26.10.2012 Views

Strained Hydrocarbons - Developers

Strained Hydrocarbons - Developers

Strained Hydrocarbons - Developers

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Strained</strong> <strong>Hydrocarbons</strong><br />

Edited by Helena Dodziuk


Further Reading<br />

A. V. Demchenko (Ed.)<br />

Handbook of Chemical Glycosylation<br />

2008<br />

ISBN: 978-3-527-31780-6<br />

P. G. Andersson, I. J. Munslow (Eds.)<br />

Modern Reduction Methods<br />

2008<br />

ISBN: 978-3-527-31862-9<br />

L. Kollár (Ed.)<br />

Modern Carbonylation Methods<br />

2008<br />

ISBN: 978-3-527-31896-4<br />

E. M. Carreira, L. Kvaerno (Eds.)<br />

Classics in Stereoselective Synthesis<br />

2008<br />

ISBN: 978-3-527-29966-9<br />

M. M. Haley, R. R. Tykwinski (Eds.)<br />

Carbon-Rich Compounds<br />

From Molecules to Materials<br />

2006<br />

ISBN: 978-3-527-31224-5


<strong>Strained</strong> <strong>Hydrocarbons</strong><br />

Beyond the van’t Hoff and Le Bel Hypothesis<br />

Edited by Helena Dodziuk


The Editor<br />

Prof. Dr. Helena Dodziuk<br />

Institute of Physical Chemistry<br />

Polish Academy of Sciences<br />

Kasprzaka 44<br />

01-224 Warsaw<br />

Poland<br />

Cover illustration<br />

Cover picture kindly provided by<br />

Dr. K. S. Nowinski<br />

All books published by Wiley-VCH are carefully<br />

produced. Nevertheless, authors, editors, and<br />

publisher do not warrant the information contained<br />

in these books, including this book, to be free of<br />

errors. Readers are advised to keep in mind that<br />

statements, data, illustrations, procedural details or<br />

other items may inadvertently be inaccurate.<br />

Library of Congress Card No.: applied for<br />

British Library Cataloguing-in-Publication Data<br />

A catalogue record for this book is available from<br />

the British Library.<br />

Bibliographic information published by<br />

the Deutsche Nationalbibliothek<br />

Die Deutsche Nationalbibliothek lists this<br />

publication in the Deutsche Nationalbibliografie;<br />

detailed bibliographic data are available in the<br />

Internet at http://dnb.d-nb.de.<br />

� 2009 WILEY-VCH Verlag GmbH & Co. KGaA,<br />

Weinheim<br />

All rights reserved (including those of translation<br />

into other languages). No part of this book may<br />

be reproduced in any form – by photoprinting,<br />

microfilm, or any other means – nor transmitted<br />

or translated into a machine language without<br />

written permission from the publishers. Registered<br />

names, trademarks, etc. used in this book, even<br />

when not specifically marked as such, are not to be<br />

considered unprotected by law.<br />

Printed in the Federal Republic of Germany<br />

Printed on acid-free paper<br />

Typesetting Manuela Treindl, Laaber<br />

Printing betz-druck GmbH Darmstadt<br />

Bookbinding Litges & Dopf GmbH, Heppenheim<br />

ISBN: 978-3-527-31767-7


Contents<br />

Foreword V<br />

Preface VII<br />

List of Contributors XVII<br />

1 Introduction 1<br />

1.1 Initial Remarks 1<br />

Helena Dodziuk<br />

1.2 <strong>Hydrocarbons</strong> with Unusual Spatial Structure: the Need to<br />

Finance Basic Research 5<br />

Helena Dodziuk<br />

1.3 Computations on <strong>Strained</strong> <strong>Hydrocarbons</strong> 12<br />

Andrey A. Fokin and Peter R. Schreiner<br />

1.4 Gallery of Molecules That Could Have Been Included in<br />

This Book 18<br />

Helena Dodziuk<br />

1.4.1 Introductory Remarks 18<br />

1.4.2 Saturated <strong>Hydrocarbons</strong> 18<br />

1.4.3 Distorted Double Bonds 21<br />

1.4.4 Benzene Rings with Nontypical Spatial Structures 22<br />

1.4.5 Cumulenes 25<br />

1.4.6 Acetylenes 26<br />

References 27<br />

2 Distorted Saturated <strong>Hydrocarbons</strong> 33<br />

2.1 Molecules with Inverted Carbon Atoms 33<br />

Kata Mlinari�-Majerski<br />

2.1.1 Introduction 33<br />

2.1.2 Small-ring Propellanes: Computational and Physicochemical<br />

Studies 35<br />

2.1.3 Small-ring Propellanes: Experimental Results 38<br />

2.1.3.1 Preparation and Reactivity of [1.1.1]Propellane 38<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

IX


X Contents<br />

2.1.3.2 Preparation and Reactivity of [2.1.1]Propellane and<br />

[2.2.1]Propellane 41<br />

2.1.3.3 [1.1.1]Propellane as the Precursor for the Synthesis of Other<br />

Unusual Molecules 42<br />

2.1.4 New Hypothetical Molecules with Inverted Carbon Atoms 43<br />

2.2 Molecules with Planar and Pyramidal Carbon Atoms 44<br />

Helena Dodziuk<br />

2.3 A Theoretical Approach to the Study and Design of Prismane<br />

Systems 49<br />

Tatyana N. Gribanova, Vladimir I. Minkin and Ruslan M. Minyaev<br />

2.3.1 Introduction 49<br />

2.3.2 Prismanes 49<br />

2.3.3 Expanded Prismanes 52<br />

2.3.3.1 Asteranes 52<br />

2.3.3.2 Ethynyl-expanded Prismanes 54<br />

2.3.4 Dehydroprismanes 55<br />

2.3.5 Polyprismanes 56<br />

2.3.5.1 Cubane Oligomers 56<br />

2.3.5.2 Fused Prismanes 57<br />

2.3.6 Conclusions 58<br />

2.4 (CH) 2n Cage Structures, ‘in’-‘out’ Isomerism in Perhydrogenated<br />

Fullerenes and Planar Cyclohexane Rings 59<br />

Helena Dodziuk<br />

2.4.1 (CH) 2n Cage Structures 59<br />

2.4.1.1 Tetrahedrane 61<br />

2.4.1.2 Triprismane 62<br />

2.4.1.3 Cubane 61, Cuneane 100 and Octabisvalene 101 C 8 H 8 62<br />

2.4.1.4 C 10 H 10 Saturated Cages 63<br />

2.4.1.5 C 12 H 12 Saturated Cages 63<br />

2.4.1.6 Higher [n]Prismanes, Dodecahedrane 64<br />

2.4.1.7 ‘In’-‘out’ Isomerism in Perhydrogenated Fullerenes C 60 H 60 64<br />

2.4.1.8 Summary 67<br />

2.4.2 Planar Cyclohexane Rings 67<br />

2.5 Ultralong C–C Bonds 70<br />

Takanori Suzuki, Takashi Takeda, Hidetoshi Kawai and Kenshu Fujiwara<br />

2.5.1 Introduction 70<br />

2.5.2 Ultralong C–C Bonds Confi ned in a Stiff Molecular Frame 72<br />

2.5.3 Tetraphenylnaphthocyclobutene as a Scaffold to Produce Ultralong<br />

C–C Bonds 73<br />

2.5.4 ‘Clumped’ Hexaphenylethane Derivatives with Elongated and<br />

Ultralong C–C Bonds 74<br />

2.5.5 HPE Derivatives with a Super-ultralong C–C Bond 78<br />

2.5.6 ‘Expandability’ of the Ultralong C–C Bond:<br />

Conformational Isomorphs with Different Bond Lengths 79<br />

2.5.7 Future Outlook 82


2.6 Ultrashort C–C Bonds 82<br />

Vladimir Y. Lee and Akira Sekiguchi<br />

2.6.1 Introduction 82<br />

2.6.2 Tricyclo[2.1.0.0 2,5 ]pentanes: Ultrashort Endocyclic Bridging<br />

C–C Bonds 83<br />

2.6.3 Coupled Cage Compounds: Ultrashort Exocyclic Intercage<br />

C–C Bonds 86<br />

2.6.4 Sterically Congested in-Methylcyclophanes: Ultrashort C–C(Me)<br />

Bonds 91<br />

2.6.5 Conclusions 91<br />

References 92<br />

Contents<br />

3 Distorted Alkenes 103<br />

3.1 Nonplanar Alkenes 103<br />

Dieter Lenoir, Paul J. Smith and Joel F. Liebman<br />

3.1.1 Introduction and Context 103<br />

3.1.2 Bridgehead Alkenes 103<br />

3.1.2.1 t-Butyl-substituted Ethylenes 104<br />

3.1.2.2 Investigations of t-Butylated Ethylenes and Other Acyclic<br />

Alkenes 106<br />

3.1.2.3 Cyclo and Bicycloalkenes … and on to Polycyclic Analogs 107<br />

3.1.2.4 Adamantylideneadamantane and its Derivatives 108<br />

3.1.2.5 t-Butyl-substituted and Cyclic Stilbenes 108<br />

3.1.3 Multiply Unsaturated Bicycloalkenes, Homoaromaticity and<br />

Cyclophanes 109<br />

3.1.3.1 The Most Distorted Ethylenes and Seemingly Simple Analogs 111<br />

3.2 Small Ring and Cage Structures Involving Nonplanar C=C Bonds<br />

112<br />

Athanassios Nicolaides<br />

3.2.1 Pyramidalized Alkenes 112<br />

3.2.1.1 Tricyclo[3.3.11.0 3,7 ]undec-3(7)-ene 38 115<br />

3.2.1.2 Tricyclo[3.3.10.0 3,7 ]dec-3(7)-ene 39 and tricyclo[3.3.9.0 3,7 ]non-3(7)-ene 40<br />

117<br />

3.2.1.3 Tricyclo[3.3.0.0 3,7 ]oct-1(5)-ene 41 119<br />

3.2.1.4 (Ph 3 P) 2 Pt Complexes 119<br />

3.2.2 Conclusions 121<br />

3.3 <strong>Strained</strong> Cyclic Allenes and Cumulenes 122<br />

Richard P. Johnson and Kaleen M. Konrad<br />

3.3.1 Introduction 122<br />

3.3.2 Allene � Bond Deformations and Strain Estimates 123<br />

3.3.3 Four- and Five-membered Ring Allenes 124<br />

3.3.4 1,2-Cyclohexadienes 125<br />

3.3.5 1,2,4-Cyclohexatrienes 127<br />

3.3.5.1 6-Methylene-1,2,4-Cyclohexatrienes and Related Structures 131<br />

3.3.6 Seven-membered Ring Allenes 131<br />

XI


XII Contents<br />

3.3.6.1 Cycloheptatetraenes 132<br />

3.3.7 Eight-membered Ring Allenes 134<br />

3.3.8 Polycyclic Allenes 135<br />

3.3.9 Cyclic Bisallenes 136<br />

3.3.10 Cyclic Butatrienes 136<br />

3.3.10.1 Butatriene � Bond Deformations and Strain Estimates 137<br />

3.3.10.2 Five- to Nine-membered Ring Cyclic Butatrienes 137<br />

3.3.11 Conclusions 139<br />

References 140<br />

4 <strong>Strained</strong> Aromatic Molecules 147<br />

4.1 Nonstandard Benzenes 147<br />

Paul J. Smith and Joel F. Liebman<br />

4.1.1 Introduction and Context 147<br />

4.1.2 Alkylated Aromatics 148<br />

4.1.3 Helicenes 148<br />

4.1.4 [n]Circulenes 149<br />

4.1.5 Cyclophanes 150<br />

4.2 Distorted Cyclophanes 153<br />

Henning Hopf<br />

4.2.1 Introduction 153<br />

4.2.2 The [n]Cyclophanes 154<br />

4.2.2.1 [n]Paracyclophanes 154<br />

4.2.2.2 [n]Metacyclophanes 160<br />

4.2.3 The [m.n]Paracyclophanes 161<br />

4.2.4 Distorted Aromatic Rings and ‘Aromatic Character’ 164<br />

4.2.5 NMR Characteristics of Cyclophanes 165<br />

4.3 Helicenes 166<br />

Ivo Starý and Irena G. Stará<br />

4.3.1 Introduction 166<br />

4.3.2 Synthesis of Helicenes 166<br />

4.3.3 Nonracemic Helicenes 171<br />

4.3.4 Intriguing Helicene Structures 172<br />

4.3.5 Physicochemical Properties and Applications 173<br />

4.3.6 Theoretical Studies 175<br />

4.3.7 Outlook 176<br />

4.4 Cycloproparenes 176<br />

Brian Halton<br />

4.4.1 Introduction 176<br />

4.4.2 Synthetic Considerations 177<br />

4.4.3 Chemical Considerations 183<br />

4.4.4 Heteroatom Derivatives 187<br />

4.4.5 Physicochemical and Theoretical Considerations 188<br />

References 193


5 Fullerenes 205<br />

5.1 Introduction 205<br />

Helena Dodziuk<br />

5.2 Chemistry Infl uenced by the Nontypical Structure: Modifi cation of<br />

[60]Fullerene 208<br />

Takuma Hara, Takashi Konno, Yosuke Nakamura and Jun Nishimura<br />

5.2.1 Introduction 208<br />

5.2.2 General Overview 209<br />

5.2.3 Modifi cation Reactions 215<br />

5.2.3.1 Reduction and Oxidation 215<br />

5.2.3.2 Alkylation 217<br />

5.2.3.3 Cycloadditions 218<br />

5.2.4 Conclusions 224<br />

5.3 Physicochemical Properties and the Unusual Structure of<br />

Fullerenes 225<br />

5.3.1 Single-crystal X-ray Structures of Fullerenes and Their<br />

Derivatives 225<br />

Olga V. Boltalina, Alexey A. Popov and Steven H. Strauss<br />

5.3.1.1 Introduction 225<br />

5.3.1.2 Disorder 226<br />

5.3.1.3 Nonplanar Steric Strain 226<br />

5.3.1.4 Nonplanar Steric Strain Parameters 229<br />

5.3.1.5 Are Non-IPR Fullerenes Sterically Unstable? 232<br />

5.3.1.6 Long and Short C(sp 2 )–C(sp 2 ) Bonds in Fullerene Cages 232<br />

5.3.1.7 Steric Strain in C 60 (X) n Isomers 236<br />

5.3.2 Vibrational and Electronic Spectra 238<br />

Alexey A. Popov<br />

5.3.2.1 Introduction 238<br />

5.3.2.2 Vibrational Spectra of Fullerenes 239<br />

5.3.2.3 The Orbital Picture of Fullerenes: High-energy Electronic<br />

Spectra 243<br />

5.3.2.4 Electronic Excitations. UPS, UV/Vis/NIR Absorption and<br />

Fluorescence Spectroscopy 246<br />

5.3.3 Nuclear Magnetic Resonance 250<br />

Toni Shiroka<br />

5.3.3.1 Introduction 250<br />

5.3.3.2 NMR of Fullerenes 251<br />

5.3.3.3 Concluding Remarks 259<br />

5.3.4 Electrochemistry 259<br />

Renata Bilewicz and Kazimierz Chmurski<br />

5.3.4.1 Electronic Properties of Fullerenes 259<br />

5.3.4.2 Electrochemical Properties of Soluble Fullerene Derivatives 263<br />

5.3.4.3 Electrocatalytic Activity of Fullerenes 270<br />

5.3.4.4 Conclusions and Outlook 272<br />

5.4 Fullerene Aggregates 273<br />

Contents<br />

XIII


XIV<br />

Contents<br />

Tommi Vuorinen<br />

5.4.1 Film Preparation Methods 274<br />

5.4.2 Fullerene Film Properties 277<br />

5.4.3 Conclusions 282<br />

5.5 Endohedral Fullerenes with Neutral Atoms and Molecules 282<br />

Sho-ichi Iwamatsu<br />

5.5.1 Introduction 282<br />

5.5.2 Preparation 282<br />

5.5.2.1 Direct Approach Using an Existing Fullerene 282<br />

5.5.2.2 Molecular Surgery Approach via an Open-cage Fullerene 284<br />

5.5.2.3 Open-cage Fullerenes, Reversible Molecular Incorporations<br />

and Ejections 285<br />

5.5.3 Properties 287<br />

5.5.3.1 Host Fullerenes 287<br />

5.5.3.2 Guest Substrates 288<br />

5.5.4 Binding Energies, Theoretical Investigations 290<br />

5.5.5 Summary 291<br />

5.6 Hydrogenated Fullerenes 291<br />

Mark S. Meier<br />

5.6.1 Synthesis and Structure 291<br />

5.6.2 C 70 Chemistry 295<br />

5.6.3 Higher Fullerenes 297<br />

5.6.4 Reactivity of Hydrogenated Fullerenes 297<br />

5.7 Applications of Fullerenes 299<br />

Rossimiriam Pereira de Freitas and Jean-François Nierengarten<br />

5.7.1 Introduction 299<br />

5.7.2 Applications in Materials Science 299<br />

5.7.2.1 C 60 Derivatives for Optical Limiting Applications 299<br />

5.7.2.2 C 60 Derivatives for Photovoltaic Applications 304<br />

5.7.3 Biological Applications 310<br />

5.7.4 Conclusions 314<br />

References 315<br />

6 Carbon Nanotubes 335<br />

6.1 The Structure and Properties of Carbon Nanotubes 335<br />

Anke Krueger<br />

6.1.1 Introduction 335<br />

6.1.2 The Structure of Single-walled Carbon Nanotubes 335<br />

6.1.3 The Structure of Multi-walled Carbon Nanotubes 342<br />

6.1.4 The Aromaticity of Carbon Nanotubes 345<br />

6.1.5 Conclusions 347<br />

6.2 The Functionalization of Carbon Nanotubes 347<br />

Anke Krueger<br />

6.2.1 Introduction 347<br />

6.2.2 Functionalization of the Nanotube Tips 348


6.2.3 Non-covalent Functionalization of Carbon Nanotubes 349<br />

6.2.4 Covalent Side-wall Functionalization of Carbon Nanotubes 352<br />

6.2.5 Endohedral Functionalization of Carbon Nanotubes 355<br />

6.2.6 Conclusions 356<br />

6.3 Applications of Carbon Nanotubes 356<br />

Marc Monthioux<br />

6.3.1 Introduction 356<br />

6.3.2 Properties of CNTs 357<br />

6.3.2.1 Which CNT for Which Application? 357<br />

6.3.2.2 Why is ‘Nano’ Beautiful? 358<br />

6.3.2.3 Potential Problems Related to the Use of CNTs 360<br />

6.3.3 Applications of CNTs 362<br />

6.3.3.1 Prospective Applications 362<br />

6.3.3.2 Applications Under Development 364<br />

6.3.3.3 Applications on the Market 366<br />

6.3.4 Conclusions 367<br />

References 368<br />

7 Angle-strained Cycloalkynes 375<br />

Henning Hopf and Jörg Grunenberg<br />

7.1 Introduction 375<br />

7.2 Cyclopropyne and Cyclobutyne: Speculations and Calculations<br />

on Non-isolable Cycloalkynes 376<br />

7.2.1 Cyclopropyne and Related Systems 376<br />

7.2.2 Cyclobutyne 378<br />

7.3 Cyclopentyne, Cyclohexyne, Cycloheptyne: from Reactive<br />

Intermediates to Isolable Compounds 379<br />

7.3.1 Cyclopentyne and its Derivatives 379<br />

7.3.2 Cyclohexyne and its Derivatives 382<br />

7.3.3 Cycloheptyne and its Derivatives 384<br />

7.4 The Isolable Angle-strained Cycloalkynes: Cyclooctyne,<br />

Cyclononyne, and Beyond 385<br />

7.4.1 Cyclooctyne and its Derivatives 385<br />

7.4.2 Cyclononyne and Cyclodecyne 386<br />

7.5 Cyclic Polyacetylenes 387<br />

7.6 Spectroscopic Properties of Angle-strained Cycloalkynes 392<br />

References 393<br />

8 Molecules with Labile Bonds:<br />

Selected Annulenes and Bridged Homotropilidenes 399<br />

Richard V. Williams<br />

8.1 Introduction 399<br />

8.2 Annulenes 399<br />

8.2.1 Cyclobutadiene 399<br />

8.3 Cyclooctatetraene 403<br />

Contents<br />

XV


XVI<br />

Contents<br />

8.4 Bond Shifting, Ring Inversion and Antiaromaticity 405<br />

8.5 Valence Isomerization 409<br />

8.6 Ions Derived from COT 410<br />

8.7 The Higher Annulenes 411<br />

8.8 Bridged Homotropilidenes 413<br />

8.9 Recent Developments 415<br />

8.9 Conclusions 419<br />

References 420<br />

9 Molecules with Nonstandard Topological Properties: Centrohexaindane,<br />

Kuratowski’s Cyclophane and Other Graph-theoretically Nonplanar<br />

Molecules 425<br />

Dietmar Kuck<br />

9.1 Introduction 425<br />

9.1.1 Is All This Trivial? 425<br />

9.2 Topologically Nonplanar Graphs and Molecular Motifs 427<br />

9.2.1 The Centrohexaquinacene Core 427<br />

9.2.2 The Nonplanar Graphs K 5 and K 3,3 and Some Molecular<br />

Representatives 428<br />

9.3 Centrohexaindane 430<br />

9.3.1 Centrohexaindane and Structural Regularities of the Centropolyindane<br />

Family 431<br />

9.3.2 Syntheses of Centrohexaindane 433<br />

9.3.3 Multiply-functionalized Centrohexaindanes 436<br />

9.4 K 5 versus K 3,3 Molecules 438<br />

9.4.1 Topologically Nonplanar Polyethers and Other K 3,3 Compounds 438<br />

9.5 Kuratowski’s Cyclophane 441<br />

9.5.1 Synthesis of Kuratowski’s Cyclophane 441<br />

9.5.2 The Structure of Kuratowski’s Cyclophane 443<br />

9.6 Conclusions 444<br />

References 445<br />

10 Short-lived Species Stabilized in ‘Molecular’ or ‘Supramolecular Flasks’ 449<br />

Helena Dodziuk<br />

References 456<br />

11 Concluding Remarks 459<br />

Helena Dodziuk<br />

References 461<br />

Index 463


Foreword<br />

Chemistry is the truly anthropic science. The molecules we make can heal us, and<br />

they can hurt us, because they are on the scale of the molecules that make up our<br />

bodies. And our synthetic creations interact, even react, with the molecules that<br />

nature – our enzymes, the environment – put into us.<br />

Molecular science is also anthropic (male and female, of course) because it<br />

presents a challenge to human intelligence that is just right, commensurate with<br />

our intellect. The exciting story this book develops bears testimony on every page<br />

to that anthropic cognitive nature of organic chemistry.<br />

Let me explain: our remarkable neural system is steered by a complex brain.<br />

That brain has prejudices for sure; it tends to simplify things, falling at every<br />

proffered opportunity for beautiful equations, simple mechanisms, Platonic solids<br />

and the honeyed simplicities of politicians. But when challenged, we can deal with<br />

substantial complexity. Indeed, the brain relishes being stretched: by rich sensual<br />

inputs, by patterns, by puzzles.<br />

Along comes a science, our chemistry. It offers in its molecular structures, a<br />

game that is at first sight deceptively simple. Take hydrocarbons (most of the<br />

molecules in this book are in this category) – what could be simpler? Two elements,<br />

C and H, that by a transparent rule of intercombination form four bonds, and<br />

one bond, respectively. You are well aware of the manifestation of these rules and<br />

combinatorics – a chemical universe of incredible diversity.<br />

These molecules can not only be thought up, they can also be synthesized in a<br />

human span – roughly the time it takes for a graduate student to get a Ph.D. We<br />

are not making a ladybug, nor a spiral galaxy; we are making a paracyclophane.<br />

The complexity of the challenge is on the human scale. And so are the possibilities:<br />

What can I do to string eight carbons across the para positions of a benzene? Can<br />

I reduce the bridging carbons to seven? Will I make it easier if the eight carbons<br />

are partially in a benzene ring themselves? The questions just flow one after the<br />

other; it takes no talent to ask them, just a normal curious human being, privy to<br />

the structural codes of chemistry.<br />

So the game itself, the game of chemical structure, is exciting. Chess pales<br />

by comparison. Add to that ludic challenge potential utility, and also the natural<br />

human desire to probe limits (just how far can I distort that double bond out of<br />

its planar normalcy?), and you have all the makings of intense interaction, part<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

V


VI Foreword<br />

intellectual, part emotional, between a human being and an object of his or her<br />

creation.<br />

The object of our intense contemplation – a compound macroscopically, a<br />

molecule microscopically – is complex enough not to be boring, yet not unpredictably<br />

chaotic. The strained molecule is just right for some of us to exercise our<br />

creativity in thinking up these strange beasts, others in coming up with ingenious<br />

ways of making them (for molecules are real!), all of us admiring the complexity,<br />

simplicity and function all rolled into one.<br />

Enjoy reading this book! Roald Hoffmann


Preface<br />

<strong>Strained</strong> hydrocarbons represent an amazing domain. About 80 years after<br />

the formulation of van’t Hoff and Le Bel’s hypothesis new, exciting molecules<br />

representing in Hoffmann and Hopf formulation (R. Hoffmann, H. Hopf, Angewandte<br />

Chemie, submitted), what is probably too much of anthropomorphization,<br />

‘molecular sadism’ were synthesized. Paraphrasing D. J. Cram, one could<br />

say that such molecules elicit wonder, stimulate the imagination and challenge<br />

both synthetic talents and interpretive instincts. Up to the early 1990s the field of<br />

strained hydrocarbons was a kind of elitist area in which only the best synthetic<br />

and theoretical chemists were active. It was a playground of few, characterized<br />

by vivid interactions between synthetic and theoretical chemistry allowing one to<br />

propose plausible synthetic targets on the basis of model calculations. On the other<br />

hand, it allowed Bader, Wiberg and their followers to refine the definition of the<br />

chemical bond. The situation in the domain of distorted molecules changed after<br />

the discovery of fullerenes and nanotubes which attracted numerous researchers.<br />

These molecules, having nonplanar systems of conjugated bonds, are not<br />

hydrocarbons but their derivatives are numerous. Thus, they have been included<br />

into this volume in view of the rapid development of these areas and, still largely<br />

unfulfilled, prospects of their applications.<br />

Several researchers helped me in this project. First of all, I would like to thank<br />

all contributors to this volume. I would like also to acknowledge the support I have<br />

obtained from Professors T. Marek Krygowski, Jay S. Siegel and Henning Hopf<br />

in the initial stage. Finding contributors was sometimes a difficult task. The help<br />

of Professors A. de Meijere, E. Osawa, F. Diederich, C. Thilgen, W. T. Borden,<br />

J. Cioslowski and H. Kuzmany in the search for coauthors is gratefully acknowledged.<br />

I am deeply obliged to my colleague, Dr. K. S. Nowinski, for designing the<br />

cover picture. On the other hand, I owe a deep apology to the authors of many<br />

interesting papers on strained hydrocarbons which could not be presented in this<br />

book or were insufficiently covered due to space limitations.<br />

The question: ‘To what extent can a bond be distorted without being broken?’<br />

is fascinating. This book is devoted to the presentation of distorted hydrocarbons.<br />

It is an effort to counteract, in this limited volume, overspecialization by showing<br />

not only syntheses, physicochemical studies and theoretical calculations of these<br />

molecules, but also the prospects of their applications. <strong>Strained</strong> molecules are<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

VII


VIII<br />

Preface<br />

exciting objects for studies per se. With several novel hypothetical molecules<br />

waiting to be synthesized on the one hand, and with the possibility of obtaining<br />

fascinating supramolecular complexes with distorted hydrocarbons as building<br />

blocks on the other, this domain will remain enthralling.<br />

Warsaw, January 2009 Helena Dodziuk


List of Contributors<br />

Renata Bilewicz<br />

University of Warsaw<br />

Department of Chemistry<br />

Pasteura 1<br />

02093 Warsaw<br />

Poland<br />

Olga V. Boltalina<br />

Colorado State University<br />

Department of Chemistry<br />

Fort Collins, CO 80523<br />

USA<br />

Kazimierz Chmurski<br />

University of Warsaw<br />

Department of Chemistry<br />

Pasteura 1<br />

02-093 Warsaw<br />

Poland<br />

Helena Dodziuk<br />

Polish Academy of Sciences<br />

Institute of Physical Chemistry<br />

Kasprzaka 44/52<br />

01-224 Warsaw<br />

Poland<br />

Andrey A. Fokin<br />

Kiev Polytechnic Institute<br />

Department of Organic Chemistry<br />

Spr. Pobedy 37<br />

03056 Kiev<br />

Ukraine<br />

Kenshu Fujiwara<br />

Hokkaido Univeresity<br />

Faculty of Science<br />

Department of Chemistry<br />

N10W8, North-ward<br />

Sapporo 060-0810<br />

Japan<br />

Tatyana N. Gribanova<br />

Southern Federal University<br />

Institute of Physical and Organic<br />

Chemistry<br />

194/2 Stachka Ave<br />

344090 Rostov on Don<br />

Russia<br />

Jörg Grunenberg<br />

Carolo Wilhelmina Technical<br />

University of Braunschweig<br />

Institute of Organic Chemistry<br />

Hagenring 30<br />

38106 Braunschweig<br />

Germany<br />

Brian Halton<br />

Victoria University of Wellington<br />

School of Chemical & Physical<br />

Sciences<br />

PO Box 600<br />

Wellington 6140<br />

New Zealand<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

XVII


XVIII<br />

List of Contributors<br />

Takuma Hara<br />

Evonik Degussa Japan Co. Ltd.<br />

Business Line Catalysts<br />

Tsukuba Minami Daiichi Kogyo<br />

Danchi<br />

21 Kasuminosato Ami-machi<br />

Inashiki-gun Ibaraki-ken 300-0315<br />

Japan<br />

Henning Hopf<br />

Technical University of<br />

Braunschweig<br />

Carolo Wilhelmina<br />

Institute of Organic Chemistry<br />

Hagenring 30<br />

38106 Braunschweig<br />

Germany<br />

Sho-ichi Iwamatsu<br />

Nagoya University<br />

Graduate School of Environmental<br />

Studies<br />

Chikusa Ku<br />

Nagoya, Aichi 464-8601<br />

Japan<br />

Richard P. Johnson<br />

University of New Hampshire<br />

Department of Chemistry<br />

Durham, NH 03824<br />

USA<br />

Hidetoshi Kawai<br />

Hokkaido University<br />

Faculty of Science<br />

Department of Chemistry<br />

N10W8, North-ward<br />

Sapporo 060-0810<br />

Japan<br />

Takashi Konno<br />

Gunma University<br />

Graduate School of Engineering<br />

Department of Chemistry and<br />

Chemical Biology<br />

Tenjincho 1-5-1<br />

Kiryu, Gunma 376-8515<br />

Japan<br />

Kaleen M. Konrad<br />

Merck Research Laboratories<br />

33 Avenue Louis Pasteur<br />

Boston, MA 02115<br />

USA<br />

Anke Krüger<br />

University of Kiel<br />

Otto-Diehls-Institute of Organic<br />

Chemistry<br />

Otto-Hahn-Platz 3<br />

24098 Kiel<br />

Germany<br />

Dietmar Kuck<br />

Universität Bielefeld<br />

Fakultät für Chemie<br />

Postfach 100131<br />

33501 Bielefeld<br />

Germany<br />

Vladimir Y. Lee<br />

University of Tsukuba<br />

Graduate School of Pure and<br />

Applied Science<br />

Department of Chemistry<br />

Tsukuba, Ibaraki 305-8571<br />

Japan<br />

Dieter Lenoir<br />

GSF-Research Center Neuherberg<br />

Institute of Ecological Chemistry<br />

Postfach 1129<br />

85758 Neuherberg<br />

Germany


Joel F. Liebman<br />

University of Maryland,<br />

Baltimore County<br />

Department of Chemistry &<br />

Biochemistry<br />

1000 Hilltop Circle<br />

Baltimore, MD 21250<br />

USA<br />

Mark S. Meier<br />

University of Kentucky<br />

Department of Chemistry<br />

Lexington, KY 40506<br />

USA<br />

Vladimir I. Minkin<br />

Southern Federal University<br />

Institute of Physical and Organic<br />

Chemistry<br />

194/2 Stachka Ave<br />

344090 Rostov on Don<br />

Russia<br />

Ruslan M. Minyaev<br />

Southern Federal University<br />

Institute of Physical and Organic<br />

Chemistry<br />

194/2 Stachka Ave<br />

344090 Rostov on Don<br />

Russia<br />

Kata Mlinarić-Majerski<br />

Rudjer Boškovi� Institute<br />

Department of Organic Chemistry<br />

and Biochemistry<br />

Bijeni�ka 54, P.O. Box 180<br />

10002 Zagreb<br />

Croatia<br />

Marc Monthioux<br />

Université Toulouse III<br />

CEMES, UPR-8011 CNRS<br />

BP 94347<br />

29 rue Jeanne Marvik<br />

31055 Toulouse Cedex 4<br />

France<br />

Yosuke Nakamura<br />

Gunma University<br />

Graduate School of Engineering<br />

Department of Chemistry and<br />

Chemical Biology<br />

Tenjincho 1-5-1<br />

Kiryu, Gunma 376-8515<br />

Japan<br />

Jean-François Nierengarten<br />

Université de Strasbourg<br />

Laboratoire de Chemie des Matériaux<br />

Moléculaires (UMR 7509)<br />

25 rue Becquerel<br />

67087 Strasbourg Cedex 2<br />

France<br />

Athanassios Nikolaides<br />

University of Cyprus<br />

Department of Chemistry<br />

P.O. Box 20537<br />

Nicosia 1678<br />

Cyprus<br />

List of Contributors<br />

Jun Nishimura<br />

Gunma University<br />

Graduate School of Engineering<br />

Department of Nano-Material<br />

Systems<br />

Tenjincho 1-5-1<br />

Kiryu, Gunma 376-8515<br />

Japan<br />

XIX


XX<br />

List of Contributors<br />

Rossimiriam Pereira de Freitas<br />

Universidade Federal de Minas<br />

Gerais<br />

Departamento de Química<br />

Av. Antônio Carlos 6627<br />

Belo Horizonte<br />

312-70901, MG<br />

Brazil<br />

Alexey A. Popov<br />

Leibniz-Institute for Solid State<br />

and Materials Research IFW<br />

Group of Electrochemistry and<br />

Conducting Polymers<br />

Helmholtzstrasse 20<br />

01069 Dresden<br />

Germany<br />

Peter R. Schreiner<br />

Justus-Liebig University<br />

Institut für Organische Chemie<br />

Heinrich-Buff-Ring 58<br />

35392 Giessen<br />

Germany<br />

Akira Sekiguchi<br />

University of Tsukuba<br />

Department of Chemistry<br />

Graduate School of Pure and<br />

Applied Sciences<br />

Tsukuba, Ibaraki 305-8571<br />

Japan<br />

Toni Shiroka<br />

Paul Scherrer Institut<br />

Laboratory for Muon-Spin<br />

Spectroscopy<br />

CH-5232 Villigen PSI<br />

Switzerland<br />

Paul J. Smith<br />

University of Maryland Baltimore<br />

County<br />

Department of Chemistry &<br />

Biochemistry<br />

1000 Hilltop Circle<br />

Baltimore, MD 21250<br />

USA<br />

Irena G. Stará<br />

Academy of Sciences of the Czech<br />

Republic<br />

Institute of Organic Chemistry &<br />

Biochemistry<br />

Center for Biomolecules & Complex<br />

Molecular Systems<br />

Flemigovo nám. 2<br />

16610 Prague 6<br />

Czech Republic<br />

Ivo Starý<br />

Academy of Sciences of the Czech<br />

Republic<br />

Institute of Organic Chemistry &<br />

Biochemistry<br />

Center for Biomolecules & Complex<br />

Molecular Systems<br />

Flemigovo nám. 2<br />

16610 Prague 6<br />

Czech Republic<br />

Steven H. Strauss<br />

Colorado State University<br />

Department of Chemistry<br />

Fort Collins, CO 80523<br />

USA<br />

Takanori Suzuki<br />

Hokkaido University<br />

Faculty of Science<br />

Department of Chemistry<br />

N10W8, North-ward<br />

Sapporo 060-0810<br />

Japan


Takashi Takeda<br />

Hokkaido University<br />

Faculty of Science<br />

Departmeht of Chemistry<br />

N10W8, North-ward<br />

Sapporo 060-0810<br />

Japan<br />

Tommi Vuorinen<br />

Tampere University of Technology<br />

Institute of Materials Chemistry<br />

P.O. Box 541<br />

33101 Tampere<br />

Finland<br />

Richard V. Williams<br />

University of Idaho<br />

Department of Chemistry<br />

P.O. Box 442343<br />

Moscow, ID 83844-2343<br />

USA<br />

List of Contributors<br />

XXI


1<br />

Introduction<br />

1.1<br />

Initial Remarks<br />

Helena Dodziuk<br />

Let us start with a bit of history. Today, it is hard to imagine how difficult it was<br />

to develop basic concepts and ideas of chemistry in the second half of the 19th<br />

century. The story about Kekulé’s fight for his benzene structure shows that not<br />

all the arguments he used in its favor are valid today [1]. His idea could not be supported<br />

by the poor experimental instrumentation of that time. There was no X-ray<br />

analysis, no modern spectroscopic techniques and no calorimetry. The idea of the<br />

constitution of molecules, that is building them from a certain number of different<br />

types of atoms, was established, as well as several experimental findings which<br />

demanded rationalization. Among them were optical activity and the existence of<br />

a number of different molecules with the same constitution. Pasteur foresaw that<br />

the former phenomenon could be related to the positioning of atoms in space but<br />

only the van’t Hoff [2] and Le Bel [3] hypotheses on the tetrahedral arrangement of<br />

substituents on the tetravalent carbon atom explained most observations known<br />

at that time. Interestingly, the independently proposed models differed slightly:<br />

that of van’t Hoff was based on a regular tetrahedron, while in the second one used<br />

an irregular tetrahedron to represent the carbon atom. This difference was not<br />

significant but, remarkably, the more idealized van’t Hoff approach was generally<br />

accepted. An illustration from the 1908 German edition of van’t Hoff’s book, showing<br />

two stereoisomers of the tetrasubstituted ethane molecule CR1CR2rCR3CR4r<br />

shows the way in which molecules were depicted at that time (Figure 1.1).<br />

The van’t Hoff and Le Bel hypothesis was met with strong criticism, not always<br />

expressed in impartial scientific language. The renowned chemist and editor of the<br />

