking
did lead to isolation of metal sites.
Also determined by this method are some interatomic distances (see
Table 3). Metal-halide and metal-phosphorus distances found by EXAFS
for (Ph 3 PhRhCI are in reasonable agreement with those determined crystal-
lographically.

®·PPh,

®- PPh 2 RhH 2 ClPPh 3


3

Table 3. Summary of Interatomic Distances (in A)


Wilkinson's 1 Hydrogenated
Polymer-bound polymer-bound
X ray EXAFS Wilkinson's 2 Wilkinson's 3

Rh-Cl 2.376 2.35 2.33 2.29


Rh-P J 2.214 2.23 2.23 2.20
Rh-P 2 2.326 2.35 2.16 2.38
454 NORMAN L. HOL Y

Electron Spin Resonance (ESR). ESR has received but limited use to
establish metal oxidation states. (7.52-56) It was shown, for example, that
norbornadiene complexes of Rh(I) on a phosphinated polystyrene-DVB
formed low levels of Rh(II) during the course of the hydrogenation of
ketones and olefins. Supported RhCh was assigned a structure ®-
PPh 2 Rh(II)Ch following ESR and ESCA characterization. (47)
X-Ray Microprobe Analysis. This technique aids in determining the
distribution of metal and ligand in a polymer-bound catalyst. (57.58) In highly
cross-linked beads, it was observed that while chloromethylation proceeded
homogeneously throughout the polymer beads, phosphine and metal
attachment depended upon polymer morphology. Polymers with large pores
(1300 A) allowed phosphination and complex formation to proceed
throughout the polymer beads. In polymers with small pores «50 A),
penetration was limited to the outer bead areas. (58) In attaching the metal
to the support, it was found that the metal was distributed in the outer
portions of the bead if a deficiency of the metal was used. With excess
metal the distribution was similar to that of the phosphine groups. (57)

4. INFLUENCE OF THE SUPPORT

In designing a catalyst it is certainly intended that the support will be


inert at the conditions of the desired reaction. Normally this has been
observed. This does not mean, however, that the support does not play an
integral role in establishing the activity of the metal center. To state the
obvious, the support is part of the ligand structure and exerts significant
influence on performance at the phosphine-metal centers. This is not at

Table 4. Influence of the Support in the


Hvdrogenation of a-Olefins. (87) Rh(IJ-phos-
phines
Relative
Phosphinated support rate

None a 1.0-2.3
Amberlite XAD-2 0.14
Biobeads SX-2 0.05
Dowex 50 0.20
PVC 0.15

a Homogeneous catalyst.
POL YMER-BOUND PHOSPHINE CATAL YSTS 455

Table 5. Influence of the Support in the


Hydrogenation of Cyclopentene. (4) Rh(l)-
phosphines

Relative
Phosphinated support activity

1.0
Soluble polystyrene b 0.91
Insoluble macroreticular
Polystyrene 0.74

a Homogeneous catalyst.
b M = 250,000.

all surprising since we know from homogeneous catalysis that even seem-
ingly minor changes in either electronic or steric factors can impact dramati-
cally on reaction rates and/ or product distributions. It would be anticipated,
therefore, that the polymer matrix in polymer-bound catalysts would have
a significant role in determining its properties (see Tables 4 and 5).

4.1. Changes in Selectivity


4.1.1. Regioselectivity

If a reagent must diffuse into a polymer to react with metal sites, it is


to be expected that the framework of the catalyst might exert some influence
upon the orientation in which the reagent approaches the catalytic site.
This concept was tested with steroids(59) having two potential sites of
reactivity:

Limiting the reaction to the addition of one mole of H2 showed that the
heterogenized catalyst was 2-4 times more selective toward side-chain
hydrogenation [polystyrene 2% DVB, anchored RhCI(PPh 3 hJ in benzene
than the homogeneous rhodium complex. In the solvent benzene: ethanol
(1: 1), a poorer swelling solvent, the selectivity was enhanced.
456 NORMAN L. HOL Y

4.1.2. Size

The steric influence exerted by a polymer matrix is also manifested in


the hydrogenation of ole fins of varying sizes. (9) Cyclohexene was reduced
32 times faster than the bulky olefin a2 -cholestene with polymer-bound
RhCI(PPh 3 h, whereas in the homogeneous reaction the rate difference was
less than 2. The support in this case was 200-400 mesh polystyrene beads
(1.8% DVB) with 10% of the aromatic rings chloromethylated.
This size selectivity may not be entirely attributable to macroscopic
polymer properties, e.g., small pore channels. It may also be a matter of
the bulkiness of the ligand itself, the ligand in this case being represented
not just by the immediate environment around the phosphine, but also the
more distant polymeric backbone. Evidence for a steric influence of distant
groups comes from studies on phosphines of the type(10):

H(CH 2 CH 2 CH 2 CH 2 J. OPPh 2 (n = 0, 1, 2, 3)

In this case, the tail simulates the polymer backbone. When the Wilkinson's
catalyst analog of this ligand was prepared, it was found that the rate of
hydrogenation of cyclohexene increased when n was increased from n = 0
to n = 3. This increase was ascribed to the electron-donating effect of the
para-alkyl group. Use of the bulkier olefin camphene instead of cyclohexene
shows a slowing of the reaction with increasing length of the tail, lending
support to the notion that quite distant groups may exert a measurable
steric influence. Application of this finding to polymeric materials should
be disciplined by the understanding that polymers are conformationally
less flexible than the examples just discussed.

4.1.3. Chemical Selectivity

The selectivity of catalysts with respect to various functional groups


may change upon attachment of a metal complex to a support. With soluble
PdClz(PPh 3 h. conjugated dienes are more readily hydrogenated to
monoenes than isolated dienes. The reverse is observed for the supported
catalyst(1o,61l; the basis for this reversal is not yet clear.

4.1.4. Polarity

A nonpolar reagent in a mixture of a polar solvent and a hydrophobic


catalyst will not be distributed uniformly throughout the mixture. Because
of the polarity difference between catalyst support and the solvent, the
nonpolar reagent will be concentrated to some extent within and near the
POL YMER-80UND PHOSPHINE CA TAL YSTS 457

polymer. A polar reagent, on the other hand, will be dispersed in the


solvent. Capitalizing upon the inequality in distribution and the proximity
of the nonpolar reagent to the catalytic sites, it has been shown that the
relative rate of cyclohexene hydrogenation was enhanced over that of allyl
alcohol(62); another study found the same trend for cyclopentene and
acrylonitrile. (4)

4.2. Asvmmetric Induction

While the choice of ligand for inducing high-optical yields has been
something of a matter of chance, (63) success is known to depend heavily
upon steric factors.(64) One intuitively would expect that a polymer-bound
catalyst would be more sterically hindered than a homogeneous analog,
thus providing higher optical yields than are available from soluble catalysts.
Optical yields often comparable to those achieved with soluble catalysts
were observed in the hydrogenation of a- amino acid
precursors(54,37,65,66,67,68,69) and the hydrosilylation of ketones.(67,78) Much
poorer optical yields were observed in the hydrogenation of a-substituted
styrene,(67) unsaturated carboxylic acids,(71a) and ketones,(71b) and during
hydroformylation. (3a,46,72)
The simplest explanation for the generally lesser optical yields from
polymer-bound catalysts lies in two considerations. The polymer backbone
would increase the steric hindrance at the posterior of a ligand, but posterior
variations have little effect upon optical yields. (73) What is more important
are the steric influences in the immediate vicinity of the metal-phosphine
coordination. The motion of the mop phosphines is known to be limited
considerably in a polymer, thus reducing the volume swept per second.(46)
This, in turn, reduces the effective bulkiness of the ligand, a factor of
importance in inducing asymmetry.

4.3. Matrix Isolation

One facet of structure that has aroused considerable attention is the


matter of the flexibility of the polymeric chains and the pendant phosphines.
The possibility of interaction between two or more sites along the same
or different chains is partially dependent upon this. Intuitively, one would
think that as the degree of cross-linking increased the flexibility of the
chains would decrease, effecting a restriction in interactions between sites.
This has important consequences both in synthesis and catalysis. In catalysis,
site-site interactions can lead to, for example, dimeric metal species that
may have much reduced activities in comparison to those of monomeric
complexes. One obviously would seek to design a matrix that would avoid
site-site interaction for such catalysts. Quite a number of experiments have
458 NORMAN L. HOL Y

been designed that test the conditions under which site-site interactions
occur.
The growing interest in the synthesis of macrocyclic ring systems has
resulted in novel synthetic methods. A severe restriction in macro cyclic
ring formation is that precursors to large rings often react intermolecularly
instead of intramolecularly. Polymers were recognized to provide the
opportunity of isolating such precursors within a matrix, thus forcing the
ring closures to proceed intramolecularly. However, when such experiments
were carried out the enhancement in intramolecular ring closures was often
disappointing and the main product was still derived from intermolecular
(site-site) reactions. (74)

WCN+O o CN
o
II base
®-CH 2 S-C-(CH 2 )sCN ----.

CN
minor major

The example shown above is by no means atypical. Treatment of a polycar-


boxylic acid with dicyclohexylcarbodiimide (DeC) under forcing conditions
effected dehydration even with a 20% DVB-polystyrene polymer.(75)

Dee
---+

There is other evidence, however, which demonstrates that matrix


isolation does reduce or even prevent site-site interaction. One very
impressive experiment dealt with the generation of benzyne on a poly-
styrene support having but 1% cross-linking.(76)
When generated in solution the coupling of benzyne is diffusion con-
trolled, but when benzyne was generated on a polymer, the lifetime was
extended to over one minute. Moreover, there was no evidence for ben-
zyne-benzyne coupling.
Another demonstration of the isolation of sites, in this case involving
metal centers, came from the EXAFS experiments by Reed.(51) On a 2%
POL YMER-80UND PHOSPHINE CATAL YSTS 459

cross-linked polymer, supported Rh(PPh 3 hBr was present in the dimeric


state, whereas on a polymer with 20% cross-linking the monomer prevailed.
The seeming contradiction between these various experiments may be
resolved if site isolation is considered to be partially a kinetic phenomenon.
For slow reversible reactions, such as anhydride formation or condensation
reactions, the polymer chains have sufficient time to exercise several degrees
of freedom, permitting site-site interaction. For reactions that occur over
shorter periods, such as the decomposition of benzynes, the probability of
close contact between reactive centers is greatly reduced. Metal dimer
formation was obviated simply because there was no important thermody-
namic driving force.
The principle of matrix isolation is important in catalysis if a catalyst
can decompose via site-site interactions. In the dimerization-methoxyla-
tion of butadiene, catalyst decomposition occurs through metal agglomer-
ation(77):

+
isomers

A polystyrene with just 1 % cross-linking was reasonably effective in reduc-


ing catalyst decomposition in this reaction.

4.4. Coordinative Unsaturation


It is desirable at times to create coordinative un saturation in metal
complexes to either permit the coordination of species that are very poor
ligands and thus cannot readily displace ones already in place, or for reasons
of rate enhancement. Examples of the former effect are seen spectacularly
in the works of Crabtree(78) and Bergman,<79) in which alkanes are coordin-
ated by transiently unsaturated metal centers. (There are no comparable
studies of heterogenized systems.)
The case for coordinative unsaturation on polymers is rather weak
and comes principally from the enhanced rates of hydrogenation observed
for polymeric catalysts. Kaneda et al. supported PdCh on a linear phosphin-
ated polystyrene, and the hydrogenation of nonconjugated dienes was
examined. (49) The supported catalyst was active even under conditions
where PdCh(PPh 3 h was inert. ESCA studies established a 2+ oxidation
state for the metal. The increase in hydrogenation activity was considered
to possibly have its origin in un saturation at palladium.
A second observation of hydrogenation activation superior to that of
the homogeneous situation was found for Ir(CO)CI(PPh 3 h on phosphinated
460 NORMAN L. HOL Y

polystyrene-l % DVB with P IIr ratios varying between 3 and 22.(80.81) At


high ratios the homogeneous and supported complexes displayed similar
activities, but at low ratios, i.e., 3 or 4, the supported catalyst displayed
rates much enhanced over those of the soluble catalyst. This difference
was attributed to the lesser availability of phosphine from the polymer,
permitting a shift in the equilibrium between (4) and (5) toward the
unsaturated (5). Such a shift would indeed enhance the rate, since it is (5)
that adds hydrogen in the rate-determining step:

®-PPh 2 hIrCI(CO)(olefin) ~ ®-PPh 2 + ®-PPh 2 IrCI(CO)(o\efin)


4 5

A third experiment in which the rates of reaction for polymer-


supported catalyst were more rapid than those of the soluble analog was
one already mentioned, the dimerization-methoxylation of butadiene. (77)
Rates were particularly fast with catalysts having low-ligand loadings (on
a polystyrene 1% DVB) and low P IPd ratios.
While coordinative unsaturation is not widely documented for poly-
mer-bound catalysts, it could be argued that few authors have looked for
it. Few investigations have systematically varied P Imetal ratios, for
example, so the phenomenon may be more widespread than suspected.
Spectroscopic examinations would be of value. It is a bit curious that the
instances in which coordinative unsaturation was argued employed poly-
mers either lacking or having very low levels of cross-linking! One certainly
would expect the phosphines in these polymers to be highly mobile. The
fact that published accounts of coordinative unsaturation employed non-
cross-linked or lightly cross-linked supports may not reflect a requirement
for the observation of coordinative unsaturation, i.e., the use of low-cross-
linked polymers, but rather may simply reflect the fact that most studies
have employed low-cross-linked polymers. Based upon the better docu-
mented concept of matrix isolation, one would anticipate that more highly
cross-linked polymers would more likely display coordinative unsaturation.

5. REACTIONS OF OLEFINS AND DIENES

5.1. Isomerization

Numerous catalysts that are used for hydrogenation or hydroformyla-


tion are also active isomerization catalysts. These catalysts require the
presence of hydrogen to be activated, probably to form a metal hydride.
Carbonyls of cobalt,(13) iron,(8) and osmium(82) are activated in this manner
and effect hydrogen shifts.
POL YMER-BOUND PHOSPHINE CATAL YSTS 461

Irradiation of Fe(CO)5 and Fe3(CO)12 supported on phosphinated


styrene-DVB at 25°C created isolated, coordinatively unsaturated metal
atoms that are catalytically active for olefin isomerization and hydrosilyla-
tion. (8) Additional observations include: 1) the extent of phosphine substitu-
tions affects both the initial catalytic activity and the isomeric ratio of the
products; and 2) the photocatalysis required continuous irradiation.
Several potentially interesting reactions apparently have not received
attention. Several unsaturated steroids are obtained via homogeneous
isomerization with RhCh. (83) The interest in examining a similar isomeriz-
ation with polymer-bound catalysts lies in the potential of an altered
isomeric ratio.
The greatest potential for industrial applications may lie with nickel
hydride catalysts. Homogeneous catalysts may be prepared from
Ni[P(OR 3hJ4 or by reduction of NiS 2 (PR 3h with alkyl aluminum com-
pounds. (84,85) Carbon skeleton rearrangements are known with these
complexes. For example, cis-l,4- hexadiene is converted to trans- 2- methyl-
1,3-pentadiene and 2,4-hexadienes at room temperature(84):

~--~+~
Apparently there are no similar studies with polymer-bound catalysts.

5.2. Hvdrogenation

Mechanistically, hydrogenation places but limited demands on metals


and ligands, so it is not at all surprising that a great variety of complexes
have demonstrated activity. Before considering some specific examples of
this chemistry, a few more general characteristics are listed.
a. Polymer-bound catalysts are fashioned after homogeneous analogs.
Those homogeneous catalysts having the highest activities also
display high activities when bound to a polymer.
b. Polymer-bound catalysts are generally less active than
homogeneous analogs. The lesser activity may be related to reduced
catalyst accessibility and/or diffusion control.
c. Hydrogenation studies have focused on simple hydrocarbon
unsaturation (alkenes, dienes, trienes). Ketone and arene hydro-
genation are known.
Interest in the hydrogenation reaction is not, however, limited to a
documentation of the organic transformations that are possible. This reac-
tion has also been used more than any other to probe the characteristics
of polymer-bound catalysts to determine the details of the effects of catalyst
attachment.
462 NORMAN L. HOL Y

5.2.1. Rhodium Catalysts

The most commonly supported catalyst is the heterogenized version


of Wilkinson's catalyst: RhCI(PPh 3 h. The mechanistic details of this highly
active homogeneous catalyst are so well established that the influences
supporting the catalyst are readily accessible. Many of these effects have
been discussed already.
The activity and selectivity of the bound Wilkinson catalyst parallels
in many ways the homogeneous system. Terminal ole fins are hydrogenated
more rapidly than internal ole fins, cis- ole fins react faster than trans- olefins,
and olefins are reduced faster than acetylenes. (86) There were some differen-
ces with the sterically hindered 1(7)-p-menthene: the polymeric catalyst,
for example, produced appreciably more isomerized material. Furthermore,
the attached complex was more selective in the hydrogenation of 1,3-
cyclooctadiene.
The activity of a catalyst depends very much on the type of support
employed. This was documented for several types of phosphinated polymers
to which RhCI(PPh 3 h was bound. All the polymers listed in Table 4 possess
small pore sizes: The macroreticular XAD-2 has 90 A pores, the others
lack permanent pore structures. For these catalysts, activity is an order of
magnitude lower than for the homogeneous catalyst. When a polymer with
large pores (1000 A) was employed as a support, the rate was almost up
to that of the homogeneous complex. It was concluded from these observa-
tions that the lesser activity of the catalysts possessing low cross-linkings
was attributable to diffusional effects. Alternative explanations are possible.
It could be argued that in low-crass-linked polymers the form of the metal
complex was different from that in more rigid materials, that in the former
the metal was present to some extent as dimer. Justification for this
argument comes fram the EXAFS study by Reed et al. In homogeneous
reactions, the dimer of Wilkinson's catalyst displays a lower activity than
the monomer. One could also argue that in the macroreticular support
there is a degree of coordinative un saturation caused by the rigidity of the
polymeric structure. But the simplest explanation, that the lesser activity
is based upon diffusional factors, would seem to be the most reasonable
choice. This is suggested by the very similar rates for three very different
catalysts: homogeneous RhCI(PPh 3 h, a noncross-linked polystyrene, and
a macroreticular polystyrene with 1000 A pores. In these cases the polymers
would permit ready access to catalytic sites. (It should be added that
macroreticular supports with smaller pores are more prone to diffusional
limitations. )
High activities can be obtained for supported catalysts having a low
degree of crass-linking if the particle size is kept quite small(12\ use of
small particles minimizes diffusional effects.
POL YMER-80UND PHOSPHINE CA TAL YSTS 463

One of the important factors that contributes to such things as catalyst


activity and metal loss is the ratio of phosphine to metal. In a study by
Bernard et at. it was shown with soluble polystyrene that upon reaching a
P/Rh ratio of 4/1 the catalyst was deactivated.(4) The behavior resembles
that of the homogeneous catalyst in its sensitivity to excess phosphine and
suggests that the support was sufficiently flexible for the metal to become
coordinated to three phosphines; the presence of the fourth phosphine
simply inhibits the dissociation of the complex so that hydrogen might add.
On the other hand, there was little change in activity for a macroreticular
support having P/Rh between 2/1 and 10/1.
Metal loss from the polymers may occur through various means includ-
ing via the oxidation of phosphine to its oxide. This is a facile process and
is the reason that feedstocks must be purified so carefully of oxygen. Since
the resulting phosphine oxide is a poor ligand, the metal is released from
the oxidized binding site. If the polymer contains a high P /metal ratio, the
metal is more likely to be retained at another site within the polymer. For
macro reticular polymers, which tend to concentrate the metal near the
outer portion of the support as opposed to being evenly dispersed
throughout the polymer, metal loss is more likely.
As we have just seen, the activities of polymer-bound catalysts are
generally somewhat less than those of corresponding homogeneous com-
plexes, but there are exceptions beyond those already discussed. Two
reports of enhanced activities are claimed using phosphinated silica. By
one report, attaching [Rh(C2H 4hCl]2 resulted in a catalyst displaying hydro-
genation activities of up to fifty times those for homogeneous
RhCl(PPh 3 h. (88) The behavior of catalysts prepared from supporting cobalt,
nickel, rhodium, or palladium salts on a silica phosphinated with
Ph 2 P(CH 2 hSi(OMeh was reported by Kochloefl et al.(89) Hydrogenation
rates for cyclohexene were 102 _10 4 times faster than for the unattached
salts. As a test for the presence of metal the CO uptake was measured and
found to be very low.
The origin of these enhanced activities is not clear. Some investigators
argue that all cases of enhanced activity are based upon the presence of
small amounts of zero-valent metal. This point is extremely difficult to
prove either way. Given the number of reports of enhanced activity in a
rather broad variety of reactions, using both phosphine and other ligands,
it would appear that such an argument is inadequate.
There is mixed evidence with regard to the question of whether
supported catalysts are more immune to poisoning than are the soluble
complexes. After supporting [RhCI(COD)h on a phosphinated silica, the
activity for I-hexene hydrogenation was lowered slightly in the presence
of thiophene, but more markedly by n- butyl mercaptan. (20) A further effect
of the mercaptan was the stabilization of the rhodium to prevent metal
464 NORMAN L. HOL Y

formation. By contrast, the activity of RhCl(PPh 3h was greatly reduced


by the presence of mercaptans. In another study by this same group at
British Petroleum, rhodium(I) carboxylates were anchored to phosphinated
polymers. This catalyst displayed as much sensitivity to sulfur as the
homogeneous catalyst. (90)
Just how important the support can be in establishing activity is
illustrated by the example of phosphite polymers of polymethallylalcohol!
CH 3 CH 3
~ CH 2 - 1 CH 2- <[}n
o 0
I I
PPh 2 PP 2
'Rh/
oc/ 'Cl
While catalysts derived from the atactic polymer were active, those from
the syndiotactic material were not. (91) Homogeneous rhodium-phosphite
complexes are not active hydrogenation catalysts.
More polar functional groups such as ketones are hydrogenated
effectively with homogeneous cationic rhodium(92) and ruthenium(93) com-
plexes. Polymer catalysts active for both olefin and ketone hydrogenation
activity were synthesized from ionic precursors such as
Rh(norbornadiene)(acac) + HCI0 4 or Rh(norbornadiene)(PEt3h +
CI0 4 -. Two types of coordination were observed, depending on the method
of attachment(53-55):