German Journal für praktische Chemie, Prof. Adolf Kolbe wrote: ‘A Dr. H. van ’t Hoff<br />

of the Veterinary School at Utrecht has no liking, apparently, for exact chemical<br />

investigation. He has considered it more comfortable to mount Pegasus (apparently<br />

borrowed from the Veterinary School) and to proclaim in his ‘La chimie<br />

dans l’éspace’ how the atoms appear to him to be arranged in space, when he is<br />

on the chemical Mt. Parnassus which he has reached by bold fly’.<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

1


2 1 Introduction<br />

Figure 1.1 The representation of two stereoisomers of a tetrasubstituted ethane molecule<br />

CR1CR2rCR3CR4r, as published 100 years ago in van’t Hoff’s book [197].<br />

In spite of such a strong attack and the difficulties associated with the lack of<br />

modern physicochemical methods, the idea of the tetrahedral arrangement of<br />

substituents around a tetravalent carbon atom was generally accepted and van’t<br />

Hoff became a first recipient of the Nobel prize for Chemistry in 1901, interestingly<br />

not for his stereochemical ideas but ‘in recognition of the extraordinary<br />

services he has rendered by the discovery of the laws of chemical dynamics and<br />

osmotic pressure in solutions’.<br />

Remarkably, an early idea of Sachse on the cyclohexane conformations (today<br />

known under the names chair and twist-boat) [4, 5] could not be proved at that<br />

time and was not accepted. Then, it took almost 80 years to understand all the<br />

consequences of the van’t Hoff and Le Bel concepts which not always were based<br />

on justified assumptions. For instance, the van’t Hoff understanding of the C–C<br />

bond implied free rotation around it. This assumption was only shown to be invalid<br />

by Pitzer’s work [6, 7] on the hindering of the rotation and preferred orientations<br />

of substituents on the C–C bond, started in 1936, which marked an important<br />

step in development of stereochemistry [8]. The combination of the ideas on<br />

tetrahedral orientation of substituents on a tetravalent carbon atom and of the<br />

hindered rotation around the C–C bond resulted in rationalization of the cyclohexane<br />

conformations and the number of isomers of its derivatives summarized<br />

in the Hassel [9, 10] and Barton [11] studies which were also honored by a Nobel<br />

Prize ‘for their contributions to the development of the concept of conformation and<br />

its application in chemistry’. Analogous studies of the spatial structure of alkenes,<br />

alkynes and aromatic compounds followed.<br />

With these achievements, the basis of the organic stereochemistry seemed to<br />

be laid, and models could be built, of spatial structures of molecules from welldefined<br />

rigid fragments. Eaton’s report on the synthesis of cubane 1 in 1964 [12]<br />

and especially the Wiberg synthesis of [1.1.1]propellane 2 [13] have shown that,<br />

in addition to the small-ring cycloalkanes, well-known since the second half of<br />

the 19th century, that exhibit Bayer strain [14], hydrocarbons having structures<br />

strongly departing from that suggested by van’t Hoff and Le Bel can exist. This book<br />

is devoted to such nonstandard structures. Let us first define what the standard<br />

hydrocarbons are: first, these are saturated hydrocarbons with the arrangement<br />

of substituents on the carbon atoms close to tetrahedral; then, double bonds and<br />

aromatic rings lying in a plane with its substituents and, last but not least, linear


1.1 Initial Remarks<br />

acetylenes. Also of interest are bond lengths that depart far from the standard<br />

value of 154 pm. Of course, fullerenes and carbon nanotubes in their idealized<br />

form are not hydrocarbons, but these conjugated aromatic systems are nonplanar<br />

and they are definitely the most widely studied distorted aromatic systems today.<br />

They also offer a unique possibility of investigating the effect of nonplanarity on<br />

structure, physicochemical properties and reactivity. Actually, this book should<br />

also be understood as part of my private campaign against too deep specialization,<br />

from which we all suffer. Thus, showing the influence exerted by molecular<br />

distortions on various physicochemical properties using fullerenes as an example<br />

seemed to be of importance.<br />

Highly distorted hydrocarbons are sometimes considered to be of no importance<br />

in view of their lack of practical applications. The significance of investigating such<br />

systems is discussed in Section 1.2. They are studied by both experimental and<br />

theoretical methods that, as discussed in Section 1.3, are of special significance<br />

in this domain. As shown by several examples (among other 1 [15] and heptacyclo[6.4.0.0<br />

2,4 .0 3,7 .0 5,12 .0 6,10 .0 9,11 ]dodecane 3 [16, 17] which have first been studied<br />

theoretically then synthesized [13, 18]), to propose novel plausible synthetic targets<br />

on the basis of molecular modeling is a reliable aim of calculations.<br />

We are experiencing such a rapid development of this science that it is not<br />

possible to discuss all the unusual hydrocarbons. Therefore, a selection, by no<br />

means considered to be exhaustive, of interesting molecules which did not find<br />

a place in other chapters is presented in Section 1.4.<br />

Simple strained saturated hydrocarbons are presented in Chapter 2. Both<br />

known, such as [1.1.1]propellane 2, and hypothetical molecules having inverted<br />

carbon atoms are discussed in Section 2.1. The importance of these molecules<br />

is emphasized by the discussion of the existence of the central bond in 2 which<br />

permitted precise definition of a bond in quantum chemistry [19]. The fascinating<br />

Hoffmann idea of planar carbon atom lying in a plane with its four substituents<br />

[20] was only realized in silico [21] and, as described in Section 2.2, still awaits<br />

realization. Prismanes (one of which is cubane 1) and asteranes are discussed<br />

mainly from the theoretical point of view in Section 2.3 but the influence of<br />

molecular distortions on the properties of the known systems is presented there,<br />

too. Saturated hydrocarbon cages and planar cyclohexanes (Section 2.4) as well<br />

as molecules with ultralong (Section 2.5) and ultrashort (Section 2.6) C–C bonds<br />

are also discussed in Chapter 2.<br />

In Chapter 3 which is devoted to alkenes, energetic aspects of distorted double<br />

bonds are presented in Section 3.1, small cage alkenes are discussed in Section 3.2<br />

while Section 3.3 is devoted to strained cumulenes.<br />

3


4 1 Introduction<br />

Analogously, in Chapter 4 energetical considerations concerning strained<br />

aromatic systems are discussed in Section 4.1, while bridged aromatic rings<br />

(i.e. cyclophanes), helicenes like 4, and cycloproparenes like 5 are presented in<br />

Sections 4.2 to 4.4, respectively.<br />

As discussed earlier fullerenes, to which the longest Chapter 5 is devoted,<br />

are actually not hydrocarbons but their extended closed nonplanar conjugated<br />

aromatic systems deserve to be discussed in this book. After a short introduction<br />

in Section 5.1 their chemistry is presented in Section 5.2, and physicochemical<br />

properties reflecting their distorted structure (X-ray in Section 5.3.1, UV/Vis<br />

spectra in Section 5.3.2, NMR spectra in Section 5.3.3 and electrochemistry in<br />

Section 5.3.4) are shown. Next, fullerene films (Section 5.4), endohedral fullerene<br />

complexes (Section 5.5), an exciting application of NMR to study the structure of<br />

hydrogenated fullerenes (Section 5.6) and, mostly prospective, fullerene applications<br />

(Section 5.7). Unfortunately, I did not succeed in finding a specialist willing<br />

to present theoretical fullerene studies and their limitations due to the size of<br />

these huge cage molecules. This a serious drawback in spite of the inclusion of<br />

some fullerene calculations in other chapters.<br />

The inclusion of nanotubes into this monograph was based on similar arguments<br />

as those advocating the inclusion of fullerenes. In Chapter 6, structure, chemistry<br />

and also mostly prospective applications are shown (Sections 6.1–6.3, respectively).<br />

There has been a fascinating development in the financing of carbon nanotubes:<br />

people expected to get really big money from investing in CNTs about five years ago<br />

and then realized that the returns do not come immediately. Some of the applications<br />

introduced or expected to be introduced soon in large-scale manufacturing<br />

have not been successful. For instance, using polymer nanocomposites containing<br />

small amount of CNTs for electropainting of cars has been abandoned by General<br />

Motors, and Korean plans to build a factory for displays involving CNTs few<br />

years ago have not been fulfilled. Moreover, a recent observation on the carcenogenicity<br />

of multiwalled nanotubes may further slow down the development of<br />

CNT applications [22]. Even if we cannot see their rapid introduction, I am sure<br />

they will be important in the longer run.<br />

Cyclic alkynes with nonlinear triple bonds are discussed in Chapter 7, while<br />

molecules with labile bonds are presented in Chapter 8. They include rigid<br />

cyclo butadiene 6, highly mobile molecules like cyclooctatetraene 7 and shortlived<br />

species that could only be trapped at very low temperature in matrices.<br />

The fascinating discussion of interconversions between some of these species is


1.2 <strong>Hydrocarbons</strong> with Unusual Spatial Structure: the Need to Finance Basic Research<br />

presented in some detail from both theoretical and experimental (mainly NMR,<br />

IR) points of view.<br />

<strong>Hydrocarbons</strong> with nonplanar graphs, discussed in Chapter 9, conform to van’t<br />

Hoff and Le Bel stereochemistry but they are so unusual that they were considered<br />

to fit into this monograph. Such molecules having distinct topological properties<br />

represent an exciting border area between chemistry and mathematics. The<br />

molecules presented in this chapter have been obtained using elegant methods of<br />

traditional organic chemistry, while other systems exhibiting nontrivial topological<br />

properties (catenanes, rotaxanes, knots, etc.) could mostly be obtained by taking<br />

advantage of methods typical of supramolecular chemistry [23].<br />

In the last Chapter 10, novel ways of obtaining and stabilizing unstable highly<br />

strained species in ‘molecular flasks’ are presented. The latter method makes use<br />

of supramolecular chemistry enabling, for instance, storage of the highly unstable<br />

cyclobutadiene 6 for a month at room temperature [24].<br />

1.2<br />

<strong>Hydrocarbons</strong> with Unusual Spatial Structure: the Need to Finance Basic Research<br />

Helena Dodziuk<br />

By reporting the results on hypothetical hexahydrosuperphane 8 [25], a highly<br />

strained molecule with a planar cyclohexane ring, we have been confronted with<br />

the question as to whether this molecule had been chosen for study because of<br />

its future practical applications. The question was posed by a computer scientist<br />

who had no knowledge of stereochemistry, but it corresponds to a general attitude<br />

governed by grant funding for scientific research. Namely, any grant application<br />

has to show its immediate usefulness. However, in reality hardly any grants that<br />

use such justification will bring marketable results at all. For the applications<br />

of others we will probably wait 10 or more years and only a few of such grants<br />

will find their way to industry soon. Let’s inspect some examples in some detail.<br />

Liquid crystals today are commonly applied in displays but several other uses (as<br />

surface thermometers showing temperature distribution over a body, in optical<br />

5


6 1 Introduction<br />

imaging, etc., http://plc.cwru.edu/tutorial/enhanced/files/textbook.htm) are<br />

foreseen for them. Discovered in 1888 by a botanist, Reinitzer [26], they were for<br />

almost 100 years considered to be the physicists’ toys. In 1966, an article entitled<br />

‘Liquid crystals – an area of research of little use?’ appeared in the German journal<br />

Nachrichten für Chemie, Technik und Labor [27]. Then in 1972 the first liquid crystal<br />

display was built giving birth to a thriving branch of industry.<br />

Exciting results reported in a series of works carried out by the Stoddart group<br />

on so-called molecular machines show that there is a long path from a concept<br />

to a marketable device [28]. By using rotaxane systems like 9a [29] a family<br />

of versatile systems has been created which can be used as sensor, switches,<br />

‘molecular abacus’ or nanomotor. In addition to the spacer 9e, the axle consists<br />

of the �-electron acceptor groups 9c, 9d which, upon the imposed conditions,<br />

can selectively bind the �-electron-donating macrocycle 9e shuttling it between<br />

the position shown in the formula 9a and that around the 3,3�-dimethyl-4,4�bipydidinium<br />

unit 9c. These studies, combining sophisticated syntheses with<br />

physicochemical methods, certainly provide an example of an important direction<br />

for nanotechnology research in the next few years. However, it is questionable<br />

whether they will bring marketable results within this time. This does not mean<br />

that such studies are not worth pursuing. Impractical studies elucidating the wave<br />

or corpuscular character of light and matter have been carried out since the famous<br />

dispute between Newton and Huygens for almost 300 years and, as the diffraction<br />

experiments on fullerene C 60 10 (this group of molecules is discussed in detail in<br />

Chapter 5) from the Zeilinger group show [30], this topic is still vivid. Of course,<br />

choosing objects for a project out of more than 27 500 000 known molecules is<br />

a hard task, but its expeditious applicability should not be decisive. Most basic


1.2 <strong>Hydrocarbons</strong> with Unusual Spatial Structure: the Need to Finance Basic Research<br />

research is driven by curiosity not practicality but somehow in the long run it pays<br />

off. No-one thought about chemical applications when the foundations of a new<br />

branch of mathematics – topology – were formulated in the 1820s, long before<br />

such abstract concepts as links (catenanes) like olympiadane 11, Möbius strip,<br />

and knots, like double knot 12, were shown to be of use in chemistry [31] (some<br />

hydrocarbons with distinct topological properties are presented in Chapter 9).<br />

7


8 1 Introduction<br />

Moreover, the discovery in the 1990s that circular DNAs in the living organisms<br />

[32–34] form links, knots and other systems with nontrivial topological properties,<br />

will have consequences which cannot be anticipated today. We begin to understand<br />

the mechanism of their formation but the role they play in nature is still unclear.<br />

Similarly, Einstein’s work on the photoelectric effect published in 1905, for which<br />

he was awarded the Nobel prize, certainly did not seem to have any practical significance<br />

at that time. Nor were cosmic studies carried out to develop kevlar and<br />

teflon! These materials were spin-offs of the space journeys.<br />

For examples closer to chemistry let us look at the fullerene 10 applications,<br />

much praised in the 1990s [35]. The Krätschmer method [36] which produces a<br />

significant amount of the substances spurred numerous proposals for their application.<br />

They were thought to exhibit superconductivity, serve as a drug carrier,<br />

its derivative C 60F 60 was anticipated to be an ideal lubricant, etc [35].<br />

None of these promises was fulfilled. Superconductivity of fullerene derivatives<br />

is exhibited only at very low temperatures [37]; C 60F 60 was synthesized but<br />

it turned out to decompose in air with the HF formation [38]; and, in spite of<br />

promising reports, to our best knowledge no drug involving fullerene is on<br />

market. One of few fullerene applications today consists in their use as AFM tips<br />

(http://www.foresight.org/Updates/Update27/Update27.3.html). Few other are<br />

presented in Section 5.0 while those discussed in Section 5.7 still await marketable<br />

applications. This does not mean that fullerenes should not be intensively studied<br />

and that they will not finally be of practical use. At present these elegant-looking,<br />

highly symmetrical molecules are exciting and worth studying simply because of<br />

their unusual properties. They (1) form the nonplanar system of conjugated bonds;<br />

(2) have a hollow space inside that can accommodate other, smaller molecules,<br />

ions or even an elementary particle [39]; (3) with cations inside they form unusual<br />

salts since the fullerene cage assumes the negative charge, thus the salt can be<br />

dissociated only by its destruction; and, last but not least, (4) their formula merely<br />

look beautiful or, in other words, they are aesthetically appealing.<br />

The last point, that is the beauty of molecular formulae as the driving force<br />

for studying a molecule, was strongly denied by Jansen and Schön in their essay<br />

provocatively entitled ‘Design in chemical synthesis – an illusion?’ [40]. Their<br />

argument, which deserves much longer comment, contraposes purely aesthetical<br />

Gropius’ teapot designs (apparently not constrained by the properties of materials<br />

from which the objects were to be made) to the design of molecules whose structure<br />

is unequivocally defined by the energy hypersurface. No objection: molecular<br />

structure must obey the basic rules of chemistry and physics. However, within<br />

these limits there is plenty of room for designing molecules with predefined<br />

desirable properties. In addition to the complicated and not always successful field<br />

of drug design, hydrocarbons with unusual spatial structure present numerous<br />

examples of molecular design which was not usually aimed at marketability.<br />

For instance, let us look at cubane 1 (discussed in Sections 2.3 and 2.4) synthesized<br />

almost 50 years ago. The molecule was obtained by Eaton [41] not because<br />

of its immediate applicability. Its synthesis presented a considerable challenge<br />

and 1 was aesthetically pleasing (the aspect denied by Jansen and Schön [40]). The


1.2 <strong>Hydrocarbons</strong> with Unusual Spatial Structure: the Need to Finance Basic Research<br />

molecule turned out to have untypical properties, due to its nonstandard structure,<br />

e.g. 1 exhibits unusual rearrangement reactions such as the rearrangement of<br />

cubane to cuneane 13. In addition, NMR spectra of cubane allow one to explore<br />

the Karplus dependence of the 3 J coupling constants [42]. And as for the cubane<br />

applications not looked for by Eaton? For more than 30 years secret studies were<br />

carried out by the American Army on nitro-derivatives of 1 because the high<br />

energy content of the cubane core magnified by the nitro-substituents suggested<br />

that these materials might have extraordinary explosive properties. (The synthesis<br />

of octanitrocubane 14 was eventually reported a few years ago [43].) There were<br />

also attempts to use cubane derivatives as therapeutic agents [44, 45].<br />

However, the emerging applications of some hydrocarbons with unusual spatial<br />

structure should not deceive us. The main goal of studying them is not their<br />

marketing but to deepen our understanding of the chemical bond. Interestingly,<br />

until recently this very fruitful concept, on which all chemistry is based, was not<br />

anchored in quantum chemistry. We could carry out calculations on the molecule<br />

as a whole but, without using artificial approximate constructs, were unable to<br />

analyze properties of specific bonds within it. In particular, studying molecules<br />

with bonds which are very different from the standard is indispensable to understand<br />

the limits of the very concept of the chemical bond. The question as to what<br />

extent a chemical bond can be distorted without breaking is thought provoking.<br />

Moreover, in certain cases even the mere existence of a bond between two carbon<br />

atoms has been questioned. This was the case encountered in [1.1.1]propellane 2<br />

(discussed in some detail in Sections 1.3 and 2.1) [46]. The synthesis of this exciting<br />

molecule, preceded by the calculations supporting its feasibility and predicting<br />

the propellane properties, serves as a fascinating example of a mutually fruitful<br />

interaction of theoretical and experimental studies [47]. 2 represents one of the<br />

most amazing examples from the point of view of organic stereochemistry since,<br />

contrary to van’t Hoff [48] and Le Bel [49] hypothesis, all four substituents on its<br />

bridgehead atoms lie in one hemisphere. Such atoms bearing the name ‘inverted<br />

carbons’, are also present in other small-ring propellanes (discussed in detail in<br />

Section 2.1) such as a derivative of [4.1.1]propellane 15 and that of [1.1.1]propellane<br />

16 [50]. The discussion of the existence of the bridgehead–bridgehead bond<br />

in small-ring propellanes is remarkable [51]. The distance between the atoms in<br />

these molecules is about 1.6 Å [52], which is significantly longer than the typical<br />

C–C bond of 1.54 Å (some authors [50] consider this difference small but the<br />

energy required for such bond lengthening is considerable). However, 1.6 Å or<br />

even longer bonds have been encountered in several hydrocarbons [53]. In spite<br />

9


10 1 Introduction<br />

of the reliable bond length of the former bond in small ring propellanes, the differential<br />

electron density maps for 15 [54] and 16 [50] measured in X-ray studies<br />

have not shown any build-up of the differential electron density between the<br />

bridgehead propellane atoms which should accompany the bond between them.<br />

The former finding and the possibility of a biradical structure for 14 without the<br />

central bond triggered a discussion on the reliability of the maps as the criterion of<br />

the bonding. On the one hand, the formulation of the limitations of this criterion<br />

resulted, stating that the lack of the build-up of the differential electron density is<br />

an artifact of the promolecule density distribution not reflecting the relative properties<br />

of the charge distributions [47]. The quantum calculations for 2 carried by<br />

Wiberg, Bader and Lau [47] showed that there is the bond-critical point along the<br />

line connecting bridgehead atoms in this molecule, thus proving the existence of<br />

the bond between the atoms. These calculations also revealed that the exceptional<br />

stability of this molecule is not due to the typical two-center integrals describing<br />

chemical bonds but is the result of the operation of the three-center integrals. In<br />

addition, other criteria for the existence of the central bond in 2 modeling [1.1.1]<br />

propellane have been checked [50].<br />

Helvetane 17 and israelane 18 appeared as a joke in a 1st April issue of Nouveau<br />

Journal de Chimie [55]. These highly strained hypothetical molecules belong to<br />

a very interesting class of (CH) 2n cage compounds to which cubane 1, dodecahedrane<br />

C 20 H 20 19 and hypothetical perhydrogenated fullerene C 60 H 60 [56]<br />

(discussed in Sections 2.3 and 2.4) belong. These molecules were shown to be<br />

of much higher energies than other members of the C 20 H 20 family [57] which<br />

should be much more easy to synthesize. Nevertheless, they have been calculated<br />

by several theoreticians who pointed out than removing symmetry constraints<br />

would significantly lower the energy of 17 and 18. Then such molecules being<br />

members of a large group of isomers without interesting properties would seem<br />

to be of no specific interest.


1.2 <strong>Hydrocarbons</strong> with Unusual Spatial Structure: the Need to Finance Basic Research<br />

It should be stressed that a molecule dismissed as purely hypothetical today can<br />

be a plausible synthetic target tomorrow. Herzberg, later awarded Nobel Prize,<br />

stated in his seminal ‘Infrared and Raman Spectra of Polyatomic Molecules’ in<br />

1945 that it is not likely that molecules of I h symmetry will ever be found [58].<br />

It took several years of hard work for the Paquette group to synthesize the first<br />

molecule of such a high symmetry, aforementioned dodecahedrane 19, about<br />

40 years later [59]. (In our opinion this synthesis deserves the name molecular<br />

design vigorously discredited by Jansen and Schön [40].) Today, the best known<br />

such molecules are fullerene 10 (discussed in Chapter 10), parent fullerene C 20<br />

20 [60] and perfluorinated fullerane C 60F 60 (which, as discussed in Chapter 10,<br />

similar to other short-lived species could be stabilized in molecular flasks) as well<br />

as the most symmetrical isomers of their higher homologs, like C 240, C 540, C 960,<br />

etc and some nested fullerenes formed by carbon cage compounds [61] belonging<br />

to the latter group.<br />

A kind of laborious play, that seems not to promise serious consequences<br />

but bears all the attributes of a standard synthetic work, has been reported by<br />

Chanteau and Tour [62]. They described the syntheses of nanoputanes, like 21,<br />

the anthropomorphic molecules named after the Jonathan Swift lilliputanes. With<br />

meticulously described syntheses, the authors showed not only the way to obtain<br />

the nanokid 21 but also got a ‘dancing’ nanoputanes layer on a surface 22.<br />

11


12 1 Introduction<br />

To summarize, the choice of a molecule for studies is not simple. Most standard<br />

systems are trivial and unworthy of serious consideration in basic research. Other<br />

untypical molecules can also have their pitfalls. Keeping in mind that what is impossible<br />

to synthesize today can be realizable tomorrow, one should nevertheless<br />

exhibit caution when choosing an object to study which should be of a serious<br />

scientific interest. However, it should be stressed once more that immediate applications<br />

should not be the reason for financing basic research.<br />

1.3<br />

Computations on <strong>Strained</strong> <strong>Hydrocarbons</strong><br />

Andrey A. Fokin and Peter R. Schreiner<br />

Despite some potential applications as high-energy materials and specialty<br />

polymers, highly strained compounds mostly play a conceptual and educational<br />

role. Over 10 000 chemical papers contain the key words ‘strained hydrocarbon’<br />

and more than 18 000 ACS papers alone the terms ‘strain energy.’ Despite the<br />

fact that our rationalistic and thrifty ages leave lesser space for exotic molecules,<br />

the aesthetic beauty of the cages such as cubane 1, tetrahedrane 23, octahedrane<br />

24 or dodecahedrane 25 still fascinate organic chemists and represent the artistry<br />

of organic synthesis. Nature also uses highly strained compounds: The recent<br />

discovery of ladderanes 26 and 27 as membrane lipids of certain anaerobic bacteria<br />

[63] and natural antifungal oligocyclopropane antibiotic 28 [64] underline the<br />

importance of such structures (Scheme 1.1).<br />

It is generally considered that highly strained compounds are difficult and<br />

expensive to make and, sometimes, also to store. However, once a challenging<br />

molecule has been prepared, the development of a simpler way for its synthesis<br />

is impending. The most recent example is highly strained octahedrane 24, which<br />

was first prepared in 1993 by an expensive and elaborate procedure [65, 66]. Now<br />

some octahedrane derivatives such as 30 can be prepared by one-step photochemical<br />

dimerization of readily available aromatic cyclophane 29 [67]. A simple<br />

preparation of octacyclopropylcubane 32 by an effective two-step condensation of<br />

four dicyclopropylacetylenes 31 is another remarkable example (Scheme 1.2) [68].<br />

The stability of strained compounds is not necessarily a concern: cubane and its<br />

derivatives are stable even at high temperatures because the strain is uniformly<br />

distributed throughout the molecule and orbital symmetry forbids the cleavage<br />

of two C–C bonds at the same time. Some unstable and highly strained hydro-


1.3 Computations on <strong>Strained</strong> <strong>Hydrocarbons</strong><br />

Scheme 1.1 Highly strained hydrocarbons and some natural compounds containing strained<br />

moieties.<br />

Scheme 1.2 Highly strained derivatives of octahedrane and cubane prepared recently through<br />

short and simple procedures.<br />

carbons have been successfully encapsulated and stored at room temperature as<br />

guest molecules in hemicarceplexes [69].<br />

The chemistry of strained organic molecules probably began with the realization<br />

of cyclopropane derivatives by Perkin [70] that was almost immediately<br />

followed by the development of strain theory [71]. It was soon recognized that<br />

Baeyer’s angular strain is the main contributor to the potential energies of organic<br />

molecules. The quantitative description of strain was first proposed in the mid<br />

13


14 1 Introduction<br />

1940s by Hill [72] and was developed further by Westheimer into molecular<br />

mechanics [73]. As with many other useful chemistry concepts like conjugation,<br />

aromaticity, chemical bonding, etc., strain itself is not defined exactly, but can be<br />

expressed well quantitatively by ‘strain energy,’ which is, however, not measurable<br />

experimentally. Its value is calculated as the difference between the experimental<br />

enthalpy of formation of the molecule of interest and that of a hypothetically strainfree<br />

structure. The enthalpies of formation of strain-free reference compounds<br />

are calculated through group additivity schemes [74] based on the ‘averaged’ contributions<br />

of groups (CH 3, CH 2, CH, etc.) from the straight-chain hydrocarbons.<br />

The group’s contributions are derived from thermochemical measurements for<br />

which, however, equilibrium conformer distributions are difficult to take into<br />

account. Group equivalent schemes also require experimental thermochemical<br />

data on the molecule of interest but these are equally error-prone. The most recent<br />

example is the heat of formation of cubane that was reinterpreted based on the<br />

corrected value of its sublimation enthalpy [75]. Unluckily, cubane has already<br />

been used for the parameterization of some molecular mechanical methods [76]<br />

that now require re-parameterization. New additivity schemes [77, 78] demonstrate<br />

excellent accuracy, but only for moderately strained hydrocarbons.<br />

The computations of �H f ° through atomization energies and bond/group separation<br />

reactions are more trustworthy; atomization energies are more useful for<br />

computation of the enthalpies of formation of small molecules [79–81]. Bond separation<br />

(isodesmic) equations proposed by Pople [82] for which the bonds between<br />

the non-hydrogen atoms are separated into strain-free reference molecules, give<br />

the strain energy directly (cf. Equation 1.1 for the evaluation of the strain energy<br />

of cubane). Alternatively, homodesmotic [83] equations (such as Equation 1.2),<br />

that contain an equal number of groups and bonds on both sides of the same<br />

type, largely cancel systematic computational errors. These two approaches lead<br />

to different strain energies (Scheme 1.3).<br />

The choice of strain-free reference compounds is problematic. The generally<br />

accepted strain energy of cubane (164.8 kcal mol –1 , without the newest correction<br />

for its enthalpy of sublimation) was computed through homodesmotic Equation<br />

1.2. However, the strain energy evaluation in Equation 1.2 is not properly<br />

balanced because the eight isobutane molecules are stabilized by twelve additional<br />

1,3-interactions (protobranching) [84] relative to cubane. Thus, isoalkanes<br />

have ‘negative strain’ relative to n-alkanes, and cannot be used as references for<br />

strain energy evaluations: using branched alkanes artificially increases the strain<br />

Scheme 1.3 Isodesmic (1.1) and homodesmotic (1.2) equations to determine the strain energy<br />

(E st ) of cubane.


1.3 Computations on <strong>Strained</strong> <strong>Hydrocarbons</strong><br />

of the molecules of interest. On the other hand, Equation 1.1 is properly balanced<br />

because it is based only on strain-free molecules – methane and ethane. The<br />

strain energy of cubane calculated via Equation 1.1 is 102.9 kcal mol –1 [85] and<br />

this value should be used. Equally, other strained hydrocarbons appear to be less<br />

strained than originally assumed. Concerning the validity of isodesmic Equation<br />

1.1, one can argue that ethane is also not a strain-free molecule because of van<br />

der Waals’ contacts while methane is destabilized by Pauli repulsions. As these<br />

effects are present in all organic molecules these two hydrocarbons are still the<br />

best candidates as strain-free reference hydrocarbons; there is no conformational<br />

problem either.<br />

Computations are the only way to evaluate the strain energies of molecules for<br />

which the experimental thermochemistry is not available. Chemically accurate<br />

computations for small molecules nowadays are inexpensive, fast, and they can<br />

be used with ease. However, there are many sources of systematic as well as nonsystematic<br />

errors in computational chemistry modeling. Some of them are of<br />

general character, others are typical only for strain energy evaluations. For instance,<br />

due to the large number of reference molecules used in the above equations, the<br />

errors in zero-point vibrational energy (ZPVE) corrections, which are usually<br />

derived from a crude harmonic approximation model, do not effectively cancel.<br />

The choice of a proper computational method especially for ‘unusual’ strained<br />

molecules is critical. Computational chemistry estimates the contributions of<br />

angular strain well, even at the level of molecular mechanics. Another source<br />

of strain, nonbonding attractions/repulsions, is more challenging to compute<br />

correctly as only very expensive state-of-the-art computational methods are able<br />

to describe them accurately. Popular density functional theory (DFT) methods<br />

offer numerous functionals with different empiric exchange-correlation terms.<br />

While some of them are especially designed to describe certain types of interactions<br />

properly, virtually all of them systematically underestimate or completely<br />

neglect weak interactions [86]. DFT methods, such as the most popular B3LYP<br />

functional, give rise to various unsystematic errors and, worse, these increase<br />

dramatically with the size of the molecules [87, 88]. DFT methods may also exhibit<br />

some artifacts like electron self-exchange, which affects the electron energies considerably.<br />

Medium-range electron correlation, which contributes to the energies<br />

of saturated systems significantly, is poorly described both by local and hybrid<br />

functionals [89]. All of the above problems lead to unacceptable DFT errors for<br />

unstrained [86, 89] and, especially, strained [88, 90] molecules with more than ten<br />

heavy (= non-hydrogen) atoms – the ones for which DFT methods currently are used<br />

most often [91]. The use of new DFT formulations [92, 93] or a posteriori corrections<br />

(MP2 or coupled cluster) together with accurate thermochemical methods<br />

(Gn [94], Wn [95], or CBS [96]) significantly improves the quality of strain energy<br />

evaluations [79]. Nevertheless, most of the computational errors in strain energy<br />

evaluations are smaller than the discrepancies resulting from the arbitrary choice<br />

of strain-free reference states.<br />

Chemists targeting the preparation of a potentially strained compound are faced<br />

with the problem of predicting its stability and reactivity. The ‘strain energy per<br />

15


16 1 Introduction<br />

heavy atom’ or ‘per bond’ is a good starting point but it reveals little about the<br />

kinetic stability of a molecule. If an appreciable reaction pathway for lowering the<br />

potential energy does not exist, the molecule may be stable despite being highly<br />

strained. The extraordinary thermal stabilities of prismane and cubane are associated<br />

with the feature that breaking just one C–C bond causes only minimal<br />

changes in the remaining part of the molecular structure. Stabilization of tetrahedrane<br />

with bulky groups, which hinder rearrangements due to the ‘corset effect,’<br />

is another example [97, 98].<br />

Computational chemistry can help predict the behavior of strained compounds,<br />

and there are many inspiring examples (for selection see Scheme 1.4). The unusual<br />

stability of highly strained [1.1.1]propellane 2 [99–101] (discussed in Section 2.1)<br />

protonated pyramidane derivatives 33 [102, 103], and the T d -1,2-dehydro-5,7-adamantanediyl<br />

dication 34 [104] first predicted computationally, initiated successful<br />

attempts to prepare these highly strained systems.<br />

Scheme 1.4 Some highly strained molecules, whose anomalous stability was predicted<br />

computationally before preparation.<br />

The prediction of the thermal stabilities of strained compounds is a routine,<br />

albeit elaborate, procedure now and involves computations on the barriers<br />

of the crucial bond breaking pathway (see, for instance, a recent study on the<br />

kinetic stability of tetrahedrane) [98]. As the kinetic stability is a relative value,<br />

i.e. it depends on the reaction partner and the conditions, the barriers for the<br />

attack of radicals on strained compounds not only allow one to analyze their<br />

potential stability, but also to choose a proper reagent for their derivatization. For<br />

instance, cubane [105] and octahedrane [66] were found to be highly sensitive to<br />

the nature of the attacking radicals and they follow either C–C addition or C–H<br />

substitution paths. Computations on the reactions of strained compounds with<br />

electrophiles are more difficult, because the carbocationic species that form after<br />

primary electrophilic attack are largely prone to rearrangements to release the<br />

strain. Even reproducing experimental proton affinities is difficult, especially for<br />

a system as strained as cubane [106]. Cyclopropane is an exception because the<br />

edge-protonated form is a minimum and the downhill ring-opening path has a<br />

relatively high barrier [107].<br />

Highly strained compounds quite often represent intriguing bonding situations,<br />

which modern computational methods are able to describe well. They offer not<br />

only accurate energies and geometries of strained compounds, but also provide<br />

information about electron density distributions, molecular orbitals, bond critical<br />

points and so forth [108]. One of the examples is the unusual bonding situation<br />

between inverted carbons of [1.1.1]propellane 2. Twenty years ago theory predicted


1.3 Computations on <strong>Strained</strong> <strong>Hydrocarbons</strong><br />

[109] that there is a bond critical point between the central carbon atoms of 2<br />

despite the fact that the electron density does not accumulate in this region.<br />

Recent synchrotron experiments on derivatives of 2 confirmed that this remarkable<br />

computational description indeed is correct [110]. The examples of m-benzyne<br />

35 and m-dehydrocubane 36 demonstrate the borderline between proper C–C<br />

bonding and open-shell singlet biradical states (Scheme 1.5). The latter is favored<br />

for m-benzyne if dynamic electron correlation is included exhaustively: the fundamental<br />

frequencies of the m-benzyne singlet biradical computed at CCSD(T)<br />

[111] perfectly agree with the experimental IR spectra [112]. m-Dehydrocubane<br />

forms a singlet state and is predicted to exhibit an extremely long C–C bond<br />

(1.844 Å from REKS-B3LYP/6-31G* data) [113]. Unusually short bonds were<br />

found for the dimers of strained compounds. In 1989 the shortest single C–C<br />

bond (1.438 Å) was computed [114] for bis-tetrahedrane 37; recently this value was<br />

confirmed experimentally [115]. The properties of highly pyramidalized alkenes<br />

are difficult to study experimentally as only some matrix IR-spectra are available<br />

[116] and computational results are difficult to validate. A real breakthrough in<br />

this area was achieved recently when it was found that the computed proton affinities<br />

and heats of hydrogenation of 1,5-dehydroquadricyclane 38 agreed well<br />

with experiment [117].<br />

Scheme 1.5 Selected strained molecules that represent unusual C–C bonding situations.<br />

Most importantly, computational chemistry can not only predict the properties<br />

of molecules, but also help to discover new classes of strained compounds that<br />

may challenge experimentalists (Scheme 1.6). Molecules with planar tetracoordinated<br />

carbon: fenestranes 39 [118], tricyclo[2.1.0.0 1,3 ]pentane 40 [119, 120], and<br />

tetracyclo[3.1.0.0 1,3 .0 3,5 ]hexane 41 [121]; with inverted geometries around the<br />

carbon atoms: pyramidane 42 [122] and bowlane 43 [123]; as well as with highly<br />

twisted double bonds: orthogonene 44 [124] and tetra-t-butyl ethylene 45 [125],<br />

Scheme 1.6 Highly strained molecules computationally predicted to be isolable.<br />

17


18 1 Introduction<br />

were computationally predicted to be isolable and thus represent challenging yet<br />

realistic synthetic targets.<br />

The development of the chemistry of strained compounds has been closely<br />

connected to the progress of computational chemistry for the last three decades.<br />

We have finally reached a situation where geometries, electronic and thermodynamic<br />

properties as well as the reactivity of highly strained and ‘unusual’ small<br />

molecules may be estimated computationally with chemical accuracy.<br />

1.4<br />

Gallery of Molecules That Could Have Been Included in This Book<br />

Helena Dodziuk<br />

1.4.1<br />

Introductory Remarks<br />

The aim of this monograph is to present the richness of the domain of hydrocarbons<br />

with unusual spatial structure, not only in chemistry but also in their<br />

physicochemical properties, and not only in experimental studies but also in model<br />

calculations that play an increasingly important role in this domain. Clearly, such<br />

a broad scope, to be understood as a protest against the narrow specialization<br />

from which we all suffer, could not be fully covered in this limited volume. Thus,<br />

for various reasons, not all molecules deserving incorporation in this book could<br />

even be mentioned. To counteract this situation, in this chapter several fascinating<br />

molecules that have not been presented in other chapters will simply be listed<br />

with short notes showing why they are of interest. This is of particular importance<br />

since at least some of them merit further, more detailed studies.<br />

1.4.2<br />

Saturated <strong>Hydrocarbons</strong><br />

Of the family of bridged spiropentanes 46, the known [4.1.0.0 1,6 ]tricycloheptane<br />

(n = 2) 46a is stable and exhibits a considerable widening of the C2C1C7 angle up<br />

to about 160° [126, 127]. There is NMR evidence of [2.1.0.0 1,3 ]tricyclopentane 46b<br />

for which ab initio calculations yielded a pyramidal configuration on the central<br />

carbon atom [128, 129]. On the basis of ab initio quantum chemical calculations,<br />

QC, exciting tricyclo[3.1.0 1,3 ]hexane 46c has been found to exhibit almost linear<br />

arrangement of the formally Csp 3 –Csp 3 bonds with the


1.4 Gallery of Molecules That Could Have Been Included in This Book<br />

178° [130]. Wiberg and Snoonian [131] reported the synthesis of the derivative 47<br />

observed at 10 K having the highly reactive ketene group. Thus, until now the<br />

highly unusual spatial structure of 46c could not be proven.<br />

One of the largest member of the trangulane family is branched C 2v-[15]triangulene<br />

48. A shortening of the central C–C bond in this molecule has been interpreted<br />

in terms of a considerable change in hybridization of the two central spirocarbon<br />

atoms due to severe steric strain [132]. Smaller, but also overcrowded, triangulanes<br />

also studied by the de Meijere group exhibited some unusual reactivity [133].<br />

As shown by boat–twist conformation of the central C 6 ring in trispirocyclopropanated<br />

cyclohexane 49 [134–136] and by the boat conformation of these rings<br />

in tetraasterane 50 [137, 138] discussed in Section 2.3, the cyclohexane ring does<br />

not necessarily have to assume the chair conformation.<br />

Recently synthesized octacyclopropylcubane 51 is not very stable: it has a half-life<br />

of 3 h at 250 °C and has ‘tremendous overall strain’ of 390 kcal mol –1 [139]. In the<br />

crystal it exhibits quite rare C 4h symmetry. The average length of C–C bonds in<br />

the cubane core of 158.3 pm have been found to be slightly but distinctly longer<br />

than that in cubane (156.5 pm in the gas phase and 155.1 in the crystal).<br />

Three examples of interesting stereochemical and/or structural phenomena<br />

will be given at the end of this subchapter.<br />

19


20 1 Introduction<br />

Figure 1.2 Bicyclic hydrocarbons.<br />

A rare but interesting observation, i.e. in, out isomerism in bicyclic hydrocarbons<br />

(Figure 1.2) [140–142], has also been noticed in several natural products [143]. For<br />

larger values of k, l and m the molecules exhibit dynamic equilibrium.<br />

Steric strain may also cause ‘squeezing’ a molecule leading to a very close<br />

distance between the nonbonded atoms. Since X-ray analysis is not the most<br />

reliable tool for the determination of H atoms’ position, indirect arguments are<br />

sometimes used for their estimation as was done for 52 [144].<br />

As found for cubane 1 [145] and C 60 53, highly symmetrical structures may<br />

exhibit interesting dynamic behavior in the solid state [146, 147]. In the latter case,<br />

it also leads to the structural diversity of host–guest and intercalation complexes<br />

of the fullerene as studied by X-ray technique [148].<br />

It took more than 20 years to synthesize dodecahedron 54 [149] which, due to<br />

its strain, exhibits quite unusual rearrangement reactions. Remarkably, both 53<br />

and 54 are of I h symmetry, that is every carbon atom (and hydrogen, respectively)<br />

in these molecules is identical with the other.