Modification of the phosphine to compare activities of ®-PPh 2 with ®-


PPhMe demonstrates that the hydrogenation of ketones is faster with the
latter, more basic, phosphine. Alkene hydrogenation is slower with these
more basic ligands. Incorporation of a very bulky alkyl group, P-
PPh(menthyl), results in a dramatic increase in the loss of rhodium. Similar
losses upon the incorporation of a bulky ligand into the phosphine do not
occur for olefin hydrogenation; the losses during the ketone reduction are
partially due to the high-coordinating ability of the product alcohol. Hydro-
genation of cyclohexene is highly solvent dependent and is rapid in THF,
slow in CH 2Ch. Some Rh(II) is formed during the reaction, but this was
considered to be inactive. During olefin hydrogenation there was an induc-
tion period, and this was ascribed to the hydrogenation of the initially
coordinated diene.
POL YMER-80UND PHOSPHINE CA TAL YSTS 465

Rhodium trichloride on a noncross-linked phosphinated polystyrene


was characterized by esr and ESCA techniques as having a ®-PPh 2 Rh(II)C}z
structure.(47) While a variety of monoenes and conjugated and nonconju-
gated dienes were reduced, there was but slight activity with alkynes, cyano,
keto, and nitro groups. Catalyst activities were highly solvent dependent;
the highest rates were in alcohols, the least activity was in hydrocarbon or
chlorinated hydrocarbon solvents.

5.2.2. Iridium

Vaska's complex IrCI(CO)(PPh3 h was attached to microporous poly-


styrene (1 % DVB) with different phosphine loadings and P jIr ratios. (80,81,94)
1,5-Cyclooctadiene was hydrogenated at the rather high temperature of
170°C. The results showed that when P jIr was <1 the rates were appreciably
faster with the supported catalyst; when the ratio reached or exceeded ¥
the homogeneous catalyst was more active. When the rate of hydrogenation
at 170° was compared with that at 80°, it was observed for some loadings
that the rate was actually lower at the higher temperature. At higher
temperatures, it was considered that the effects of matrix isolation were
overcome and more phosphine became available for coordination, reducing
the concentration of the 4-coordinate intermediate involved in the catalytic
cycle.

5.2.3. Palladium and Platinum

Use of rhodium catalysts to effect the hydrogenation of polyenes results


in high conversions to saturated products. On the other hand, if palladium
or platinum is employed, the hydrogenation can be stopped at the monoene
stage. This has important implications, particularly for the partial hydro-
genation of vegetable oils. The linoleic and linolenic residues in vegetable
oils are prone to oxidation, so from the standpoint of stability it is desirable
to hydrogenate them. But to avoid the build-up of low-density cholesterols
in humans, it is important to retain cis- monoene units. It is not surprising
that much of the research on palladium and platinum complexes has been
in the context of vegetable oil hydrogenation.(1o,61) The complexes most
commonly supported are PdC}z(PhCNh and PtC}z(PhCNh, Hydrogenation
of soybean oil methyl ester proceeds fastest in alcohol solvents, but the
highest selectivity to monoenes occurs in dichloromethane.
It is very clear that differences exist between the attached and the
homogeneous model of the attached complex, Pd(PPh3 hCI 2 • Ir studies
show that the polymer-attached bis(phosphine)dichloropalladium is in a
trans configuration, whereas the homogeneous complex exists as both cis
and trans isomers. (10) Furthermore, the rate of hydrogenation by the
466 NORMAN L. HOL Y

homogeneous complex is not accelerated by alcoholic solvents. Also,


whereas stannous chloride activates the homogeneous complex by produc-
ing coordinative unsaturation, the polymeric version is deactivated by
SnClz. Analogous (supported) platinum complexes are activated by stan-
nous chloride. Homogeneous Pd(PPh3hClz hydrogenates nonconjugated
dienes faster than conjugated ones, while the supported version displays
the opposite selectivity.
Generally, the activities of polymer-bound palladium catalysts are less
than those of unsupported ones, but here too there are exceptions. Kaneda
et at. reported that PdClz, on phosphinated polystyrene was more active
than the homogeneous situation.(49) Rates were very solvent dependent-
hydrogenation of styrene was slow in dimethyl sulfoxide; optimum activity
was obtained in solvents of moderate coordinating ability (see Table 6).

5.3. Dim eriza tion, Oligomerization, and Polymerization

5.3.1. Dimerization

Investigations with polymer-bound catalysts are very limited and are


patterned generally after homogeneous analogs. Thus, the linear dimeriz-
ation of butadiene was effected by the homogeneous catalyst
NiBr2(PPh3h/NaBH4.(40) The polymer-bound version gave essentially the
same results, though its activity ceased after about 1500 turnovers.

__
lO_OO_C_...
, ~
EtOH-THF
95%

Nickel-based Ziegler catalysts are very effective for propylene dimeriz-


ation. (116) The unsupported catalyst is prepared by mixing NiClz, Et3Al,
and butadiene in chlorobenzene to yield a C 12 1T- allyl complex of nickel.
A phosphine is then added, followed by liquid propylene at 15 atm pressure.
A mixture of n- hexenes, 2-methylpentenes, and 2,3-dimethylbutenes in
85-90% yield are produced rapidly at 30-40°C. Apparently this reaction
is not reported for a polymer-bound version. This would be of interest
since it is observed that the product distribution is quite sensitive to the
nature of the phosphine.

5.3.2. Cyclodimerization and Cyclotrimerization

The nickel carbonyl catalyst, Ni(CO)z(PPh 3h. is excellent for the


cyclooligomerization of butadiene, providing a mixture of 4-vinyl-
Table 6. Hydrogenation with Phosphine-Bound Catalysts
~
r-
Metal complex
Support Ligand supported Illustrative reactant Product Reference ~
~
Soluble ®--PR 2 CoCI 2 30 tlJ
a
polystyrene c::
®--PR 2 PdCI 2 1,5-COD COD 49 ~
®--PR 2 [RhCI(C 2H 4 h]2 Cyclopentene Cyclopentene 4 ~
aC/)
®-PR 2 [RhCI(CsH 14 hJz Cyclohexene 95
®--PR 2 [RhCI(COhJz ~
."
®--PR 2 Rh(acac )(COh Cyclopentene 3 ~
®--PR 2 RhCI(PPh 3 h
i;!
r-

Styrene/DVB ®-PR 2 NiCIz/NaMH4 96 ~


til
®--PR 2 H2PtCI 6 1-Heptene 16
®--PR 2 PdCI 2(PhCNh Soybean oil methyl ester Selective to monoene 10,97
®--PR 2 RhCI 3 Olefin 82,98
®--PR 2 RhCI(PPh 3 h Cyclolefin trienes, dienes 14,12,9,62,100,99
®--PR 2 [Rh(diene)(PR 3 ht Olefins, ketones 4,43,53,54,55,101
®--PR 2 [RhCI(olefinhJz
®--PR 2 Rh(acac)(olefinh Alkenes 43,53,102
®--PR 2 Rh(acac)(COh 102
®--PR 2 RuCIz(COh(PPhhh 1,5-COD Cyclooctene 103
®--PR 2 Ir(COD)(acac) Cyclohexene 53
0)
'-.j
""
®-PR 2 IrCI(CO)(PPh 3 h Diene 80,81,94
.....
&l

Table 6. (Continued)

Metal complex
Support Ligand supported Illustrative reactant Product Reference

Styrene/DVB ®-PR 2 Rh 4(CO)12 Arenes 104


®-PR 2 Rh6(COlt6 Arenes, olefins 104-106
®-PR 2 H 4Ru4(CO)12 Ethylene 11,107
®-PR 2 Ir4(CO)12 108, 109
®-PR 2 Fe2Pt(C°lt2 Ethylene
®-PR 2 RuPt 2(CO)s Ethylene 11
®-PR 2 Pd(PPh 3 )4, Pd(OAch Un sat. esters 110
®-PR 2 Ni(COh(PPh 3 h + Dienes, trienes 14,111
RhCl(PPh 3 h <:
Pdy(imino- ®-PR 2 [RhCl( olefinhh Cyclohexene 112 ~
methylene ~
<:
Poly-l,6-bis ®-PR 2 [RhCl( olefinhh 1, 3-COD, l-octene, benzene 35 !"
(p-toluene
sulfonate)- ~
,....
2, 4-hexadiyne
""
Polyphenylene- ®-PR 2 RhCl(PPh 3 h Olefin 17
~
r-
isophthalamide
~
Soluble poly- ®-PR 2 [RhCl(olefinhh 1-0ctene 18 ~
ttJ
siloxanes C
Butadiene/maleic ®-PR 2 PdCl 2 113
anhydride H 2PtCl 6
~
36,37,65
~
Chiral ®-DIOP [RhCI(C)D)h a- Acetamidoacrylic C
C/)
methacrylate acid
®-DIOP [RhCl(COD)h a- Acetamidoacrylic 65,38,66,68,114
acid
i
~
Polymethallyl ®-OPR 2 [RhCI(COhJ2 Terminal ole fins 115 i;!
alcohol r-

Si0 2 <@-PPh 2 [RhCl(COD)]2 1-Hexene 20,22


~
~
<@-PPh 2 Alkenes 19
<@-PPh 2 1-Hexene 23
®>-PPh 2 RhCl(PPh 3 h Alkenes 86
®l-PPh 2 Cyclohexene 26

""
m
470 NORMAN L. HOL Y

cyclohexene, cycloocta-1,5-diene, and cyclodeca-1,5,9-triene. For the sup-


ported complex, prepared through phosphine exchange, the product distri-
bution was similar to that of the soluble catalyst but the rate was only
one-third that of the soluble complex. (12,13,94) Both types of catalysts respon-
ded similarly to changes in CO pressure. After 1100-1200 catalytic cycles,
the polymeric catalyst was no longer active-an infrared spectrum revealed
no CO absorptions. There was a slow leaching of metal, but at a rate slower
than that of catalyst deactivation.
A similar Ni( 0) species is derived from bis- (cyclooctadiene) nickel
and a phosphinated polystyrene. (43) The catalyst has little intrinsic activity
in butadiene cyclodimerization to cyclooctadiene and vinylcyclohexene,
but this was enhanced to a level of about 60-100 g-product/g-Ni/hr by
the addition of AlEt 2(OEt). Cyclododecatriene was not produced, indicat-
ing coordination of a phosphine to the nickel throughout the process.
Polymer-bound Ni(COh(PPh3h is active in the cyclotrimerization of
acetylenes in benzene or THF. The lifetime of this catalyst is rather short,
and there are apparently changes of product distribution during the course
of the reaction. (12)
The Shell Higher Oletins, a large-scale process, is a nickel-based
oligomerization of ethylene to linear a-oletinsY17l The catalyst, a
homogeneous one, is a nickel hydride generated by reduction of a nickel
salt in the presence of a chelating ligand such as diphenylphosphinoacetic
acid. A polymer-bound catalyst for this reaction would be attractive if the
oletins in the C lO-C 18 range could be increased above the 40% produced
by the homogeneous version. The homogeneous reaction is performed in
ethylene glycol and, for efficient operation with a polymer-bound catalyst,
it might be necessary to use a hydrophilic polymer.

5.3.3. Polymerization

Work in this area has focused on nonphosphine supports.(2) Polymeriz-


ation of ethylene was reported with TiCl4 on a homopolymer or copolymer
of vinylphenylphosphine; AIR3 was added to activate the catalyst. With a
nickel salt on a phosphinated polystyrene, addition of NaBH4 resulted in
a catalyst active in the oligomerization and polymerization of acctylenic
monomers(118) (see Table 7).

5.4. Addition

Polymer-attached catalysts facilitate addition to oletins:


RCH = CH z + H Y ---. RCHzCH zY + RCH - CH 3
I
Y
H Y = HSiR;, HSi(OR)" HeN
C3
r-
~
~
Table 7. Dimerization, Oligomerization and Polymerization
~
c::
Metal complex
Reactant Product Reference
~
Support Ligand supported
~
C
Styrene/DVB ®-PR 2 NiCI 2/NaBH 4 118 C/)

®-PR 2 NiBr2(PPh3h/NaBH4 Butadiene Linear dimerization 40 ~


~
III
®-PR 2 Pd(PPh 3l4 1,3-Butadiene Dimerization - 77
+MeOH methylation ~
~
r-
®-PR 2 PdCIz Styrene, alkenes Cotrimers 119, 120
Pd(OAcjz Butadiene Octadienyl acetates 121
~
®-PR 2 (fj
®-PR 2 Pd(OAch, PdCIz(PhCNh 1,3-Butadiene Linear dimers 110
Pd(PPh 3l4
®-PR 2 [PdCI( 112 --pinenyl))z 122
®-PR 2 Ni(CODh/ AlEt 2 0Et 1,3-Butadiene Cyclodimers 43
®-PR 2 Ni(CODhl Al 2Et 3Cl 3 123
®-PR 2 Ni(COh(PPh 3h Butadiene Cyclooligomers 12,13,94
Pol y( yin y\Chloride 1 ®-PR 2 CoCI 2/NaBH4 124
Polyglycolphosphate ®-OP(ORh Ni(COl 4 125

-I:>.
~
472 NORMAN L. HOL Y

The additions illustrated in the equation are of commercial interest. Hydro-


silylation is used for the preparation of silicone polymers. Silicone rubbers
are "cured" through addition of silanes, a process that converts the rubber
to a hard material, suitable, for example, as dental cement. The usual
homogeneous catalyst is chloroplatinic acid.(126) For supported RhCh,
conversions for HSiEt 3 addition were very poor when polystyrene was the
support, but improved when the support was a phosphinated allyl chloride-
DVB. (16) Addition of HSi(OR h to I-hexene with this catalyst was efficient.
In a study with dimethylaminomethyl polystyrene!DVB polymer, a
close correlation was observed between pore size and catalytic activityY27)
Polymers with low cross-linking, i.e., those with virtually no large pores,
displayed no catalytic activity. A catalyst with 60% DVB, having a high
percentage of pores over 300 A, was most active. Catalysts with intermedi-
ate cross-linkings displayed intermediate activities. Metal loss was par-
ticularly high for the high-cross-linked support.
The synthesis of adiponitrile commercialized by DuPont requires the
high regioselective addition of two moles of HCN to butadiene:

These reactions are effected by Ni[P(OArhJ4.(128) Apparently no studies


of this reaction with polymer-bound catalysts are reported (see Table 8).

5.5. Reactions with CO


5.5.1. Hydroformylation

The essential mechanistic steps in hydroformylation are addition of a


metal hydride to an olefin, CO insertion into the alkyl-metal bond, and
reaction with hydrogen. Hydroformylation of ole fins with homogeneous
catalysts is a major industrial process:

Butyraldehyde, for example, is produced from propylene and synthesis gas


on a scale of about three million tons per year worldwide. Also commercially
~
r-

~
~
~
c::
~

Table 8. Olefin Hydrosilylation

Metal complex
Support Ligand supported Addition Reference
;
~
);!
Styrene/DYB ®-PPh 2 RhCl 3 1-Hexene + triethoxysilane 16, 129 r-

Styrene/DYB ®-PPh2 PtCIz, H 2PtCl 6 1-Hexene + HSiEt3 16 ~


Cil
Styrene/DYB ®-PPh 2 Fe(CO)s, Fe3(COh2 1-Pentene + triethylsilane 8,130
Polymethyl-acrylate ®-PPh 2 RhCI 3, H 2PtCl 6 1-Hexene + HSi(OEth 16
allyl-chloride/DYB
Styrene/DYB ®-DIOP [RhCl(C 2H 4 hh Acetophenone + HSi(OEth or H 2SiPh 2 67,131

..".
~
474 NORMAN L. HOL Y

important is the production of fatty alcohols via hydroformylation for use


in biodegradable detergents:
CHO
co I
~CHO + /'-...
~
normal branched

Two types of catalysts are of greatest commercial interest: those based on


cobalt and those based on rhodium. Commercial operations are based
typically upon CO 2 (CO)s, even though both higher temperatures and press-
ures are required than for rhodium catalysts. Furthermore, cobalt catalysts
produce greater percentages of the usually less desirable branched products
than do rhodium-catalyzed reactions.
Yet in spite of these negative aspects, the cobalt catalysts are pre-
dominant in industrial applications. The reason is very simple: the cost
of rhodium. Given its cost, it is imperative that the catalyst be recovered
"completely." This consideration was partially responsible for the very
early and intense interest in hydroformylation with polymer-supported
catalysts.
From this broad experience several generalizations are apparent.
a. The activities of polymer-bound complexes are generally less than
activities displayed by soluble catalysts, though the difference narrows
at higher temperatures.
b. As the P: metal ratio increases the n/ b ratio increases. This trend
parallels that observed in homogeneous catalysis and is explained in
terms of an equilibrium between tetra- and pentaccoordinate species 6
and 7. Both electronic and steric factors favor more selective addition
of 7 to form terminal metal intermediates. Within a polymer matrix,
bis-phosphine species analogous to 7 are favored by high P /metal ratios
(see Scheme 2).

~
j
RCH 2 CH 2 Rh(CO)z(PPh 3)z + RhCHCH 3

1
kh(CO)z(PPh3)z
Hz, co H 2 • co

RCH 2 CH 2 CHO RCHCH 3 RCHCH 3


I I
CHO CHO

Scheme 2
POL YMER-80UND PHOSPHINE CATAL YSTS 475

c. To restrict the loss of metal from the support: 1) Keep feeds free of
oxygen. Oxygen converts phosphines to phosphine oxides, thereby
reducing the phosphine content in the support. Phosphine oxides are
poor ligands for rhodium. 2) Use resins with high P: Rh ratios. This
favors bis-phosphine complexes of the attached complexes. 3) Keep
olefin conversions at modest levels. Aldehyde is able to coordinate
rhodium relatively well, so at high conversions metal losses increase.
4) Avoid the use of polar, coordinating solvents. 5) Operate at high
temperatures. Coordination of the metal with phosphine is thermo-
dynamically more favorable, with respect to CO coordination, at higher
temperatures. 6) Use a polymer with low cross-linking. In macroreticular
beads the metal is concentrated at the surface. 7) Use a chelating ligand.
d. Bidentate phosphines give low nib ratios.
e. Functionalized silicas have given slightly lower n/ b ratios than those
derived from polystyrene.
Pittman has found unusual selectivities in phosphinated polystyrenes
with attached RhH(CO)(PPh 3 h. At low phosphine loadings (8% phenyl
rings phosphinated) at low (3.5) P /Rh ratios, the n/ b values at 400 psi
were 3 or 2 at 40° and 110°. These values are comparable to those of the
homogeneous catalyst at these conditions. With a very high P /Rh ratio(19)
and a very high phosphine loading (40% phenyl rings phosphinated) the
n/ b ratio increases to 6 at 100 psi and even greater (up to 15) upon
increasing the pressure to 400 psi. The homogeneous catalyst does not
display a similar trend. These very high selectivities can be understood to
be a consequence of species 7 being favored in the polymer because of the
proximity of phosphine.

5.5.2. Carboxylation

In the hydroformylation reaction, the acyl metal intermediate reacts

E
with hydrogen to form aldehyde. If, however, a nucleophile is present, the
course of the reaction may be altered, producing carboxylic acids or deriva-
tives:

o H20
CH 3CH 2C0 2H
II ROH
CH 3 CH 2 C-M CH 3 CH 2 C0 2 R

R2NH
CH 3 CH 2 CONR 2

Investigations in this area with polymer-bound catalysts are rather limited.


This modest interest is due, in part, to the fact that excellent industrial
routes to large-volume carboxylic acids already exist. One aspect that would
476 NORMAN L. HOL Y

be of interest would be the effect of a polymeric catalyst on the n/ b ratio.


In the ethoxycarbonylation of 1-pentene:

co
EtOH
PdCI 2

it was found that for the same P: Pd ratios, the bound catalyst gave higher
n/ b ratios than the soluble one. (139)
The Monsanto process for acetic acid production from methanol has
been a very successful application of homogeneous catalysis with
rhodium. (132)

When the polymer-bound version of the catalyst was investigated, (44) it


was discovered that the supported catalyst was not very active and, further-
more, that the metal was rapidly lost from the support. Reactions of CO
in a polar medium, in this case methanol, should be expected to lead to
loss of metal (see Table 9).

5.5.3. Decarboxylation

Insertion of CO into an alkyl-metal bond is a reversible process.


Through abstraction of CO from an aldehyde, alkanes may be synthesized:
RCHO ---. RH + co
This is a well-known reaction for homogeneous catalysts, but apparently
there are no examples using polymer-bound catalysts.