1.4.3<br />

Distorted Double Bonds<br />

1.4 Gallery of Molecules That Could Have Been Included in This Book<br />

A detailed discussion of several routes that were expected to lead to highly strained<br />

tetrakis-t-butylethene 55 but were unsuccessful is given in detail in Ref. [150].<br />

Hypothetical bicyclo[1.1.0]-1(4)-pentene 56a and bicyclo[1.1.0]-1(3)-butene 56b<br />

remain unknown but according to ab initio calculations such molecules should<br />

have considerably pyramidalized formally C sp2 carbon atoms [151, 152].<br />

Pyramidalized carbon atoms should be exhibited in known bridged bicyclobutane<br />

57 (n = 3) [153–156] and 58 which has probably been observed [157].<br />

The smallest synthesized [m][n]betweenanene 59 has m = n = 8 [158, 159]. To<br />

the best of our knowledge, no X-ray structure determination exists but simple<br />

MM modeling indicates significant distortions from standard geometry [160].<br />

The larger (m = 22, n = 10) not highly strained betweenanenes have been expected<br />

to exhibit interesting dynamic effects involving ‘a jumping of the longer chain’<br />

around [161].<br />

Tricyclo[4.2.2.2 2,5 ]dodeca-1,5-diene 60 [162] and its tetraaryl-substituted derivative<br />

61 [163, 164], both with strongly pyramidalized C sp2 carbon atoms, are<br />

known, while diene 62, also with very close distance between double bonds, is still<br />

unknown [165]. Noteworthy, the aromatic rings in 61 are planar.<br />

21


22 1 Introduction<br />

The name Dewar benzene of 63 is thought to have no sound basis [166]. The<br />

system is stable in the form of its tri-t-butyl [167, 168] or hexamethyl [169, 170]<br />

derivatives. Although the highly reactive 63 was obtained more than 40 years<br />

ago and attracted the attention of several theoreticians [171], the reactivity of this<br />

molecule and/or its derivatives is still the subject of studies today [172].<br />

Out of two possible 64a and 64b diastereomers of [2.2]cyclooctatetraenophane,<br />

which are present as a (d,l) mixture, the former was synthesized and found to<br />

isomerize to the latter [173].<br />

1.4.4<br />

Benzene Rings with Nontypical Spatial Structures<br />

Typical structure of aromatic systems consists in the planarity of the aromatic ring<br />

and its substituents and in close to 120° value of all bond angles. Steric hindrance<br />

can force another spatial structure. To the best of our knowledge, highly strained<br />

hexa-t-butylbenzene 65 (X = C) is not known. However, a less strained derivative<br />

(due to longer C–Si than C–C bonds), hexakis(trimethylsilyl) derivative 65, exhibits<br />

an unusual distorted chair conformation [174]. Another, less symmetrical type of<br />

nonplanar distortion of the aromatic ring is provided by 1,2,3-tri-t-butylnaphthalene<br />

66 [175]. Considerable strain in the latter molecule allowed for the freezing of<br />

internal rotations of the methyl groups in 1- and 2-positions at 193 K. Interestingly,<br />

in disagreement with molecular mechanics [176], modeling the barrier for<br />

the rotation of the groups at the 3-position was the smallest. It is also noteworthy<br />

that due to its large size, 66 did not form the inclusion complex with �-cyclodextrin<br />

but rather ‘sat’ on top of the macrocyclic sugar.


1.4 Gallery of Molecules That Could Have Been Included in This Book<br />

Twisted acenes are schematically presented in formula 67 [177]. Additional<br />

phenyl groups as in 68 stabilize the molecule and have allowed Pascal to achieve<br />

a formidable twist of 144° between the planes of terminal aromatic rings [178].<br />

Interestingly, 68 has been resolved into enantiomers. Such highly twisted aromatic<br />

systems can find an application as porous solids. They are also expected to have<br />

chiroptical properties and have been incorporated into light emitting diodes (www.<br />

cnsi.ucla.edu/arr/paper?paper_id=193298). According to DFT calculations, 68 is<br />

a disjointed radical exhibiting exciting electronic structure [179].<br />

Latos–Grazynski group reported the synthesis of di-p-benzhexaphyrin that is in<br />

dynamic equilibrium of two forms: ‘standard’ 69a and 69b representing a topologically<br />

nontrivial Möbius strip [180]. (Other topological nontrivial molecules are<br />

presented in Chapter 9.) In the solid state, only the latter was found to exist.<br />

Cyclophanes [181] (covered in Section 4.2) have been studied mainly because<br />

of their non-standard structure and a strong �–� interaction between close-lying<br />

aromatic rings [182] manifesting itself in UV/Vis [183] and NMR spectra [184].<br />

23


24 1 Introduction<br />

Syntheses of impressive layered para-cyclophanes called cochins having up<br />

to six aromatic rings as in 70 and 71 were reported by Otsubo and coworkers<br />

[185–187] while those of three isomers of four-layered [2.2]metacyclophanes 72–74<br />

were published by Umemoto [188, 189]. In analogy with the smaller [2.2]metacyclophane<br />

75 discussed in detail in Section 4.2, protons of C–H bonds situated<br />

between two bridges exhibit very unusual values of chemical shifts since they lie<br />

above (or below) the plane of the neighboring aromatic ring.<br />

In spite of its high strain, superphane 76 [190, 191] is relatively stable, even [2 6 ]<br />

(1,2,3,4,5,6)cyclophane-1-ene 77 with an additional double bond has been reported<br />

[191]. The benzene C sp2 carbon atoms in 76 all lie in the respective planes but its<br />

spatial structure is untypical since not all substituents on the aromatic rings lie<br />

in the plane of the rings [192]. 78 and 79 still await their syntheses [190].


1.4 Gallery of Molecules That Could Have Been Included in This Book<br />

Corannulene 80 has the shape of a bowl because it includes a five-membered<br />

ring, and is known to invert rapidly [193]. In addition to its nonstandard geometry<br />

and dynamic behavior, the molecule attracted a lot of interest since it has been considered<br />

as an important building block that should enable the organic chemistry<br />

synthesis of C 60 53. Corrannulene derivatives also exhibit interesting packing<br />

behavior in the solid state [193]. As discussed in detail in Kawase and Kurata<br />

review [194] not only bowl-shaped but also ball- and belt-shaped aromatic systems<br />

provide an exciting opportunity to explore the concave–convex �–� interactions<br />

by studying their complexation.<br />

Another type of revealing distortion in aromatic rings consists in differentiation<br />

of their bond lengths achieved by fusing cyclobutane or cyclopentane rings to them,<br />

resulting in the remarkable differences in the C–C bond lengths for 81 [195] and<br />

82 [196]. Such systems are indispensable for studying the limits of aromaticity.<br />

1.4.5<br />

Cumulenes<br />

Interestingly, the cumulenes’ structure was predicted by van’t Hoff [197] who stated<br />

that in cumulenes with an even number of double bonds the four substituents<br />

must be placed in two perpendicular planes while for the odd-numbered series<br />

the substituents must lie in one plane with the double bonds. The cumulenes<br />

discussed in Section 3.3 are distorted from such arrangements. No X-ray structure<br />

of bicyclic allene 83 [198] and triene 84 [199] have been published but, according<br />

to MM modeling, the planes of respective bonds are at angles different from<br />

zero and 180° [160]. Similarly, to the best of our knowledge no structural data for<br />

hexaene 85 [200] and pentaene 86 [201] have been published but the molecules,<br />

with t-butyl groups added to increase stability, definitely do not have the standard<br />

structure.<br />

25


26 1 Introduction<br />

1.4.6<br />

Acetylenes<br />

Permethylated [5]pericyclyne 87 (R = Me) and larger analogs are known [202,<br />

203]. Interestingly, in analogy with cyclopentane the central ring in 87 adopts the<br />

envelop conformation even in the solid state while model calculations indicate<br />

that in permethylated [6]pericyclyne the central ring [204] can adopt either the<br />

most stable chair conformation or boat or twist-boat ones. Aromaticity and the<br />

role of conjugation in 88 and other analogous carbocycles have been studied by<br />

Lepetit [205]. X-ray spectra of an octaphenyl derivative of 89 [206] reveal the planar<br />

structure of the bicyclic core with considerable bond angle distortions.<br />

Only a few of many exciting distorted hydrocarbons could be mentioned in this<br />

chapter. It should be stressed, however, that with a new domain of macrocyclic<br />

host molecules rapidly developing this area will expand further since not all<br />

large macrocycles are strain-free. For instance, 90 can host C 60 53 into its cavity<br />

assuming C 6v symmetry [194, 207].


References<br />

1 Berson, J. A. In Chemical Discovery and<br />

the Logicians’ Program; Wiley-VCH:<br />

Weinheim, 2003, p. 47.<br />

2 van’t Hoff, J. H. Nederl. Sci. Exactes Nat.<br />

1874, 445.<br />

3 Le Bel, J. A. Bull. Soc. Chim. Fr. 1874, 22,<br />

337.<br />

4 Sachse, H. Z. Phys. Chem. 1892, 10, 203.<br />

5 Sachse, H. Ber. 1890, 23, 1363.<br />

6 Pitzer, K. S. Annu. Rev. Phys. Chem.<br />

1987, 38, 1.<br />

7 Pitzer, K. S. Thermodynamics; McGraw-<br />

Hill: New York, 1995.<br />

8 Kemp, J. D.; Pitzer, K. S. J. Chem. Phys.<br />

1936, 4, 391.<br />

9 Hassel, O. Quart. Revs. 1953, 7, 221.<br />

10 Hassel, O. Tidsskr. Kjemi Bergvesen Met.<br />

1948, 3, 32.<br />

11 Barton, D. H. R. Experientia 1950, 6,<br />

316.<br />

12 Eaton, P. E.; Cole, T. W. J. Am. Chem.<br />

Soc. 1964, 86, 962, 3157.<br />

13 Wiberg, K. B.; Walker, F. H. J. Am.<br />

Chem. Soc. 1982, 104, 5239.<br />

14 Bayer, A. Ber. 1885, 18, 2269.<br />

15 Wiberg, K. B. Chem. Rev. 1989, 89, 975.<br />

16 Dodziuk, H.; Nowinski, K. Bull. Pol.<br />

Acad. Sci. Chem. 1987, 35, 195.<br />

17 Dodziuk, H. Bull. Pol. Acad. Sci. Chem.<br />

1990, 38, 11.<br />

18 de Meijere, A.; Lee, C. H.; Kuznetsov,<br />

M. A.; Gusev, D. V.; Kozhushkov, S. I.;<br />

References<br />

Fokin, A. A.; Schreiner, P. R. Chem. Eur.<br />

J., 2005, 11, 6175.<br />

19 Wiberg, K. B.; Bader, R. F. W.;<br />

Lau, C. D. H. J. Am. Chem. Soc. 1987,<br />

109, 985.<br />

20 Hoffmann, R.; Alder, R. W.; Wilcox, C. F.<br />

J. Am. Chem. Soc. 1979, 92, 4992.<br />

21 Radom, L.; Rasmussen, D. R. Angew.<br />

Chem. Int. Ed. 1999, 38, 2876.<br />

22 Zhu, L.; Chang, D. W.; Dai, L.; Hong, Y.<br />

Nanolett. 2007, 7, 3592.<br />

23 Dodziuk, H. Introduction to supramolecular<br />

Chemistry; Kluwer: Dordrecht, 2002.<br />

24 Tanner, M. E., Knobler, C. B., Cram, D. J.<br />

Angew. Chem., Int. Ed. 1991, 30, 1924.<br />

25 Dodziuk, H.; Ostrowski, M. Eur. J. Org.<br />

Chem. 2006, 5231.<br />

26 Reinitzer, F. Monatsh. Chem. 1888, 9,<br />

421.<br />

27 Nachr. Chem. Techn. Lab. 1966, 14, 29.<br />

28 Stoddart, J. F. Pure Appl. Chem. 2005, 77,<br />

1089.<br />

29 Ashton, P. R.; Ballardini, R.;<br />

Balzani, V.; Credi, A.; Dress, K. R.;<br />

Ishow, E.; Kleverlaan, C. J.; Kocian, O.;<br />

Preece, J. A.; Spencer, N.; Stoddart, J. F.;<br />

Venturi, M.; Wenger, S. Chem. Eur. J.<br />

2000, 6, 3558.<br />

30 Nairz, O.; Arndt, M.; Zeilinger, A.<br />

Am. J. Phys. 2003, 71, 319.<br />

31 Dodziuk, H.; Nowinski, K. Tetrahedron<br />

1998, 54, 2917.<br />

27


28 1 Introduction<br />

32 Shekhtman, E. M.; Wasserman, S. A.;<br />

Cozarelli, N.; Solomon, M. J. New J.<br />

Chem. 1993, 17, 757.<br />

33 Berger, J. M.; Gamblin, S. J.;<br />

Harrison, S. C.; Wang, J. C. Nature 1996,<br />

379, 225.<br />

34 Chen, J.; Rauch, C. A.; White, J. H.;<br />

Englung, P. T.; Cozarelli, N. R. Cell 1995,<br />

80, 61.<br />

35 Stoddart, J. F. Angew. Chem. Int. Ed.<br />

Engl. 1991, 30, 70.<br />

36 Krätschmer, W.; Fostiropoulos, K.;<br />

Huffmann, D. Chem. Phys. Lett. 1990,<br />

170, 167.<br />

37 Margadonna, S.; Prassides, K. J. Solid<br />

State Chem. 2002, 168, 639.<br />

38 Taylor, R.; Avent, A. G.; Dennis, T. J.;<br />

Hare, J. P.; Kroto, H. W.; Holloway, J. H.;<br />

Hope, E. G.; Langley, G. J. Nature 1992,<br />

355, 27.<br />

39 Prassides, K.; Dennis, T. J. S.;<br />

Christides, C.; Roduner, E.; Kroto, H. W.;<br />

Taylor, R.; Walton, D. M. R. J. Phys.<br />

Chem. 1992, 96, 10600.<br />

40 Jansen, M.; Schön, J. C. Angew. Chem.<br />

Int. Ed. Engl. 2006, 45, 3406.<br />

41 Eaton, P. E.; Cole, T. W. J. Am. Chem.<br />

Soc. 1964, 86, 962, 3157.<br />

42 Karplus, M. J. Am. Chem. Soc. 1963, 85,<br />

2870.<br />

43 Zhang, M. X.; Eaton, P. E.; Gilardi, R.<br />

Angew. Chem. Int. Ed. Engl. 2000, 39, 401.<br />

44 Mahkam, M.; Sanjani, N. S.;<br />

Entezami, A. A. J. Bioact. Compat. Pol.<br />

2000, 15, 396.<br />

45 Mahkam, M. J. Biomed. Mater. Res. B<br />

2005, 75B, 108.<br />

46 Wiberg, K. B.; Walker, F. H. J. Am.<br />

Chem. Soc. 1982, 104, 5239.<br />

47 Wiberg, K. B.; Bader, R. F. W.;<br />

Lau, C. D. H. J. Am. Chem. Soc. 1987,<br />

109, 985.<br />

48 van’t Hoff, J. H. Nederl. Sci. Exactes Nat.<br />

1874, 447.<br />

49 Le Bel, J. A. Bull. Soc. Chim. Fr. 1874, 22,<br />

337.<br />

50 Messerschmidt, M.; Scheins, S.;<br />

Grubert, L.; Patzel, M.; Szeimies, G.;<br />

Paulmann, C.; Luger, P. Angew. Chem.<br />

Int. Ed. Engl. 2005, 44, 3925.<br />

51 Dodziuk, H. In Modern Conformational<br />

Analysis. Elucidating Novel Exciting<br />

Molecular Structures; VCH Publishers,<br />

Inc.: New York, 1995, p. 159.<br />

52 Levin, M. D.; Kaszynski, P.; Michl, J.<br />

Chem. Rev. 2000, 100, 169.<br />

53 Kaup, G.; Boy, J. Angew. Chem. Int. Ed.<br />

Engl. 1997, 36, 48.<br />

54 Chakrabati, P.; Seiler, P.; Dunitz, J. D.;<br />

Schluter, A.-D.; Szeimies, G. J. Am.<br />

Chem. Soc. 1981, 103, 7378.<br />

55 Dinsburg G.; (Ginsburg, G.,) New J.<br />

Chem. 1982, 6, 175.<br />

56 Dodziuk, H.; Nowinski, K. Chem. Phys.<br />

Lett. 1995, 249, 406.<br />

57 Dodziuk, H. In Modern Conformational<br />

Analysis. Elucidating Novel Exciting<br />

Molecular Structures; VCH Publishers,<br />

Inc.: New York, 1995, p. 172.<br />

58 Herzberg, G. Infrared and Raman Spectra<br />

of Polyatomic Molecules; Lancaster Press:<br />

Lancaster, 1945.<br />

59 Ternansky, R. J.; Balogh, D. W.;<br />

Paquette, L. J. Am. Chem. Soc. 1982, 104,<br />

4503.<br />

60 Prinzbach, H.; Weller, A.;<br />

Landenberger, P.; Wahl, F.; Worth, J.;<br />

Scott, L. T.; Gelmont, M.; Olevano, D.;<br />

von Issendorff, B. Nature 2000,<br />

407(6800), 60.<br />

61 Iijima, S. J. Cryst. Growth 1980, 50, 675.<br />

62 Chanteau, S. H.; Tour, J. M. J. Org.<br />

Chem. 2003, 68, 8750.<br />

63 Sinninghe Damste, J. S.; Strous, M.;<br />

Rijpstra, W. I. C.; Hopmans, E. C.;<br />

Geenevasen, J. A. J.; van Duin, A. C. T.;<br />

van Niftrik, L. A.; Jetten, M. S. M. Nature<br />

2002, 419, 708–712.<br />

64 Yoshida, M.; Ezaki, M.; Hashimoto, M.;<br />

Yamashita, M.; Shigematsu, N.;<br />

Okuhara, M.; Kohsaka, M.; Horikoshi, K.<br />

J. Antibiot. 1990, 43, 748–754.<br />

65 Lee, C.-H.; Liang, S.; Haumann, T.;<br />

Boese, R.; de Meijere, A. Angew. Chem.<br />

1993, 105, 611–613.<br />

66 de Meijere, A.; Lee, C.-H.; Kuznetsov,<br />

M. A.; Gusev, D. V.; Kozhushkov, S. I.;<br />

Fokin, A. A.; Schreiner, P. R. Chem. Eur.<br />

J. 2005, 11, 6175–6184.<br />

67 Okamoto, H.; Satake, K.; Ishida, H.;<br />

Kimura, M. J. Am. Chem. Soc. 2006, 128,<br />

16508–16509.<br />

68 de Meijere, A.; Redlich, S.; Frank, D.;<br />

Magull, J.; Hofmeister, A.; Menzel, H.;<br />

König, B.; Svoboda, J. Angew. Chem. Int.<br />

Ed. 2007, 46, 4574–4576.<br />

69 Roach, P.; Warmuth, R. Angew. Chem.<br />

Int. Ed. 2003, 42, 3039–3042.


70 Perkin, W. H., j. Ber. 1884, 17, 323–329.<br />

71 Baeyer, A. Ber. 1885, 18, 2269–2281.<br />

72 Hill, T. L. J. Chem. Phys. 1946, 14, 465.<br />

73 Westheimer, F. H.; Mayer, J. E. J. Chem.<br />

Phys. 1946, 14, 733–738.<br />

74 Cohen, D.; Benson, S. W. Chem. Rev.<br />

1993, 93, 2419–2438.<br />

75 Bashir-Hashemi, A.; Chickos, J. S.;<br />

Hanshaw, W.; Zhao, H.; Farivar, B. S.;<br />

Liebman, J. F. Thermochimica Acta 2004,<br />

424, 91–97.<br />

76 Allinger, N. L.; Yuh, Y. H.; Lii, J.-H.<br />

J. Am. Chem. Soc. 1989, 111, 8551–8566.<br />

77 Gronert, S. J. Org. Chem. 2006, 71,<br />

1209–1219.<br />

78 Wodrich, M. D.; Schleyer, P. v. R.<br />

Org. Lett. 2006, 8, 2135–2138.<br />

79 Cheng, M.-F.; Li, W.-K. J. Phys. Chem.<br />

2003, 107, 5492–5498.<br />

80 Tasi, G.; Izsak, R.; Matisz, G.;<br />

Csaszar, A. G.; Kallay, M.; Ruscic, B.;<br />

Stanton, J. F. Chem. Phys. Chem. 2006, 7,<br />

1664–1667.<br />

81 Notario, R.; Castan, O.; Abboud, J.-L. M.;<br />

Gomperts, R.; Frutos, L. M.; Palmeiro, R.<br />

J. Org. Chem. 1999, 64, 9011–9014.<br />

82 Hehre, W. J.; Ditchfield, R.; Radom, L.;<br />

Pople, J. A. J. Am. Chem. Soc. 1970, 92,<br />

4796–4801.<br />

83 George, P.; Trachtman, M.; Bock, C. W.;<br />

Brett, A. M. Tetrahedron 1976, 32,<br />

317–323.<br />

84 Wodrich, M. D.; Corminboeuf, C.;<br />

Schleyer, P. v. R. Org. Lett. 2006, 8,<br />

3631–3634.<br />

85 Wodrich, M. D.; Wannere, C. S.;<br />

Mo, Y.; Jarowski, P. D.; Houk, K. N.;<br />

Schleyer, P. v. R. Chem. Eur. J. 2007, 13,<br />

7731–7744.<br />

86 Wodrich, M. D.; Corminboeuf, C.;<br />

Schleyer, P. v. R. Org. Lett. 2006, 8,<br />

3631–3634.<br />

87 Curtiss, L. A.; Raghavachari, K.;<br />

Redfern, P. C.; Pople, J. A. J. Chem. Phys.<br />

2000, 112, 7374–7383.<br />

88 Schreiner, P. R.; Fokin, A. A.;<br />

Pascal, R. A.; de Meijere, A. Org. Lett.<br />

2006, 8, 3635–3638.<br />

89 Grimme, S. Angew. Chem. Int. Ed. 2006,<br />

45, 4460–4464.<br />

90 Schreiner, P. R. Angew. Chem. Int. Ed.<br />

2007, 46, 4217–4219.<br />

91 Check, C. E.; Gilbert, T. M. J. Org. Chem.<br />

2005, 70, 9828–9834.<br />

References<br />

92 Zhao, Y.; Truhlar, D. G. Org. Lett. 2006,<br />

8, 5753–5755.<br />

93 Schwabe, T.; Grimme, S. Phys. Chem.<br />

Chem. Phys. 2006, 8, 4398–4401.<br />

94 Pople, J. A.; Head-Gordon, M.; Fox, D. J.;<br />

Raghavachari, K.; Curtiss, L. A. J. Chem.<br />

Phys. 1989, 90, 5622–5629.<br />

95 Martin, J. M. L.; de Oliveira, G. J. Chem.<br />

Phys. 1999, 111, 1843–1856.<br />

96 Ochterski, J. W.; Petersson, G. A.;<br />

Montgomery, J. A. J. J. Chem. Phys. 1996,<br />

104, 2598–2619.<br />

97 Maier, G.; Pfriem, S.; Schäfer, U.;<br />

Matusch, R. Angew. Chem. Int. Ed. Engl.<br />

1978, 17, 520–521.<br />

98 Nemirowski, A.; Reisenauer, H. P.;<br />

Schreiner, P. R. Chem. Eur. J. 2006, 12,<br />

7411–7420.<br />

99 Wiberg, K. B.; Walker, F. H. J. Am.<br />

Chem. Soc. 1982, 104, 5239–5240.<br />

100 Wiberg, K. B. J. Am. Chem. Soc. 1983,<br />

105, 1227–1233.<br />

101 Semmler, K.; Szeimies, G.; Belzner, J.<br />

J. Am. Chem. Soc. 1985, 107, 6410–6411.<br />

102 Stohrer, W.-D.; Hoffmann, R. J. Am.<br />

Chem. Soc. 1972, 94, 1661–1668.<br />

103 Masamune, S.; Sakai, M.; Ona, H.;<br />

Jones, A. J. J. Am. Chem. Soc. 1972, 94,<br />

8956–8958.<br />

104 Bremer, M.; Schleyer, P. v. R.; Schotz, K.;<br />

Kausch, M.; Schindler, M. Angew. Chem.<br />

Int. Ed. 1987, 26, 761–763.<br />

105 Fokin, A. A.; Lauenstein, O.;<br />

Gunchenko, P. A.; Schreiner, P. R.<br />

J. Am. Chem. Soc. 2001, 123, 1842–1847.<br />

106 Fokin, A. A.; Tkachenko, B. A.;<br />

Gunchenko, P. A.; Schreiner, P. R.<br />

Angew. Chem. Int. Ed. 2005, 44,<br />

146–149.<br />

107 Chiavarino, B.; Crestoni, M. E.;<br />

Fokin, A. A.; Fornarini, S. Chem. Eur. J.<br />

2001, 7, 2916–2921.<br />

108 Coppens, P. Angew. Chem. Int. Ed. 2005,<br />

44, 6810–6811.<br />

109 Wiberg, K. B.; Bader, R. F. W.;<br />

Lau, C. D. H. J. Am. Chem. Soc. 1987,<br />

109, 985–1001.<br />

110 Messerschmidt, M.; Scheins, S.;<br />

Grubert, L.; Pätzel, M.; Szeimies, G.;<br />

Paulmann, C.; Luger, P. Angew. Chem.<br />

Int. Ed. 2005, 44, 3925–3928.<br />

111 Smith, C. E.; Crawford, T. D.; Cremer, D.<br />

J. Chem. Phys. 2005, 122, 17430901–<br />

17430913.<br />

29


30 1 Introduction<br />

112 Marquardt, R.; Sander, W.; Kraka, E.<br />

Angew. Chem. Int. Ed. 1996, 35, 746–748.<br />

113 de Visser, S. P.; Filatov, M.; Schreiner,<br />

P. R.; Shaik, S. Eur. J. Org. Chem. 2003,<br />

4199–4204.<br />

114 Schleyer, P. v. R.; Bremer, M. Angew.<br />

Chem. Int. Ed. 1989, 28, 1226–1228.<br />

115 Tanaka, M.; Sekiguchi, A. Angew. Chem.<br />

Int. Ed. 2005, 44, 5821–5823.<br />

116 Radziszewski, J. G.; Yin, T.-K.;<br />

Renzoni, G. E.; Hrovat, D. A.;<br />

Borden, W. T.; Michl, J. J. Am. Chem.<br />

Soc. 1993, 115, 1454–1456.<br />

117 Hoenigman, R. L.; Kato, S.;<br />

Bierbaum, V. M.; Borden, W. T. J. Am.<br />

Chem. Soc. 2005, 127, 17772–17777.<br />

118 Keese, R. Chem. Rev. 2006, 106,<br />

4787–4808.<br />

119 Wiberg, K. B.; McMurdie, N.;<br />

McClusky, J. V.; Hadad, C. M. J. Am.<br />

Chem. Soc. 1993, 115, 10653–10657.<br />

120 Dodziuk, H.; Leszczynski, J.;<br />

Nowinski, K. S. J. Org. Chem. 1995, 60,<br />

6860–6863.<br />

121 Dinadayalane, T. C.; Priyakumar, U. D.;<br />

Sastry, G. N. J. Phys. Chem. A 2004, 108,<br />

11433–11448.<br />

122 Kenny, J. P.; Krueger, K. M.; Rienstra-<br />

Kiracofe, J. C.; Schaefer III, H. F. J. Phys.<br />

Chem. A 2001, 105, 7745–7750.<br />

123 Dodziuk, H. J. Mol. Struct. 1990, 239,<br />

167–172.<br />

124 Lewars, E. G. J. Phys. Chem. A 2005, 109,<br />

9827–9830.<br />

125 Lenoir, D.; Wattenbach, C.; Liebman, J. F.<br />

Struct. Chem. 2006, 17, 419–422.<br />

126 Smith, Z.; Andersen, B.; Bunce, S.<br />

Helv. Chim. Acta A 1977, 31, 557.<br />

127 Boese, R.; Blaeser, D.; Gomann, K.;<br />

Brinker, U. H. J. Am. Chem. Soc. 1989,<br />

111, 1501.<br />

128 Wiberg, K. B.; McMurdy, N.;<br />

McClusky, J. V.; Hadad, C. M. J. Am.<br />

Chem. Soc. 1993, 115, 10653.<br />

129 Wiberg, K. B.; McClusky, J. V.<br />

Tetrahedron Lett. 1987, 28, 5411.<br />

130 Dodziuk, H.; Leszczy�ski, J.;<br />

Nowi�ski, K. J. Org. Chem. 1995, 60, 6860.<br />

131 Wiberg, K. B.; Snoonian, J. R. J. Org.<br />

Chem. 1998, 63, 1390.<br />

132 Yufit, D. S.; Howard, A. J. K.;<br />

Kozhushkov, S. I.; Kostikov, R. R.;<br />

de Meijere, A. Acta Crystallogr. 2001,<br />

C57, 968.<br />

133 de Meijere, A.; von Seebach, M.;<br />

Zollner, S.; Kozhushkov, S. I.;<br />

Belov, V. N.; Boese, R.; Haumann, T.;<br />

Yufit, D. S.; Howard, A. J. K. Chem. Eur.<br />

J. 2001, 7, 4021.<br />

134 Fitjer, L.; Scheuermann, H.-J.;<br />

Klages, U.; Wehle, D.; Stephenson, D. S.;<br />

Binsch, G. Chem. Ber. 1986, 119, 1144.<br />

135 Wulf, K.; Klages, U.; Rissom, B.;<br />

Fitjer, L. Tetrahedron 1997, 53, 6011.<br />

136 Weiser, J.; Golan, O.; Fitjer, L.; Biali, S. E.<br />

J. Org. Chem. 1996, 61, 8277.<br />

137 Hutmacher, H. M.; Fritz, H.-G.;<br />

Musso, H. Angew. Chem. Int. Ed. 1975,<br />

14, 180.<br />

138 Fritz, H.-G.; Hutmacher, H. M.;<br />

Musso, H.; Åhlgren, G.; Akermark, B.;<br />

Karlsson, R. Chem. Ber. 1976, 3781, 180.<br />

139 de Meijere, A.; Redlich, S.; Frank, D.;<br />

Magull, J.; Hofmeister, A.; Menzel, H.;<br />

Kenig, B.; Svoboda, J. Angew. Chem. Int.<br />

Ed. 2007, 46, 4574.<br />

140 McMurry, J. E.; Hodge, C. N. J. Am.<br />

Chem. Soc. 1984, 106, 6450.<br />

141 McMurry, J. E.; Lectka, T.; Hodge, C. N.<br />

J. Am. Chem. Soc. 1989, 111, 8867.<br />

142 McMurry, J. E.; Lectka, T. J. Am. Chem.<br />

Soc. 1993, 115, 10167.<br />

143 Alder, R. W.; East, S. P. Chem. Rev. 1996,<br />

96, 2097.<br />

144 Bodige, S. G.; Sun, D. L.;<br />

Marchand, A. P.; Namboothiri, N. N.;<br />

Shukla, R.; Watson, W. H. J. Chem.<br />

Crystall. 1999, 29, 523.<br />

145 Yildirim, T.; Gehring, P. M.;<br />

Neumann, D. A.; Eaton, P. E.; Emrick, T.<br />

Carbon 1998, 36, 809.<br />

146 Moret, R. Acta Crystallogr. A 2005, 61,<br />

62.<br />

147 Blank, V. D.; Buga, S. G.;<br />

Dubitsky, G. A.; Serebryanaya, N. R.;<br />

Denisov, V. N.; Marvin, A. V. I. B. N.;<br />

Popov, M. Y. Mol. Cryst. Liquid Cryst. Sci.<br />

Technol. C 1996, 7, 251.<br />

148 Makha, M.; Purich, A.; Raston, C. L.;<br />

Sobolev, A. N. Eur. J. Inorg. Chem. 2006,<br />

3, 507.<br />

149 Ternansky, R. J.; Balogh, D. W.;<br />

Paquette, L. A. J. Am. Chem. Soc. 1982,<br />

104, 4503.<br />

150 Hopf, H. Classics in Hydrocarbon<br />

Chemistry; Wiley-VCH, 2000, p. 138.<br />

151 Hehre, W. J.; Pople, J. A. JACS 1975, 97,<br />

6941.