5.6. Metathesis

2R'CH=CHR 2 ¢ R'CH=CHR' + R 2 CH=CHR 2

During the late 1960s and the 1970s, olefin metathesis was one of the
most intensively investigated organometallic reactions. This reaction was
intriguing mechanistically and presented the possibility of selectively pro-
ducing attractive olefins and polymers. Ultimately, the mechanism was
proven to be a carbene process. There has been but modest activity using
phosphine-bound catalysts.
Basset investigated the catalytic properties of Mo(CO)6 attached to
a phosphinated polystyrene-DVB resinY48) Upon treatment with ethyl-
aluminum dichloride and oxygen, this slightly active system gave a conver-
sion of cis-2-pentene of 3.4% in 20 min at room temperature.
Table 9. Hvdroformvlation

Illustrative conditions "'IJ


Q
Metal complex r-
Support Ligand supported P/M Temp Pressure Reactant nib Reference ~
S}
Styrene/DYB ®-PR 2 Co 2(CO)s 150 1000 psi 1-Pentene 61/24 13 ~
Q
Styrene/DYB ®-PR 2 RhCI 3 /NaBH 4 133
Styrene/DYB ®-PR 2 [RhCI(COhJ2 100 1 atm Ethylene 43,134 ~
Styrene/DYB ®-PR 2 RhCI(CO)(PPh 3 h 3.7% Rh 90 20 atm Methanol 44 ~
Q
(f)
3.3% P
Styrene/DYB ®-PR 2 RhH(CO)(PPh 3 h 4.4 62 1000 psi 1-Pentene 4.4 42,12,14,
135-138 i
Styrene/DYB ®-PR 2 Ru(COh(PPh 3 h 3.1 140 1000 psi 1-Pentene 3.7 103
~
~
r-
Styrene/DYB ®-PR 2 PdCh/EtOH 139
~
Styrene/DYB ®-PR 2 [RhCI(COhh 130 15 atm Ethylene 140 CI!
Styrene/DYB ®-PPhCH 2CH 2PPh 2 RhH(CO)(PPh 3 h 14.3 80 800 psi Styrene 1/6.9 137
Styrene/DYB ®-DlOP RhH(CO)(PPh 3 h 4.0 40 400 psi Styrene 1/4.8 141,142
®-DlOP Rh(CO)PPh 3 )(acac) 22 80atm 1-Pentene 3b
Poly(vinyl ®-PR 2 [Rh(COh(acac)h 4.8% P 80-90 42 atm 1-Hexene 1.5-2.5 21
chloride) 7.9% Rh
Poly(butadiene) ®-PR 2 Co 2(CO)s 145
Poly(dimethyl- RhCI(CO)(PPh 3 h 0.50% P 100 1000 psi 1-Hexene 0.9 144
siloxane)
0.79% Rh
Polystyrenes Chelating RhH(CO)(PPh 3 h 1.9 20 1 atm 1-Hexene 6.4 145
pho~phines .t..
'I
'I
Styrene resins ®-P(OMeh [RhCI(COhh 146
Si0 2 @l-PPh 2 [RhCI(COhh 0.7 147
478 NORMAN L. HOL Y

Table 10. Metathesis

Metal complex
Support Ligand supported Reactant Reference

Styrene/DVB ®-PR z Mo(PPh 3h(NOhCl z 1, 7 -Octadiene 15


Styrene/DVB ®-PR z Mo(CO)6/ AIEtzCI + O 2 cis- 2- Pentene
Styrene/DVB ®-PR 2 W(CO)6/iBu3AI + Oz trans- 3 - Heptene 149

A similar catalyst based upon W(CO)6 was studied by Warwel and


Buschmeyer. (149) Adding isobutylaluminum and oxygen to the catalyst
effected a 50% conversion of trans- 3- heptene. Upon recycling the catalyst,
the activity decreased dramatically. The tungsten was displaced from the
polymer by the alkylaluminum compound, and the tungsten complex then
performed as a homogeneous catalyst (see Table 10).

6. TRENDS

6.1. Catalyst Characterization

One of the fruitful endeavors that might be continued in the future is


that of catalyst characterization. Surface properties such as pore size and
distribution are inadequately considered when catalytic activity is inter-
preted. There is still but a poorly formed concept of long-range interactions
between the support and the reaction center. We know, for example, that
the structures of conventional heterogeneous catalysts are often crucial in
establishing their activity. This is seen perhaps most dramatically in the
zeolite catalysts, where their electrostatic environments are largely respon-
sible for their chemistry. It is clear that polymeric supports affect reactions,
but the precise influences of cavity size, edge effects (surface vs. interior),
solvation, and polymer flexibility are unknown. It appears that coordination
patterns closely parallel those seen in homogeneous catalysis. There is no
direct spectroscopic evidence yet that polymers may force unusual bond
angles at coordinated metal centers. Are polymers, even highly cross-linked
ones, so flexible that strained metal coordinations are not formed?
Upon examination of the literature of homogeneous catalysis one is
struck by the remarkable paucity of kinetic studies. The same comment
could be made about polymer-bound catalysis. It is clear that in any
relatively new field the normal initial thrust is to define the possible reaction
chemistry. This search for interesting reactions certainly has not become
POL YMER-80UND PHOSPHINE CA TAL YSTS 479

passe, but there is a need to mature the field by learning more of the details
of the catalytic cycles.
It is also of note that a virtual void exists in evaluating the influence
of the polymer on stereoselectivity. Furthermore, isotope studies are largely
lacking. We might also expect additional supports, ones mechanically and
thermally more stable than polystyrene, to gain increased use in the future.
Phosphines have been attached to electrical insulators; would the use of
semiconducting or conducting supports provide access to catalysis more
akin to conventional metal heterogeneous catalysts? Could one usefully
study classical heterogeneous catalysts by attaching small aggregates to
polymer supports? A rigid support might provide the matrix isolation
necessary to examine the chemistry of small units.

6.2. Reactions
Since the researcher normally looks to the chemistry of soluble com-
plexes in designing polymer-bound catalysts, it is notable that some areas
that have proven fruitful in homogeneous catalysis have been omitted from
investigations using polymer-bound catalysts. One of these areas concerns
the reactions of arenes. Benzene, for example, may be hydrogenated with
homogeneous cobalt phosphite and ruthenium phosphine complexes, but
the corresponding supported versions are not reported. Aryl halides may
be carboxylated in the presence of a soluble palladium catalyst:

Similarly, the coupling of aryl halides and ole fins with soluble palladium
catalysts has but one polymer-bound analog.(150)

O ' / Q /
~
\\
~
II
X+
H
/
c=c"- - /
c=c "-

Soluble catalysts are also employed for coupling reactions between


aryl halides and metal alkyls. The most effective catalysts are those derived
from NiCh(PPh 3 h.

OX+RM-OR
Rearrangement of the alkyl moiety may occur, and the product distribution
is highly dependent upon the nature of the ligand.
480 NORMAN L. HOL Y

When we wish to reduce a compound catalytically, we typically think


of using hydrogen. The same transformation can sometimes be effected
using hydrogen already found in organic compounds. Hydrogen transfer
from alcohols (e.g., isopropyl alcohol), using soluble catalysts, is highly
effective for the reduction of ketones, olefins, and even nitro compounds.
Apparently there are no published examples of hydrogen transfer using
polymer-bound catalysts.
For all the examples discussed thus far, it is assumed that the metal
complexes remain attached to the polymer during the course of the reaction.
This is not the only type of catalyst, however, that would be attractive.
Equally attractive is the situation in which the metal complex decoordinates
from the polymer under certain conditions, e.g., high temperature, and
functions as a normal homogeneous catalyst, then is recoordinated when
the conditions are changed. The basic feature of catalyst recovery would
not be lost if a catalyst were to operate in this manner. There is an account
of this concept, using pyridine as the ligand. Apparently there are no
examples in phosphine chemistry.

REFERENCES

1. C. U. Pittman, Jr., Polymer-Supported Reactions in Organic Synthesis, edited by P.


Hodge and D. C. Sherrington (John Wiley and Sons, New York, 1980), p. 249.
2. Y. Chauvin, D. Commereuc, and F. Dawans, Prog. Polym. Sci. 5, 95-226 (1977).
3. a. E. Bayer and V. Schurig. Chem. Techno!. 1976, 212. b. E. Bayer and V. Schurig,
Angew. Chem. Inti. Edn. 14, 493-494 (1975).
4. G. Bernard, Y. Chauvin, and D. Commereuc, Bull. Soc. Chim. 1163, 1168-72 (1976).
5. W. Heitz, Adv. Pdym. Sci. 23, 1 (1977).
6. M. J. FarralI and J. M. Frechet, 1. Org. Chem. 43, 2618 (1978).
7. K. A. Abdula, N. P. AlIen, A. H. Badran, J. Dwyer, C. A. McAuliffe, and N. D. A.
Toma, Chem. Ind. 273 (1976).
8. R. D. Sanner, R. G. Austin, M. S. Wrighton, W. D. Honnick, and C. U. Pittman, Jr.,
"Interfacial Photoprocesses: Energy Conversion and Synthesis," ACS Adv. in Chem.
Ser. 184, 1980, 13; C. U. Pittman, Jr., W. D. Honnick, M. S. Wrighton, R. D. Sanner,
R. G. Austin, Fundamental Research in Homogeneous Catalysis, edited by M. Tsutsui
(Plenum Publishing, 1979), p. 603.
9. R. H. Grubbs and L. C. KrolI, 1. Am. Chem. Soc. 93, 3062 (1971).
10. H. S. Bruner and J. C. BaiIar, Inorg. Chem. 12, 1465 (1973).
11. R. Pierantozzi, K. J. McQuade, B. C. Gates, M. Wolf, H. Knozinger, and W. Ruhmann,
1. Am. Chem. Soc. 101, 5436 (1979).
12. C. U. Pittman, Jr., L. R. Smith, and R. M. Hanes, 1. Am. Chem. Soc. 97, 1742 (1975).
13. G. O. Evans, C. U. Pittman, Jr., R. McMillan, R. T. Beach, and R. Jones, 1. Organometal.
Chem. 67, 295 (1974).
14. c. U. Pittman, Jr. and L. R. Smith, 1. Am. Chem. Soc. 97, 1749 (1975).
15. R. H. Grubbs, S. Swetnick, and S. C. H. Su, 1. Mol. Catal. 3,11-15 (1977-1978).
POL YMER-80UND PHOSPHINE CATAL YSTS 481