152 Wagner, H.-U.; Szeimies, G.;<br />

Chandrasekhar, J.; Schleyer, P. v. R.;<br />

Pople, J. A. JACS 1978, 100, 1210.<br />

153 Szeimies, G.; Harnisch, J.; Stadler, K. H.<br />

Tetrahedron Lett. 1978, 243.<br />

154 Szeimies, G.; Harnisch, J.; Baumgärtel,<br />

O. JACS 1977, 99, 5183.<br />

155 Harnisch, J.; Legner, H.; Szeimies-<br />

Seebach, U.; Szeimies, G. Tetrahedron<br />

Lett. 1978, 3683.<br />

156 Harnisch, J.; Baumgärtel, O.;<br />

Szeimies, G.; Meerssche, v. M.;<br />

Declerq, J.-P. JACS 1979, 101, 3370.<br />

157 Schlüter, A. D.; Belzner, J.; Heywang, H.;<br />

Szeimies, G. Tetrahedron Lett. 1983, 24,<br />

891.<br />

158 Nakazaki, M.; Yamamoto, K.; Yanagi, J.<br />

J. Am. Chem. Soc. 1979, 101, 147.<br />

159 Marshall, J. A.; Flynn, K. E. J. Am. Chem.<br />

Soc. 1983, 105, 3360.<br />

160 Dodziuk, H., Molecular modelling of<br />

[8.8]betweenane and other cumulenes,<br />

unpublished result.<br />

161 Marshall, J. A. Acc. Chem. Res. 1980, 13,<br />

213.<br />

162 Viavattene, R. L.; Greene, F. D.;<br />

Cheung, L. D.; Majeste, R.;<br />

Trafonas, L. M. J. Am. Chem. Soc. 1974,<br />

96, 4342.<br />

163 Wiberg, K. B.; Matturro, M. G.;<br />

Okarma, P. J.; Jason, M. E. J. Am. Chem.<br />

Soc. 1984, 106, 2194.<br />

164 Wiberg, K. B.; Matturro, M. G.;<br />

Adams, R. D. J. Am. Chem. Soc. 1981,<br />

103, 1600.<br />

165 Ostrowski, M.; Dodziuk, H., to be<br />

published.<br />

166 Baker, W.; Rouvray, D. H. J. Chem. Ed.<br />

1978, 55, 645.<br />

167 van Tamelen, E. E.; Pappas, S. P. J. Am.<br />

Chem. Soc. 1963, 85, 3297.<br />

168 van Tamelen, E. E.; Pappas, S. P.;<br />

Kirk, K. L. J. Am. Chem. Soc. 1971, 93,<br />

6092.<br />

169 Schäfer, W.; Hellmann, H. Angew. Chem.<br />

Int. Ed. 1967, 6, 518.<br />

170 Schäfer, W. Angew. Chem. Int. Ed. 1966,<br />

5, 669.<br />

171 Balaban, A. T.; Banciu, M.; Ciorba, V.<br />

In: Annulenes, Benzo-, Hetero-, Homo-<br />

Derivatives, and their Valence Isomers,<br />

Vol. 2, pp. 5–21; CRC: Boca Raton, 1987.<br />

172 Norton, J. E.; Olson, L. P.; Houk, K. N.<br />

J. Am. Chem. Soc. 2006, 128, 7835.<br />

References<br />

173 Paquette, L. A.; Kesselmayer, M. A.<br />

J. Am. Chem. Soc. 1990, 112, 1258.<br />

174 Sakurai, H.; Ebata, K.; Kabuto, C.;<br />

Sekiguchi, A. J. Am. Chem. Soc. 1990,<br />

112, 1799.<br />

175 Dodziuk, H.; Sybilska, D.; Miki, S.;<br />

Yoshida, Z.; Sitkowski, J.;<br />

Asztemborska, M.; Bielejewska, A.;<br />

Kowalczyk, J.; Duszczyk, K.; Stefaniak, L.<br />

Tetrahedron 1994, 50, 3619.<br />

176 Osawa, E.; Musso, H. Angew. Chem, Int.<br />

Ed. Engl. 1983, 22, 1.<br />

177 Pascal, R. A. J. Chem. Rev. 2006, 106,<br />

4809.<br />

178 Lu, J.; Ho, D. M.; Vogelaar, N. J.;<br />

Kraml, C. M.; Pascal, R. A. J. J. Am.<br />

Chem. Soc. 2004, 126, 11168.<br />

179 Norton, J. E.; Houk, K. N. J. Am. Chem.<br />

Soc. 2004, 127, 4162.<br />

180 Stepien, L.; Latos-Grazynski, L.;<br />

Sprutta, N.; Chwalisz, P.; Szterenberg, L.<br />

Angew. Chem, Int. Ed. 2007, 46, 7869.<br />

181 In: Modern cyclophane chemistry;<br />

Gleiter, R.; Hopf, H., eds.; Wiley-VCH:<br />

Weinheim, 2004.<br />

182 Misumi, S.; Otsubo, T. Acc. Chem. Res.<br />

1978, 11, 251.<br />

183 Rademacher, P. In: Modern cyclophane<br />

chemistry; Gleiter, R., Hopf, H., Eds.;<br />

Wiley-VCH: Weinheim, 2004, p. 275.<br />

184 Ernst, L.; Ibrom, K. In: Modern cyclophane<br />

chemistry; Gleiter, R., Hopf, H.,<br />

Eds.; Wiley-VCH: Weinheim, 2004,<br />

p. 381.<br />

185 Otsubo, T.; Horita, H.; Misumi, S.<br />

Synth. Commun. 1976, 6, 591.<br />

186 Otsubo, T.; Tozuka, Z.; Mizogami, S.;<br />

Sakata, Y.; Misumi, S. Tetrahedron Lett.<br />

1972, 297.<br />

187 Otsubo, T.; Mizogami, S.; Otsubo, I.;<br />

Tozuka, Z.; Sakagami, A.; Sakata, Y.;<br />

Misumi, S. Bull. Chem. Soc. Japan 1973,<br />

46, 3519.<br />

188 T. Umemoto; Otsubo, T.; Misumi, S.<br />

Tetrahedron Lett. 1974, 1573.<br />

189 T. Umemoto; Otsubo, T.; Sakata, Y.;<br />

Misumi, S. Tetrahedron Lett. 1973,<br />

593.<br />

190 Gleiter, R.; Roers, R. In: Modern cyclophane<br />

chemistry; Gleiter, R., Hopf, H.,<br />

Eds.; Wiley-VCH: Weinheim, 2004,<br />

p. 105.<br />

191 Sekine, Y.; Boeckelheide, V. J. Am. Chem.<br />

Soc. 1981, 103, 1777.<br />

31


32 1 Introduction<br />

192 Hanson, A. W.; Cameron, T. S. J. Chem.<br />

Res. S1980, 336.<br />

193 Y.-T. Wu; Siegel, J. S. Chem. Rev. 2006,<br />

106, 4843.<br />

194 Kawasa, T.; Kurata, H. Chem. Rev. 2006,<br />

106, 5250.<br />

195 Holms, D.; Kumaraswamy, S.;<br />

Matzger, A. J.; Vollhardt, K. P. C. Chem.<br />

Eur. J. 1999, 5, 3399.<br />

196 Bürgi, H.-B.; Baldrige, K. K.;<br />

Hardcastle, K.; Frank, N. L.; Gantzel, P.;<br />

Siegel, J. S.; Ziller, J. Angew. Chem, Int.<br />

Ed. 1995, 34, 1454.<br />

197 van’t Hoff, J. H. La chimie dans l’éspace;<br />

Bazendijk: Rotterdam, 1875.<br />

198 Nakazaki, M.; Yamamoto, K.; Maeda, M.;<br />

Sato, O.; Tsutsui, T. J. Org. Chem. 1982,<br />

47, 1435.<br />

199 Macomber, S. S.; Hemling, C. T.<br />

J. Am. Chem. Soc. 1986, 108, 343.<br />

200 Negi, T.; Kaneda, T.; Mizuno, H.;<br />

Toyoda, T.; Sakata, Y.; Misumi, S. Bull.<br />

Chem. Soc. Japan 1974, 47, 2398.<br />

201 Negi, T.; Kaneda, T.; Sakata, Y.;<br />

Misumi, S. Chem. Lett. 1972, 703.<br />

202 Utimoto, K.; Tanaka, M.; Kitai, M.;<br />

Nozaki, H. Tetrahedron Lett. 1978,<br />

2301.<br />

203 Scott, L. T.; DeCicco, D. J.; Hyun, J. L.;<br />

Reinhardt, G. J. Am. Chem. Soc. 1985,<br />

107, 6546.<br />

204 Houk, K. N.; Scott, L. T.; Rondan, N. G.;<br />

Spellmeyer, D. C.; Reinhardt, G.;<br />

Hyun, J. L.; DeCicco, D. J.; Weiss, R.;<br />

Chen, M. H. M.; Bass, L. S.; Clardy, J.;<br />

Jorgensen, F. S.; Eaton, T. A.; Sarkozi, V.;<br />

Petit, C. M.; Ng, L.; Jordan, K. D. J. Am.<br />

Chem. Soc. 1985, 107, 6556.<br />

205 Lepetit, C.; Godard, C.; Chauvin, R.<br />

New J. Chem. 2001, 25, 572.<br />

206 Gholami, M.; Melin, F.; McDonald, R.;<br />

Ferguson, M. J.; Echegoyen, L.;<br />

Tykwinski, R. R. Angew. Chem, Int. Ed.<br />

2007, 46, 9081.<br />

207 Kawasa, T.; Oda, M. Pure and Appl.<br />

Chem. 2006, 78, 831.


2<br />

Distorted Saturated <strong>Hydrocarbons</strong><br />

2.1<br />

Molecules with Inverted Carbon Atoms<br />

Kata Mlinari�-Majerski<br />

2.1.1<br />

Introduction<br />

The tetrahedral geometry at a saturated carbon atom has been known as the<br />

only possible geometry for almost a century [1]. However, in the late 1960s<br />

theoretical predictions and some experimental results suggested other possibilities<br />

such as planar, pyramidal or even inverted geometry at the carbon [2]. The<br />

interest in nontetrahedral saturated carbon atoms has been growing very rapidly.<br />

Nontetrahedral geometries at a tetravalent carbon atom are defined by the four<br />

interatomic vectors emanating from the saturated carbons. Deformation of the<br />

conventional tetrahedral arrangement (Figure 2.1a) via an umbrella motion leads<br />

to the inverted geometry in which all of the four substituents lie in the same<br />

hemisphere (Figure 2.1b) [3a].<br />

As established by microwave spectra [4a], the simplest molecule having the<br />

carbon atoms with inverted geometry is bicyclobutane (Figure 2.1c) but the<br />

existence of this unusual feature was only recognized later by Paddon-Row and<br />

coworkers [4b]. The possible existence of this unusual geometry was explored by<br />

Wiberg who carried out ab initio calculations at the 6–31G* level, first predicted<br />

Figure 2.1 (a) Tetrahedral geometry at carbon; (b) Inverted geometry at carbon;<br />

(c) bicyclobutane; (d) [k.l.m]propellanes.<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

33


34 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

properties of small-ring propellanes, synthesized the molecules and confirmed<br />

the predictions [3, 5a].<br />

In particular, the ultimate member of the small-ring propellanes, extensively<br />

studied by Wiberg [3, 5], [1.1.1]propellane 1 was found to be much more stable<br />

than higher members of the series [k.l.m]propellanes with k = 2, d; l, m = 1, 2.<br />

This remarkable molecule represents a landmark in the quest for highly strained<br />

small-ring organic molecules.<br />

Propellanes are defined as the tricyclic systems (Figure 2.1d) in which three rings<br />

are fused together by a common central, bridgehead–bridgehead C–C bond [6].<br />

Large-ring propellanes (Figure 2.1d; k, l, m � 3) behave chemically like ‘normal’<br />

polycyclic hydrocarbons and their bridgehead carbon atoms have tetrahedral configuration<br />

[7]. However, reduction of the ring size as in 1–7 leads to the small-ring<br />

propellanes possessing two inverted carbon atoms in the central bond [3]. Such a<br />

nonstandard structure leads to several peculiarities. For instance, the small-ring<br />

propellanes 1–7 are remarkably reactive toward electrophiles and free radicals;<br />

2, 3 and 5 have remarkable thermal lability while 1 is stable [3].<br />

Oxa[3.2.1]propellane 8, is the first reported small-ring propellane [8], and the<br />

synthesis of the parent hydrocarbon 7 soon followed [9, 10]. The X-ray structure<br />

analysis of dichloro[3.2.1]propellane 7a proved the inverted configuration on the<br />

bridgehead carbon atoms [11]. Interestingly, in spite of the distance between the<br />

bridgehead atoms of 1.60 Å in 1 [5a], the mere existence of the central bond has<br />

been questioned for long time [2a] and the small-ring propellanes have become<br />

the subject of many theoretical studies and a great challenge to synthetic chemists.<br />

Until discovery of the molecules having inverted carbon atoms, in quantum<br />

chemistry there was no definition of such a useful concept as the chemical bond.


2.1 Molecules with Inverted Carbon Atoms<br />

One could only carry out calculations for a molecule as a whole. As discussed<br />

below, introduction of bond critical points allowed the researchers to refine the<br />

definition [34b, 34c]. This topic is discussed in Section 1.3.<br />

Several other small-ring propellanes 5, 6 and 7 containing a total of six atoms in<br />

their bridges as in [2.2.2]propellane [12], [3.2.1]propellane [9–11, 13, 14], and [4.1.1]<br />

propellane [15–17], respectively, have been prepared. Of these, particular interest<br />

was paid to [2.2.2]propellane 5 and many theoretical studies have been undertaken<br />

[18–21]. Prior to the synthesis of 5, Stohrer and Hoffmann had predicted a facile<br />

cleavage of its central C–C bond [20]. Davidson [21] calculated the rearrangement<br />

of 5 by applying a variety of theoretical methods (CASSCF, PUHF, MUMP2, DFT,<br />

UDFT, CI) suggested as methods for obtaining a reliable potential energy surface<br />

for diradicals. The results obtained suggested that [2.2.2]propellane 5 can exist<br />

and would have a substantial barrier for opening to the bicyclic diradical or rearrangement<br />

to dimethylenecyclohexane. However, monosubstituted [2.2.2]propellane<br />

proved highly unstable [12a]. Recently, the extremely stable perfluoro[2.2.2]<br />

propellane has been prepared [12c].<br />

A more strained geometry than that in 5 was expected in [3.1.1]-, [2.2.1]-, [2.1.1]-,<br />

and particularly in [1.1.1]propellanes (4–1, respectively). As mentioned earlier,<br />

even the possibility of their existence was questioned [2a, 18]. However, for the<br />

[1.1.1]propellane 1 the calculations led to the conclusion that 1 should be more<br />

stable than the corresponding diradical [19, 20]. Wiberg and Walker found that<br />

high energy of 65 kcal mol –1 is needed to break the central bond of 1 to obtain the<br />

bicyclo[1.1.1]pentyl diradical [5a]. Amazingly, that calculation result also suggested<br />

that 1 should be the most easily prepared and the most stable small-ring propellane.<br />

This prediction proved correct [5].<br />

Indeed, the small-ring propellanes, [3.1.1]propellane 4 [15c, 15d, 22, 23] and<br />

[1.1.1]propellane 1 [5] or their derivatives [24–26] have been synthesized. On the<br />

other hand, their homologs [2.2.1]propellane 3 [27–29] and [2.1.1]propellane 2<br />

[30, 31] were observed either at low temperature, captured in the argon matrix,<br />

or as inter mediates which were subsequently trapped by reagents to give stable<br />

products.<br />

2.1.2<br />

Small-ring Propellanes: Computational and Physicochemical Studies<br />

In view of the exceptional role that calculations have played for these compounds,<br />

this Section will begin with a discussion of theoretical results. As mentioned earlier,<br />

the unusual properties of the inverted carbon atoms and the mere existence and<br />

nature of the central bond that connects the two inverted bridgehead atoms in the<br />

small-ring propellanes have been the subject of many theoretical investigations<br />

[19, 20, 27, 30, 32–41], X-ray [11, 15b, 42] and electron-diffraction analyses [15b,<br />

43, 44], vibrational [45], photoelectron [46] and electron impact [47] spectroscopic<br />

and NMR studies [23b, 48–52].<br />

The experimental and theoretical results of many different research groups have<br />

led to the apparently opposite conclusions on the small-ring propellane structure.<br />

35


36 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

Some researchers even questioned the very existence of the central bond assuming<br />

that the molecules have biradical structure with lone-pairs directed outside the<br />

cage [19a, 20, 32]. For instance, according to the early low-level ab initio molecular<br />

orbital studies on [1.1.1]propellane 1 [19a] there is no evidence for a central bond<br />

in terms of charge distribution since the sp 4 hybrid orbitals forming the bond are<br />

directed away from each other and with the zero overlap population in this bond.<br />

Jackson and Allen [32a] pointed out that 1 is electron-deficient because its HOMO<br />

is nonbonding or slightly antibonding and there is low electronic density between<br />

the bridgehead carbon atoms C1 and C3. The C1–C3 bond is considered to be<br />

formed by three-center two-electron molecular orbitals and termed ‘�-bridged �<br />

bond’. These theoretical results are supported by an experimental investigation<br />

of the differential electron density of two [1.1.1]propellane derivatives [44]. On<br />

the basis of ab initio quantum chemical calculations, Feller and Davidson [35]<br />

came to the conclusion that 1 is just a strained cage with negligible bridgehead to<br />

bridgehead through-space covalent bonding. Wiberg, Bader and Lau [34b, c] came<br />

to a different conclusion by analyzing the bond critical points. Their reasoning<br />

was based on the assumption that the electronic charge density is a physical<br />

property of the system and as such it should be model-independent. On the basis<br />

of an analysis of the second derivative of the electron density determined by all<br />

occupied orbitals, the authors concluded (1) that the exceptional stability of 1 is<br />

due to specific three-center integrals and (2) that bonding was found between the<br />

bridgehead atoms since there is an appreciable accumulation of charge between the<br />

bridgehead atoms. The results of quasi ab initio PRDDO calculation additionally<br />

support the conclusion that the bridgehead atoms in 1 are weakly bonded in the<br />

ground state [38]. Recently, the electron density and bonding between the inverted<br />

carbon atoms have been studied experimentally on the [1.1.1]propellane derivative<br />

9 by Luger [39] who discussed different criteria for the existence of a bond.<br />

The authors discussed the quantitative results obtained by a synchrotron<br />

experiment and model calculations concluding that ‘the bond critical point was<br />

found between the bridgehead atoms (C1–C3) in 9. This bond is unusual according to<br />

topological analysis: it has a bond path with a bond critical point of significant density,<br />

as is characteristic for covalent bond, but no charge accumulation is evident at the bond<br />

critical point where the Laplacian is positive’. A bond order n was derived from the<br />

electron density at the critical point and was found to be 0.71 which is close to<br />

0.73 as determined by Wiberg [34]. A similar low bond order was determined for<br />

[2.1.1]propellane 2, while in [2.2.1]propellane 3 bond order was 1.0 and in [2.2.2]<br />

propellane 5 it was 1.3 [34b]. The nature of bonding in 1 is also discussed on the<br />

basis of localized MOs and bond indices [40]. LMO and bond indices analyses<br />

indicate that there is a significant interaction between the bridgehead carbon


2.1 Molecules with Inverted Carbon Atoms<br />

Figure 2.2 The strain energy (SE, kcal mol –1 ) of small-ring propellanes 1–3, 5 and bicyclo[1.1.1]<br />

pentane (10) at the 6-31G*/6-31G* level [34a].<br />

atoms in 1 and that about 80% of the total contribution of the central bond<br />

comes from the HOMO. The authors have also found out that in the series of the<br />

small-ring propellanes the total bond index increases in the order [1.1.1] < [2.1.1]<br />

< [2.2.1] < [2.2.2]propellane whereas the HOMO contribution remains almost<br />

constant [40]. Therefore, the difference is in the electron density of the outer<br />

envelope of the central propellane bond which most probably influences their<br />

reactivity. However, the strain energies of these propellanes (of 103 kcal mol –1 ,<br />

106 kcal mol –1 , 109 kcal mol –1 , 97 kcal mol –1 and 67 kcal mol –1 for 1, 2, 3, 5, 10,<br />

respectively) are very similar (Figure 2.2) [34a]. Although they are all sufficiently<br />

strained to permit facile reaction, some are quite reactive and other quite stable,<br />

as is [1.1.1]propellane 1. The high stability of 1 is due to the low exothermicity of<br />

the reactions of breaking the C1–C3 bond to give bicyclo[1.1.1]pentane 10, still<br />

highly strained structure. The strain release is less than a third of the strain energy<br />

and breaking the side bond is forbidden by symmetry [34].<br />

Several studies of the strain energy of small-ring propellanes have been reported<br />

[34, 37]. They point to the charge withdrawal from the neighboring groups to the<br />

bridgehead region with the increasing strain [34c].<br />

A photoelectron (PE) spectroscopic study of [1.1.1]propellane 1 revealed that<br />

there should be only minute change in the geometry between 1 and its radical<br />

cation [46a]. This was attributed to the nonbonding or slightly antibonding<br />

character of the HOMO of 1. PE investigation on the less strained [3.1.1]propellane<br />

4 [46b] was indicative of a slightly bonding HOMO, while the PE spectra of<br />

several [n.1.1]propellanes [46c] showed that the energy of the first band depends<br />

very strongly on n. For n = 1 the ionization energy is around 9 eV, while for n = 3<br />

and 4 it is around 8 eV.<br />

The central C1–C3 bond in small-ring propellanes are longer (e.g. 159.6 ±<br />

0.05 pm in 1) than normal C–C bond (154 pm) whereas the length of the bridge–<br />

bridgehead C–C bonds of 152.5 pm were found to be close to the bond length in<br />

a cyclopropane ring (151 pm) [43–45b]. The results of natural bond orbital (NBO)<br />

analysis of the central bond in [1.1.1]propellane and some [1.1.1]heteropropellanes<br />

37


38 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

suggested that in these compounds the C1–C3 bond lengths are closely controlled<br />

by p character of the hybrid orbitals [53]. While Jarret and Cusumano [49], on the<br />

basis of NMR measurements, suggested a low p character of the central propellane<br />

bond other authors assigned a very high p character to this central bond on<br />

the basis of coupling constants [50]. Recently, the high-level ab initio calculations<br />

yielded the values of all NMR shielding and spin–spin coupling constants in 1<br />

[51]. The computed NMR parameters agree with the experimental values, except<br />

for the C1–C3 coupling constants. The ab initio studies using the equations of<br />

motion approach have also been reported J(CC) of the bridgehead bond in other<br />

small-ring propellanes [52].<br />

The controversy regarding the central bond and the nature of bonding interactions<br />

in the small-ring propellanes still revolves around different explanations<br />

of the similar experimental and theoretical results. Recently, the comprehensive<br />

work concerning [1.1.1]propellane and the other compounds containing the<br />

bicyclo[1.1.1]pentane skeleton was reviewed by Michl [54]. The work on the<br />

propellane addresses the very important issue of the nature of central (C1–C3)<br />

bond. The results recently reported by Luger [39] represent a culmination of the<br />

experimental studies of charge density and, as pointed out by Coppens [55], the<br />

characterization of this chemical bond is not a closed subject.<br />

2.1.3<br />

Small-ring Propellanes: Experimental Results<br />

2.1.3.1 Preparation and Reactivity of [1.1.1]Propellane<br />

As predicted by Wiberg [5a], the synthesis of [1.1.1]propellane 1 was successfully<br />

carried out by reduction of dibromide 11 with butyllithium. However, 1 had<br />

remained rather inaccessible because of the long and tedious synthesis of precursor<br />

11, until Szeimies [24a] synthesized 1 in a single step (Scheme 2.1), starting from<br />

the readily available 1,1-dibromo-2,2-bis(chloromethyl)cyclopropane 12. The latter<br />

procedure made 1 the most easily prepared small-ring propellane.<br />

Scheme 2.1<br />

The method of ring closure with 2 equiv. of alkyllithium was also used to prepare<br />

the bridged [1.1.1]propellanes 13 and the related propellane 14 [24a, 24b, 25b].<br />

In addition, the propellane 13 was also prepared by intramolecular addition of<br />

carbene to a double bond [25a].


2.1 Molecules with Inverted Carbon Atoms<br />

Wiberg and McMurdie [56] reported the formation of 1 by nucleophilic attack<br />

on 1,3-diiodobicyclo[1.1.1]pentane 15, as shown on Scheme 2.2.<br />

Scheme 2.2<br />

The availability of 1 enabled the study of properties and reactivity of that remarkable<br />

molecule. Most of the properties of 1, first predicted by calculation, have been<br />

confirmed experimentally. For example, the reaction of 1 with acetic acid led to<br />

ring opening and formation of 3-methylenecyclobutyl acetate (16, Scheme 2.3).<br />

The enthalpy of that reaction was measured [45a] and the enthalpy of formation<br />

of 16 was determined to be 85 kcal mol –1 in good agreement with the theoretically<br />

estimated value of 89 kcal mol –1 [34a, 57].<br />

Scheme 2.3<br />

[1.1.1]Propellane 1 is known to be stable at the room temperature but it rearranges<br />

to methylenecyclobutene 17 [5a] at 114 °C or to 1,2-dimethylenecyclopropane<br />

18 at 370 °C (Scheme 2.4) [24b]. At first, the reason for this difference<br />

Scheme 2.4<br />

39


40 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

was not understood. Recently, Jarosch, Walsh and Szeimies [58] investigated the<br />

kinetics of thermal rearrangement of 1 by gas-phase pyrolysis in the stationary<br />

system and found that the unimolecular reaction leads to dimethylenecyclopropane<br />

18 and its thermal isomerization product ethenylidenecyclopropane 20. The<br />

ab initio and DFT calculations of the potential energy surface indicated that the<br />

isomerization follows an asynchronous reaction path in which two side bonds of<br />

1 are broken with activation barrier of 40.0 kcal mol –1 . Furthermore, the authors<br />

showed that the minor product methylenecyclobutene 17 and its thermal isomerization<br />

product 1,2,4-pentatriene 19 resulted from the side reaction catalyzed by<br />

the reaction vessel surface (Scheme 2.4).<br />

Wiberg [59] studied a wide variety of the free radical reactions of 1 and compared<br />

its reactivity to those of bicyclo[1.1.0]butane 21 and bicyclo[2.1.0]pentane 22.<br />

Having observed that the reactions of 1 with free radicals were faster then that<br />

of 21, whereas 22 was relatively inert, the authors concluded that the reactivity<br />

was not determined by the strain energy relief or the HOMO energy of these<br />

compounds, but rather by the local charge distribution that could be a more<br />

important factor.<br />

A free radical addition represents the most effective route of preparing various<br />

1,3-disubstituted bicyclo[1.1.1]pentane derivatives 23 (Scheme 2.5) [5b, 24a,<br />

59–61].<br />

Scheme 2.5<br />

In some cases these reactions lead to oligomers [5b]. The free radicals were found<br />

to react more readily with 1 than with styrene [62]. Furthermore, it was found that<br />

the electronic energy transfer to 1 occurs with a rate constant significantly below<br />

the diffusion-controlled limit (for instance, triplet benzophenone was quenched<br />

by 1 with a bimolecular rate constant of 9.9 � 10 6 M –1 s –1 ) [62b].<br />

In addition, the reactions of 1 with electron deficient alkenes and alkynes have<br />

been studied and compared with the corresponding reactions of 21 and 22 [5b,<br />

63]. In these reactions relative reactivity of 1 and 21 varied considerably with the<br />

used reagent while 22 was again unreactive.<br />

The bridged [1.1.1]propellanes (for example 13 and 14) behave in a similar<br />

fashion as the parent propellane 1 [24a, 24b, 25].


2.1 Molecules with Inverted Carbon Atoms<br />

2.1.3.2 Preparation and Reactivity of [2.1.1]Propellane and [2.2.1]Propellane<br />

There are few reports on experimental observation of [2.1.1]propellane 2 [30–31]<br />

and [2.2.1]propellane 3 [27–29]. These two propellanes are the only members of<br />

the small-ring propellanes that could not be isolated as the stable molecules but<br />

were observed in the argon matrix at low temperature (~30 K) by IR spectroscopy.<br />

The propellanes 2 [30] and 3 [27] were obtained by gas-phase dehalogenation of 24<br />

and 25, respectively, with potassium. When the matrix was warmed up to about<br />

50 K the compounds polymerized. When bromine was introduced to the matrix<br />

of 3, 1,4-dibromonorbornane 26 was obtained (Scheme 2.6).<br />

Scheme 2.6<br />

It is also suggested that [2.2.1]propellane 3 is an intermediate in the electrolytic<br />

reduction of the corresponding 1,4-dihalonorbornanes 25 and 26 [28].<br />

Experimental results for 2 and 3 and their derivatives [29, 31b, 31c] showed their<br />

high reactivity toward free radicals and some organometalic reagents. Recently,<br />

Jarosch and Szeimies [64] studied thermal rearrangement of 2 by using density<br />

functional and the ab initio molecular orbital calculations and found that the low<br />

energy isomerization path proceeds via a retro-carbene reaction to give carbene<br />

27 (Scheme 2.7).<br />

Scheme 2.7<br />

The most favorable consecutive reactions were found to be 1,2-hydrogen<br />

shifts leading to dienes 28 and 29. A few higher propellanes are discussed in<br />

Section 3.3.<br />

41


42 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

2.1.3.3 [1.1.1]Propellane as the Precursor for the Synthesis of Other Unusual<br />

Molecules<br />

The feasibility of radical addition across the central bond of [1.1.1]propellane<br />

1 enables polymerization of 1 into functionalized [n]staffanes [n]1. Michl [65]<br />

reported the synthesis of [n]staffanes (Scheme 2.8) as a new family of the endfunctionalized,<br />

inert, transparent and straight rods with a van der Waals radius<br />

of 2.3 Å and a length increment of 3.35 Å, which could be used as a construction<br />

element [54, 65]. Schlüter reported the formation of the same structures by<br />

spontaneous polymerization of neat 1 [66]. Also, he reported the first anionicallyinduced<br />

polymerization of the bridged [1.1.1]propellane 13 [67]. The [n]staffane<br />

skeleton is expected to be of importance for prospective applications as an electrical<br />

insulator [54, 68].<br />

Scheme 2.8<br />

[1.1.1]Propellane 1 also served as the precursor for the preparation of an<br />

unnatural amino acid, 3-aminobicyclo[1.1.1]pentane-1-carboxylic acid which has<br />

been incorporated into linear and cyclic peptides [69].<br />

Next, the truly remarkable molecule of high symmetry, [1.1.1.1]paddlane 30 that<br />

should have two pyramidal carbon atoms, was reported to be one of the products<br />

(although only in 7% of yield) in the direct photolysis of 1 in the presence of<br />

diazomethane [70] (Scheme 2.9).<br />

Scheme 2.9<br />

The authors reported the NMR spectrum to be compatible with the proposed<br />

structure. However, in view of high strain in this molecule, which was estimated to<br />

be 456 kcal mol –1 [71], the claim that they obtained 30 is extremely surprising.


2.1.4<br />

New Hypothetical Molecules with Inverted Carbon Atoms<br />

2.1 Molecules with Inverted Carbon Atoms<br />

On the basis of simple model arguments an infinite family of [k.1.1]propellanes<br />

was found to possess inverted carbon atoms independent of the k value [36a]. This<br />

is not surprising in view of the bicyclobutane structure. Dodziuk and coworkers<br />

[36, 72], on the basis of molecular mechanics [36a] and ab initio quantum chemical<br />

calculations [72], proposed small ring geminanes 31–33 as another group of<br />

molecules that should have inverted carbon atoms.<br />

The latter molecules have not yet been obtained but they seem to be plausible<br />

synthetic targets and challenge for the synthetic chemist.<br />

Acknowledgments<br />

I am sincerely grateful to my late husband Professor Zdenko Majerski who<br />

triggered my interest and love for chemistry of the strained ring compounds.<br />

I gratefully acknowledge financial support of the Ministry of Science, Education<br />

and Sport of the Republic of Croatia (grant 098-0982933-2911).<br />

43


44 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

2.2<br />

Molecules with Planar and Pyramidal Carbon Atoms<br />

Helena Dodziuk<br />

About 90 years after its formulation, van’t Hoff and Le Bel’s hypothesis [73, 74]<br />

on the tetrahedral orientation of substituents on a tetravalent carbon atom has<br />

proven its validity and predictive power. In particular, it revealed the molecular<br />

foundations of chirality and the number of isomers of cyclohexane derivatives. The<br />

idea was commonly accepted in the 1960s. Therefore, it took the great intellectual<br />

courage of Roald Hoffmann and his collaborators to propose ‘planar methane’, that<br />

is a tetravalent carbon atom lying in a plane with its four substituents [75, 76] as a<br />

part of an plausible organic structure. On the basis of very simple model quantum<br />

chemical calculations, QC, the authors analyzed prospective hydrocarbons that<br />

could eventually have an atom exhibiting this configuration.<br />

Several reviews on the planar and pyramidal carbon atoms have been published<br />

[77–83]. Most studies in this area dealt with molecules in which heteroatoms took<br />

part in forcing the planar configuration on a carbon atom [81, 84, 85]. Since this<br />

monograph is devoted to hydrocarbons, the studies dealing with this mechanism<br />

of planarization will not be discussed here.<br />

Interestingly, QC have shown that the planar methane configuration does not<br />

correspond to a minimum on the potential energy surface [86, 87]. However, the<br />

latter observation did not hamper the development and the idea of the so called<br />

planar carbon turned out to be very fruitful resulting, in particular, in vigorous<br />

research in small ring fenestranes 34 [88, 89] and paddlanes 2 [90, 91] which<br />

were thought to have carbon atom(s) with the planar configuration. Concerning<br />

the former group of compounds, QC have shown that [4.4.4.4]fenestrane 34a<br />

(k = l = m = n = 1), also called windowpane, assumes a nonplanar arrangement<br />

of substituents on the central atom that is called half-planar (bisphenoidal) configuration<br />

[92]. This configuration was close to that determined by X-ray for the<br />

smallest fenestrane synthesized [4.4.4.5]fenestrane 34 (k = 2, l = m = n = 1) [93].<br />

As concerns paddlanes 35 [89], using model calculations Wiberg showed that the<br />

smallest, [1.1.1.1]- and [2.2.2.2]-paddlanes 35 (k = l = m = n = 1 and 2, respectively),<br />

were too strained to exist [91] (a published synthesis of the former molecule briefly<br />

discussed in the former section seems unreliable [94]) while the smallest synthesized<br />

paddlane 36 (k = 10, l = m = n = 2] understandably had a close to tetrahedral<br />

configuration on the ternary carbon atoms [90].


2.2 Molecules with Planar and Pyramidal Carbon Atoms<br />

When analyzing the possibility of planar configuration at a tetravalent carbon<br />

atom Hoffmann group used extended Hückel, EHT, and CNDO/2 calculations<br />

[75, 76] to estimate the energy difference between the planar and tetrahedral<br />

configurations of methane<br />

� �E = E planar – E tetrahedral (2.1)<br />

equal to 5.5 eV and 10.8 eV, respectively, and discussed two ways in which the value<br />

can be lowered. The former configuration could be forced either by destabilizing<br />

the tetrahedral arrangement (for instance by steric hindrance in the system as in<br />

34 with k, l, m, n smaller than 3) or by stabilizing the planar configuration on the<br />

carbon atom. With this purpose in mind he discussed the bonding in square planar<br />

methane in the valence bond framework. Within this model two sp 2 hybrids at the<br />

carbon are engaged in normal two-electron two-center bonds with two hydrogen<br />

atoms making use of two out of the four carbon valence electrons. The third of<br />

the sp 2 hybrids forms a two-electron three-center bond with the remaining two<br />

hydrogens using only the hydrogen electrons, while the remaining two valence<br />

electrons of carbon are placed in the 2p orbital perpendicular to the molecular<br />

plane. Equivalence of all CH bonds results from resonance among equivalent<br />

structures with different positioning of three-center and two-center CH bonds.<br />

Such a procedure or analysis of the planar methane orbitals lead to the following<br />

conclusions:<br />

1. All CH bonds in planar methane are weaker than in tetrahedral one with the<br />

average bond order of 3/4 in the former case. In terms of MO, there are only<br />

six bonding electrons in the planar configuration while there are eight in the<br />

tetrahedral one leading to a significant � bond weakening.<br />

2. There is a considerable electron transfer from hydrogens to carbon locating<br />

considerable electron density on the latter atom.<br />

3. There is a pure 2p electron lone pair perpendicular to molecular plane.<br />

4. Two deformations of the tetrahedral methane (T d � D 2 � D 4h and T d � D 2d �<br />

D 4h ) which represent a symmetry allowed process can transform it to the planar<br />

one.<br />

Conclusions 2 and 3 formed the basis of the strategy that should lead to a carbon<br />

atom lying in a plane with its four substituents. The substitution of hydrogen atoms<br />

by good electron acceptors, like C�N, should cause delocalization of the lone pair<br />

reducing �E value (EHT method) to 3.4 eV. An alternative procedure consisted in<br />

45


46 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

incorporation of the lone pair to form (4n + 2) �-electron system as in the �-cation<br />

of an aromatic anion 37 or benzonium ion 38 resulting in �E (EHT) values of 4.2<br />

and 2.9 eV, respectively. Substitution of hydrogen atoms by less electronegative<br />

groups yielded further �E lowering to 1.8 eV in C(BH 2) 4 and 2.9 eV for C(SiH 3) 4<br />

when 3d orbitals on Si were included. In the latter case Si simultaneously acted<br />

as a �-donor and �-acceptor. Both factors acting in the direction of favoring the<br />

planar configuration at the central carbon operate in hypothetical tetrasila[5.5.5.5]<br />

fenestrane 39. The effect of electronegativity differential should be even stronger<br />

for Li substitution or replacement of carbon by N + .<br />

Purely qualitative criteria were invoked to reject unsaturated fenestranes 40–42<br />

as candidates for an effective �-stabilization while 43–45 were found to be especially<br />

promising.<br />

As mentioned before, vigorous activity in the domains of fenestranes and<br />

paddlanes did not produce hydrocarbons having either pyramidal or planar carbon<br />

atom. Even if those attempts were unsuccessful, nevertheless they show that a<br />

incorrect prediction can lead to exciting chemistry.<br />

By playing with molecular Dreiding models Dodziuk has found that cyclooctane<br />

in the crown conformation could accommodate a carbon atom in the planar<br />

configuration. The resulting molecule was dubbed bowlane [95]. Subsequent<br />

molecular mechanics, MM, [96, 97] calculations [95] yielded a structure 46 of C 4v<br />

symmetry with the pyramidal carbon atom. A whole family of such molecules


2.2 Molecules with Planar and Pyramidal Carbon Atoms<br />

46–51 was then studied using MM and semiempirical quantum chemical calculations<br />

[98] (yielding intermediate configurations between the distorted pyramidal<br />

and tetrahedral ones) and together with two ‘dimers’ 52 and 53. The former was<br />

expected to have the planar configuration on the central carbon atom since the<br />

upper and lower ‘halves’ were thought to force the desired bonds configuration.<br />

However, these expectations were not fulfilled, resulting in C 4v symmetry and<br />

both 52 and 53 were found to be highly strained.<br />

On the other hand, ab initio quantum chemical calculations for the parent 46<br />

using STO-3G and 6-31G* basis sets carried out by McGrath et al. [99] yielded a<br />

structure of a lower C 2v symmetry as an energy minimum with the bonds connecting<br />

the apical carbon atom with its neighbors departing only slightly from the<br />

planar arrangement. Similarly, HF/6-31G* calculations for 52, named octaplane,<br />

carried out by the latter group [100] showed that the S 4 structure corresponding<br />

to a local minimum represents ‘the closest approach to planarity for a tetracoordinated<br />

carbon atom in a neutral saturated hydrocarbon reported today’. The most<br />

interesting features of the calculated minimum structure of 52 were considerable<br />

bond length distortions (up to 161.5 pm), moderate energy cost for achieving<br />

planarity of ca. 70 kJ mol –1 calculated at MP2/6-31G*//HF/6-31G* level and the<br />

fact that the highest occupied molecular orbital is essentially a pure lone pair orbital<br />

localized on the quaternary carbon atom. By capping the central fragment with<br />

two cyclobutanes 55, cyclohexanes in the boat conformation 56, or cyclooctane<br />

47


48 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

in the crown conformation 52 (that is earlier proposed bowlane ‘dimer’ [98]) the<br />

authors arrived at the family of alkaplanes [100, 101].<br />

The aim of designing theoretically a neutral hydrocarbon with a planar tetracoordinated<br />

carbon atom was achieved by Radom and Rasmussen [101] when<br />

dimethanospiro[2.2]octaplane 54 was obtained by linking two opposite pairs of<br />

adjacent carbon atoms � to the central one and adding methylene bridges connecting<br />

the top and bottom caps. The calculations at the MP2 level using 6-31G(d)<br />

basis set resulted in D 2h symmetrical minimum structure having the planar configuration<br />

on the central carbon atom with the longest C–C bond of 159.1 pm and<br />

the shortest one in cyclopropane rings of 144.5 pm. By preliminary exploration<br />

of a number of decomposition pathways corresponding to the lowest vibrational<br />

modes, the authors checked that, in spite of high degree of strain, the structure<br />

found for 54 appeared to lie in a relatively deep potential well. Similarly to 46, the<br />