16. M. Capka, P. Svoboda, M. Kraus, and J. Hetfiejs, Chem. Ind., 650-651 (1972).
17. T. H. Kim and H. F. Rase, Ind. Eng. Chem. Prod. Res. Dev. 15,249-254 (1976).
18. M. Czakova and M. Capka, J. Mol. Catal. 11, 313-322 (1981).
19. K. G. Allum, R. D. Hancock, S. McKenzie, and R. C. Pitkethly, Proceedings of the
Fifth International Congress on Catalysis 1972, edited by J. W. Hightower (North
Holland, Amsterdam, 1973), Vol. 1, p. 447.
20. R. D. Hancock, 1. V. Howell, R. C. Pitkethly, and P. J. Robinson, Catalysis
Heterogeneous and Homogeneous, edited by B. Delmon and G. Jannes (Elsevier, Amster-
dam, 1975), pp. 349-359.
21. K. G. Allum, R. D. Hancock, 1. V. Howell, R. C. Pitkethly, and P. T. Robinson, J.
Catalysis 43,322-330 (1976).
22. K. G. Allum, R. D. Hancock, 1. V. Howell, T. E. Lester, S. McKenzie, R. C. Pitkethly,
and P. J. Robinson, J. Catalysis 43,331 (1976).
23. K. G. Allum, R. D. Hancock, 1. V. Howell, T. E. Lester, S. McKenzie, R. C. Pitkethly,
and P. J. Robinson, J. Organometallic Chem. 107, 393 (1976).
24. K. G. Allum, R. D. Hancock, I. V. Howell, S. McKenzie, R. C. Pitkethly, and P. J.
Robinson, J. Organomet. Chem. 87,203 (1975).
25. J. M. Moreto, J. Albaiges, and F. Camps, Catalysis Heterogeneous and Homogeneous,
edited by B. Delmon and G. Jannes (Elsevier, Amsterdam, 1975), pp. 339-347.
26. K. Kochloefi, W. Liebelt, and H. Knozinger, J. Chem. Soc., Chem. Comm. 1977, 510.
27. M. Capka and J. Hetfiejs, Coll. Czech. Chem. Comm. 39, 154 (1974).
28. Badische Anilir und Soda-Fabrik, A. G., French Patent No.2 053 300.
29. K. G. Allum, R. D. Hancock, British Patent No.1 287 566.
30. J. Manassen, Israel J. Chem. 8, 5p (1970).
31. s. V. McKinley and J. W. Rakshys, United States Patent No.3 708462 (to Dow
Chemical Company).
32. Y. Nonaka, S. Takahashi, and N. Hagihara, Mem. Inst. Sci. Ind. Res., Osaka Univ. 31,
23 (197~).
33. A. J. Naaktgeboren, R. J. M. Nolte, and W. Drenth, Reel. Trav. Chim. Pays-Bas 97,
112(1978).
34. A. J. Naaktgeboren, R. J. M. Nolte, and W. Drenth, J. Amer. Chem. Soc. 102, 3350
(1980).
35. J. Kiji, S. Kadoi, and J. Furukawa, Angew. Mak;omol. Chem. 46, 163 (1975).
36. N. Takaishi, H. Imai, C. A. Bertelo, and J. K. Stille, J. Am. Chem. Soc. 98, 5400 (1976).
37. N. Takaishi, H. Imai, C. A. Bertelo, and J. K. Stille, J. Am. Chem. Soc. 100, 264 (1978).
38. G. L. Baker, S. J. Fritschel, and J. K. Stille, J. Org. Chem. 46, 2960-2965 (1981).
39. I. Dietzmann, D. Tomanova, and J. Hetfiejs, Coll. Czech. Chem. Comm. 39,135 (1974).
40. C. U. Pittman, Jr. and L. R. Smith, J. Amer. Chem. Soc. 97, 341 (1975).
41. C. U. Pittman, Jr. and R. F. Felis, J. Organometallic Chem. 72, 399 (1974).
42. c. U. Pittman, Jr. and R. M. Hanes, J. Amer. Chem. Soc. 98, 5402 (1976).
43. K. G. Allum, R. D. Hancock, 1. V. Howell, R. C. Pitkethly, and P. J. Robinson, J.
Organometallic Chem. 87, 189 (1975).
44. M. S. Jarrell and B. C. Gates, J. Catalysis 40,255 (1975).
45. D. Doskocilova, B. Schneider, and J. Jakes, J. Magn. Reson. 29, 79 (1978).
46. s. J. Fritschel, J. J. H. Ackerman, T. Keyser, and J. K. Stille, J. Org. Chem. 44, 3152
(1979).
47. T. Imanaka, K. Kaneda, S. Teranishi, and M. Terasawa, Proceedings of the Sixth
International Congress on Catalysis, 1976, edited by G. C. Bond, P. B. Wells, and F.
C. Tompkins (The Chemical Society, London, 1977), Vol. 1, p. 509.
48. N. Takahashi, 1. Okura, and T. Keii, J. Amer. Chem. Soc. 97, 7489 (1975).
49. K. Kaneda, M. Terasawa, T. Imamaka, and S. Teranishi, Chem. Lett. 1005-1008 (1975).
482 NORMAN L. HOL Y

50. N. L. Holy, Fundamental Research in Homogeneous Catalysis 3, edited by M. Tsutsui


(Plenum Publishing, 1979), pp. 691-706.
51. J. Reed, P. Eisenberger, B.-K. Teo, and B. M. Kincaid, f. Amer. Chem. Soc. 99,
5217-5218 (1977); 100,2375-2378 (1978).
52. L. D. Rollmann, f. Amer. Chern. Soc. 97, 2132 (1975).
53. F. Pinna, M. Bonivento, G. Strukul, M. Graziani, E. Cernia, and N. Palladino, f. Mol.
Catalysis 1,309 (1975-1976).
54. G. Strukul, M. Bonivento, M. Graziani, E. Cernia, and N. Palladino, Inorg. Chim. Acta
12, 15 (1975).
55. M. Graziani, G. Strukul, M. Bonivento, F. Pinna, E. Cernia, and N. Pallidino, Catalysis
Heterogeneous and Homogeneous, edited by B. Delmon and G. Jannes (Elsevier, Amster-
dam, 1975), pp. 331-338.
56. G. Strukul, P. Dolimpio, M. Bonivento, F. Pinna and M. Graziani, f. Mol. Catalysis
2,179 (1977).
57. D. Tatarsky, D. H. Kohn, and M. Cais, f. Polym. Sci., Polym. Chern. Ed. 18, 1387-1397
(1980).
58. R. H. Grubbs and E. M. Sweet, Macromolecules 8,241-242 (1975).
59. R. H. Grubbs, C. Gibbons, L. C. Kroll, W. D. Bonds, and C. H. Brubaker, Jr., f. Amer.
Chem. Soc. 95, 2373 (1973).
60. J. Manassen, Reference 20, 293-306.
61. M. S. Bruner and J. C. Bailor, f. Amer. Oil Chem. Soc. 49, 533 (1972).
62. R. H. Grubbs, L. C. Kroll, and E. M. Sweet, f. Macromol. Sci., Chem. A 7, 1047 (1973).
63. J. D. Morrison, W. F. Master, and M. F. Neuberg, Adv. Catal. 25, 81 (1976).
64. B. D. Vineyard, W. S. Knowles, M. J. Sabacky, G. L. Bachman, and D. J. Weinkauff,
f. Am. Chern. Soc. 99, 5946-5952 (1977).
65. T. Masada and J. K. Stille, f. Am. Chem. Soc. 100,268 (1978).
66. G. C. Baker, S. J. Fritschel, J. R. Stille, and J. K. Stille, f. Org. Chern. 46, 2954-2960
(1981).
67. W. Dumont, J. C. Poulin, T. P. Dang, and H. B. Kagan, f. Am. Chem. Soc. 95,
8295-8299 (1973).
68. K. Achiwa, Chem. Lett. 905 (1978).
69. M. E. Wilson and G. M. Whitesides, f. Am. Chern. Soc. 100,306 (1977).
70. F. Ciardelli, E. Chiellini, C. Carlini, and R. Nocci, Polym. Preprints 17,188 (1976).
71. a. K. Ohkubo, K. Fujimori, and K. Yoshinaga, Inorg. Nucl. Chem. Lett. 15, 231-234
(1979).
b. K. Ohkubo, M. Setoguchi, and K. Yoshinaga, Inorg. Nucl. Chem. Lett. 15, 235-238
(1979).
72. C. U. Pittman, Jr., A. Hirao, J. J. Yang, Q. Ng, R. Hanes, and C. C. Lin, Preprints Div.
Petrol. Chem. (ACS) 22, 1196 (1977).
73. C. F. Hobbs and W. S. Knowles, f. Org. Chem. 46, 4422-4427 (1981).
74. J. I. Crowley and H. Rapoport, Acc. Chem. Res. 9,135-144 (1976).
75. L. T. Scott, J. Rebek, L. Oysyanko, and C. L. Sims, f. Am. Chern. Soc. 99, 625-626
(1977).
76. P. Jayalekshmy and S. Mazur, f. Am. Chem. Soc. 98, 6710-6711 (1976).
77. C. U. Pittman, Jr. and Q. Ng, f. Organometal. Chem. 153,85 (1978).
78. R. H. Crabtree, M. F. Mellea, J. M. Miheicic, and J. M. Quirk, f. Am. Chem. Soc. 104,
107-113 (1982).
79. A. H. Janowicz and R. G. Bergman, f. Am. Chern. Soc. 104,352 (1982).
80. S. Jacobson, W. Clements, H. Hiramoto, and C. U. Pittman, Jr., f. Mol. Catal. 1,
73 (1975).
81. c. U. Pittman, Jr., S. E. Jacobson, and H. Hiramoto, f. Am. Chem. Soc. 97, 4774 (1975).
82. B. C. Gates and J. Lieto, Chem. Technol. 248 (1980).
POL YMER-80UND PHOSPHINE CA TAL YSTS 483

83. J. Andrieux, D. H. R. Barton, and H. Patin, I. Chern. Soc. Perkin I, 359-000 (1977).
84. C. A. Tolman, I. Am. Chern. Soc. 94, 2994-2999 (1972).
85. R. G. Miller, P. A. Pinke, R. D. Stauffer, H. J. Golden, and D. J. Baker, I. Am. Chern.
Soc. 96, 4211-4220 (1974); P. A. Pinke and R. G. Miller, op. cit., 4221-4229; P. A.
Pinke, R. D. Stauffer, and R. G. Miller, op. cit. 4229-4234.
86. J. M. Moreto, J. Albaiges, and F. Camps, Ref. 20, 339-348.
87. P. A. Gosselain, Reference 20,107-132.
88. J. Conan, M. Bartholin, and A. Guyot, I. Mol. Catalysis 1,375 (1975-1976).
89. K. Kochloefl, W. Liebelt, and H. Krozinger, 1.C.S. Chern. Cornrnun. 510 (1977).
90. I. V. Howell and R. D. Hancock, British Patent No.1 408013.
91. W. R. Cullen, D. J. Patmore, A. J. Chapman, and A. D. Jenkins, I. Organornetal. Chern.
102, C12 (1975).
92. R. R. Schrock and J. Osborn, 1.C.S., Chern. Cornrnun. 567 (1970).
93. G. Pez, R. A. Grey, and J. Corsi, I. Am. Chern. Soc. 103, 7528 (1981).
94. S. E. Jacobson and C. U. Pittman, Jr., I. Chern. Soc. Chern. Cornrn. 187 (1975).
95. M. H. J. M. DeCroon and J. W. E. Coenen, I. Mol. Catal. 11, 301-311 (1981).
96. S. Lecolier, French Patent No.2 270 238.
97. M. Terasawa, K. Kaneda, T. Imanaka, and S. Teranishi, I. Catal. 51,406 (1978).
98. K. Kaneda, M. Terasawa, T. Imanaka, and S. Teranishi, Chern. Lett. lapan 995 (1976).
99. R. H. Grubbs, L. C. Kroll, and E. M. Sweet, Polymer Preprints 13,828 (1972).
100. R. H. Grubbs, E. M. Sweet, and S. Phisanbut, Catalysis In Organic Synthesis, edited
by P. Rylander and H. Greenfield (Academic Press, New York, 1975), p. 153.
101. A. Guyot, C. Graillat, and M. Bartholin, I. Mol. Catal. 3, 39 (1977-1978).
102. K. G. Allum, R. D. Hancock, and R. C. Pitkethly, British Patent No.1 295 675.
103. C. U. Pittman, Jr. and G. Wilemon, Ann. N. Y. Acad. Sci. 333, 67 (1980).
104. J. P. Collman, L. S. Hegedus, M. P. Cooke, J. R. Norton, G. Go\cetti, and D. N.
Marquart, I. Am. Chern. Soc. 94, 1789 (1972).
105. M. S. Jarrell, B. C. Gates, and E. D. Nicholson, I. Am. Chern. Soc. 100, 5727 (1978).
106. M. S. Jarrel and B. C. Gates, I. Catal. 54, 81 (1978).
107. Z. Otero-Schipper, Z. Lieto, and B. C. Gates, I. Catal. 63, 175 (1980).
108. J. Lieto, J. J. Rafalko, and B. C. Gates, I. Catal. 62, 149 (1980).
109. J. J. Rafalko, J. Lieto, B. C. Gates, and G. L. Schrader, Jr., I. Chern. Soc., Chern.
Cornrnun. 540 (1978).
110. C. U. Pittman, Jr., S. K. Wuu, and S. E. Jacobson, I. Catal. 44, 87 (1976).
111. C. U. Pittman, Jr., L. R. Smith, and S. E. Jacobson, Catalysis: Heterogeneous and
Homogeneous, edited by B. Delmon and G. Jannes (Elsevier, Amsterdam, 1975), p. 393.
112. A. J. Naaktgeboren, R. J. M. Notte, and W. Drenth, I. Mol. Catal.ll, 343-351 (1981).
113. R. G. Muratova, R. Z. Khairullina, S. V. Shulyndin, B. E. Ivanov, and R. I. Izmailov,
Kin. and Catal. 15,115 (1974).
114. G. L. Baker, S. J. Fritschel, and J. K. Stille, Polymer Preprints 22(1),155 (1981).
115. W. R. Cullen, D. J. Patmore, A. J. Champan, and A. D. Jenkins, I. Organornet. Chern.
102, C12 (1975).
116. B. Bogdanovic, H. Biserka, H. G. Karmann, H. G. Nussel, D. Walter, and G. Wilke,
Ind. Eng. Chern. 62, 34-38 (1970). .
117. R. S. Bauer, H. Chung, P. W. Glockner, and W. Keirn, United States Patent No.
3644563 (1972); R. F. Mason, United States Patent No.3 737 475 (1973).
118. Badische Anilin und Soda Fabrik A.G., French Patent No.2 053 300.
119. K. Kaneda, M. Terasawa, T. Imanaka, and S. Teranishi, Tetrahedron Lett. 2957 (1977).
120. K. Kaneda, T. Uchiyama, M. Terasawa, T. Imanaka, and S. Teranishi, Chern. Lett.
lapan 449 (1976).
121. C. U. Pittman, Jr. and S. E. Jacobson, I. Mol. Catal. 3,293-297 (1977-1978).
122. F. Hobjabri, Polymer 17,58 (1976).
484 NORMAN L. HOL Y