HOMO in 54 has predominantly p-type lone pair character localized at the central<br />

quaternary carbon atom and surrounded by the hydrocarbon cage. Interestingly,<br />

the latter molecule has an extremely low ionization energy of ca. 5 eV comparable<br />

to those of alkali metals Li and Na.<br />

Bowlane 46 also formed the basis of a hypothetical hemispiroalkaplanes family<br />

57–59 consisting, among other of cyclohexane in the boat conformation, norbornane<br />

or cyclooctane in the crown conformation fragments capped with bent<br />

spiro[3.3]pentane unit [102]which also contained a pyramidal carbon atom.<br />

To summarize, up to the present theoretical studies of molecules that should<br />

have planar carbon atom(s) have been carried out mostly at the HF level using


2.3 A Theoretical Approach to the Study and Design of Prismane Systems<br />

rather small basis sets. The MP2 calculations have been carried out at best. It<br />

should be stressed that DFT calculations seem not to be fully reliable, in particular<br />

when all the rearrangement/dissociation pathways are analyzed.<br />

Until now no neutral hydrocarbon with a planar carbon atom has been synthesized.<br />

Hoffmann’s bright idea of molecules having such unusual spatial structure<br />

still awaits its realization.<br />

2.3<br />

A Theoretical Approach to the Study and Design of Prismane Systems<br />

Tatyana N. Gribanova, Vladimir I. Minkin and Ruslan M. Minyaev<br />

2.3.1<br />

Introduction<br />

The aesthetically attractive highly symmetrical molecular structure and unusual<br />

properties of prismanes ensure that they continue to draw the attention of both<br />

experimentalists and theoreticians [103–105]. The starting point for the extensive<br />

studies of these compounds was the synthesis of cubane [106, 107] which initiated<br />

the progress of a specific area of these highly strained, albeit kinetically stable<br />

compounds [107–111]. In contrast with cubane, the experimental study of two<br />

other currently synthesized members of this class, triprismane and pentaprismane,<br />

is much scarcer. Theoretical investigations of prismane systems come to<br />

the fore more and more actively. Whereas in the first years of the development<br />

of prismane chemistry the theoretical work was mostly limited to the analysis of<br />

geometries and steric strain of the synthesized compounds [104, 109], the modern<br />

quantum chemical calculations of prismanes based on more sophisticated and<br />

reliable methods, and more powerful computers are focused on the design of new<br />

compounds with interesting properties. This review concentrates on the most<br />

important and interesting results of the computational chemistry of prismane<br />

systems. In some cases the calculations performed served as essential additions<br />

to the experimental data and their explanation. In other cases, they led to the<br />

formulation of new experiments and opened the way to new structural motifs.<br />

Special attention is given to the studies related to the theoretical design of new<br />

hypothetical prismanes.<br />

2.3.2<br />

Prismanes<br />

The molecules of C 2nH 2n prismanes, formed by two parallel regular n-gons<br />

connected by rectangular faces, are highly strained cage systems in which the bond<br />

angles differ considerably from the tetrahedral angle of sp 3 -hybridized carbon.<br />

The first three members of the prismane series have so far been synthesized:<br />

triprismane 60 [112], cubane 61 [106, 107] and pentaprismane 62 [113]. Attempts<br />

49


50 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

to synthesize hexaprismane 63 have not yet been successful although certain<br />

progress in this area has been reported [104, 114]. At the same time, according<br />

to ab initio [115], DFT [116] and molecular mechanics [117] calculations, hexaprismane<br />

is one of the least stable members of C 12H 12 family with low probability<br />

of being synthesized [117].<br />

Prismane systems provoked the interest of researchers by their aesthetically<br />

attractive highly symmetrical structure and extraordinary kinetic stability. The<br />

structural and electronic properties of cubane were investigated using various<br />

experimental techniques and quantum chemical calculations and the results<br />

achieved in both directions have been reviewed [103–105, 108–111]. A novel<br />

focus of the recent work is the insight into structural and dynamic properties of<br />

solid cubane [118]. The structural parameters of substituted triprismanes and<br />

substituted pentaprismane were determined by X-ray crystallography and for the<br />

parent hydrocarbons with help of quantum chemical calculations (see [119–122]<br />

and references cited therein). There is a good agreement between the calculated<br />

and experimental results (Figure 2.3).<br />

The higher prismanes remain a subject of only theoretical study [119, 122, 127,<br />

128]. The primary challenges are elucidation of the stability limits of the prismatic<br />

structures with D nh -symmetry and the trends in the geometry and energy characteristics<br />

with increase in n. According to higher level B3LYP/6-311G(2df,p)<br />

computational results [122], the highest member of the [n]prismane family is represented<br />

by decaprismane C 20 H 20 . The principal components of strain energy of<br />

prismanes are angular strain in the n-membered cycles and in the four-membered<br />

faces. The calculated SE of cubane (Figure 2.3) is close to the experimental value<br />

[129] of 157 kcal mol –1 (see, however, discussion in Section 1.3). Penta prismane<br />

62 is characterized by the lowest strain energy and heat of formation (�H f ) and<br />

is the most stable molecule within the family [105, 119, 122, 128]. For the other<br />

members of the series, an increase in n is accompanied by an increase in the<br />

SE and �H f values and correlates with an increase in the number of strained<br />

tetragonal faces. Destabilization of the higher prismanes is caused by influence<br />

of the lateral four-membered cycles as well as by the eclipse of vicinal hydrogens<br />

[119, 122].<br />

A significant contribution to the strain energy of prismanes is their �-antiaromaticity<br />

[130]. Analysis of �-aromaticity of 60–64 by NICS indices shows that cubane<br />

61 may be considered to be a ‘super �-antiaromatic’ system [130]. The high strain<br />

of prismanes determines their low thermodynamic stability: all members of<br />

the family are strongly destabilized relative to other isomeric forms. The main<br />

reason for the high kinetic stability of prismanes is that their rearrangement and<br />

decomposition reactions are forbidden by the rules of conservation of orbital<br />

symmetry. Stability of the D nh-symmetrical prismane structures is determined<br />

by the stabilized orbirtal � �–� � interactions between the parallel n-membered<br />

fragments [131].<br />

The CH bonds in prismanes are shortened because the carbon AO has a high<br />

percent of s-character; this is caused by rehybridization of the sterically strained<br />

carbon center [113]. The increase in ‘n’ leads to weakening of the CH bonds.


2.3 A Theoretical Approach to the Study and Design of Prismane Systems<br />

Figure 2.3 Structural characteristics and strain energies (SE) of prismanes calculated [122]<br />

by the B3LYP/6-311G(2df,p) method. The experimental data obtained by X-ray crystallography<br />

(italic) for the derivatives of 60 [123] and 62 [124] and by electron diffraction (italic) [125] and<br />

microwave spectroscopy (bold italic) [126] for 61. Here and in other figures bond lengths are<br />

given in angström, angles in degrees, strain energies in kcal mol –1 .<br />

51


52 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

Figure 2.4 Calculated [136] and determined experimentally [134, 138] structural characteristics<br />

of nitrocubane 68 and octanitrocubane 69.<br />

According to quantum-chemical calculations [132, 133], substitution of hydrogen<br />

atoms in prismanes by substituents possessing �-donor and �-acceptor properties<br />

(Li, BeH, BH 2 ) provide, as a rule, for the lowering of the strain energy. The<br />

�-donating substituents (such as NH 2 ) exert a similar effect.<br />

As discussed in Section 1.2 nitrocubanes are of great interest as high-energy<br />

systems suitable for the use as explosives and fuels [108]. Many nitrocubanes<br />

including hepta- and octanitro derivatives have been synthesized (see [134]). The<br />

important factor governing the structure and relative stability of nitrocubanes is<br />

the repulsive interaction of the nitro-groups, which resuls in a decrease in stability<br />

upon increase in the degree of replacement [135, 136]. According to RHF and<br />

B3LYP calculations [135], the highest nitrocubanes are characterized by highest<br />

values of SE and �H f . The calculations of octanitrocubane lead to the conclusion<br />

on very low rotation barrier (0.006 kcal mol –1 ) of nitro groups, which can be<br />

described as cooperative disrotatory motion of two subgroups located in the apexes<br />

of the carbon tetra hedrons [136, 137]. The calculated structural parameters of<br />

nitrocubanes agree well with those determined by an X-ray study (Figure 2.4).<br />

2.3.3<br />

Expanded Prismanes<br />

2.3.3.1 Asteranes<br />

The steric strain of prismanes can be lowered considerably by expanding their<br />

cages by introducing additional atomic groups. Examples of such expanded<br />

prismanes are asterane molecules formed by inclusion of CH 2-groups in the<br />

lateral edges of prismanes.


2.3 A Theoretical Approach to the Study and Design of Prismane Systems<br />

Only [3]-asterane 70 [139] and [3]-asterane 71 [140] have been synthesized. Also<br />

bi[4]asterane representing the class of fused compounds (see Section 5.2) has been<br />

preparatively isolated [141]. Like prismanes, asteranes are characterized by high<br />

kinetic stability: for example, 4-asterane is stable at temperature higher than 300 °C<br />

[141]. According to the B3LYP/6-311G(2df,p) calculations [122], only asterane with<br />

n = 3–7 are characterized by stable structure of D nh-symmetry (Figure 2.5). The<br />

calculated geometric parameters of 70 agree well with the data obtained by the<br />

electron diffraction study [142]. The calculated SE values of 70–74 (Figure 2.5)<br />

are notably lower than those for the prismane analogs 60–64. Changes of the<br />

strain energies in the asterane family are similar to the trends which have been<br />

found for the prismane systems. In contrast to the prismanes which experience<br />

the destabilizing influence of the sterically strained lateral four-membered cycles,<br />

the main factor destabilizing higher asteranes is repulsive interaction between<br />

hydrogen atoms of the adjacent methylene groups [122,143].<br />

Figure 2.5 Structural parameters and strain energies of asteranes 70–74 calculated [122] by the<br />

B3LYP/6-311G(2df, p) method. Experimental data for 70 (italic) taken from Ref. [142].<br />

53


54 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

The dehydrogenated asterane frameworks may be used for the formation of<br />

hypothetical hydrocarbon cages like 75, 76 containing hypercoordinated centers<br />

[144–146]. The stability of such systems is governed by 8e rule [131]. The regulation<br />

of charge may occur by variation of basal and bridged atoms and allow the<br />

design of various non-classical structures [145, 146].<br />

2.3.3.2 Ethynyl-expanded Prismanes<br />

Another interesting type of expanded prismane is ethynyl (77, 78) and diethynylexpanded<br />

prismanes (79, 80) of which only a derivative of 80 has been synthesized<br />

[147]. These molecules fall into the category of nonlinear acetylenes discussed in<br />

detail in Chapter 7. Unlike cubane, synthesized diethynyl-expanded cubane is<br />

very unstable. Preparative yield of this compound is extremely low, which made<br />

measurement of its physical properties and calorimetric study difficult or even impracticable.<br />

However, many important characteristics of the expanded prismanes<br />

have been obtained by quantum chemical calculations [148–150].<br />

According to DFT and MP2 calculations [148, 150] of 77 and 78, the inclusion<br />

of ethynyl groups into the C–C bonds of triprismane and cubane provides for<br />

expansion of the valence angles at tetravalent carbon atoms, which values are<br />

close to that of a tetrahedral angle. The flexibility of the diine fragments leads to<br />

the formation of the structures with bent bonds. The strain energies of 77 and 78,<br />

calculated at the MP2/6-31+G* level, are 58.5 and 33.3 kcal mol –1 , respectively,<br />

and are much lower than those of the parent hydrocarbons. Another interesting<br />

feature of the expanded systems is the increase in their acidic properties compared<br />

with parent prismanes; this is caused by the presence of triple bonds enhancing<br />

the acidity of adjacent protons. The further expansion of the prismane framework<br />

(systems 79, 80) results in lowering of free energies of the deprotonation and steric<br />

strain. The structural distortion of the diethynyl fragments leads to a significant<br />

rise of their reactivity and can be considered as one of the factors of the kinetic<br />

instability of preparatively accessible diethynyl-expanded cubane [149]. Expanded


2.3 A Theoretical Approach to the Study and Design of Prismane Systems<br />

prismanes offer a great opportunity to create systems with specific absorption and<br />

interesting optoelectronic properties [148–150]. In contrast to prismanes, whose<br />

small size and high steric strain make the formation of endohedral complexes<br />

difficult [151], the inner cavity of the expanded prismanes allows them to form<br />

complexes with alkaline and alkaline-earth metals and even with small neutral<br />

molecules [148–150].<br />

2.3.4<br />

Dehydroprismanes<br />

An interesting type of extraordinary cage compounds, whose structure is based<br />

on prismanes is prismenes – the compounds of highly pyramidalized olefin class.<br />

The possibility of the existence of 1,2–dehydrocubane (cubene) 81 was discussed<br />

on the basis of ab initio calculations [152, 153] which showed that, in spite of the<br />

presence of the highly pyramidalized double bond, preparative isolation of this<br />

compound is possible. This conclusion was corroborated by the experimental<br />

detection of cubene [154].<br />

The calculated [153] characteristics of 81 (olefin strain energy (OSE) =<br />

58.9 kcal mol –1 ; heat of hydrogenation = 82.5 kcal mol –1 ) are in a good agreement<br />

with experimental [155] data (OSE = 63±4 kcal mol –1 ; heat of hydrogenation<br />

= 90±4 kcal mol –1 ). The presence of the strongly distorted double bond in 81<br />

leads to the increase in the strain energy (227±4 kcal mol –1 ) compared with<br />

cubane. The calculations indicate that there is sufficient overlap of the p-orbitals<br />

to consider cubene as an olefin and not as a biradical. According to experimental<br />

data, exceptionally reactive cubene is capable of reactions typical of olefins [156].<br />

In agreement with expectations, ab initio and DFT calculations [157, 158] testify<br />

for the highest stability of ortho-cubene 81 compared with meta- and para-isomers,<br />

82 and 83. In all three cases the ground singlet states are well separated from<br />

the corresponding lowest lying triplet states. Like 81, 83 was generated upon<br />

treatment of dihalocubanes with organolithium compounds [159, 160] and it has<br />

been suggested that 82 might be preparable by the same type of reaction [157].<br />

According to theoretical and experimental data, 83 does not contain a diagonal<br />

C–C bond, the most probable state of 83 is a singlet biradical [157, 159, 160].<br />

The highly pyramidalized and even more strained dehydrotriprismane (triprismene)<br />

was studied by ab initio methods [161]. Of two possible triprismene isomers<br />

the most stable is the structure 84, which can be considered as an olefin. A tentshaped<br />

isomer 85 is slightly destabilized relative to 84. The calculations indicate<br />

kinetic stability of triprismene and the possibility of its experimental detection.<br />

However, attempts [162] at such detection have so far been unsuccessful.<br />

55


56 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

2.3.5<br />

Polyprismanes<br />

2.3.5.1 Cubane Oligomers<br />

An essential property of cubane derivatives is their tendency to form multidimensional<br />

structures [163–175]. The series of polycubanes have been obtained<br />

by nucleophilic addition to 1,2-dehydrocubane and 1,4-dehydrocubane [164–166].<br />

The simplest system of such type is cubylcubane 86 (Figure 2.6). According to<br />

experimental and theoretical data [165], a remarkable structural feature of 86<br />

is the unusually short C–C bond linking two cubane fragments; such systems<br />

are discussed later in Section 2.6. The high s-character of the exocyclic orbitals<br />

of cubane leads to the formation of more compact bonds compared with those<br />

formed by sp 3 -hybridized carbon atoms.<br />

The p-[n]cubyl oligomers are sufficiently rigid rod-type systems (analogous to<br />

staffands presented in Section 2.1.3.3) retaining geometrical parameters of the cubane<br />

fragments since stretching or angle distortions of the intercage bonds require<br />

significant energy costs [164]. According to PWSCF [168] and B3LYP/6-311G**<br />

[169] calculations on the polycubane one-dimensional chain and two-dimensional<br />

network, such properties as ionization potential and electron affinity are only slightly<br />

dependent on the number of building blocks involved. Cubane chains maintain<br />

the electronic properties of insulator; at the same time they have very interesting<br />

elastic qualities and are of interest in the development of electronic nanodevices<br />

[168]. As shown by DFT/PW91 calculations, inclusion of electron-enriched functional<br />

groups (NO, NS) in cubane chains favors electronic conduction [170].<br />

An example of a three-dimensional polymeric cubane derivative is supercubane<br />

(C 8) n, proposed as an allotrope of carbon [171–175]. The simplest non-periodic<br />

supercubane model is percubylcubane 87 (Figure 2.6) which was studied by<br />

Figure 2.6 Calculated structural parameters of cubylcubane 86 and percubylcubane 87.<br />

Experimental data for 86 taken from Ref. [165].


2.3 A Theoretical Approach to the Study and Design of Prismane Systems<br />

HF/STO-3G [174] and PM5 [175] calculations. New structures termed cubanoids<br />

were computationally designed on the basis of percubylcubane [175].<br />

2.3.5.2 Fused Prismanes<br />

The formation of fused systems joined at n-membered faces serves as another<br />

way for the formation of polymeric prismane compounds, which may be termed<br />

poly[n]prismanes [176–179] (such as bi[n]prismanes 88–91). The unusual structural<br />

feature of poly[n]prismanes is the presence of carbon centers with nonclassic<br />

bisphenoidal configuration of bonds.<br />

Bi[n] and tri[n]prismanes (n = 3–6) have been studied using DFT calculations<br />

[177]. Although poly[n]prismanes are strongly destabilized with respect to their<br />

‘classic’ isomeric forms, calculations indicate their kinetic stability and possible<br />

experimental detection. The origin of their relatively high stability is to be found in<br />

the strong � � –� � orbital interaction between the symmetry adapted frontier MOs<br />

of the external annulene (CH) n and the inner C n rings. Increase in the number of<br />

fused cages leads to increase in strain compared with prismanes. Maximal steric<br />

strain is observed for the first members (n = 3, 4) of each family and for each n the<br />

strain increases with increase in the size of the structure. The decrease in stability<br />

of poly[n]prismanes on passing from prismanes to their bi, and tri-congeners<br />

also manifests itself in progressive decrease of the energy gap between frontier<br />

orbitals along this set of compounds. The increase in the number of fused cages<br />

is accompanied by the further lowering of the stability [178]. The interesting<br />

characteristic of poly[n]prismanes is their auxetic property, i.e. the ability of these<br />

molecules to become thicker by longitudinally stretching and thinner by compression.<br />

According to quantum-chemical calculations, poly[n]-prismanes are the first<br />

systems that manifest auxetic behavior at the molecular level [179].<br />

‘Mixed’ polycage prismane derivatives 92–94 containing pentaprismane and<br />

dodecahedron cages fused by five-membered faces were proposed on the basis<br />

57


58 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

of B3LYP/6-31G* calculations [180–182]. Increase in the number of fused cages<br />

leads to increase in strain and decrease in stability of these systems compared<br />

with pentaprismane and dodecahedrane. The possible experimental preparation<br />

of fused prismanes is supported by successful synthesis of bi[4]asterane 95 and its<br />

substituted derivatives [141, 183]. In contrast with [4]asterane 71 which is stable at<br />

temperatures over 300 °C, the transformation of 95 occurs at a lower temperature<br />

(270 °C) [141] confirming the theoretical conclusions about the decrease of stability<br />

of fused systems upon increase in the number of building blocks.<br />

It is also possible to form fused polyprismane structures by fusing fourmembered<br />

faces of the cage systems. In contrast with cape-fused prismanes<br />

containing nonclassical bisphenoidal carbon centers, the face-fused prismanes are<br />

characterized by the presence of carbon atoms with nonclassic inverted (umbrella)<br />

configuration of bonds. Such molecules called propellanes are discussed in detail<br />

in Section 2.1. The existence of bipentaprismane 96 was predicted using ab initio<br />

and DFT calculations [184, 185]. Even more unusual example of face-fused poly[5]<br />

prismanes 97 is a hypothetical product of multiple pentaprismane splicing, in<br />

which the [2+2] cycloreversions of cyclobutane rings, leading to the series of<br />

isomers with pairs of parallel �-bonds interconverting by Cope rearrangement<br />

may be realized [186]. By analogy with ‘mixed’ cape-fused polyprismanes the<br />

‘mixed’ face-fused systems can be formed. The stable system 98 investigated [187]<br />

by B3LYP/6-31G* method is derived from cubane and two pentaprismane cages<br />

fused by two four-membered faces. Based on the results of the computational<br />

studies of the fused prismanes, one can conclude that an increase in the number<br />

of fused cages is accompanied by increase in steric strain, decrease in lowest<br />

vibration frequencies and HOMO-LUMO energy gap. All these effects lead to<br />

lowering in the stability of designed complicated systems.<br />

2.3.6<br />

Conclusions<br />

The comparison of theoretical and experimental data indicates that theoretical<br />

methods can be a reliable tool for comprehensive investigation and design of<br />

various molecules constructed on the basis of prismanes. Prismane systems,<br />

which in themselves are unique molecules with nonstandard stereochemistry and<br />

properties, can form series of novel unusual systems such as dehydroprismanes<br />

representing a class of pyramidalized olefins; supercubane as an allotrope of<br />

carbon; fused prismanes, including nonclassical bisphenoidal and inverted carbon


2.4 (CH) 2n Cage Structures, ‘in’-‘out’ Isomerism in Perhydrogenated Fullerenes<br />

centers, as well as asterane derivatives, containing hypercoordinated centers, etc.<br />

Various modifications of prismanes lead to the appearance of new intriguing properties.<br />

Poly[n]prismanes are expected to show auxetic behavior; ethynyl-expanded<br />

prismanes can reveal useful optoelectronic characteristics; nitro-prismanes are<br />

good candidates for powerful, high-density explosives; prismanes networks are<br />

of interest in the development of nanoarchitectures.<br />

The variety of the systems generated on the basis of prismanes is not exhausted<br />

(for example, an extensive class of heteroprismanes [128]) and requires further<br />

studies, which open the way to create new materials with unusual properties.<br />

Although the majority of reported studies correspond to cubane derivatives,<br />

available theoretical data reveal similar tendencies in behavior and properties of<br />

other members of the family. The principles of formation of the lowest prismane<br />

derivatives can be extrapolated to the area of more complex compounds. Synthesis<br />

of the molecules considered in this chapter seems to be a great challenge for<br />

chemists and the success in the field of preparation of numerous cubane derivatives<br />

(in particular, of octanitrocubane [138]) is an impressive manifestation of<br />

progress and rich opportunities of modern experimental techniques. The data<br />

presented show that many hypothetical systems can be suitable objects for the<br />

synthetic quest and we believe that results of quantum-chemical research help to<br />

invigorate new experimental efforts in this field.<br />

Acknowledgments<br />

This work was supported by Russian Foundation for Basic Research (grant<br />

07-03-00223), RF Ministry of Industry and Science (grant 4849.2006.3).<br />

2.4<br />

(CH) 2n Cage Structures, ‘in’-‘out’ Isomerism in Perhydrogenated Fullerenes and<br />

Planar Cyclohexane Rings<br />

Helena Dodziuk<br />

2.4.1<br />

(CH) 2n Cage Structures<br />

Studies pertaining to saturated cage compounds of a general formula (CH 2 ) n can<br />

be traced back to the ancient Greeks since the ideal polyhedra: the tetrahedron, the<br />

cube and the dodecahedron, which represent the carbon skeletons of the molecules<br />

with n = 2, 4 and 10, were discussed by Plato and Pythagoras [188]. The molecules<br />

are formed by 2n carbon atoms each of which is connected to three carbon and one<br />

hydrogen atoms. For n = 2 and 3 there is only one possible isomer: tetrahedrane<br />

99a and triprismane 60, respectively, while for larger n the number of isomers<br />

increases. Cubane 61 (discussed in detail in Section 2.3), cuneane 100 and octabis-<br />

59


60 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

valene 101 are the three isomers for n = 4 and the number of possible structures<br />

increases rapidly with the increase of n; for instance, it is equal to 9 for n = 5, 32<br />

for n = 6, etc. Of the higher members of the series only pentaprismane 62, and<br />

diademane 102 with n = 5, heptacyclo[6.4.0.0 2,4 .0 3,7 .0 5,12 .0 6,10 .0 9,11 ]dodecane 103<br />

(n = 6), dodecahedrane 104 (n = 10) and a derivative of perhydrogenated fullerene<br />

C 60 X 60 105 (n = 30, X = F) have been synthesized. These molecules are highly<br />

strained hydrocarbons exhibiting, among other, nontypical rearrangement [189,<br />

190] (Figure 2.7) and cyclization reactions [191]. The rigid structure of saturated<br />

cage compounds means that they can serve as ideal models in studies both of the<br />

angular dependence of the coupling constants, and of dependence of J( 13 C–H) on<br />

the hybridization (see Section 2.6).<br />

Figure 2.7 Untypical rearrangement reactions of cubane into cuneane (a) and pagodane into<br />

dodecahedrane (b).


2.4 (CH) 2n Cage Structures, ‘in’-‘out’ Isomerism in Perhydrogenated Fullerenes<br />

61<br />

2.4.1.1 Tetrahedrane<br />

The elegant molecule tetrahedrane, 99a C 4 H 4 (n = 2), has attracted the attention<br />

of synthetic and theoretical chemists for almost 100 years [192, 193] and the<br />

first review on this molecule [194] was published almost at the same time as the<br />

synthesis of its tetra-t-butyl derivative 99b [195]. Then for a long time, the latter<br />

molecule was the only tetrahedrane derivative known [196]. A tetrasiladerivative<br />

99c and other siladerivatives with much smaller substituents 106 (X = Li, CH 3 ,<br />

H, C(SiMe 3 ) 3 ) as well as tetrahedryltetrahedrane 107 (X = SiMe 3 ) [197] followed.<br />

Syntheses of these molecules have recently been reviewed by de Meijere and<br />

coworkers [198]. Physicochemical studies of 99b included X-ray [199], spectral [200]<br />

and NMR [201] measurements. Its high kinetic stability was partly ascribed by<br />

Minkin [202] to unfavorable steric repulsions in its rearrangement product 108b.


62 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

Other theoretical studies ranged from molecular mechanics, MM, [203] prediction<br />

of T symmetry of 99b [204] to high level quantum chemical, QC, calculations<br />

indicating that 99a should be isolable [205]. However, unlike cyclobutadiene 108a<br />

[206], discussed in Chapter 10, unsubstituted tetrahedrane remains unknown.<br />

The NMR coupling constants in 99a, 60, 61, 62, 102 and hexaprismane 63 were<br />

calculated by Krivdin [207], while of several calculations analyzing reasons which<br />

cause shortening of the linking C–C bond in tetrahedranyltetrahedrane 107<br />

(Y = H), recent work by Mo should be mentioned [208].<br />

2.4.1.2 Triprismane<br />

Triprismane 60 C 6 H 6 (n = 3) [209], also called [3]prismane, is discussed in<br />

Section 2.3 together with other prismanes.<br />

2.4.1.3 Cubane 61, Cuneane 100 and Octabisvalene 101 C 8H 8<br />

Cubane 61 is discussed mainly from theoretical point of view in Section 2.3, so here<br />

only few aspects of its structure and properties will be covered. This highly strained<br />

molecule is remarkably stable since it corresponds to a very deep narrow minimum<br />

on the potential energy surface. Its chemistry, physicochemical properties and<br />

theoretical studies have been reviewed by Griffin and Marchand [210] and by Hassenrück<br />

and coworkers [211]. Its high symmetry allowed the Hedberg group [212]<br />

to compare the accuracy of electron diffraction [213], microwave [214] and X-ray<br />

[215] determinations of C–C bond length. The unusual skeletal rearrangement of<br />

61 into 100 (Figure 2.7) was mentioned above. An interesting model of cubane<br />

hydrogenolysis was discussed by Stober et al. [216]. Several interesting cubane<br />

derivatives are discussed in Section 2.3. They include known propella[3 4 ]prismane<br />

109 [217] and cubene 81 [218] (which lies outside the scope of this chapter but was<br />

presented in the former section). The trimethylene bridges in the former molecule<br />

are very flexible leading to only three signals in 13 C NMR spectrum [217].


2.4 (CH) 2n Cage Structures, ‘in’-‘out’ Isomerism in Perhydrogenated Fullerenes<br />

63<br />

Cuneane 100 and octabisvalene 101 have been much less studied [189, 211,<br />

219, 220] and the existence of inverted carbon atoms (discussed in Section 2.1)<br />

in the latter molecule, which can be considered as a bicyclobutane dimer, has<br />

been mostly overlooked [219]. The X-ray structures of octamethylcubane and<br />

octamethylcuneane have been determined by Irngartinger [221]. A large group<br />

recently carried out a calorimetric, crystallographic and computational study of<br />

61, 100 and their carboxylates [222].<br />

2.4.1.4 C 10 H 10 Saturated Cages<br />

Pentaprismane 62 [223] and diademane 102 [224, 225] are the only known members<br />

of this family. For the former molecule, only photoelectron spectra [226] and X-ray<br />

analysis [227] are known. For the latter, several centrally bridgehead-substituted<br />

derivatives have been obtained recently [228]. According to semiempirical [229]<br />

and ab initio [230] calculations, the cyclohexane ring in 102 should be planar, as<br />

are such rings in other molecules having the cis-tris-�-homobenzene fragment 110<br />

presented later in this chapter. Chemistry of 102 and some of its derivatives and<br />

heteroanalogs has recently been discussed by de Meijere and coworkers [198].<br />

2.4.1.5 C 12 H 12 Saturated Cages<br />

The highly symmetrical hexaprismane 63 and truncated tetrahedrane 111 have been<br />

long pursued but, contrary to the common practice not to publish unsuccessful<br />

attempts, only failed syntheses of them or larger such molecules have been reported<br />

[231–234]. The synthesis of a derivative of heptacyclo[6.4.0.0 2,4 .0 3,7 .0 5,12 .0 6,10 .0 9,11 ]<br />

dodecane 103 went unnoticed [235] as there was a considerable puckering of the<br />

central cyclohexane ring in the parent compound when it was synthesized by de<br />

Meijere’s group [236]. A more detailed study of the reactivity of 103 was recently<br />

published by the same group [237] and its chemistry reviewed [198]. Several<br />

members of this family have been analyzed using MM [203] model calculations<br />

yielding 103 as the most stable and hexaprismane 63 as the least stable molecules<br />

among those studied. Both the MM study and QC calculations [238] for 111<br />

yielded a highly symmetrical T d structure and close values of heat of formation<br />

(ca 91 kcal mol –1 by QC [239] and ca 87 kcal mol –1 by the MM method [240]).


64 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

According to MM modeling [240] (in the QC studies symmetry constraints were<br />

imposed), 63 and 111 are characterized by planar C 6 rings [240] while the central<br />

cyclohexane ring in 103 was found to be considerably more puckered than in the<br />

standard cyclohexane structure, first on the basis of modeling [219, 240] and then<br />

experimentally [237].<br />

Of several theoretical and experimental studies discussing preparation of<br />

hexaprismanes, papers from the Mehta group aiming at the synthesis of [6]- and<br />

[7]prismanes [241–243] and the paper by Chou [244] should be mentioned. In<br />

spite of this massive effort, obtaining hexaprismane 63 remains a great unsolved<br />

problem in synthetic organic chemistry [245]. Maybe reactions in molecular flasks,<br />

as presented in Chapter 10, will enable us to obtain 63.<br />

2.4.1.6 Higher [n]Prismanes, Dodecahedrane<br />

Hypothetical higher [n]prismanes were discussed in Section 2.3.<br />

Of still higher prismanes, israelane 112 and helvetane 113 can be mentioned. Following<br />

a humorous idea of Eschenmoser, these C 24 H 24 molecules were discussed<br />

by Ginsburg in the 1st April issue of 1982 of the Nouveau Journal de Chimie [246].<br />

The family to which they belong consists of numerous molecules and 112 and 113<br />

would definitely be of very high energy, precluding their existence. Thus, some<br />

calculations of their structure and energy discussed in [247] seem insignificant.<br />

It took the Paquette group almost 20 years to synthesize dodecahedrane C 20 H 20<br />

104 [248]. It was the first molecule of the I h symmetry synthesized in a laboratory,<br />

contrary to the Herzberg prediction that such molecules would be unlikely to exist<br />

[249]. It is the only known member of the family under scrutiny. Dodecahedrane<br />

chemistry has been discussed in detail [250] but its physicochemical properties<br />

received little attention.<br />

2.4.1.7 ‘In’-‘out’ Isomerism in Perhydrogenated Fullerenes C 60 H 60<br />

As fullerenes are discussed at length in Chapter 5, only the exciting ‘in’-‘out’<br />

isomerism in hypothetical fully hydrogenated fullerenes C 60 H 60 (named fullerane)<br />

proposed by Saunders [251] will be considered here. On the basis of MM calculations,<br />

he found that of all possible ‘in’ and ‘out’ isomers, a nonsymmetrical one<br />

with 10 C–H bonds pointing inside should be the most stable. ‘In’-‘out’ isomerism<br />

in hydrogenated fullerenes has also been studied using the MM [252] and/or QC<br />

[252–254] methods. Two isomers with two C–H bonds pointing inside (other C–H<br />

bonds are omitted for the sake of clarity) are shown in Figure 2.8. Let us note that,


2.4 (CH) 2n Cage Structures, ‘in’-‘out’ Isomerism in Perhydrogenated Fullerenes<br />

65<br />

Figure 2.8 Two geometrical isomers of C 60 H 60 with two CH bonds pointing inside.<br />

as shown in the latter figure, the carbon cage in the in-isomers undergoes considerable<br />

distortions and such isomers have a lower symmetry. It should be stressed that<br />

for higher number n of bonds pointing inside a real challenge in the calculations<br />

was to generate sufficiently large set of the ‘in’ structures to be sure that the one<br />

of the lowest energy for the given n represents the lowest energy structure.<br />

Dodziuk and Nowinski noticed that two isomers with a C–H bond pointing<br />

inside and outside the cage are topological isomers [255] since in order to come<br />

from one of them to another, one has to break C–C bonds of the C 60 cage. Dodziuk<br />

also studied the ‘in’-‘out’isomerism in C 60 H 60 using a different procedure for<br />

generation of isomers than that applied by Saunders [251] and different parameterization<br />

and found n = 10 CH bonds pointing inside the C 60 cage for the<br />

lowest energy structure. The latter result agreed with that of Saunders, however,<br />

as shown on the right of Figure 2.9, the optimum structure for n = 10 was of high<br />

symmetry. Comparison of the results by Saunders and those by Dodziuk group<br />

shows the limitations of the applicability of the simple MM model and only the<br />

results that do not depend on the force field used are reliable. Dodziuk group<br />

has also shown that not only several C–H bonds pointing inside the cage lower<br />

its energy but there is sufficient space in the cage for a methyl, ethyl or n-propyl<br />

substituent inserted inside and these isomers are considerably more stable than<br />

their fully ‘out’ counterparts. For iso-propyl and n-butyl substituents the energy of<br />

‘in’ and ‘out’ isomers assumed comparable values while for more bulky substituents<br />

the ‘all-out’ isomer was more stable. Moreover, two methyl groups situated<br />

on the opposite ends of the cage in 1,51-dimethyl fullerane (see Figure 2.10 for<br />

computer-generated atom numbering) were also calculated to be more stable than<br />

their isomers with one or both methyl groups pointing out [256]. Interestingly, not<br />

only was a methyl group ‘in’ more stable than that with the group ‘out’ but also<br />

its energy, independently of the force field applied, was very close to the energy<br />

of the C 60H 60 with one hydrogen ‘in’ [256].<br />

Saunders [251] and Yoshida [252] groups claimed that one could obtain ‘in’<br />

isomers of hydrogenated fullerenes by heating or inversion on a protonated carbon<br />

atom while Dodziuk group postulated that, in view of high barrier that must be<br />

overcome to move a hydrogen atom from an outside to an inside position, the<br />

aimed synthesis must be carried out to obtain the ‘in’ isomers.


66 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

Figure 2.9 Two projections of the calculated structures of ‘all-out’ C 60 H 60 (left) and the most<br />

stable ‘ten-in’ isomer (right). For clarity only carbon and ‘in’ hydrogen atoms are shown.<br />

The projections along a C 5 symmetry axis are shown in the upper part while the lower ones<br />

are along one of the C 6 axes.<br />

Figure 2.10 Atom numbering in C 60 determined by Rose (as cited in Kroto [258]).


2.4 (CH) 2n Cage Structures, ‘in’-‘out’ Isomerism in Perhydrogenated Fullerenes<br />

67<br />

As discussed in Chapter 5, in the early days of fullerene chemistry numerous,<br />

sometimes peculiar, applications of this molecule were proposed [257]. The use<br />

of C 60F 60 as an ideal lubricant was one of them. The molecule was synthesized<br />

in all-‘out’ form but proved highly unstable when exposed to air, producing HF<br />

as a decomposition product [258]. The instability of C 60F 60 can be, at least partly,<br />

related to the high strain found in its hydrogenated analog [251, 259].<br />

2.4.1.8 Summary<br />

As discussed in Chapter 1, it has become fashionable today to show practical applications<br />

of the molecules under scrutiny. Let us repeat once more, Eaton tried<br />

to obtain cubane because synthesis of this attractive highly symmetrical molecule<br />

was a great challenge. However, in agreement with the above-mentioned trend,<br />

two interesting applications will be mentioned here. Cage compounds discussed<br />

in this chapter are too small to host other molecules. However, the possibility that<br />

they could play the role of host for atomic hydrogen or ions has been explored<br />

theoretically [260] and the use of a cubane derivative has been proposed in a drug<br />

delivery system [261].<br />

2.4.2<br />

Planar Cyclohexane Rings<br />

Let us recognize that the statement ‘a cyclohexane ring can be planar’ can cost a<br />

student a failure of examination [247]. However, molecules 114–116 and some<br />

complexes of 117 with cis-fusion of C 3 rings have been proven to exhibit such<br />

structures [262–265]. As discussed in the former chapter, diademane 102 [224,<br />

225], hypothetical hexaprismane 63 and truncated tetrahedrane 111 are predicted<br />

to have planar rings [240]. Cyclophanes 118 and 119 synthesized by the de Meijere<br />

group [266, 267] should also possess the planar C 6 ring. Similarly, according to<br />

model calculations a hypothetical hexahydrosuperphane 120 [268] should also<br />

have such a ring. As will be briefly discussed below, planarization can result<br />

either from fusion of at least three three-membered rings to the saturated sixmembered<br />

one [269] or it can be forced by the cage structure in which such a<br />

ring is built [240, 268].<br />

As discussed in detail in Stereochemistry of Organic Compounds [270] the first<br />

notion of chair and boat conformations of the ring stemming from the tetra hedral<br />

arrangement of substituents on a tetravalent carbon atom were expressed by<br />

Sachse as early as in 1890 [271]. For long time physicochemical methods were too<br />

crude to enable the observation of nonplanarity of these conformations. Although<br />

several experimental results pointed to the chair shape of the six-membered ring,<br />

only a pioneering paper by Barton [272] brought a full comprehension of the<br />

physical and chemical consequences of such a structure. For this and subsequent<br />

studies he was awarded a Nobel Prize for Chemistry in 1969.<br />

The ring inversion leading to axial–equatorial equilibrium of substituents in<br />

monosubstituted cyclohexanes is a rapid process (of ca. 2 � 10 5 s –1 ) the observation<br />

of which depends on the timescale of experimental technique applied. For


68 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

instance, at room temperature two bands corresponding to the stretching C–Br<br />

vibrations of bromocyclohexane are observed in the IR spectra [273] while only<br />

one, average signal can be seen in the corresponding spectra of the hydrogen<br />

attached to the bromine substituted carbon in the NMR spectrum. This signal is<br />

split into two at about –50 °C when the ring inversion is frozen [274, 275]. Thus,<br />

in many cases only a combination of different experimental techniques can give<br />

a real insight into the molecular structure.<br />

Cyclohexanes with flattened rings have been known for long time [276–278]<br />

but, as mentioned earlier, the existence of molecules with a planar saturated C 6<br />

ring is not fully recognized even today. As concerns hydrocarbons, according to<br />

Engelhardt and Lüttke [279] trans-tris-�-homobenzene 110a should have such a<br />

ring, although no sound argument in favor of the planarity has been given. On<br />

the other hand, a close to planar C6 ring in the hexamethyl derivative of 117 [280]<br />

substantiated this claim. A historical account of the studies of tris-�-homobenzene<br />

has been published by Rücker [281]. A planar cyclohexane ring was proposed for<br />

known [224, 225] diademane 102 on the basis of computer modeling [229, 230]<br />

and, as discussed in Section 2.3, for hypothetical [6]asterane 121 [203]. A planar<br />

central ring was also found in the X-ray study of the tribenzoderivative of tetracyclo[8.2.0.0<br />

2,5 0 6,9 ]dodecane 122a [282] but on the basis of model calculations<br />

[269] discussed below, the planarity cannot be due to the effect of fusion of fourmembered<br />

rings to a cyclohexane ring. It should rather be ascribed to crystal<br />

forces, in particular those operating among the stacked aromatic rings. A recent<br />

determination of the nonplanarity of the central ring in 122b supports the latter<br />

conclusion. On the other hand, the central ring in 123 and 124 has been proven<br />

to be planar [283].<br />

The influence of a cis fusion of with a smaller ring on the planarization of<br />

the C 6 ring was systematically studied by Dodziuk [269] using the MM method<br />

by changing the number from 1 to 6 and size of the fused ring(s) 110, 125–136<br />

(some of them, like 132, 134 and 136 purely hypothetical). A detailed comparison<br />

of the calculated geometry with the experimental one proved difficult since<br />

X-ray analysis is mainly carried out, often for a derivative of the molecule under


2.4 (CH) 2n Cage Structures, ‘in’-‘out’ Isomerism in Perhydrogenated Fullerenes<br />

69<br />

scruting, in the solid state while the calculations refer to the isolated molecule.<br />

Moreover, the crystals sometimes contain solvent molecules influencing the<br />

results obtained for the molecule under investigation. However, the comparison<br />

of the calculated and available experimental data showed that the primitive MM<br />

model used in 1987 reproduced semiquantitatively torsional angles allowing one<br />

to analyze the planarity of the cyclohexane ring. The calculations led to the conclusion<br />

that fusion of smaller and/or more rings causes larger distortion of the C 6<br />

ring toward planarity. In consequence, cis- and trans- tris-�-homobenzenes 110<br />

should have a planar central ring. A comparison of the results of calculations for<br />

130 [269, 284], that is the saturated core of 122a and 122b, with the experimental<br />

data for the latter molecules indicate that, as mentioned earlier, the planarity<br />

of the central ring in the 122a is forced by intermolecular forces present in the<br />

crystalline state. Similarly, on the basis of the same type of modeling, planar rings<br />

should be present in 110a [269], 63, and 111 [240]. For 102, the same conclusion<br />

was drawn from semiempirical [229] and ab initio [230] QC. The planarity of the<br />

C 6 ring built into a cyclophane cage in 120 was obtained on the basis of both ab<br />

initio and DFT calculations [268].<br />

Interestingly, both factors forcing planarity are present in cyclophanes 118 and<br />

119 synthesized by the de Meijere group [228]. Thus, the molecules should have<br />

a planar cyclohexane ring although no proof of this has been reported.


70 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

2.5<br />

Ultralong C–C Bonds<br />

Takanori Suzuki, Takashi Takeda, Hidetoshi Kawai and Kenshu Fujiwara<br />

2.5.1<br />

Introduction<br />

The bond length between the two sp 3 carbons is one of the fundamental parameters<br />

in chemistry. However, there are cases [285] where the bond length d significantly<br />

exceeds the standard value of 1.54 Å [286], as in sterically congested molecules such<br />

as t-Bu 3 CH [1.611(5) Å] [287]. Furthermore, it has been experimentally verified<br />

that a small number of compounds possesses the ultralong C–C bond (d > 1.70 Å)<br />

[288–290]. This chapter describes how the Csp 3 –Csp 3 bond can be elongated to<br />

such an extreme length and predicts a limit for the ultralong C–C bond.<br />

The most popular method to precisely determine the values of the length<br />

in crystalline materials is single-crystal X-ray analysis, although Raman [289]<br />

and 13 C NMR spectra [291] were also found effective in some cases. Diffraction<br />

measurement is often carried out at low temperature to attain high accuracy by<br />

suppressing thermal motion of atoms in crystal. However, the X-ray methodology<br />

dose not always give the right answer. It has been pointed out [292] that some of<br />

the ultralong C–C bonds in the earlier reports [293] were obtained as the result<br />

of unintentional overestimation by crystallographic artifacts, caused by incorporation<br />

of the valence isomer with a large separation of atoms in the crystal<br />

(Scheme 2.10a). In another case, the length was determined to be much smaller


Scheme 2.10<br />

2.5 Ultralong C–C Bonds<br />

than the actual d value due to positional disorder [294] or internal motion in the<br />

crystal [295] (Scheme 2.10b). Despite the opposite outcome, the thermal ellipsoids<br />

for the atoms in question are mis-shaped to indicate the presence of anomalies<br />

in both cases. Another common feature is the large estimated standard deviation<br />

(esd) for the bond length and other structural parameters. To avoid being misled<br />

by these crystallographic artifacts, only X-ray structures with a suitable accuracy<br />

(esd for d < 0.01 Å) are considered in this chapter [296].<br />

The representative ultralong bonds can be found in Toda’s naphthocyclobutene<br />

derivatives 137 [288] [1.712(5)–1.734(5) Å] and the benzannulated caged hydrocarbon<br />

138 [1.713(2) Å] reported by Herges [289]. Before recent results by the authors<br />

were reported [290], diiodonaphthocyclobutene 137-I [1.734(5) Å] had been considered<br />

as the world record holder in terms of the length of a conventional C–C<br />

bond [297]. According to the linear correlation between bond dissociation energy<br />

(BDE) and d proposed by Zavitsas (d/Å = 1.748–0.002371 � BDE/kcal mol –1 ) [298],<br />

the experimentally determined values for 137 and 138 are just at the edge of a<br />

maximum bond length limit of 1.75 Å. In the next section, the pioneering work<br />

on these compounds is described to show how the syntheses of these elegant<br />

molecules with such a ultralong C–C bond were successfully realized.<br />

71


72 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

2.5.2<br />

Ultralong C–C Bonds Confined in a Stiff Molecular Frame<br />

Tricyclo[4.2.2.2 2,5 ]dodeca-1,3,5,7,9,11-hexaene 139 is the �-bond shift isomer of<br />

[0.0]paracyclophane 139�. This hydrocarbon is highly reactive because it has a<br />

strained molecular framework with the bridgehead unsaturated bonds [299]. Its<br />

tetrabenzannulated derivative 140 is much more stable than 139 [300], yet it still<br />

can undergo thermal [2+2] cycloaddition with benzyne. In this way, Herges and<br />

co-workers synthesized the caged hydrocarbon 138 [289], in which the ultralong<br />

bond in question faces the strongly pyramidalized double bond (Scheme 2.11).<br />

X-ray analysis at –130 °C revealed that 138 has one of the longest C–C bonds of<br />

1.713(2) Å among hydrocarbons. The extremely large value of d is not related to<br />

a contribution from the bond-dissociated isomer 141 because the stiff molecular<br />

frame prevents symmetry-allowed conrotatory ring opening into the o-quinodimethane<br />

derivative. The presence of an ultralong C–C bond in 138 was also<br />

predicted theoretically (B3LYP/6-31G*), where the calculated value of d = 1.721 Å is<br />

comparable to the experimental value [289]. Even when the calculation started with<br />

the geometry of ring-opened isomer 141, the high-level optimization converged<br />

into the structure of ring-closed form 138, since the structure of 141 has been not<br />

found to be a stationary point at the DFT level.<br />

Scheme 2.11<br />

In the Raman spectrum of 138, the C–C stretching vibration was observed at a<br />

lower wavelength (687 cm –1 ) than the standard value (995 cm –1 for ethane). This<br />

observation is in accordance with the extraordinary bond expansion that causes<br />

considerable decrease in strength of the C–C bond. Its force constant (133.1 N m –1 )<br />

was calculated as less than one-third of that for ethane (455.7 N m –1 ). Successful<br />

isolation and characterization of this caged hydrocarbon show that the stiff<br />

molecular framework effectively maintains the weakened ultralong C–C bond<br />

in 138.


2.5 Ultralong C–C Bonds<br />

2.5.3<br />

Tetraphenylnaphthocyclobutene as a Scaffold to Produce Ultralong C–C Bonds<br />

Based on their own discovery that thermal cyclization of diallenes proceeds<br />

smoothly in the solid state [301], Toda, Tanaka, and co-workers prepared a variety<br />

of cyclobutene-fused aromatic compounds from 1,2-diallenylarenes [288d],<br />

some of which have an ultralong C–C bonds. Among them, 3,8-dichloro-1,1,2,2tetraphenylcyclobuta[b]naphthalene<br />

137-Cl [1.720(4) Å at 25 °C] [288a] is the most<br />

well-studied species, whose structural analyses [1.710(2) Å at –183 °C] [288b] and<br />

high-level calculations [302] were also carried out by several research groups. All<br />

the results support the intrinsic nature of the ultralong bond of 137-Cl.<br />

The origin of the bond elongation is primarily explained by a classical steric<br />

argument without the need to resort to orbital interaction. The (�-�*)-type<br />

through-bond orbital interaction had previously been proposed as a plausible<br />

reason to account for the bond elongation [303], yet several recent experimental<br />

studies did not support its importance [304]. The through-bond orbital interaction<br />

[305] is mainly caused by (�-�*)-type filled–filled coupling [306], and perturbation<br />

through the (�-�*)-type coupling was proven to be negligible according to highlevel<br />

theoretical calculation [307].<br />

Especially indicative that the (�-�*)-type coupling is unimportant are the values<br />

of d experimentally determined in a series of cyclobutaarenes [288c, 288d]. For<br />

example, the X-ray structural analyses were carried out on two isomeric compounds<br />

142, in both of which the two t-butyl groups are substituted for the phenyls in<br />

137-Cl. The d value in gem-di(t-butyl) derivative gem-142 [1.729(2) Å] is much<br />

larger than the corresponding value in trans-1,2-di(t-butyl) derivative trans-142<br />

[1.686(5) Å], although the latter must be quite suitable for the (�-�*)-type coupling.<br />

The larger d in the former molecule is accounted for mainly by the more severe<br />

steric repulsion in gem-142 than in trans-142, which also induces outward deviation<br />

of bonding electron-density on the molecular plane around the carbon with the<br />

two t-Bu groups attached.<br />

Compounds 137 and 138 with an ultralong C–C bond have a 1,1,2,2-tetraaryl<br />

benzocyclobutene skeleton in common, which can be regarded as a sort of<br />

‘condensed’ derivative [308] of hexaphenylethane (HPE). The parent hydrocarbon<br />

143 was shown to undergo easy interconversion into 7,7,8,8,-tetraphenylo-quinodimethane<br />

144, which is too labile to be isolated, as it undergoes further<br />

73


74 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

Scheme 2.12<br />

pericyclic reaction to dihydroanthracene (Scheme 2.12a) [309]. The successfully<br />

isolated dibenzo analog of o-quinodimethane 146 [310] exhibits no signs of conversion<br />

into the cyclobutene-type isomer 145 (Scheme 2.12b) probably because<br />

of the easy cleavage of the weakened bond in 145, if formed. Thus, the successful<br />

observation of ultralong C–C bond in 137 and 138 relies on the suppression of their<br />

ring-opening pericyclic reaction to o-quinodimethanes: its formation is disfavored<br />

by the delocalization energy loss for the 2,3-naphthoquinodimethane moiety in<br />

137 and by the stiff molecular framework in 138. In the following sections, other<br />

members of HPE-type molecules are described which have a longer ‘ethane’ bond<br />

than the standard.<br />

2.5.4<br />

‘Clumped’ Hexaphenylethane Derivatives with Elongated and Ultralong C–C Bonds<br />

Since the pioneering work by Mislow [291, 311], HPEs with three bulky aryl groups<br />

at both ends of ethane bond have been used as excellent models to observe the<br />

elongated C–C bond because the steric repulsion between substituents is the<br />

most important and powerful factor to expand the C–C bond. One of the major<br />

obstacles in studying HPEs is the bond-dissociation equilibrium generating trityl<br />

radicals, which is also related to �,p-dimer formation or high reactivity toward<br />

oxygen [312].When the two trityls are connected by a suitable bridge, intramolecular<br />

bond-formation process giving �,�-dimer (HPE) becomes much more facile<br />

(Scheme 2.13a), thus endowing the ‘cross-clumped’ HPEs such as 147-H with high


Scheme 2.13<br />

2.5 Ultralong C–C Bonds<br />

stability despite the elongated C–C bond (calc. 1.64 Å) [311, 313]. Another type of<br />

clump (‘back-clump’), by which two trityl parts are not connected as in the case of<br />

9,9�-diphenyl-9,9�-bifluorenyl, does not prevent cleavage of the weakened bond so<br />

effectively (Scheme 2.13b). Since BDEs decrease with the increase in C–C bond<br />

length [298], the bridging modification is essential to obtain thermodynamicallystabilized<br />

HPEs that can allow investigation on the weakened and easily-dissociated<br />

long C–C bond.<br />

‘Cross-clumped’ HPE 147-H with the dihydrophenanthrene skeleton was<br />

prepared from the bis(triphenylmethanol) 148-H via the dianionic species<br />

(Scheme 2.14a) [313]. When the electron-donating substituents are attached<br />

to the aryl groups, acid treatment of diols 148-NMe 2 and 148-OMe gave stable<br />

dications, [biphenyl-2,2�-diyl bis(diarylmethylium)s], which were transformed<br />

to HPEs 147-NMe 2 and 147-OMe smoothly upon reduction [314]. The reductive<br />

generation of HPEs via the dicationic species has synthetic merit for generating<br />

highly congested HPEs 147 from the less-hindered precursors through short-step<br />

transformation. Dispiro-HPEs (147-spiro-N and 147-spiro-O) were prepared by this<br />

protocol (Scheme 2.14b), whose ‘ethane’ bond lengths are as long as 1.635(2) Å<br />

[314c] and 1.656(5) Å [314d], respectively.<br />

Not only the dihydrophenanthrene-type compounds, but also HPEs with an<br />

acenaphthene skeleton were available by this method. Thus, naphthalene-1,8diyl<br />

bis(diarylmethylium)s could be converted into 1,1,2,2-tetraarylacenaphthene<br />

derivatives 149 (Scheme 2.15) [315], whose ‘ethane’ bond is as long as 1.70 Å<br />

[1.701(3) Å for 149-H [315b] and 1.707(2), 1.705(2) Å for 149-F (two independent<br />

75


76 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

Scheme 2.14


Scheme 2.15<br />

2.5 Ultralong C–C Bonds<br />

molecules in crystal) [315c]. Torsional fixation in the five-membered ring is the<br />

key feature for producing the larger ‘front strain’ [311a] resulting in elongation of<br />

the C–C bonds in 149 compared with that in 147, since the latter can relieve the<br />

strain by adopting a chair-like conformation of the central six-membered ring.<br />

Although the estimated BDE for the long C–C bond of 1.70 Å is only 20 kcal mol –1 ,<br />

the acenaphthene-type HPEs 149 could be isolated with no signs of decomposition<br />

under ambient conditions. The remarkable stability of 149 can be rationalized<br />

by considering that the bond-dissociated species with the non-Kekulé structure<br />

(naphthalene-1,8-diyl), if formed, only undergo bond-reformation to give 149. This<br />

is in sharp contrast to the case of tetraarylbenzocyclobutene 143 [309], which is<br />

transformed into the dihydroanthracene derivative via the bond-dissociated isomer,<br />

tetraaryl-o-quinodimethanes 144, as discussed previously (Scheme 2.12a).<br />

It is clear from the above results that the fate of the bond-dissociated species<br />

is the major determinant factor for successful isolation and characterization<br />

of a compound with an ultralong C–C bond (d > 1.70 Å). By rational design to<br />

prevent the bond-dissociated species from intramolecular reaction other than<br />

bond-reformation, and from intermolecular reactions to give the adduct [316],<br />

the molecules with a super-ultralong C–C bond (d > 1.75 Å) would be obtained,<br />

whose bond length comes closer to the shortest nonbonded 1,3 C–C contact<br />

(1.80–1.90 Å) [317].<br />

77


78 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

2.5.5<br />

HPE Derivatives with a Super-ultralong C–C Bond<br />

The acenaphthene scaffold in HPEs 149 is one of the ideal skeletons for further<br />

exploration, yet the drawback is that the ‘rigid’ naphthalene-fused five-membered<br />

ring is not so rigid as expected. The experimentally determined d values for the<br />

‘ethane’ bond in HPEs 149 exhibit considerable variation [1.633(3) Å for 149-Cl<br />

[315b] and 1.670(3) Å for 149-MeO [315a]], and the smaller values were observed<br />

when the acenaphthene skeleton adopted the skewed conformation with the<br />

larger torsion angle for C 8a -C 1 -C 2 -C 2a as in 149-Cl [27.9(3)°] [315b] and 149-MeO<br />

[20.6(2)°]. Apparently, the HPEs 149-H [0.0(2)°] and 149-F [3.2(2)°, 3.6(2)°] with<br />

the eclipsed conformation suffer from the larger front strain, which expands the<br />

polyarylated C 1 –C 2 bond. Based on our recent results, the electronic effects of the<br />

substituents on the aryl groups do not determine the preference for the eclipsed/<br />

skewed conformation [315c]. Thus, the torsion angle would be much affected by<br />

crystal packing force.<br />

When the acenaphthene skeleton is modified to increase its rigidity, skewing<br />

deformation would become more costly in energy than expanding the long C–C<br />

bond. Thus, the pyracene-type HPEs 150 with the extra five-membered ring are<br />

better candidates to exhibit an ultralong C–C bond. Furthermore, the additional<br />

bridge induces expansion of the C 1–C 2 bond [318] in the parent hydrocarbons<br />

[calculated by B3LYP/6-31G* [319]: 1.586 Å for pyracene and 1.569 Å for acenaphthene,<br />

respectively]. Since the bond-elongation effect is strongly enhanced for the<br />

prestrained bond [303a, 320], the steric repulsion in the pyracene skeleton must<br />

enlarge the d value for the ‘ethane’ bond more effectively in 150 (Scheme 2.16).<br />

This is also the heart of the authors’ molecular design to realize the super-ultralong<br />

C–C bond, whose d is larger than 1.75 Å.<br />

Various diarylketones were reacted with 5,6-dilithioacenaphthene generated in<br />

situ from 5,6-dibromide. Upon treatment with acid under dehydrating conditions,<br />

the resultant diols are smoothly converted to a series of acenaphthene-5,6-diyl<br />

bis(diarylmethylium)s, from which the desired HPEs 150 were generated upon<br />

reduction [321]. X-ray analyses have shown that they all adopt the eclipsed conformation<br />

with the torsion angle of C 8a-C 1-C 2-C 2a less than 3°, and consequently, all of<br />

Scheme 2.16


2.5 Ultralong C–C Bonds<br />

Figure 2.11 X-ray structure of 1,1,2,2-tetraphenylpyracene 150-H with a super-ultralong C 1 -C 2<br />

bond length [1.754(2) Å] (C2/c, Z = 4, R = 5.5%, T = 123 K). The molecule is located on the<br />

crystallographic 2-fold axis.<br />

the ‘ethane’ bonds are longer than 1.7 Å. The super-ultralong bond (d > 1.75 Å) is<br />

found in several of them, such as 150-H [1.754(2) Å] and 150-F [1.761(4) Å] [321].<br />

Though largely separated, these carbon atoms must form ordinary covalent bonds.<br />

The two carbons are hybridized in the sp 3 manner, as judged by the tetrahedral<br />

coordination determined by X-ray analysis (Figure 2.11). Despite the marginal<br />

down-field shift, the 13 C signal (78.73 ppm for 150-H in CDCl 3 ) stays within the<br />

region typical of the sp 3 carbons. Theoretical calculation (B3LYP/6-31G*) gave the<br />

d values of 1.758 Å for 150-H and 1.762 Å for 150-F, respectively, which correspond<br />

well with the observed values [321]. Through this achievement, the authors have<br />

proposed a general scheme to design the HPE derivatives with a super-ultralong<br />

C–C bond by taking advantage of: (1) increase in front strain by torsional fixation<br />

and skeletal rigidity; (2) non-Kekulé structure for the bond-dissociated species<br />

that can only undergo bond-reforming without giving any by-products. The latter<br />

condition is essential to isolate and characterize the HPE containing a weakened<br />

ethane bond.<br />

2.5.6<br />

‘Expandability’ of the Ultralong C–C Bond:<br />

Conformational Isomorphs with Different Bond Lengths<br />

Based on the proposal for a linear relationship between bond lengths and BDEs<br />

[298], the ‘ethane’ bond in 150-H and 150-F would have ‘negative’ BDEs, which<br />

is unrealistic. There should be a nonlinear relationship or some special effects in<br />

the region of the ultralong and the super-ultralong C–C bonds. During further<br />

examination of the extremely long C–C bond in the pyracene-type HPEs, the<br />

authors have encountered a special case of a ‘soft’ long C–C bond. That is, the<br />

79


80 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

same C–C bond in a certain HPE exhibits several different values (1.71–1.77 Å)<br />

depending on the requisites from crystal packing.<br />

Compared with the conformational polymorphs where different conformers<br />

crystallize separately in different crystal forms, conformational isomorphs have<br />

seldom been observed, in which the different conformers are packed orderly in<br />

the same crystal. The latter is the case for the pyracene-type HPE 151 having two<br />

spiro(10-methylacridan) units. According to the X-ray analysis at –180 °C, there are<br />

four crystallographically independent molecules in the crystal of 151. Two adopt<br />

the eclipsed conformation [torsion angle: 3.4(2), 9.4(2)°] and other two are skewedshaped<br />

[23.4(1), 24.7(1)°] (Figure 2.12). Adoption of two different conformations<br />

is similar to the case of acenaphthene-type HPEs (149-H, 149-F: eclipsed; 149-Cl,<br />

149-MeO: skewed), yet both conformers of 151 do coexist in the same crystal. As<br />

expected, the eclipsed conformers possess much longer ‘ethane’ bond [1.771(3)<br />

and 1.758(3) Å] than the skewed molecules [1.712(2) and 1.707(2) Å]. By measuring<br />

the diffraction data at the elevated temperature, the longest bond length becomes<br />

as large as 1.781(6) Å at –40 °C. This value of d is one of the largest ever reported<br />

for the C–C bond length determined with sufficient accuracy [290].<br />

The molecular structures in conformational isomorphs often differ significantly<br />

in torsional angles but not in bond angles; changes in bond lengths are found less<br />

frequently. The last cases are usually related to the presence of differently charged<br />

molecules [322] or tautomerism in the crystal [323]. Even when the molecule<br />

crystallizes in several forms (polymorph), the chemically equivalent bonds in<br />

general exhibit the same bond length within experimental error. Thus, the crystallographically<br />

determined value of d has been considered as ‘intrinsic’ or an<br />

‘eigenvalue’ for a compound of the same charge and the same atom connectivity.<br />

This provides the basis for the validity of comparison of the bond lengths determined<br />

by X-ray analyses and to discuss the relationship between bond lengths and<br />

chemical structures. What happens when the same molecule possesses different<br />

bond lengths in polymorphs? What happens when the chemically equivalent<br />

but crystallographically independent molecules exhibit considerably different<br />

bond lengths in the same single crystal? This is the case for HPE 151 described<br />

above. So, none of the d values precisely determined [1.771(3), 1.758(3), 1.712(2),<br />

1.707(2) Å] is the eigenvalue of compound 151, but X-ray analysis just demonstrates<br />

that the bond in question can be expanded, if it likes, to 1.771(3) Å under<br />

the given circumstances, and that the bond can also remain much shorter than<br />

that. This is not a unique and special phenomenon for 151: structurally related<br />

compound 152, which has the etheno-bridge instead of ethano-bridge in 151, also<br />

crystallizes as a conformational isomorph [319]. Its crystal packing is quite different<br />

from that in 151, and there are three crystallographically independent molecules<br />

of 152: one eclipsed [bond length and torsion angle at –130 °C: 1.749(2) Å and<br />

5.5(2)°] and two skewed forms [1.726(2) Å and 21.3(2)°; 1.721(2) Å and 22.8(2)°].<br />

This dihydropyracylene-type HPE 152 is another example that shows the large<br />

‘expandability’ of the ultralong C–C bond. In contrast, HPE 153 with a less rigid<br />

acenaphthene skeleton only adopts the skewed conformation with a much smaller<br />

d value [1.696(3) Å and 18.1(3)°] [315d] than 151 or 152.


2.5 Ultralong C–C Bonds<br />

Figure 2.12 X-ray structures of two of four crystallographically independent molecules of<br />

bis(10-methylspiroacridine)-type HPE 151 with an eclipsed (a) and skewed (b) conformation<br />

(P1bar, Z = 8, R = 5.2%, T = 93 K).<br />

81


82 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

2.5.7<br />

Future Outlook<br />

Ultralong C–C bonds present intriguing facets of organic covalent bonds. For the<br />

super-ultralong bond with d > 1.75 Å, there is no longer the linear relationship<br />

between d and BDE. The super-ultralong bond can exhibit a range of d values<br />

in crystal depending on the packing arrangement since only a small energy is<br />

necessary to expand the ‘expandable’ covalent bond. Thus, the precisely determined<br />

d value for the ‘soft’ bond is not necessarily the ‘eigenvalue’ of the bond in a certain<br />

compound, but it just points out to the length it can exhibit. That value can be the<br />

higher limit in one case or the lower limit in another case.<br />

In any case, the experimentally determined large d value will attract attention. It<br />

surely shows that the bond can be longer than the standard value. Chemists can<br />

continue the game to find the further elongated C–C bond, and the outstanding<br />

stability of compounds 150 has prompted the authors to prepare the derivative<br />

that shows even larger value of d in crystal (1.79 Å) [321], which is nearly equal<br />

to the shortest nonbonded 1,3 C–C distance (1.80 Å) [317]. The authors appreciate<br />

Herges’ view [289] that ‘the stable C–C bonds even longer than the smallest<br />

nonbonding distance are conceivable’.<br />

2.6<br />

Ultrashort C–C Bonds<br />

Vladimir Y. Lee and Akira Sekiguchi<br />

2.6.1<br />

Introduction<br />

The generally accepted normal value for the single C–C bond length between<br />

two sp 3 -hybridized tetracoordinate carbon atoms, compiled from a large number<br />

of X-ray structures, is 1.54 Å [324, 325], the value close to that of the C–C bond<br />

in diamond. In a vast number of crystallographically characterized organic<br />

compounds, the length of ordinary C–C bonds may deviate slightly from this<br />

average value, albeit being still very close to it. However, the two limiting cases of<br />

nonclassical C–C bonds are of particular importance: extremely long (ultralong)<br />

and extremely short (ultrashort) C–C bonds. The first topic, ultralong C–C bonds,<br />

is covered by Suzuki (Section 2.5), whereas our goal is to discuss the second<br />

topic, ultrashort C–C bonds, bringing together the most important experimental<br />

accomplishments and theoretical considerations in this field.Basically, one can<br />

distinguish four fundamental classes of compounds featuring ultrashort C–C<br />

single bonds, which are typically markedly shorter than the border value of 1.50 Å.<br />

The first class includes those compounds in which the C–C bond is framed on<br />

both sides by either C=C or C�C multiple bonds: the most common examples are<br />

1,3-butadiene, H 2 C=CH–CH=CH 2 , and 1,3-butadiyne, HC�C–C�CH, where the<br />

central C–C bonds are considerably shortened to 1.463(3) Å [326] and 1.384(2) Å


2.6 Ultrashort C–C Bonds<br />

[327], respectively, being just intermediate between those of the normal C–C<br />

single (1.54 Å) and C=C double (1.34 Å) bond lengths. This phenomenon is<br />

typically interpreted in terms of hybridization (increased s-character of the sp 2 - or<br />

sp-hybrids used to form the C–C �-bonds) and, more importantly, conjugation<br />

(multiple bonds conjugation leading to the delocalization of �-electrons over the<br />

entire molecule and shortening of the central C–C bond due to its partial double<br />

bond character). In contrast to this first class of ultrashort C–C bond compounds,<br />

well known from standard textbooks of organic chemistry, the other three classes<br />

are much less familiar to the general chemical community. The second class of<br />

compounds, exhibiting extraordinarily short C–C bonds between the tetracoordinated<br />

C atoms, includes the derivatives of tricyclo[2.1.0.0 2,5 ]pentane, featuring<br />

exceptionally short endocyclic bridging bonds. The third class of compounds, for<br />

which ultrashort C–C bonds were experimentally found, is represented by the<br />

dimeric polyhedral compounds (e.g., bi(tetrahedranyl), bi(bicyclo[1.1.0]butyl),<br />

bi(bicyclo[1.1.1]pentyl), bi(cubyl)), exhibiting remarkably short central intercage<br />

exocyclic bonds. The fourth and newest class of compounds featuring the ultrashort<br />

C–C bonds is known more computationally than experimentally at present.<br />

In such derivatives the C–C bond, force to point inside the inert cage (like cyclophane),<br />

is compressed by the evident steric congestions. In this section, we will<br />

deal with the last three classes of compounds featuring ultrashort C–C bonds,<br />

briefly describing their structural peculiarities and particularly emphasizing the<br />

origin of the exceptional bond shortening.<br />

2.6.2<br />

Tricyclo[2.1.0.0 2,5 ]pentanes: Ultrashort Endocyclic Bridging C–C Bonds<br />

The derivatives of tricyclo[2.1.0.0 2,5 ]pentane 154a–h, known since 1964 [328, 329],<br />

particularly those of tricyclo[2.1.0.0 2,5 ]pentan-3-one 154 (R 2 , R 2 = C=O), represent<br />

the record shortening of the single bond between two tetracoordinate C atoms<br />

(Table 2.1).<br />

Indeed, the most remarkable structural feature of all tricyclo[2.1.0.0 2,5 ]pentane<br />

derivatives 154 is the exceedingly short bond between the bridgehead carbon<br />

atoms C1–C5, ranging from 1.408 to 1.509 Å (av. 1.45 Å) [330–340]. The only<br />

exceptional case is represented by the heavy group 14 elements containing<br />

2,4-disila-1-germatricyclo[2.1.0.0 2,5 ]pentane derivative 155, in which the bridging<br />

Ge–C bond of 2.242(3) Å was extraordinarily stretched, being 15% longer than<br />

83


84 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

Table 2.1 Endocyclic C 1 –C 5 bridging bonds in tricycle[2.1.0.0 2,5 ]pentane derivatives 154.<br />

Tricyclo[2.1.0.0 2,5 ]pentane<br />

derivative<br />

R 1 , R 2<br />

154a R 1 = Ph;<br />

R 2 = OCO(p-Br-C 6 H 4 )<br />

154b R 1 = Ph;<br />

R 2 ,R 2 = C=O<br />

154c R 1 = Me;<br />

R 2 ,R 2 = C=O<br />

154d R 1 = CH 2 OCOCH 3 ;<br />

R 2 ,R 2 = C=O<br />

154e R 1 = CH 2 OCOCH 3 ;<br />

R 2 ,R 2 = –OCH 2 CH 2 O–<br />

154f R 1 ,R 1 = –CH 2 N(COOMe)N(COOMe)CH 2 –;<br />

R 2 = OEt<br />

154g R 1 = COOCH 3 ;<br />

R 2 ,R 2 = –OCH 2 CH 2 O–<br />

154h R 1 = COOCH 3 ;<br />

R 2 ,R 2 = C=O<br />

r (C 1–C 5), Å Ref.<br />

1.44(5) [330]<br />

[331]<br />

1.444(2) [332]<br />

1.408(3) at 298 K<br />

(1.417(1) at 118 K)<br />

[332]<br />

[(334)]<br />

1.416(2) [333]<br />

[335]<br />

1.455(3) [336]<br />

1.509(2) [337]<br />

1.485(2) [338]<br />

1.453(2) [338]<br />

the average value, the fact interpreted in terms of the appreciable singlet biradical<br />

contribution to the nature of the Ge–C bridging bond [341].<br />

The shortest known endocyclic C–C single bonds were measured for 1,5-dimethyltricyclo[2.1.0.0<br />

2,5 ]pentan-3-one 154c (1.408(3) Å at 298 K [332] and 1.417(1) Å<br />

at 118 K [334]) and 1,5-bis(acetoxymethyl)tricyclo[2.1.0.0 2,5 ]pentan-3-one 154d<br />

(1.416(2) Å) [333, 335]. The recent DFT calculations fairly well reproduced the<br />

structural parameters of 154c: bridging bond length of 1.425 Å (1.417(1) Å by<br />

X-ray diffraction) and interplanar angle 94.96° (95.0° by X-ray diffraction) [342]. In<br />

sharp contrast, bicyclo[1.1.0]butane derivatives 156, which represent a part of the<br />

tricyclo[2.1.0.0 2,5 ]pentane skeleton, exhibit either normal (short bond isomers) or<br />

very long (long bond isomers) bridging C–C distances (Scheme 2.17).<br />

Such a striking difference between these two classes of polycyclic derivatives is<br />

due to the evident geometrical constraints in tricyclo[2.1.0.0 2,5 ]pentanes 154, in<br />

which the methylene C3-bridge between the C2 and C4 atoms ‘tightens’ the two<br />

three-membered rings, thus providing a fixed interplanar angle �, much smaller<br />

than that of the ‘nonfixed’ bicyclo[1.1.0]butanes 156: 94–99° vs. 113–130° [339].


Scheme 2.17<br />

2.6 Ultrashort C–C Bonds<br />

Consequently, such a small interplanar angle � forces the hybrids at C1 and C5<br />

atoms in 154 to form an extraordinarily short highly bent C1–C5 bond [339, 340,<br />

343]. However, this bridgehead C1–C5 bond is not a classical �-bond formed<br />

by the linear overlap of two sp 3 -hybrids. Thus, the electron density distribution<br />

study of 1,5-dimethyltricyclo[2.1.0.0 2,5 ]pentan-3-one 154c by X-ray analysis at<br />

118 K revealed that the electron density maximum of the bridging C1–C5 bond is<br />

displaced by ~0.40 Å outwards from the bond axis, which amounts to an extreme<br />

bending by 28° [334]. Consequently, the C1–C5 bridging bond is best described<br />

as a highly bent �-bond. The early calculations considered the bridging C–C<br />

bond in 154 to be a result of the �-type overlap of the two p z -orbitals from each<br />

bridgehead carbon atom, this suggestion being based on the general reactivity<br />

of this bond [344]; however, later DFT calculations objected such a viewpoint,<br />

showing that the bridging C–C bond has no � character and is to be described<br />

as a classical two-center bent �-bond [342]. The hybridization of the bridgehead<br />

carbon atoms C1 and C5 in 154 sharply deviates from the normal sp 3 -state: the<br />

hybrid orbitals of the exocyclic bonds are high in s-character (the coupling constant<br />

1 JCH for exocyclic C1–H and C5–H bonds of 2,4-dimethyltricyclo[2.1.0.0 2,5 ]pentane<br />

were measured to be as large as 212 Hz) [345], whereas the hybrids used for the<br />

formation of the endocyclic C1–C5 bond have substantial p-character (the value<br />

of sp 4.21 was calculated for the unsubstituted tricyclo[2.1.0.0 2,5 ]pentan-3-one 154<br />

(R 1 = R 2 = H) 342 ) [334, 339].<br />

The extent of the shortening of the bridging C1–C5 bond in 154 is greatly affected<br />

by two principal factors: geometrical (interplanar angle �) and electronic (influence<br />

of the �-accepting substituents). Accordingly, the bridging bond becomes longer<br />

when the interplanar angle widens, and vice versa–the bridging bond shortens<br />

upon decreasing the interplanar angle [332]. Another geometrical factor, greatly influencing<br />

the length of the bridging bond, is the external bond angle �. In complete<br />

accordance with the theoretical prediction [343], the relationship between these<br />

two geometrical parameters is inverse: the larger the bond angle �, the shorter the<br />

bridging bond, and vice versa [337]. Electronically, �-accepting substituents (like<br />

carbonyl, phenyl groups) may effectively interact with the bridging C1–C5 bond<br />

in 154, because it is made of hybrids high in p-character (vide supra); however, the<br />

direction of such influence depends on the position of the �-accepting group [339].<br />

For example, the carbonyl groups at the bridgehead atoms C1 and C5 give rise<br />

85


86 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

to a lengthening of the C1–C5 bridging bond in 154h by ~0.03 Å because of the<br />

optimal orientation of the carbonyl groups, allowing for the effective interaction<br />

of their C=O �-bonds with the bridging C1–C5 �-bond [338, 339].<br />

In contrast, the carbonyl group at the bridging position C3 (C3=O), tightening<br />

the bicyclo[1.1.0]butane fragment at the C2 and C4 positions, shows the opposite<br />

effect, shortening the bridging C1–C5 bond in 154h by the same extent of ~0.03 Å<br />

[338, 339,346].<br />

2.6.3<br />

Coupled Cage Compounds: Ultrashort Exocyclic Intercage C–C Bonds<br />

This class of compounds featuring ultrashort C–C bonds best of all fits the original<br />

definition of such bonds, lacking the highly perturbing factors of either conjugation<br />

to multiple bonds (as the representatives of the first class of compounds with<br />

ultrashort C–C bonds, 1,3-butadiene and 1,3-butadiyne) or incorporation into the<br />

highly strained tricyclic framework (as the representatives of the second class of<br />

compounds with ultrashort C–C bonds, derivatives of tricyclo[2.1.0.0 2,5 ]pentane).<br />

Indeed, in striking contrast to the highly bent nature of the endocyclic bridging<br />

C–C �-bonds of tricyclo[2.1.0.0 2,5 ]pentane derivatives, which makes the definition<br />

of the bond length somewhat ambiguous, the intercage C–C �-bonds are exocyclic<br />

and nonbent, thus experiencing minimal geometrical constraints.