123. H. Pracejus, German Patent No.2 230 739.


124. K. G. Allum and R. D. Hancock, Fifth International Congress of Catalysis, Preprint 31
(Miami, 1972).
125. R. F. Clark, C. D. Storrs, and G. B. Barnes, United States Patent No.3 364273.
126. J. L. Speier, "Homogeneous Catalysis by Transition Metals," Adv. Organornet. Chern.
17,407 (1979).
127. I. Dietzmann, D. Tomanova, and J. Hetfiejs, Coll. Czechoslov. Chern. Cornrn. 39, 123
(1974).
128. V. D. Luedeke, A dipronitrile, edited by J. J. McKetta and W. A. Cunningham,
Encyclopedia of Chemical Processing and Design (Marcel Dekker, 1977), Vol. 2, p. 146.
129. M. Capka, P. Svoboda, M. Corny, and J. Hetfiejs, Tetrahedron Lett. 4787 (1971).
130. R. D. Sanner, R. G. Austin, M. S. Wrighton, W. D. Honnick, and C. U. Pittman, Jr.,
Inorg. Chern. 18, 928 (1979).
131. J. C. Poulin, W. Dumont, T. P. Dang, and H. B. Kagan, Cornpt. Rend. Ser. C 277,41
(1973).
132. J. F. Roth, J. H. Craddock, A. Hershman, and F. E. Paulik, Chern. Tech. 1, 600 (1971);
H. D. Grove, Hydrocarbon Proc., 76 (1972); F. E. Paulih, United States Patent No.
3769329 (1973).
133. I. V. Howell and R. D. Hancock, British Patent No.1 408 013.
134. H. Arai, T. Kanedo, and T. Kunugi, Chern. Lett. lapan, 265 (1975).
135. C. U. Pittman, Jr. and W. D. Honnick, l. Org. Chern. 45, 2132 (1980).
136. c. U. Pittman, Jr., A. Hirao, C. Jones, R. M. Hanes, and Q. Ng, Ann. N. Y. Acad. Sci.
295, 15 (1977).
137. c. U. Pittman, Jr. and C. C. Lin., l. Org. Chern. 43, 4928 (1978).
138. C. U. Pittman, Jr., W. D. Honnick, and J. J. Yang, l. Org. Chern. 45, 684 (1980).
139. c. U. Pittman, Jr. and Q. Y. Ng, United States Patent No.4 258 206 (1981).
140. H. Arai, l. Catal. 51, 135-142 (1978).
141. c. U. Pittman, Jr., A. Hirao, J. J. Yang, Q. Ng, R. Hanes, and C. C. Lin, Preprints Div.
Petrol. Chern. (ACS) 22, 1196 (1977).
142. S. J. Fritschel, J. J. H. Ackerman, T. Keyser, and J. K. Stille, l. Org. Chern. 44, 3152
(1979).
143. Badische Anilin und Soda-Fabrik A.G., French Patent No.2 053 300.
144. M. O. Farrell and C. H. VanDyke, l. Organornetal. Chern. 172,367-376 (1979).
145. A. R. Sanger and L. R. Schallig, l. Mol. Catal. 3, 101 (1977-1978).
146. W. O. Haag and D. D. Whitehurst, "Catalysis," edited by J. W. Hightower, paper 29,
Vol. 1, p. 465 (1973).
147. R. D. Hancock, I. V. Howell, R. C. Pitkethly, and P. J. Robinson, Reference 20, 361-371.
148. J. M. Basset, R. Mutin, G. Descotes, and D. Sinou, Cornpt. Rend. Ser. C 280, 1181
(1975).
149. S. Warwel and P. Buschmeyer, Angew. Chem., Int. Ed. Engl. 17, 131-000 (1978).
150. M. Terasawa, K. Kaneda, T. Smanaka, and S. Teranishi, l. Organornetal. Chern. 162(3),
403 (1978).
Index

Acetylene trimerization, 207 Asymmetric induction, 41


Acrylic acid esters from alcohols, acetylene, with supported catalysts, 457
and CO,S AU2[/L-(CH2hPR2]2,199
Active catalyst, defined, 14
Acyl carbonyl complexes, structures, 124t BINAP,138
Acyl complexes, structures, 125t Binuclear diphosphine bridging geometries,
A-frame, 175 175f
A-frame complexes, hydrogenation catalytic Binuclear elimination reactions, 226
activity, 219t Bis(diphenylarsino)methane bridged com-
Aldehydes plexes, structural parameters, 181t
adduct with [Rh(dpppht, 364 Bis(2-diphenylphosphinoethyl)amine
decarbonylation, 313, 343 ligands, 244
hydrogenation, 318-319 1,5-Bis(diphenylphosphino)-3-oxopentane,
oxidative addition of, 352 complexes of, 242-243f
Alkane activation, 310 Bonding modes in diphosphine complexes,
Alkane dehydrogenation, 311 217
Alkenylphosphine ligands, 251 BPh4- as a ligand, 299
Alkyl carbonyl complexes, structures, 123t Bridging carbonyl ligands, 196t
Alkyldiazenido complexes, 418 Bridging methyl groups, 33
Alkyl olefin complexes, structures, 127t Bulky phosphines, 258
Alkyne to alkene hydrogenation, 221 Butadiene, Ni catalyzed cyclooligomeriza-
Amines from N 2, 6 tion,64-80
Aminobis (difluorophosphines), 170 Butene isomerization, 49
Aminoketones, hydrogenation of, 333-335 mechanism using Ni[P(OEth]4, 50f
Anchimeric assistance, in rate of oxidative
addition, 241 Carbon monoxide bridged complexes, 223t
Aryl halides, coupling reactions, 479 Catalyst deactivation product, 15
Asymmetric bisphosphines, 138f Catalyst precursor, definition, 14
Asymmetric hydrogenation, 9, 138 Cationic bisphosphine rhodium diene com-
deuterium isotope effects, 160 plexes, 148
energetics, 161f Cationic complexes, 303, 305, 307
of ketones, 328t Cationic complexes, synthetic methods, 303-
kinetic studies, 151 304
mechanism and stereoselectivity, 154 Celanese hydroformylation studies, 280f
optical yields, 140t Charge effects on catalysts, 297-303
pressure effects, 149-150t Chelate ring size, effect on catalytic rates, 225
reaction pathway, 159-160 Chelating bisphosphine rhodium complexes,
substrate complexes, 155 structures, 152f