2.6 Ultrashort C–C Bonds<br />

The first experimental accomplishments in this field were preceded by the prediction<br />

that the exocyclic intercage C–C bonds in the hypothetical bi(cubyl) 157<br />

and 1,1�-bi(bicyclo[1.1.1]pentyl) 159 could be markedly shortened by up to 0.08 Å<br />

compared with the normal values [347].<br />

This hypothesis was based on the general idea that the C–C bonds are expected<br />

to be appreciably shortened if they were involved in widened bond angles (and vice<br />

versa: the C–C bonds adjacent to compressed bond angles should be stretched).<br />

This was the case of compounds 157 and 159, in which the intercage C–C bonds<br />

participate in six significantly widened C–C–C bond angles [347]. A year later<br />

this fruitful idea found spectacular experimental confirmation in the synthesis<br />

and crystal structure determination of bicubyl 157 and (2-t-butylcubyl)cubane<br />

158 [348]: the intercage C–C bonds in both 157 and 158 were significantly contracted,<br />

1.458(8) and 1.464(5) Å, respectively, being in very good agreement with<br />

the predicted value of 1.46 Å [347] and very close to the value of 1.463(3) Å for the<br />

central C–C bond in 1,3-butadiene [326].<br />

Such appreciable shortening was simply explained in hybridization terms: endocyclic<br />

C–C bonds of bicubyl derivatives are richer in p-character than the normal<br />

sp 3 -hybrids, whereas the exocyclic C–C bonds are higher in s-character, which<br />

allows them to form shorter bonds to substituents than their sp 3 -counterparts<br />

do. The specific geometrical constraints of bicubyl molecules create less steric<br />

crowding than, for example, in hexasubstituted ethane, the latter factor also<br />

favoring shortening of the intercage C–C bond. In contrast to 157 and 158, in<br />

1,1�-biadamantyl 163 the intercage C–C bond is formed by nearly pure sp 3 -hybrids;<br />

consequently, this bond is not ultrashort but stretched to 1.578(2) Å [349], being<br />

0.12 Å longer than that in bicubyl 157.<br />

Other examples of coupled cage compounds featuring ultrashort intercage<br />

C–C bonds include: bi- and tri(bicyclo[1.1.1]pentyl) derivatives 160–162 and 164,<br />

87


88 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

1,1�-bihomocubyl 165 [350–353]. Thus, in keeping with expectations [347], the<br />

bi(bicyclo[1.1.1]pentyl) derivative 160 showed a short intercage bond distance of<br />

1.480(4) Å [350], whereas the tri(bicyclo[1.1.1]pentyl) derivative 164 manifested<br />

even shorter distances of 1.464(7) and 1.476(7) Å [351].<br />

The other two representatives of bi(bicyclo[1.1.1]pentyl) derivatives, 161 and<br />

162, also displayed diagnostically short intercage C–C bonds of 1.480(3) and<br />

1.469(6) Å, respectively [352]. An MNDO calculation showed that the nature of<br />

the substituents at the bridgehead 3,3�-positions (H, NH 2 , or CN) has almost<br />

no influence on the length of the intercage C–C bond, being nearly invariable<br />

(1.470–1.472 Å) in all calculated structures. Quite similar to the above cases of<br />

bicubyl 157 and (2-t-butylcubyl)cubane 158 [348], and in complete accord with the<br />

prediction [347], 1,1�-bihomocubyl 165 exhibited the short intercage C–C bond<br />

length of 1.460(1) Å, this value is very close to the corresponding values in 157<br />

(1.458(8) Å) and 158 (1.464(5) Å) [353].<br />

The 1,1�-coupled derivatives of bicyclo[1.1.0]butane were expected to manifest a<br />

further shortening of the intercage C–C bonds, because their adjacent bond angles<br />

are very wide, being larger than those in bicubyl 157 and bi(bicyclo[1.1.1]pentyl)<br />

159. Indeed, in the two structurally characterized examples of such compounds,<br />

namely, 1,1�-bi(tricyclo[4.1.0.0 2,7 ]heptyl) 166 and 1,1�-bi(tricyclo[3.1.0.0 2,6 ]hexyl) 167<br />

derivatives the exocyclic intercage C–C bonds were exceptionally short, 1.445(3)<br />

and 1.440(2) Å, respectively, being ~0.10 Å shorter than the average value of<br />

1.54 Å [354].


2.6 Ultrashort C–C Bonds<br />

Extrapolating the experimental results on the lengths of the intercage C–C bonds<br />

in coupled cage compounds, the authors deduced the shortening of ~0.10 Å for<br />

the central bond in the hypothetical bitetrahedranyl 169 [354].<br />

Such impressive experimental accomplishments raised a great deal of interest<br />

among theoreticians about the problem of extreme shortening of the intercage C–C<br />

bonds. Thus, the length of this bond in bicubyl 157 was calculated at the HF/DZ+d<br />

SCF level as 1.484 Å, this value being 0.026 Å longer than the experimental value<br />

of 1.458(8) Å [355]. Definitely, the Hartree–Fock method markedly overestimated<br />

the length of the central C–C bond in the coupled cage compounds. Likewise, the<br />

early calculations on the parent unsubstituted 1,1�-bi(tricyclo[3.1.0.0 2,6 ]hexyl) 168<br />

produced the overestimated values of 1.477 Å at the STO-3G SCF level and 1.453 Å<br />

at the DZ SCF level [356] (cf. 1.440(2) Å experimentally observed in 167 [354]).<br />

On the other hand, the semiempirical AM1 calculations gave underestimated<br />

values of the intercage C–C bond in the 1,1�-bi(tricyclo[3.1.0.0 2,6 ]hexyl) derivatives:<br />

1.409 Å for the unsubstituted model 168 and 1.411 Å for the real compound 167<br />

[357]. The recent DFT B3LYP/6-31G(d,p) calculations on the real molecules 166<br />

and 167 provided results most closely approaching the experimental X-ray data:<br />

1.456 and 1.452 Å vs. 1.445(3) and 1.440(2) Å [342]. Consequently, the hybridization<br />

of the intercage C–C bond in 166 and 167 was calculated as sp 1.31 and sp 1.26 ,<br />

which amounts to an s-contribution to the bonding orbitals of 43.2% and 44.3%,<br />

respectively. The extreme shortening of the intercage C–C bonds in both 166<br />

and 167 was further manifested in the extraordinarily large values of the 1 J(C, C)<br />

coupling constant of 99.8 and 85.9 Hz, respectively, undoubtedly reflecting the<br />

great s-character of these bonds [342].<br />

The most intriguing molecule to study computationally was bi(tetrahedranyl)<br />

169, whose structure was calculated by several research groups. (Other tetrahedrane<br />

derivatives were discussed in Section 2.4.1.1.) The earliest calculations<br />

agreed well with the above prediction of ~1.44 Å [354], giving the estimates of<br />

the intercage C–C bond in 169 as 1.438 Å at the 6-31G* [358] and 1.444 Å at the<br />

DZ+P SCF [356] levels of theory. In the latter study, the six equivalent bond angles<br />

at the central C–C bond were calculated to be 144.6°, which corresponded to a<br />

bond angle widening of 35.1° compared with a normal tetrahedral angle value of<br />

109.5° [356]. Consequently, the intercage C–C bond length in 169 was expected to<br />

be intermediate between the 1.440(2) Å of 1,1�-bi(tricyclo[3.1.0.0 2,6 ]hexyl) derivative<br />

89


90 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

167 (widening angle 130.9° – 109.5° = 21.4°) and the 1.384(2) Å of 1,3-butadiyne,<br />

H–C�C–C�C–H (widening angle 180.0° – 109.5° = 71.5°). A subsequent study<br />

by the same authors performed at the MP2 level with the same DZ+d basis set<br />

showed a slightly shorter intercage C–C bond in 169 of 1.434 Å [355]. A larger<br />

shortening of the central C–C bond in 169 was found by the semiempirical AM1<br />

calculations: 1.386 Å Å [357]. It is well-known that the NMR coupling constant<br />

1 J(C, C) between the two carbon atoms forming the intercage C–C bond is another<br />

diagnostic measure of the electronic distribution in alkane bonds, effectively indicating<br />

the alterations in the hybridization of carbon atoms and, consequently, in<br />

the C–C bond length. Thus, the extreme shortening of the C–C bond linking the<br />

two tetrahedrane units in 169 (1.444 Å) was manifested in an extraordinarily large<br />

1 J(C, C) coupling constant of 151 Hz [359]. This 1 J(C, C) value is much greater than<br />

the values of 35.0 Hz measured for isobutane (CH 3) 2CHCH 3 [360], and 53.7 Hz<br />

for the central C–C bond of 1,3-butadiene, H 2C=CH–CH=CH 2 [361], closely<br />

approaching the value of 154.8 Hz measured for 1,3-butadiyne, HC�C–C�CH<br />

[362]. Consequently, the calculated 1 J(C, C) coupling constant of the intercage<br />

C–C bond in 169 is very similar to that of a bond between the two sp-hybridized<br />

carbon atoms, and it possibly represents the highest 1 J(C, C) value ever reported<br />

for a C–C bond between two saturated tetracoordinated carbon atoms.<br />

The experimental realization of the theoretical predictions discussed above was<br />

achieved a couple of years ago, when the structure of the stable hexakis(trimethylsilyl)bi(tetrahedranyl)<br />

170 was established [363]. The intercage C–C bond in 170<br />

was exceedingly short, 1.436(3) Å, this value being very close to those predicted<br />

by calculations (vide supra) of 1.434–1.444 Å, being the shortest acyclic nonbent<br />

C–C single bond between two saturated tetracoordinated carbon atoms known<br />

thus far. The principal reason for such a great bond shortening was ascribed to<br />

the high s-character of this exocyclic C–C bond in 170: the hybridization of this<br />

bond was calculated (NBO analysis) to be sp 1.53 at the B3LYP/6-31G(d) level [363].<br />

However, the latest computations at the HF/6-311+G(d,p) level pointed out that<br />

apart from this evident cause, there is another reason behind the appreciable<br />

intercage bond shortening in bi(tetrahedranyl) 169, namely, vicinal hyperconjugative<br />

orbital interactions between the two tetrahedranyl units over the central<br />

C–C bond [364]. This hyperconjugation, facilitated by six electron donating silyl<br />

substituents, was calculated to be even more important than conjugation in<br />

1,3-butadiene: the hyperconjugation energy in the staggered conformation of<br />

bi(tetrahedranyl) 169 of –15.2 kcal/mol vs. conjugation energy in 1,3-butadiene<br />

of –9.9 kcal mol –1 . Such an interaction of the two tetrahedranyl units in 170<br />

raised the HOMO energy level and lowered the LUMO energy level, resulting<br />

in the decrease in the HOMO–LUMO energy gap [365]. This was reflected in<br />

the red shift of the longest wavelength absorption of 170, compared with that of<br />

tetrakis(trimethylsilyl)tetrahedrane [365–367]. Thus, the two comparable causes<br />

responsible for the extreme shortening of the intercage C–C bond in 170 were<br />

suggested: the high s-character of this exocyclic bond in bi(tetrahedranyl) (58%<br />

contribution) and vicinal hyperconjugative interactions between the tetrahedranyl<br />

groups (42% contribution) [363, 364].


2.6.4<br />

Sterically Congested in-Methylcyclophanes: Ultrashort C–C(Me) Bonds<br />

2.6 Ultrashort C–C Bonds<br />

Recently, other examples of compounds featuring ultrashort C–C bonds were<br />

computationally studied [368], however, it should be noted that at present such<br />

chemistry seems to be difficult to realize experimentally. The first class of such<br />

derivatives includes those compounds in which a particular C–C bond is pressed<br />

by encapsulation in an inert cage, either inter- (neopentane inside C 60 ) or intramolecularly<br />

(cyclophanes in which a methyl group is pressed towards the aromatic<br />

ring similarly to the recently reviewed molecular ‘iron maidens’ [369]. Until now<br />

only two examples of the compounds of such type were reported by Pascal et al.,<br />

namely, sterically congested in-methylcyclophane derivatives featuring very short<br />

C–Me bond distances of 1.475(6) and 1.495(6) Å [370]. However, it seems possible<br />

to shorten the C–C bond without pushing on it, just by covalent means (for<br />

example, C–C bonds inside the small covalent cage) [368]. Other examples of ultrashort<br />

C–C bonds are represented by the stiffened and strapped intracage bonds<br />

or tied-up bicycloalkanes (for example, interior propellanes formed by strapping<br />

bicyclo[n.n.n]alkane units together) [368]. All of these calculated molecules feature<br />

very short C–C bonds; however, the reason for such marked bond shortening is<br />

not as simple as just their increased s-character. Thus, the calculated C–C bond<br />

shortening at least by 0.1 Å exceeds the value that would be expected from hybridization<br />

alone. An additional bond shortening originates from the strain caused<br />

by the threefold symmetric geometry constraints: the computed ultrashort C–C<br />

bonds in the above systems are simply unable to achieve normal bond lengths<br />

without causing notable perturbation in the whole molecule.<br />

2.6.5<br />

Conclusions<br />

The general definition of the ‘normal’ C–C single bond length is somewhat arbitrary:<br />

the value of 1.54 Å is just an average of the vast number of available crystallographic<br />

data. However, depending on the particular case, the ‘normal’ C–C bond<br />

can be significantly stretched or shortened; consequently the two extremes, the<br />

ultralong C–C bonds and the ultrashort C–C bonds, may deviate from the average<br />

value by ca. 0.10 Å (6.5%) each, thus giving rise to the overall variation of ca. 0.20 Å<br />

(13%) with respect to the average value of 1.54 Å. In this section we have discussed<br />

in detail the two most important contributions to the field of ultrashort C–C bonds:<br />

derivatives of tricyclo[2.1.0.0 2,5 ]pentanes with squeezed endocyclic bridging C–C<br />

bonds and coupled cage compounds featuring contracted exocyclic intercage C–C<br />

bonds. The origin of the bond shortening in these two cases is clearly different,<br />

because in tricyclo[2.1.0.0 2,5 ]pentanes the short C–C bond is endocyclic and highly<br />

bent, whereas in the coupled cage compounds the short C–C bond is exocyclic<br />

and nonbent. After the characterization of bi(tetrahedranyl) 170, manifesting the<br />

shortest acyclic nonbent C–C single bond between two saturated tetracoordinate<br />

carbon atoms, the next prominent candidates for molecules featuring ultrashort<br />

91


92 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

C–C bonds would be polycyclic compounds, in which the particular C–C bond is<br />

inter- or intramolecularly involved in the cage fragment. This job has already been<br />

approached computationally–now it is the experimentalists’ turn.<br />

Acknowledgments<br />

We are greatly indebted to all our coworkers, who have made an invaluable experimental<br />

contribution to this work and whose names are listed in the references.<br />

This work was supported by a Grant-in-Aid for Scientific Research (Nos. 17550029,<br />

19105001, 19020012, 19022004, 19029006) from the Ministry of Education,<br />

Science, Sports, and Culture of Japan.<br />

References<br />

1 In 1874, van’t Hoff and Le Bel concluded<br />

independently that tetracoordinate<br />

carbon prefers tetrahedral arrangements<br />

of the four substituents.<br />

(a) van’t Hoff, J. H. Arch. Neerl. Sci.<br />

Exactes. Nat. 1874, 9, 445. (b) Le Bel,<br />

J. A. Bull. Soc. Chim. Fr. 1874, 22, 337.<br />

2 (a) Greenberg, A.; Liebmann, J. F.<br />

<strong>Strained</strong> Organic Molecules, Academic<br />

Press, New York 1978. (b) Vögtle, F.;<br />

Fascinating Molecules in Organic<br />

Chemistry, Wiley, Chichester 1992,<br />

Chapter 2. (c) Dodziuk, H. Modern<br />

Conformational Analysis. Elucidating<br />

Novel Exciting Molecular Structures; VCH<br />

Publishers, New York 1995, Chapter 7<br />

and references therein.<br />

3 The small-ring propellanes, the compound<br />

with inverted carbon atoms have<br />

been reviewed extensively by Wiberg:<br />

(a) Wiberg, K. B. Acc. Chem. Res. 1984,<br />

17, 379. (b) Wiberg, K. B. Chem. Rev.<br />

1989, 89, 975.<br />

4 (a) Cox, K. W.; Harmony, M. D.;<br />

Nelson, G.; Wiberg, K. B. J. Chem. Phys.<br />

1969, 50, 1976. (b) Paddon-Row, M. N.;<br />

Houk, K. N.; Dowd, P. Tetrahedron Lett.<br />

1981, 22, 4799.<br />

5 (a) Wiberg, K. B.; Walker, F. H.<br />

J. Am. Chem. Soc. 1982, 104, 5239.<br />

(b) Wiberg, K. B.; Waddell, S. T. J. Am.<br />

Chem. Soc. 1990, 112, 2194.<br />

6 Ginsburg, G. Acc. Chem. Res. 1972, 5, 249.<br />

7 (a) Ginsburg, D. Propellanes: Structure<br />

and Reactions, Verlag Chemie, Wein-<br />

heim, 1975. (b) Ginsburg, D. Propellanes:<br />

Structure and Reactions, Sequels I and II,<br />

Technion, Haifa 1980 and 1985.<br />

8 Wiberg, K. B.; Hiatt, J. E.; Burgmaier,<br />

G. J. Tetrahedron Lett. 1968, 5855.<br />

9 Wiberg, K. B.; Burgmaier, G. J.<br />

Tetrahedron Lett. 1969, 317.<br />

10 Gassman, P. G.; Topp, A.; Keller, J. W.<br />

Tetrahedron Lett. 1969, 1093.<br />

11 Wiberg, K. B.; Burgmaier, G. J.; Shen, K.;<br />

La Place, S. J.; Hamilton, W. C.;<br />

Newton, M. D. J. Am. Chem. Soc. 1972,<br />

94, 7402.<br />

12 (a) Eaton, P. E.; Temme, III, G. H.<br />

J. Am. Chem. Soc. 1973, 95, 7508.<br />

(b) Dannenberg, J. J.; Prociv, T. M.;<br />

Hutt, C. J. Am. Chem. Soc. 1974, 96,<br />

913. (c) He,Y.; Junk, C. P.; Cawley, J. J.;<br />

Lemal, D. M. J. Am. Chem. Soc. 2003,<br />

125, 5590.<br />

13 (a) Wiberg, K. B.; Burgmaier, G. J.<br />

J. Am. Chem. Soc. 1972, 94, 7396.<br />

(b) Wiberg, K. B.; Pratt, W. E.;<br />

Bailey, W. F. J. Am. Chem. Soc. 1977, 99,<br />

2297. (c) Wiberg, K. B.; Connon, H. A.;<br />

Pratt, W. E. J. Am. Chem. Soc. 1979, 101,<br />

6970.<br />

14 Aue, D. H.; Reynolds, R. N. J. Org.<br />

Chem. 1974, 39, 2315.<br />

15 (a) Szeimies-Seebach, U.; Szeimies, G.<br />

J. Am. Chem. Soc. 1978, 100, 3966.<br />

(b) Szeimies-Seebach, U.; Harnisch, J.;<br />

Szeimies, G.; Van Meerssche, M.;<br />

Germain, G.; Declercq, J.-P. Angew.<br />

Chem. Int. Ed. Engl. 1978, 17, 848.


(c) Szeimies-Seebach, U.; Schöffer, A.;<br />

Römer, R.; Szeimies, G. Chem. Ber.<br />

1981, 114, 1767. (d) Baumgart, K.-D.;<br />

Harnisch, H.; Szeimies-Seebach, U.;<br />

Szeimies, G. Chem. Ber. 1985, 118,<br />

2883.<br />

16 Hamon, D. P. G.; Trenerry, V. C.<br />

J. Am. Chem. Soc. 1981, 103, 4962.<br />

17 (a) Majerski, Z.; Žuani�, M.<br />

J. Am. Chem. Soc. 1987, 109, 3496.<br />

(b) Majerski, Z.; Kostov, V.; Hibšer, M.;<br />

Mlinari�-Majerski, K. Tetrahedron Lett.<br />

1990, 31, 915.<br />

18 Herr, M. L. Tetrahedron 1977, 33, 1897.<br />

19 (a) Newton, M. D.; Schulman, J. M.<br />

J. Am. Chem. Soc. 1972, 94, 773.<br />

(b) Newton, M. D.; Schulman, J. M.<br />

J. Am. Chem. Soc. 1972, 94, 4391.<br />

20 Stohrer, W.-D.; Hoffmann, R. J. Am.<br />

Chem. Soc. 1972, 94, 779.<br />

21 Davidson, E. R. Chem. Phys Lett. 1998,<br />

284, 301.<br />

22 Gassman, P. G.; Proehl, G. S.<br />

J. Am. Chem. Soc. 1980, 102, 6862.<br />

23 (a) Mlinari�-Majerski, K.; Majerski, Z.<br />

J. Am. Chem. Soc. 1980, 102, 1418.<br />

(b) Majerski, Z.; Mlinari�-Majerski, K.;<br />

Mei�, Z. Tetrahedron Lett. 1980, 21,<br />

4117. (c) Vinkovi�, V.; Majerski, Z.<br />

J. Am. Chem. Soc. 1982, 104, 4027.<br />

(d) Mlinari�-Majerski, K.; Majerski, Z.<br />

J. Am. Chem. Soc. 1983, 105, 7389.<br />

(e) Mlinari�-Majerski, K.; Majerski, Z.;<br />

Rakvin, B.; Veksli, Z. J. Org. Chem. 1989,<br />

54, 545. (f) Mlinari�-Majerski, K.; Šafar<br />

Cvitaš, D.; Majerski, Z. Tetrahedron Lett.<br />

1991, 32, 1655. (g) Mlinari�-Majerski, Z.;<br />

Šafar Cvitaš, D.; Veljkovi�, J. J. Org.<br />

Chem. 1994, 59, 2374.<br />

24 (a) Semmler, K.; Szeimies, G.; Belzner, J.<br />

J. Am. Chem. Soc. 1985, 107, 6410.<br />

(b) Belzner, J.; Szeimies, G. Tetrahedron<br />

Lett. 1986, 27, 5839. (c) Belzner, J.;<br />

Bunz, U.; Semmler, K.; Szeimies, G.;<br />

Opitz, K.; Schlüter, A.-D. Chem. Ber.<br />

1989, 122, 397. (d) Mondanaro, K. R.;<br />

Dailey, W. P. Organic Syntheses, Coll.<br />

Vol. 10, 2004, p. 658; Vol. 75, 1998,<br />

p. 98.<br />

25 (a) Belzner, J.; Szeimies, G. Tetrahedron<br />

Lett. 1987, 28, 3099. (b) Belzner, J.;<br />

Gareiss, B.; Polborn, K.; Schmid, W.;<br />

Semmler, K.; Szeimies, G. Chem. Ber.<br />

1989, 122, 1509.<br />

References<br />

26 Bothe, H.; Schlüter, A.-D. Chem. Ber.<br />

1991, 124, 587.<br />

27 Walker, F. H.; Wiberg, K. B.; Michl, J.<br />

J. Am. Chem. Soc. 1982, 104, 2056.<br />

28 (a) Carroll, W. F., Jr.; Peters, D. G.<br />

Tetrahedron Lett. 1978, 38, 3543.<br />

(b) Carroll, W. F., Jr.; Peters, D. G.<br />

J. Am. Chem. Soc. 1980, 102, 4127.<br />

29 Majerski, Z.; Veljkovi�, J.; Kaselj, M.<br />

J. Org. Chem. 1988, 53, 2662.<br />

30 Wiberg, K. B.; Walker, F. H.; Pratt, W. E.;<br />

Michl, J. J. Am. Chem. Soc. 1983, 105,<br />

3638.<br />

31 (a) Morf, J.; Szeimies, G. Tetrahedron<br />

Lett. 1986, 27, 5363. (b) Fuchs, J.;<br />

Szeimies, G. Chem. Ber. 1992, 125, 2517.<br />

(c) Ströter, T.; Szeimies, G. J. Am. Chem.<br />

Soc. 1999, 121, 7476.<br />

32 (a) Jackson, J. E.; Allen, L. C. J. Am.<br />

Chem. Soc. 1984, 106, 591. (b) Epiotis,<br />

N. D. J. Am. Chem. Soc. 1984, 106, 3170.<br />

33 (a) Zilberg, S. P.; Ioffe, A. I.;<br />

Nefedov, O. M. Izv. Akad. Nauk SSSR,<br />

Ser. Khim. 1984, 358. (b) Ushio, T.;<br />

Kato, T.; Ye, K.; Imamura, A. Tetrahedron<br />

1989, 45, 7743.<br />

34 (a) Wiberg, K. B. J. Am. Chem. Soc.<br />

1983, 105, 1227. (b) Wiberg, K. B.;<br />

Bader, R. F. W.; Lau, C. D. H.<br />

J. Am. Chem. Soc. 1987, 109, 985.<br />

(c) Wiberg, K. B.; Bader, R. F. W.;<br />

Lau, C. D. H. J. Am. Chem. Soc. 1987,<br />

109, 1001.<br />

35 Feller, D.; Davidson, E. R. J. Am. Chem.<br />

Soc. 1987, 109, 4133.<br />

36 (a) Dodziuk, H. Tetrahedron 1988, 44,<br />

2951. (b) Dodziuk, H. J. Comput. Chem.<br />

1984, 5, 571.<br />

37 Michl, J.; Radziszewski, G. J.;<br />

Downing, J. W.; Wiberg, K. B.;<br />

Walker, F. H.; Miller, R. D.; Kova�i�, P.;<br />

Jawdosiuk, M.; Bona�i�-Koutecky, V.<br />

Pure Appl. Chem. 1983, 55, 315.<br />

38 Pierini, A. B.; Reale, H. F.;<br />

Medrano, J. A. J. Mol. Struct. (Theochem)<br />

1986, 148, 109.<br />

39 Messerschmidt, M.; Scheins, S.;<br />

Grubert, L.; Pätzel, M.; Szeimies, G.;<br />

Paulmann, C.; Luger, P. Angew. Chem.<br />

Int. Ed. 2005, 44, 3925.<br />

40 Kar, T.; Jug, K. Chem. Phys. Lett. 1996,<br />

256, 201.<br />

41 (a) Lee, I.; Yang, K.; Kim, H. S.<br />

Tetrahedron 1985, 41, 5007.<br />

93


94 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

(b) Messmer, R. P.; Schultz, A. P.<br />

J. Am. Chem. Soc. 1986, 108, 7407.<br />

42 (a) Szeimies-Seebach, U.; Szeimies, G.;<br />

Van Meerssche, M.; Germain, G.;<br />

Declercq, J.-P. Nouv. J. Chim. 1979,<br />

3, 357. (b) Chakrabarti, P.; Seiler, P.;<br />

Dunitz, J. D.; Schülter, A.-D.;<br />

Szeimies, G. J. Am. Chem. Soc. 1981,<br />

103, 7378. (c) Dunitz, J. D.; Seiler, P.<br />

J. Am. Chem. Soc. 1983, 105, 7056.<br />

43 Hedberg, L.; Hedberg, K. J. Am. Chem.<br />

Soc. 1985, 107, 7257.<br />

44 Seiler, P.; Belzner, J.; Bunz, U.;<br />

Szeimies, G. Helv. Chim. Acta 1988, 71,<br />

2100.<br />

45 (a) Wiberg, K. B.; Dailey, W. P.;<br />

Walker, F. H.; Waddell, S. T.;<br />

Crocker, L. S.; Newton, M. J. Am.<br />

Chem. Soc. 1985, 107, 7247.<br />

(b) Wiberg, K. B.; Waddell, S. T.;<br />

Rosenberg, R. E. J. Am. Chem. Soc.<br />

1990, 112, 2184. (c) Bistri�i�, L.;<br />

Baranovi�, G.; Šafar Cvitaš, D.; Mlinari�-<br />

Majerski, K. J. Phys. Chem. A 1997,<br />

101, 941. (d) Jensen, J. O. J. Mol. Struct.<br />

(Theochem) 2004, 673, 51.<br />

46 (a) Honegger, E.; Huber, H.;<br />

Heilbronner, E.; Dailey, W. P.;<br />

Wiberg, K. B. J. Am. Chem. Soc. 1985,<br />

107, 7172. (b) Eckert-Maksi�, M.;<br />

Mlinari�-Majerski, K.; Majerski, Z.<br />

J. Org. Chem. 1987, 52, 2098.<br />

(c) Gleiter, R.; Pfeifer, K.-H.;<br />

Szeimies, G.; Belzner, J.; Lehne, K.<br />

J. Org. Chem. 1990, 55, 636.<br />

47 (a) Schafer, O.; Allan, M.; Szeimies, G.;<br />

Sanktjohanser, M. J. Am. Chem. Soc.<br />

1992, 114, 8180. (b) Winstead, C.;<br />

Sun, Q.; McKoy, V. J. Chem. Phys. 1992,<br />

97, 9483. (c) Allan, M. J. Chem. Phys.<br />

1994, 101, 844.<br />

48 Orendt, A. M.; Facelli, J. C.;<br />

Grant, D. M.; Michl, J.; Walker, F. H.;<br />

Dailey, W. P.; Waddell, S. T.;<br />

Wiberg, K. B.; Schindler, M.;<br />

Kutzelnigg, W. Theor. Chim. Acta, 1985,<br />

68, 421.<br />

49 Jarret, R. M.; Cusumano, L. Tetrahedron<br />

Lett. 1990, 31, 171.<br />

50 Werner, M.; Stephenson, D. S.;<br />

Szeimies, G. Liebigs Ann. 1996, 1705.<br />

51 Pecul, M.; Dodziuk, H.; Jaszunski, M.;<br />

Lukin, O.; Leszczy�ski, J. Phys. Chem.<br />

Chem. Phys. 2001, 3, 1986.<br />

52 Galasso, V. Chem. Phys. Lett. 1994, 230,<br />

387.<br />

53 Ebrahimi, A.; Deyhimi, F.; Roohi, H.<br />

J. Mol. Struct. (Theochem) 2003, 626,<br />

223.<br />

54 Levine, M. D.; Kaszynski, P.; Michl, J.<br />

Chem. Rev. 2000, 100, 169.<br />

55 Coppens, P. Angew. Chem. Int. Ed. 2005,<br />

44, 6810.<br />

56 Wiberg, K. B.; McMurdie, N. J. Am.<br />

Chem. Soc., 1991, 113, 8995.<br />

57 Wiberg, K. B. J. Comput. Chem. 1984, 5,<br />

197.<br />

58 Jarosch, O.; Walsh, R.; Szeimies, G.<br />

J. Am. Chem. Soc. 2000, 122, 8490.<br />

59 Wiberg, K. B.; Waddell, S. T.; Laidig, K.<br />

Tetrahedron Lett. 1986, 27, 1553.<br />

60 (a) Kaszynski, P.; Michl, J. J. Org.<br />

Chem. 1988, 53, 4593. (b) Kaszynski, P.;<br />

McMurdie, N. D.; Michl, J. J. Org. Chem.<br />

1991, 56, 307.<br />

61 Bunz, U.; Polborn, K.; Wagner, H.-U.;<br />

Szeimies, G. Chem. Ber. 1988, 121, 1785.<br />

62 (a) McGarry, P. F.; Johnston, L. J.;<br />

Scaiano, J. C. J. Org. Chem. 1989, 54,<br />

6133. (b) McGarry, P. F.; Scaiano, J. C.<br />

Can. J. Chem. 1998, 76, 1474.<br />

63 Wiberg, K. B.; Waddell, S. T. Tetrahedron<br />

Lett. 1987, 28, 151.<br />

64 Jarosch, O.; Szeimies, G. J. Org. Chem.<br />

2003, 68, 3797.<br />

65 (a) Kaszynski, P.; Michl, J. J. Am. Chem.<br />

Soc. 1988, 110, 5225. (b) Hassenrück, K.;<br />

Murthy, G. S.; Lynch, V. M.; Michl, J.,<br />

J. Org. Chem. 1990, 55, 1013.<br />

66 Schlüter, A.-D. Polymer Commun. 1989,<br />

30, 34.<br />

67 (a) Schlüter, A. D. Angew. Chem. Int. Ed.<br />

Engl. 1988, 27, 296. (b) Schlüter, A.-D.<br />

Macromolecules 1988, 21, 1208.<br />

68 (a) Friedli, A. C.; Kaszynski, P.;<br />

Michl, J. Tetrahedron Lett. 1989, 30, 455.<br />

(b) Kaszynski, P.; Friedli, A. C.; Michl, J.<br />

J. Am. Chem. Soc. 1992, 114, 601.<br />

69 Patzel, M.; Sanktjohanser, M.; Doss, A.;<br />

Henklein, P.; Szeimies, G. Eur. J. Org.<br />

Chem. 2004, 493.<br />

70 Lee, W. B.; Oh, D. W. Bull. Korean Chem.<br />

Soc. 1999, 20, 629.<br />

71 Wiberg, K. B. Tetrahedron Lett. 1985, 26,<br />

5967.<br />

72 Dodziuk, H.; Leszczy�ski, J.;<br />

Jackowski, K. J. Org. Chem. 1999, 64,<br />

6177.