485
486 INDEX

Chiral ferrocenyldiphosphine ligand, 253f Dinitrogen to ammonia, 406


Chiral shift reagents, 40 (-)-DIOP, 9,137
CHIRAPHOS, 42, 138 DIOP complexes, conformational analysis,
Cluster hydrides, 309 142f
C-N bond formation from coordinated N2, DIOXOP,143
414 Dioxygen complexes, reactions with e1ec-
Co-catalysts, 14 trophiles, 380-382
Coordinative unsaturation in supported ketones, 381£
catalysts, 459--460 DIPAMP, 138, 139
Co-oxidation reactions 2(Diphenylphosphino )pyridine, 182
of alkenes and hydrogen, 392 Diphosphine bridge binuclear complexes,
of alkenes and tertiary phosphines, 390- catalytic reactions, 206-208
391£ Diphosphines
of isocyanides and CO, 392f attached to organic polymers, 288-289
of triphenylphosphines and CO, 393 bridging ligands, 170
CO 2[IL-PNPh(COh,l72f dppb,360,368
Counter ion effects, 302 dppe, 360, 361
CS 2 coupling reaction, 203 dppm, 360, 368-369
CTSRDS,18 dppp,359-363
Cyclometalation reactions, 310 Diphosphite ligands, 170
Cyclooligomerization of butadiene Double A-frame complexes, 175
ligand concentration control maps, 69 Dppm bridged complexes, structural para-
ligand effects, 65 meters, 179t
mechanism,76f
Ni catalyzed, 64-80 E-a -benzamidocinnamic acid, asymmetric
Cyclopropanes, 126-129 hydrogenation of, 144-145
CYPHOS, 140 Enantiomer excesses in hydrogenation of a-
benzamidocinnamic acids, 145t
Deactivation of catalysts, 15,308 Enantiomeric ratio determination by polar-
Dead catalyst, 15 ity,38
Decarbonylation Enantioselective hydrogenation of ketones,
acid chlorides, 345, 347-352, 372, 348t 324t
aldehydes, 352-355 Enantioselective reduction of C=N double
aldehyde substituent effects, 363 bond,337t
catalytic, 355-372, 357t, 360t Equilibrium isotope effects, 30
diphosphine complexes, 346, 359-372 ESCA of polymer bound catalysts, 452-453t
Ir catalysts, 370t Ethane decompoisition, 21
kinetic studies, 354-355, 363-368 EXAFS, 154
ligand steric effects, 360t, 368-369 of polymer bound catalysts, 453
oxidative addition intermediates, 352
P-N complexes of Rh, 370
Face-to-face binuclear complex, 175
RhCI(PPh 3 h, 345, 347-355
tetraphenylporphyrin complex, 346 Fe2Cp2(CO)4, 169
Ferrocene-based P2 ligands, 281£
Dehydrogenation of alkanes, 311
Fischer-Tropsch reaction, 30
Deuterium isotope effects, 160
Diastereomers (NMR), 34, 37
Diastereoselection in enamide complexes, Hemilabile chelating agents, 241
162f High dilution synthesis of phosphorus
Diazenido complexes of Mo and W, 414-415 macrocycles,269-271
Dinitrogen complexes, 410-412f Hydrazido complexes of Mo and W, 414--415
reactivity with electrophiles, nucleophiles, Hydride attack on ligands, 297
and radicals, 433-434 Hydrido alkene complexes, structures, 114t
INDEX 487
Hydrido-carbonyl complexes, structures, kinetic isotope effects, 27
119t NMR,22
Hydroformylation
catalysis with chelating diphosphine Keto acids and esters, hydrogenation, 331,
ligands, 279-285 332t
effect of diphosphine, 227-229f Ketones,
effect of triphosphine, 230f asymmetric hydrogenation, 328t
Rh-catalyzed,81-99 hydrogenation, 319-324, 320f
ligand steric and electronic effects, 95 Kinetic isotope effect, decarbonylation, 350,
mechanism, 86f 355
olefin structure effects, 94t
product distribution, 97t L-DOPA, catalytic synthesis, 138
rates for various olefins, 88t Ligand properties important in catalysis, 6
spectroscopic studies, 82 Long chain diphosphines, 258
Hydroformylation catalysts, Co, Rh, and Ir
comparison, 99-102 Matrix isolation in supported catalysis, 457-
Hydrogenation 459
in aqueous solution, 288 Mechanism of catalyst system, general, 15
catalysis with chelating triphosphine Metaliacycle complexes, structures, 130-
ligands, 275-279 13lt
catalytic activity of A-frame complexes, Metal-metal bonds, insertion, 189
219t Methylaminobis( difluorophosphine) bridged
with cationic Rh and Ir complexes, 304- complexes, structural parameters, 18lt
310 Microwave spectroscopy, 50
of olefins, influence of solid support, Mixed donor ligands, PS, PAS, PN, 267-269
454t Mixed metal PN bridge dimers, 250f
using Pdz(dppmh, 207 M-M separations in binuclear complexes,
of phenyl acetylene, 221f 192-193
with phosphine bound catalysts, 467-469t Mnz(1L -dppmlz(CO)6, 202
rates as a function of diphosphine chelate MoBrz(NNHz)(triphos), 425
ring size, 225 Mo(N z)z(dppe)z,409
Hydrogen atom abstraction from toluene, MOz[1L -CH3N(PFz)Zl4C\z, 171f
252
Hydrosilation of ketones, 329 Ni(CO)z(PPh 3)z, 5
Ni[P(ORhl4, 49
Imines, asymmetric reduction of, 336 Nitrile group activation, 298
Iminophosphines, 250 Nitrite complexes
Infrared at high temperature and pressure, oxygen atom transfer, 396
81,82,100 oxygen transfer to alkenes, 398
Insertion into metal-metal bonds, 189 reaction with CO, 396-397
Ion cyclotron resonance, 312 Nitrogenase, 406
Ionic catalysts, 303 NMR
IrCI(CO)(PPh 3)z, 344, 345 determination of stereochemistry, 34
[Ir(diphosphineht, 369-372 CH and 31 p ) of IL-dppm binuclear com-
IrLz(Hlz +,311 plexes, 183
Irz(COlz(1L -dppmlz(1L -S), 199 saturation transfer studies, 155
Isocyanide bridged dimers, 198 shift reagents, 40
Isotopic labeling NOPAPHOS, 140
cross-over experiments, 23
decarbonylation, 353, 363 a -diphenylphosphinoanisole, 241
effects, 27-30 a -diphenylphosphinobenxoylpinacolone, Pt
IR.22 and Cu complexes, 248-249f
488 INDEX

Olefin carbonyl complexes, structures, 120- 31 p NMR for polymer-bound catalysts, 451-
12lt 452
o-diphenphosphinobenzoic acid, 247 31 p NMR of RhCl(PPh 3h and H2 system, 59f
Olefin hydrogenation Polydentate ligands
kinetics and mechanism with RhCl(PPh 3h, phosphines, 259f
53-61 synthesis, 261-275
rate and equilibrium constants, 58t Polyetherphosphonite ligands, 243
scheme,57f Polyethylene, 8
using RhClL 3 Polymer bound
ligand dependence on rate, 62t optically-active diphosphine ligands, 263
olefin structure dependence, 63t phosphine catalysts, 444
Olefin isomerization, 307 rhodium catalysts in ketone reduction, 321
Olefin metathesis, 476 Polynuclear complexes of N2, 428
Optical activity, aldehydes-mechanistic Polypropylene, 8
probe, 353, 361 Polystyrene as a solid support, 444--448
Optical purity by optical rotation, 39 Prochiral a(3 -unsaturated carboxylic acids,
OsBr2(PPh3h, 344 147
OS2(CO)gR 2, 23 PROPHOS, 139
Oxidation reactions Protein-bound rhodium hydrogenation
metal catalyzed, 377-378 catalyst, 245f
of aldehydes and ketones, 386-387 Protonation of NiL4 complexes, equilibrium
of alkenes, 383-385 constants, 53t
of carbon monoxide, 207, 385-386, 31p spin-lattice relaxation times of polymer-
394-395 bound catalysts, 451
of cumene, 387 Pt(C 2H 4)zL 2, 17
of isocyanides, 385 oxidative addition of Mel to, 17
of nitric oxide, 394-395 Pt0 2(PCY3)z, 382
of phosphines, 387-389 Pt 2(CH 3h(f,t-dppm)z +, 174f
Oxidative addition in diphosphine bridged Pt2(CH3)4(f.L-dppm)z, 174f
complexes, 197-200 Pt 2CIz(f,t-dppm)z, reaction chemistry, 190f
Oxygen atom transfer reactions, 393-394 Pt3(f.L-COh(PR3h, 24
Oxygen bridged binuclear complexes, 192 195 Pt NMR, 25
Oxygen complexes, 378-380
Radical
CIDNP, 29, 46
Paramagnetic species, 19 esr spectra of, 44
Pd(PEt 3h, 39 pairs and stereochemistry, 48
Pd0 2(PPh 3lz, 395 pathway in hydroformylation reactions,
Pd 2 Cl(f,t -dppm)z(SnCI 3), 177f 101
Pd 2 CI 2 (f,t -dppmlz(f,t -S02), 178f reaction pathways, 29, 43
Peracyls, 396 reactions, 21
P-H addition to vinyl compounds, 266-267 spin traps, 45
PHELLANPHOS, 140 Rate laws, 16-17, 18,20
PHEPHOS, 140 Reduction of C=O and C=N functions, 310,
Phoran, Ph(H)P(OCH 2CH 2lzN, 245 317-318
Phosphide anions, R 2P-, 171 RhCl(diop),337
PhosphinocycIopentadiene ligands, 254f RhCl(PPh 3h, 8, 112, 345
Phosphoranes as ligands, 245 RhCl(CO)(PPh 3 lz. 81
Phosphorus macrocycles, 270 RhCI 2(CO)(R)(PPh 3 )z,351
Phosphorus-nitrogen donor ligands, 371 RhCI 2(COR)(PPh 3 }z, 349
31 p NMR coupling constants in (f,t- Rh(Cl)z(dppp)(PhCO),372
dppm lzM2 complexes, 186-187t Rh(CO)(dppp)zt, 363
INDEX 489

Rh(CO)(PPh 3h +,359 Solid supports


Rh(diene)(P 2 t, 322 polymethycrylate allyl chloride, 446
Rh(diphosphineh +, 359-372 polyphenylene-isophthalamide, 447
Rh(dppeh +,225 polysiloxanes, 446
[Rh(dppp)]BF4,360 polystyrene, 444-448
RhHCl(CO)(PPh 3 h, 343 polyvinyl chloride, 446
RhH(CO)(PPh 3h, 87, 228, 280, 475 Solvent effects, 299, 303, 308, 311
RhH[(-)-diop)]2> 225 Solvent stabilized species, 19
Rh(Hh(dppph +,225 Spin -state labeling, 23
RhH(PPh 3h,112 Steric and electronic effects, 16, 77-78, 95
Rh 2(1L -Cl)(1L -CO)(COh(1L -dppmh +, 178f Steroid reductions, 309
Rh 2Cl(COh(N 3)(IL-dppmh,224 Sulfur bridged binuclear complexes, 193
Rh 2CI 2(COh(1L -dppmh, 176f Supported catalysts
Rh 2C1 2(CO lz(1L- Ph 2P( CH 2hO( CH 2)zPPh 2], hydroformylation, 472-475, 477t
175f hydrogenation, 461
Rh 2(CNR)4(IL-dppmh 2+, 195, 197 hydrosilylation, 473t
Rh 2(COhCp2(IL-dppb),182f iridium complexes, 465
Rh 2(CO)4(PPh 3)4, 386 olefin isomerization, 460-461
Rh 2(COhX 2(IL-dppmh,218 palladium, platinum complexes, 465-466
RhPd(Clh(CO)(1L -Ph 2Ppy)z, 204-206f rhodium complexes, 462-465
l03 Rh NMR, 23 Supported phosphine catalysts, 285-289
Rhodium alkyl hydride complex in asym- Surface bound catalysts, 10
metric hydrogenation, 158
Rhodium-phosphine catalyzed hydroformy-
Template synthesis of polyphosphine ligands,
lation, mechanisms, 283f, 284f
271-274
[Rh(P-Nht, 371
Transition state, composition, 18
1,2-R,R -bis( anisylphenylphosphino )-ethane
Trialkylphosphines, preparations, 2
ligand, synthesis, 240
Trimethylphosphine, 1
[Ru(bipy hCl]2(1L _dppm)2+, 182
Tri-n- butylphosphine, 7
RuCI 2(PPh 3h, 9, 112
Triphosphine attached to glass beads, 287
[Ru2CI3(PEt2Ph)6]Cl, 344
Triphosphine ligands, 260
RU2(CO)4CP2> 169
Tripod ligands, 261, 290, 291
RuH(11 2-BH4)(ttp),277
HC(PPh 2h, 208
RuHCl(PPh 3h, 9
Tris(diphenylphosphino)methane, 262
Ru(PPh 3h(tetraphenylporphyrin), 346
Tritium labeling, 147
Ru(RCNhL/+, 298
Turnover number, 15
Saturation labeling in NMR, 25
Saturation transfer, 25, 26 Wacker oxidation of ethylene to acetalde-
Selectivity, 9 hyde, 21
in aldehyde decarbonylation, 36lt, 367- Water gas shift reaction, 313-314
368 Water soluble phosphine complexes, 288
in polymer-bound catalysts, 455-456
Shift reagents, 40
Side-by-side binuclear complex, 175 Z-dehydroamino acids, asymmetric hydro-
Sixteen- and eighteen-electron rule, 19 genation of, 140f
SKEWPHOS, 141 Ziegler-Natta polymerization, 7

You might also like