73 van’t Hoff, J. H. Arch. Nederl. Sci. Exactes<br />

Nat. 1874, 445.<br />

74 Le Bel, J. A. Bull. Soc. Chim. Fr. 1874, 22,<br />

337.<br />

75 Hoffmann, R. Pure Appl. Chem. 1972, 28,<br />

181.<br />

76 Hoffmann, R.; Alder, R.; Wilcox, J.,<br />

C. F., J. Am. Chem. Soc. 1970, 92, 4992.<br />

77 Liebman, J. F.; Greenberg, A. Chem. Rev.<br />

1976, 76, 311.<br />

78 Liebman, J. F.; Greenberg, A. <strong>Strained</strong><br />

Organic Molecules; Academic Press: New<br />

York, 1978.<br />

79 Minkin, V. I.; Minyaev, R. M.; Zhdanov,<br />

Y., A., Nonclassical Structures of Organic<br />

Compounds; Mir: Moscow, 1987.<br />

80 Dodziuk, H. Top. Stereochem. 1994, 21,<br />

351.<br />

81 Sorger, K.; Schleyer, P. v. R. J. Mol.<br />

Struct. (Theochem) 1995, 338, 317.<br />

82 Minkin, V. I.; Minayev, R. M.;<br />

Hoffmann, R. Usp. Khim. 2002, 71, 989.<br />

83 Dodziuk, H. Modern Conformational<br />

Analysis. Elucidating Novel Exciting<br />

Molecular Structures; VCH Publishers:<br />

New York, 2002.<br />

84 Collins, J. B.; Dill, J. D.; Jemmis, E. D.;<br />

Apeloig, Y.; Schleyer, P. v. R.; Seeger, R.;<br />

Pople, J. A. J. Am. Chem. Soc. 1976, 98,<br />

5419.<br />

85 Su, M.-D. Inorg. Chem. 2005, 44, 4829.<br />

86 Stucky, G. D.; Eddy, M. M.; Harrison,<br />

W. H.; Lagow, R. J.; Kawa, H.; Cox, D. E.<br />

J. Am. Chem. Soc. 1990, 112, 2425.<br />

87 Pepper, M. J. M.; Shavitt, I.;<br />

Schleyer, P. v. R.; Glukhotsev, M. N.;<br />

Janoschek, R.; Quack, M. J. Comput.<br />

Chem. 1995, 16, 207.<br />

88 Venepalli, B. R.; Agosta, W. C. Chem.<br />

Rev. 1987, 87, 399.<br />

89 Keese, R. Chem. Rev. 2006, 106, 4787.<br />

90 Eaton, P. E.; Leipzig, B. D. J. Am. Chem.<br />

Soc. 1983, 105, 1656.<br />

91 Wiberg, K. B. Tetrahedron Lett. 1985, 26,<br />

5967.<br />

92 Minayev, R. M.; Minkin, V. I.;<br />

Gribanova, T. N.; Starikov, A. G.;<br />

Hoffmann, R. J. Org. Chem. 2003, 68,<br />

8588.<br />

93 Rao, V. B.; George, C. F.; Wolff, S.;<br />

Agosta, W. C. J. Am. Chem. Soc. 1985,<br />

107, 5732.<br />

94 Lee, W. B.; Oh, D. W. Bull. Krean Chem.<br />

Soc. 1999, 20, 629.<br />

References<br />

95 Dodziuk, H. J. Mol. Struct. 1990, 239, 167.<br />

96 Osawa, E.; Musso, H. Angew. Chem. Int.<br />

Ed. 1983, 22, 1.<br />

97 Allinger, N. L. Adv. Phys. Org. Chem.<br />

1976, 13, 1.<br />

98 Dodziuk, H.; Lipkowitz, K. In: 2nd<br />

WATOC Congress, Toronto, Canada,<br />

1992.<br />

99 McGrath, M. P.; Radom, L.; Schäfer,<br />

H. F., III, J. Org. Chem. 1992, 57, 4847.<br />

100 McGrath, M. P.; Radom, L. J. Am. Chem.<br />

Soc. 1993, 115, 3320.<br />

101 Radom, L.; Rasmussen, D. R. Angew.<br />

Chem. Int. Ed. 1999, 38, 2876.<br />

102 Rasmussen, D. R.; Radom, L. Chem. Eur.<br />

J. 2000, 6, 2470.<br />

103 Greenberg, A.; Liebman, J. F. <strong>Strained</strong><br />

Organic Molecules; Academic Press: New<br />

York, 1978.<br />

104 Mehta, G.; Padma, S. In Carbocyclic Cage<br />

Compounds; Chemistry and Applications;<br />

Osawa, E., Yonemitsu, O., Eds.; VCH<br />

Publishers Inc.: New York, 1992, p. 183.<br />

105 Dodziuk, H. In Topics in Stereochemistry,<br />

Eliel E. L., Wilen, S. H., Eds.; Vol. 21,<br />

p. 351; John Wiley & Sons: New York,<br />

1994.<br />

106 Eaton, P. E.; Cole, T. W., Jr. J. Am. Chem.<br />

Soc. 1964, 86, 962.<br />

107 Eaton, P. E.; Cole, T. W., Jr. J. Am. Chem.<br />

Soc. 1964, 86, 3157.<br />

108 Eaton, P. E. Angew. Chem. Int. Ed. Engl.<br />

1992, 31, 1421.<br />

109 Griffin, G. W.; Marchand, A. P. Chem.<br />

Rev. 1989, 89, 997.<br />

110 Higuchi, H.; Ueda, I.; In: Carbocyclic<br />

Cage Compounds; Chemistry and<br />

Applications; Osawa, E. J., Yonemitsu,<br />

O., Eds.; VCH Publishers Inc: New York,<br />

1992, p. 217.<br />

111 Bashir-Hashemi, A. Carbocyclic and<br />

Heterocyclic Cage Compounds and Their<br />

Building Blocks. Advances in <strong>Strained</strong> and<br />

Interesting Organic Molecules; Laali, K.,<br />

Ed.; JAI Press Inc.: Stamford, CT, 1999,<br />

p. 1.<br />

112 Katz, T. J.; Acton, N. J. Am. Chem. Soc.<br />

1973, 95, 2738.<br />

113 Eaton, P. E.; Or, Y. S.; Branca, S. J.<br />

J. Am. Chem. Soc. 1981, 103, 2134.<br />

114 Forman, M. A.; Dailey, W. P. J. Org.<br />

Chem. 1993, 58, 1501.<br />

115 Disch, R. L.; Schulman, J. M. J. Am.<br />

Chem. Soc. 1990, 112, 3377.<br />

95


96 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

116 Wu, H.-S.; Qin, X.-F.; Xu, X.-H.; Jiao, H.;<br />

Schleyer, P. v. R. J. Am. Chem. Soc. 2005,<br />

127, 2334.<br />

117 Dodziuk, H.; Nowi�ski, K. Bull. Polish<br />

Acad. Sci. 1987, 35, 195.<br />

118 Yildirim, T.; Ciraci, S.; Kılıç, Ç.;<br />

Buldum, A. Phys. Rev. B 2000, 62, 7625.<br />

119 Disch, R. L.; Schulman, J. M. J. Am.<br />

Chem. Soc. 1988, 110, 2102.<br />

120 Dinadayalane, T. C.; Priyakumar, U. D.;<br />

Sastry, G. N. J. Phys. Chem. A 2004, 108,<br />

11433.<br />

121 Jenkins, S. J.; King, D. A. Chem. Phys.<br />

Lett. 2000, 317, 381.<br />

122 Gribanova, T. N.; Minyaev, R. M.;<br />

Minkin, V. I. Doklady Chemistry 2006,<br />

411, 193.<br />

123 Maier, G.; Bauer, I.; Huber-Patz, U.;<br />

Jahn, R.; Kallfass, D.; Rodewald, H.;<br />

Irngartinger, H. Chem. Ber. 1986, 119,<br />

1111.<br />

124 Engel, P.; Eaton, P. E.; Shankar,<br />

B. K. R. Z. Kristall. 1982, 159, 239.<br />

125 Hedberg, L.; Hedberg, K.; Eaton, P. E.;<br />

Nodari, N.; Robiette, A. G. J. Am. Chem.<br />

Soc. 1991, 113, 1514.<br />

126 Hirota, E.; Endo, Y.; Fujitake, M.; Della,<br />

E. W.; Pigou, P. E.; Chickos, J. S. J. Mol.<br />

Struct. 1988, 190, 235.<br />

127 Chen, K.-H.; Allinger, N. L. J. Mol.<br />

Struct. (Theochem) 2002, 581, 215.<br />

128 Nagase, S. Acc. Chem. Res. 1995, 28, 469.<br />

129 Kybett, B. D.; Carroll, S.; Natalis, P.;<br />

Bonnell, D. W.; Margrave, J. L.; Franklin,<br />

J. L. J. Am. Chem. Soc. 1966, 88, 626.<br />

130 Moran, D.; Manoharan, M.; Heine, T.;<br />

Schleyer, P. v. R. Org. Lett. 2003, 5, 23.<br />

131 Minkin, V. I.; Minyaev, R. M.; Zhdanov,<br />

Yu. A. Nonclassical Structures of Organic<br />

Compounds; Mir: Moscow, 1987.<br />

132 Dill, J. D.; Greenberg, A.; Liebman, J. F.<br />

J. Am. Chem. Soc. 1979, 101, 6814.<br />

133 Novak, I. Chem. Phys. Lett. 2003, 380, 258.<br />

134 Gilardi, R.; Butcher, R. J. J. Chem.<br />

Crystallogr. 2003, 33, 281.<br />

135 Gejji, S. P.; Patil, U. N.; Dhumal, N. R.<br />

J. Mol. Struct. (Theochem) 2005, 718, 237.<br />

136 Hrovat, D. A.; Borden, W. T.; Eaton, P. E.;<br />

Kahr, B. J. Am. Chem. Soc. 2001, 123,<br />

1289.<br />

137 Kortus, J.; Pederson, M. R.; Richardson,<br />

S. L. Chem. Phys. Lett. 2000, 322, 224.<br />

138 Zhang, M.-X.; Eaton, P. E.; Gilardi, R.<br />

Angew. Chem., Int. Ed. 2000, 39, 401.<br />

139 Musso, H.; Biethan, U. Chem. Ber. 1967,<br />

100, 119.<br />

140 Fritz, H.-G.; Hutmacher, H.-M.;<br />

Musso, H.; Ahlgren, G.; Akermark, B.;<br />

Karlson, R. Chem. Ber. 1976, 109, 3781.<br />

141 Hoffmann, V. T.; Musso, H. Angew.<br />

Chem. Int. Ed. Engl. 1987, 26, 1006.<br />

142 Ahlquist, B.; Almenningen, A.;<br />

Benterud, B.; Trætteberg, M.; Bakken, P.;<br />

Lüttke, W. Chem. Ber. 1992, 125, 1217.<br />

143 Osawa E., Musso, H. Angew. Chem. 1983,<br />

95, 1.<br />

144 Wang, Y.; Huang, Y.; Liu, R. Chem.<br />

Eur. J. 2006, 12, 3610.<br />

145 Gribanova, T. N.; Minyaev, R. M.;<br />

Minkin, V. I. Izv. Akad. Nauk. Ser. Khim.<br />

2006, 55, 1825.<br />

146 Gribanova, T. N.; Minyaev, R. M.;<br />

Minkin, V. I. Doklady Chemistry 2007,<br />

412, 1.<br />

147 Manini, P.; Amrein, W.; Gramlich, V.;<br />

Diederich, F. Angew. Chem., Int. Ed.<br />

2002, 41, 4339.<br />

148 Bachrach, S. M. J. Phys. Chem. A 2003,<br />

107, 4957.<br />

149 Jarowski, P. D.; Diederich, F.;<br />

Houk, K. N. J. Org. Chem. 2005, 70, 1671.<br />

150 Bachrach, S. M.; Demoin, D. W.<br />

J. Org. Chem. 2006, 71, 5105.<br />

151 Moran, D.; Woodcock, H. L.; Chen, Z.;<br />

Schaefer, H. F., III; Schleyer, P. v. R.<br />

J. Am. Chem. Soc. 2003, 125, 11442.<br />

152 Schubert, W.; Yoshimine, M.; Pacansky, J.<br />

J. Phys. Chem. 1981, 85, 1340.<br />

153 Hrovat, D. A.; Borden, W. T.<br />

J. Am. Chem. Soc. 1988, 110, 4710.<br />

154 Eaton, P. E.; Maggini, M. J. Am. Chem.<br />

Soc. 1988, 110, 7230.<br />

155 Staneke, P. O.; Ingemann, S.; Eaton, P.;<br />

Nibbering, N. M. M.; Kass, S. R.<br />

J. Am. Chem. Soc. 1994, 116, 6445.<br />

156 Lukin, K.; Eaton, P. E. J. Am. Chem. Soc.<br />

1995, 117, 7652.<br />

157 Hrovat, D. A.; Borden, W. T.<br />

J. Am. Chem. Soc. 1990, 112, 875.<br />

158 Visser, S. P.; Filatov, M.; Schreiner, P. R.;<br />

Shaik, S. Eur. J. Org. Chem. 2003, 4199.<br />

159 Hassenrück, K.; Radziszewski, J. G.;<br />

Balaji, V.; Murthy, G. S.; McKinley, A. J.;<br />

David, D. E.; Lynch, V. M.; Martin, H.-D;<br />

Michl, J. J. Am. Chem. Soc. 1990, 112,<br />

873.<br />

160 Eaton, P. E.; Tsanaktsidis, J.<br />

J. Am. Chem. Soc. 1990, 112, 876.


161 Jonas, V.; Frenking, G. J. Org. Chem.<br />

1992, 57, 6085.<br />

162 Gleiter, R.; Ohlbach, F. J. Org. Chem.<br />

1996, 61, 4929.<br />

163 Barone, P. M. V. B.; Camilo, A., Jr.;<br />

Galvão, D. S. Synth. Met. 1999, 102,<br />

1410.<br />

164 Eaton, P. E.; Pramod, K.; Emrick, T.;<br />

Gilardi, R. J. Am. Chem. Soc. 1999, 121,<br />

4111.<br />

165 Gilardi, R.; Maggini, M.; Eaton, P. E.<br />

J. Am. Chem. Soc. 1988, 110, 7232.<br />

166 Eaton, P. E.; Li, J.; Upadhyaya, S. P.<br />

J. Org. Chem. 1995, 60, 966.<br />

167 Herrera, B., Toro-Labbé, A. Chem. Phys.<br />

Lett. 2001, 344, 193.<br />

168 Valencia, F.; Romero, A. H.; Kiwi, M.;<br />

Ramírez, R.; Toro-Labbe, A. J. Chem.<br />

Phys. 2004, 121, 9172.<br />

169 Herrera, B.; Valencia, F.; Romero, A. H.;<br />

Kiwi, M.; Ramírez, R.; Toro-Labbé, A.<br />

J. Mol. Struct. (Theochem) 2006, 769,<br />

183.<br />

170 Valencia, F.; Romero, A. H.; Kiwi, M.;<br />

Ramírez, R.; Toro-Labbe, A. Phys. Rev. B<br />

2005, 71, 033410.<br />

171 Burdett, J. K.; Lee, S. J. Am. Chem. Soc.<br />

1985, 107, 3063.<br />

172 Johnston, R. L.; Hoffmann, R. J. Am.<br />

Chem. Soc. 1989, 111, 810.<br />

173 Winkler, B.; Milman, V. Chem. Phys. Lett.<br />

1998, 293, 284.<br />

174 Vlahacos, C. P.; Jensen, J. O. J.<br />

Mol. Struct. (Theochem) 1996, 362, 225.<br />

175 Pichierri, F. Internet Electron. J. Mol. Des.<br />

2004, 3, 134.<br />

176 Seidl, E. T.; Schaefer, H. F., III.<br />

J. Am. Chem. Soc. 1991, 113, 1915.<br />

177 Minyaev, R. M.; Minkin, V. I.;<br />

Gribanova, T. N.; Starikov, A. G.;<br />

Hoffmann, R. J. Org. Chem. 2003, 68,<br />

8588.<br />

178 Liu, F.-L.; Peng, L. J. Mol. Struct.<br />

(Theochem) 2004, 710, 163.<br />

179 Pour, N.; Itzhaki, L.; Hoz, B.; Altus, E.;<br />

Basch, H.; Hoz, S. Angew. Chem. Int. Ed.<br />

2006, 45, 5981.<br />

180 Liu, F.-L. J. Mol. Struct. (Theochem)<br />

2004, 681, 51.<br />

181 Liu, F.-L.; Xie, Y.; Dai, L.-H., Peng, L.<br />

J. Mol. Struct. (Theochem) 2004, 710, 25.<br />

182 Liu, F.-L.; Zhai, Y.-Q.; Feng, S.; Guo,<br />

W.-L. J. Mol. Struct. (Theochem) 2005,<br />

719, 185.<br />

References<br />

183 Hoffmann, V. T.; Musso, H. Chem. Ber.<br />

1991, 124, 103.<br />

184 Schulman, J. M.; Disch, R. L. J.<br />

Mol. Struct. (Theochem) 1995, 358, 51.<br />

185 Liu, F.-L.; Feng, S.; Guo, W.-L.;<br />

Zhai, Y.-Q. J. Mol. Struct. (Theochem)<br />

2005, 725, 247.<br />

186 Tantillo, D. J.; Hoffmann, R.<br />

Angew. Chem. Int. Ed. 2002, 41, 1033.<br />

187 Liu, F.-L.; Xie, Y.; Dai, L.-H. J.<br />

Mol. Struct. (Theochem) 2004, 683, 51.<br />

188 Heath, T. A. A History of Greek<br />

Mathematics London, 1976.<br />

189 Eaton, P. E.; Cassar, L.; Halpern, J.<br />

J. Am. Chem. Soc. 1970, 92, 6366.<br />

190 Bertau, M.; Weiler, A.; Wahl, F.;<br />

Scheumann, K.; Worth, J.; Keller, M.;<br />

Prinzbach, H. Tetrahedron 1997, 53,<br />

10029.<br />

191 Cassar, L., Eaton, P. E.; Halpern, J.<br />

J. Am. Chem. Soc. 1970, 92, 6366.<br />

192 Beesly, R., M.; Thoprpe, J. F. J. Chem.<br />

Soc. 1920, 591.<br />

193 Beesly, R., M.; Thoprpe, J. F. Proc. Chem.<br />

Soc. 1913, 29, 346.<br />

194 Zefirov, I. S.; Kozmin, A. S.;<br />

Abramenkov, A. B. Usp. Khim. Engl.<br />

transl. 1978, 47, 163.<br />

195 Maier, G.; Pfriem, S.; Schäfer, U.<br />

Angew. Chem. Int. Ed. 1978, 17, 520.<br />

196 Maier, G.; Born, D. Angew. Chem. Int. Ed.<br />

1989, 28, 1050.<br />

197 Tanaka, M.; Sekiguchi, A. Angew. Chem,<br />

Int. Ed. Engl. 2005, 44, 5821.<br />

198 de Meijere, A.; Kozhushkov, S. I.;<br />

Schill, H. Chem. Rev. 2006, 106, 4926.<br />

199 Irngartinger, H.; Goldmann, A.; Jahn,<br />

R.; Nixdorf, M.; Rodewald, H.; Maier, G.;<br />

Malsch, K. D.; Emrich, R. Angew. Chem,<br />

Int. Ed. Engl. 1984, 23, 993.<br />

200 Maier, G.; Pfriem, S.; Malsch, K. D.;<br />

Kalinowski, H.-O.; Dehnicke, K.<br />

Chem. Ber. 1981, 114, 3988.<br />

201 Loerer, T.; Machinek, R.; Lüttke, W.;<br />

Frany, L. H.; Malsch, K. D.; Maier, G.<br />

Angew. Chem, Int. Ed. Engl. 1983, 22,<br />

878.<br />

202 Minkin, V. I.; Minayev, R. M.; Zhdanov,<br />

Y. A. In: Nonclassical Structures of Organic<br />

Compounds; Mir: Moscow, 1987, p. 68.<br />

203 Osawa, E.; Musso, H. Angew. Chem. Int.<br />

Ed. 1983, 22, 1.<br />

204 Hounshell, W. D.; Mislow, K. Tetrahedron<br />

Lett. 1979, 14, 1205.<br />

97


98 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

205 Nemirowski, A.; Reisenauer, H. P.;<br />

Schreiner, P. R. Chem. Eur. J. 2006, 12,<br />

7411.<br />

206 Tanner, M. E., Knobler, C. B., Cram, D. J.<br />

Angew. Chem., Int. Ed. 1991, 30, 1924.<br />

207 Krivdin, L. B. Magn. Res. Chem. 2004, 42,<br />

919.<br />

208 Mo, Y. R. Org. Lett. 2006, 8, 535.<br />

209 Katz, T. J.; Acton, N. J. Am. Chem. Soc.<br />

1973, 95, 2738.<br />

210 Griffin, G. W.; Marchand, A. P. Chem.<br />

Rev. 1989, 89, 997.<br />

211 Hassenrück. K.; Martin, H.-D.; Walsh, R.<br />

Chem. Rev. 1989, 89, 1125.<br />

212 Hedberg, L.; Hedberg, K.; Eaton, P. E.<br />

J. Am. Chem. Soc. 1991, 113, 1514.<br />

213 Almenningen, A.; Jonvil, L. T.;<br />

Martin, H. D.; Urbanek, T. J. Mol. Struct.<br />

1985, 128, 239.<br />

214 Hirota, E.; Endo, Y.; Fujitake, M.;<br />

Della, E. W.; Pigou, P. E.; Chikos, J. S.<br />

J. Mol. Struct. 1988, 190, 235.<br />

215 Fleischer, E. B. J. Am. Chem. Soc. 1964,<br />

86, 3889.<br />

216 Stober, R.; Musso, H.; Osawa, E.<br />

Tetrahedron 1986, 42, 1757.<br />

217 Gleiter, R.; Karcher, M. Angew. Chem,<br />

Int. Ed. Engl. 1988, 27, 840.<br />

218 Eaton, P. E.; Maggini, M. J. Am. Chem.<br />

Soc. 1988, 110, 7230.<br />

219 Dodziuk, H. Bull. Pol. Acad. Sci. Chem.<br />

1990, 38, 11.<br />

220 Iwamura, H.; Morio, K.; Kunii, T. L.<br />

J. Chem. Soc. D 1971, 1408.<br />

221 Irngartinger, H.; Strack, S.; Gleiter, R.;<br />

Brand, S. Acta Cryst. C 1997, 53, 1145.<br />

222 Roux, M. V.; Davalos, J. Z.; Jimenez, P.;<br />

Notario, R.; Castano, O.; Chickos, J. S.;<br />

Hanshaw, W.; Zhao, H.; Rath, N.;<br />

Liebman, J. F.; Farivar, B. S.;<br />

Bashir-Hashemi, A. J. Org. Chem. 2005,<br />

70, 5461.<br />

223 Eaton, P. E.; Or, Y. S.; Branca, S. J.<br />

J. Am. Chem. Soc. 1981, 103, 2134.<br />

224 Spielmann, W.; Fick, H.-H.;<br />

Meyer, L.-U.; de Meijere, A. Tetrahedron<br />

Lett. 1976, 4057.<br />

225 Kaufmann, D.; Fick, H.-H.;<br />

Schallner, O.; Spielmann, W.;<br />

Meyer, L.-U.; Gölitz, P.; de Meijere, A.<br />

Chem. Ber. 1983, 116, 587.<br />

226 Honegger, E.; Eaton, P. E.;<br />

Shankar, B. K. R.; Heilbronner, E.<br />

Helv. Chim. Acta 1982, 65, 1982.<br />

227 Engel, P.; Eaton, P. E.; Shankar, B. K. R.<br />

Z. Kristall., 1982, 159, 239.<br />

228 de Meijere, A.; Lee, C. H.; Bengtson, B.;<br />

Pohl, E.; Kozhushkov, S. I.;<br />

Schreiner, P. R.; Boese, R.; Haumann, T.<br />

Chem. Eur. J. 2003, 9, 5481.<br />

229 Spanget-Larsen, J.; Gleiter, R.<br />

Angew. Chem. Int. Ed. 1978, 17, 441.<br />

230 Schulman, J. M.; Miller, M. A.;<br />

Disch, R. L. J. Mol. Struct. (Theochem)<br />

1988, 169, 563.<br />

231 Brousseau, R. J.; PhD Thesis, Harvard<br />

University, 1977.<br />

232 Vedejs, E.; Wilber, W. R.; Twieg, R.<br />

J. Org. Chem. 1977, 42, 401.<br />

233 Vedejs, E.; Shepherd, R. A. J. Org. Chem.<br />

1976, 41, 742.<br />

234 Mehta, G.; Padma, S. J. Am. Chem. Soc.<br />

1987, 109, 2212.<br />

235 Horao, K. I.; Ohuchi, Y.; Yonemistu, O.<br />

J. Chem. Soc. Chem. Commun. 1982, 99.<br />

236 Lee, C. H.; Liang, S.; Haumann, T.;<br />

Boese, R.; de Meijere, A. Angew. Chem,<br />

Int. Ed. Engl. 1993, 32, 559.<br />

237 de Meijere, A.; Lee, C. H.;<br />

Kuznetsov, M. A.; Gusev, D. V.;<br />

Kozhushkov, S. I.; Fokin, A. A.;<br />

Schreiner, P. R.; Haumann, T.<br />

Chem. Eur. J. 2005, 11, 6175.<br />

238 Disch, R. L.; Schulman, J. M.<br />

J. Am. Chem. Soc. 1990, 112, 3377.<br />

239 Schulman, J. M.; Disch, R. L.; Sabio, M. L.<br />

J. Am. Chem. Soc. 1986, 108, 3258.<br />

240 Dodziuk, H.; Nowinski, K. Bull. Pol.<br />

Acad. Sci. Chem. 1987, 35, 195.<br />

241 Mehta, G.; Padma, S.; Osawa, E.;<br />

Barbiric, D. A.; Mochizuki, Y.<br />

Tetrahedron Lett. 1987, 28, 1295.<br />

242 Osawa, E.; Barbiric, D. A.; Lee, O. S.;<br />

Hitano, Y.; Padma, S.; Mehta, G.<br />

J. Chem. Soc. Perkin Trans 2 1989, 1161.<br />

243 Mehta, G.; Padma, S. Tetrahedron 1991,<br />

47, 7783.<br />

244 Chou, T. C.; Lin, G. H.; Yeh, Y. L.;<br />

Lin, K. J. J. Chin. Chem. Soc. 1997, 44,<br />

477.<br />

245 Hoffmann, R.; Hopf, H. Angew. Chem,<br />

Int. Ed. 2008, 47, 4474.<br />

246 Dinsburg, G. Nouv. J. Chim. 1982, 6, 175.<br />

247 Dodziuk, H. Top. Stereochem. 1994, 21,<br />

351.<br />

248 Ternansky, R. J.; Balogh, D. W.;<br />

Paquette, L. J. Am. Chem. Soc. 1982, 104,<br />

4503.


249 Herzberg, G., Infrared and Raman<br />

Spectra of Polyatomic Molecules, Van<br />

Nostrand Co., Priston, 1945, p. 12.<br />

250 Paquette, L. Chem. Rev. 1989, 89, 1051.<br />

251 Saunders, M. Science 1991, 253, 330.<br />

252 Yoshida, Z.-I.; Dogane, I.; Ikehira, H.;<br />

Endo, T. Chem. Phys. Lett. 1991, 201, 481.<br />

253 Dunlap, B. I.; Brenner, D. W.; Mintmire,<br />

J. W.; Mowrey, R. C.; White, C. T. J. Phys.<br />

Chem. 1991, 95, 5763.<br />

254 Guo, T.; Scuseria, G. E. Chem. Phys. Lett.<br />

1992, 191, 527.<br />

255 Dodziuk, H. Tetrahedron 1998, 54, 2917.<br />

256 Dodziuk, H.; Lukin, O.; Nowinski, K.<br />

Pol. J. Chem. 1999, 73, 299.<br />

257 Stoddart, J. F. Angew. Chem. Int. Ed.<br />

1991, 30, 70.<br />

258 Taylor, R.; Avent, A. G.; Dennis, T. J.;<br />

Hare, J. P.; Kroto, H. W.; Walton,<br />

D. R. M.; Holloway, J. H.; Hope, E. G.;<br />

Langley, G. J. Nature 1992, 355, 27.<br />

259 Dodziuk, H., Nowinski, K., Chem. Phys.<br />

Lett. 1995, 249, 406.<br />

260 Kiran, B.; Kandalam, A. K.; Jena, P.<br />

J. Chem. Phys. 2006, 124, 224703.<br />

261 Siadat, S. O. R.; Mahkam, M.;<br />

Mohammadi, R. Asian J. Chem. 2007, 19,<br />

1875.<br />

262 Kabuto, C. H.; Yagihara, M.; Asso, T.;<br />

Kitahara, Y. Angew. Chem. Int. Ed. 1973,<br />

12, 836.<br />

263 Littke, W., Druck, U., Angew. Chem. Int.<br />

Ed. 1974, 13, 539.<br />

264 Vogel, E.; Breuer, A.; Sommerfield,<br />

C.-D.; Davis, R. E.; Liu, L. K.<br />

Angew. Chem. Int. Ed. 1977, 16, 169.<br />

265 Schwesinger, R.; Piontek, K.; Littke, W.;<br />

Druck, U.; Prinzbach, H. Angew. Chem.<br />

Int. Ed. 1986, 24, 318.<br />

266 Stobbe, M.; Behrens, U.; Adividjaja, G.;<br />

Golitz, P.; de Meijere, A. Angew. Chem.<br />

Int. Ed. 1983, 22, 867.<br />

267 Erden, I.; Golitz, P.; Nader, R.;<br />

de Meijere, A. Angew. Chem. Int. Ed.<br />

1981, 20, 583.<br />

268 Dodziuk, H.; Ostrowski, M. Eur. J. Org.<br />

Chem. 2006, 5231.<br />

269 Dodziuk, H. Bull. Chem. Soc. Japan 1987,<br />

60, 3775.<br />

270 Eliel, E. L.; Wilen, S. H. Stereochemistry of<br />

Organic Compounds; Wiley-Inter science,<br />

1994, p. 686.<br />

271 Sachse, H. Ber. Stsch. Chem. Ges. 1890,<br />

23, 1363.<br />

References<br />

272 Barton, D. H. R. Experientia 1950, 6, 316.<br />

273 Larnaudie J. Phys. Radium 1954, 15, 69.<br />

274 Berlin, A. J.; Jensen, F. R. Chem. Ind.<br />

London 1960, 998.<br />

275 Reeves, L. W.; Stromme, K. O.<br />

Can. J. Chem. 1960, 38, 1241.<br />

276 Kellie, G. M.; Riddell, F. G. Top. Stereochem.<br />

1974, 8, 225.<br />

277 Vereshchagin, A. N. Usp. Khim. Engl.<br />

transl. 1983, 52, 1081.<br />

278 Eichberg, M. J.; Houk, K. N.;<br />

Lehmann, J.; Leonard, P. W.;<br />

Märker, A.; Norton, J. E.; Sawicka, D.;<br />

Vollhardt, K. P.; Whitener, G. D.; Wolff, S.<br />

Angew. Chem, Int. Ed. 2007, 46, 6894.<br />

279 Engelhardt, M.; Lüttke, W. Angew. Chem.<br />

Int. Ed. 1972, 11, 310.<br />

280 Krüger, C.; Roberts, P. J. Cryst. Struct.<br />

Commun. 1974, 3, 459.<br />

281 Rücker, C.; Müller-Bötticher, H.;<br />

Braschwitz, W.-D.; Prinzbach, H.;<br />

Reifenstahl, U.; Irngartinger, H.<br />

Liebigs Ann./Rec. 1997, 976.<br />

282 Mohler, D. L.; Vollhardt, K. P.; Wollf, S.<br />

Angew. Chem, Int. Ed. 1990, 29, 1151.<br />

283 Mohler, D. L.; Vollhardt, K. P.; Wollf, S.<br />

Angew. Chem, Int. Ed. 1995, 34, 563.<br />

284 Dodziuk, H. unpublished results 1992.<br />

285 (a) Kaupp, G.; Boy, J. Angew. Chem., Int.<br />

Ed. Engl. 1997, 36, 48. (b) Komarov, I. V.<br />

Russ. Chem. Rev. 2001, 70, 991.<br />

286 Allen, F. H.; Kennard, O.; Watson, D. G.;<br />

Brammer, L.; Orpen, A. G.; Taylor, R.<br />

J. Chem. Soc., Perkin Trans. 2 1987, S1.<br />

287 Bartell, L. S.; Burgi, H. B. J. Am. Chem.<br />

Soc. 1972, 94, 5239.<br />

288 (a) Toda, F.; Tanaka, K.; Stein, Z.;<br />

Goldberg, I. Acta Crystallogr., Sect C<br />

1996, 52, 177. (b) Baldridge, K. K.;<br />

Kasahara, Y.; Ogawa, K.; Siegel, J. S.;<br />

Tanaka, K.; Toda, F. J. Am. Chem. Soc.<br />

1998, 120, 6167. (c) Toda, F.; Tanaka, K.;<br />

Watanabe, M.; Tamura, K.; Miyahara, I.;<br />

Nakai, T.; Hirotsu, K. J. Org. Chem. 1999,<br />

64, 3102. (d) Toda, F. Eur. J. Org. Chem.<br />

2000, 1377. (e) Tanaka, K.; Takamoto, N.;<br />

Tezuka, Y.; Kato, M.; Toda, F. Tetrahedron<br />

2001, 57, 3761.<br />

289 Kammermeier, S.; Jones, P. G.;<br />

Herges, R. Angew. Chem., Int. Ed. Engl.<br />

1997, 36, 1757.<br />

290 Kawai, H.; Takeda, T.; Fujiwara, K.;<br />

Inabe, T.; Suzuki, T. Cryst. Growth Des.<br />

2005, 5, 2256.<br />

99


100 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

291 Yannoni, N.; Kahr, B.; Mislow, K.<br />

J. Am. Chem. Soc. 1988, 110, 6670.<br />

292 (a) Harada, J.; Ogawa, K.; Tomoda,<br />

S. Chem. Lett. 1995, 24, 751.<br />

(b) Battersby, T. R.; Gantzel, P.;<br />

Baldridge, K. K.; Siegel, J. S. Tetrahedron<br />

Lett. 1995, 36, 845. (c) Dougherty, D. A.;<br />

Choi, C. S.; Kaupp, G.; Buda, A. B.;<br />

Rudzińsky, J. M.; �sawa, E. J. Chem.<br />

Soc., Perkin Trans. 2 1986, 1063.<br />

293 (a) Ehrenberg, M. Acta Crystallogr. 1966,<br />

20, 182. (b) Bianchi, R.; Morosi, G.;<br />

Mugnoli, A.; Simonetta, M. Acta<br />

Crystallogr. 1973, B29, 1196.<br />

294 Stein, M.; Winter, W.; Rieker, A. Angew.<br />

Chem., Int. Ed. Engl. 1978, 17, 692.<br />

295 (a) Harada, J.; Ogawa, K.; Tomoda, S.<br />

J. Am. Chem. Soc. 1995, 117,<br />

4476. (b) Kahr, B.; Mitchell, C. A.;<br />

Chance, J. M.; VernonClark, R.;<br />

Gantzel, P.; Baldridge, K. K.; Siegel, J. S.<br />

J. Am. Chem. Soc. 1995, 117, 4479.<br />

296 The bond in a silabicyclobutane might<br />

be one of the longest C-C bond [1.781 Å]<br />

ever reported, yet its esd (0.015 Å) is too<br />

large to compare the value with other<br />

precisely determined bond lengths:<br />

Fritz, V. G.; Wartanessian, S.; Matern, E.;<br />

Hönle, W.; Schnering, H. G. v. Z. Anorg.<br />

Allg. Chem. 1981, 475, 87.<br />

297 Two-electron four-centered bond (2.9 Å)<br />

is observed in the solid state for the<br />

dyad of tetracyanoethylene anionradicals,<br />

which is in a new category of<br />

C-C bond: Novoa, J. J.; Lafuente, P.;<br />

Del Sesto, R. E.; Miller, J. S. Angew.<br />

Chem., Int. Ed. 2001, 40, 2540.<br />

298 Zavitsas, A. A. J. Phys. Chem. A 2003,<br />

107, 897.<br />

299 Tsuji, T.; Okuyama, M.; Ohkita, M.;<br />

Imai, T.; Suzuki, T. Chem. Commun.<br />

1997, 2151.<br />

300 (a) Vivattene, R. L.; Greene, F. D.;<br />

Cheung, L. D.; Majeste, R.; Trefonas,<br />

L. M. J. Am. Chem. Soc. 1974, 96, 4342.<br />

(b) Kammermeier, S.; Herges, R. Angew.<br />

Chem., Int. Ed. Engl. 1996, 35, 417.<br />

301 Toda, F.; Tanaka, K.; Tamashima, T.;<br />

Kato, M. Angew. Chem., Int. Ed. 1998, 37,<br />

2724.<br />

302 (a) Bettinger, H. F.; Schleyer, P. v. R.;<br />

Schaefer III, H. F. Chem. Commun.<br />

1998, 769. (b) Choi, C. H.; Kertesz, M.<br />

Chem. Commun. 1997, 2199.<br />

303 (a) �sawa, E.; Ivanov, P. M.; Jaime, C.<br />

J. Org. Chem. 1983, 48, 3990.<br />

(b) Dougherty, D. A.; Hounshell, W. D.;<br />

Schlegel, H. B.; Bell, R. A.; Mislow, K.<br />

Tetrahedron Lett. 1976, 17, 3479.<br />

304 (a) �sawa, S.; Sakai, M.; �sawa, E.<br />

J. Phys. Chem. A 1997, 101, 1378.<br />

(b) Suzuki, T.; Ono, K.; Nishida, J.;<br />

Takahashi, H.; Tsuji, T. J. Org. Chem.<br />

2000, 65, 4944. (c) Suzuki, T.; Ono, K.;<br />

Kawai, H.; Tsuji, T. J. Chem. Soc., Perkin<br />

Trans. 2 2001, 1798.<br />

305 (a) Hoffmann, R. Acc. Chem. Res. 1971,<br />

4, 1. (b) Gleiter, R. Angew. Chem., Int.<br />

Ed. Engl. 1974, 13, 696. (c) Paddon-<br />

Row, M. N. Acc. Chem. Res. 1982,<br />

15, 245. (d) Jordan, K. D.; Paddon-<br />

Row, M. N. Chem. Rev. 1992, 92, 395.<br />

306 Heilbronner, E., Muszkat, K. A.<br />

J. Am. Chem. Soc. 1970, 92, 3818.<br />

307 Baldridge, K. K.; Battersby, T. R.;<br />

VernonClark, R.; Siegel, J. S. J. Am.<br />

Chem. Soc. 1997, 119, 7048.<br />

308 Dyker, G.; Körning, J.; Bubentischek, P.;<br />

Jones, P. G. Liebigs Ann./Recueil 1997,<br />

203.<br />

309 Quinkert, G.; Wiersdorff, W.-W.;<br />

Finke, M.; Opitz, K.; von der Haar, F.-G.<br />

Chem. Ber. 1968, 101, 2302.<br />

310 Iwashita, S.; Ohta, E.; Higuchi, H.;<br />

Kawai, H.; Fujiwara, K.; Ono, K.;<br />

Takenaka, M.; Suzuki, T. Chem.<br />

Commun. 2004, 2076.<br />

311 (a) Hounshell, W. D.; Dougherty, D. A.;<br />

Hummel, J. P.; Mislow, K. J. Am. Chem.<br />

Soc. 1977, 99, 1916. (b) Kahr, B.; Engen,<br />

D. V.; Mislow, K. J. Am. Chem. Soc. 1986,<br />

108, 8305.<br />

312 (a) McBride, J. M. Tetrahedron 1974, 30,<br />

2009. (b) Schlenk, W.; Herzenstein, A.;<br />

Weickel, T. Chem. Ber. 1910, 43, 1753.<br />

313 Wittig, G.; Petri, H.; Justus Liebigs Ann.<br />

1933, 505, 17.<br />

314 (a) Suzuki, T.; Nishida, J.; Tsuji, T.<br />

Angew. Chem., Int. Ed. Engl. 1997,<br />

36, 1329. (b) Suzuki, T.; Nishida, J.;<br />

Tsuji, T. Chem. Commun. 1998,<br />

2193. (c) Suzuki, T.; Migita, A.;<br />

Higuchi, H.; Kawai, H.; Fujiwara, K.;<br />

Tsuji, T. Tetrahedron Lett. 2003, 44,<br />

6837. (d) Suzuki, T.; Yamamoto, R.;<br />

Higuchi, H.; Hirota, E.; Ohkita, M.;<br />

Tsuji, T. J. Chem. Soc., Perkin Trans. 2<br />

2002, 1937.


315 (a) Wang, H.; Gabbaï, F. P. Angew.<br />

Chem., Int. Ed. 2004, 43, 184.<br />

(b) Wang, H.; Gabbaï, F. P. Org. Lett.<br />

2005, 7, 283. (c) Takeda, T.; Kawai, H.;<br />

Fujiwara, K.; Suzuki, T. unpublished<br />

results. (d) Kawai, H.; Takeda, T.;<br />

Fujiwara, K.; Suzuki, T. Tetrahedron Lett.<br />

2004, 45, 8289.<br />

316 Suzuki, T.; Nishida, J.; Ohkita, M.;<br />

Tsuji, T. Angew. Chem., Int. Ed. Engl.<br />

2000, 39, 1804.<br />

317 Adcock, J. L.; Gakh, A. A.; Pollitte, J. L.;<br />

Woods, C. J. Am. Chem. Soc. 1992, 114,<br />

3980.<br />

318 (a) Balasubramaniyan, V. Chem.<br />

Rev. 1966, 66, 567. (b) Clough, R. L.;<br />

Kung, W. J.; Marsh, R. E.; Roberts, J. D.<br />

J. Org. Chem. 1976, 41, 3603.<br />

319 Kawai, H.; Takeda, T.; Fujiwara, K.;<br />

Suzuki, T. Chem. Eur. J. 2008, 14, 5780.<br />

320 Harano, K.; Ban, T.; Yasuda, M.;<br />

�sawa, E.; Kanematsu, K. J. Am. Chem.<br />

Soc. 1981, 103, 2310.<br />

321 Takeda, T.; Kawai, H.; Fujiwara, K.;<br />

Suzuki, T., unpublished results.<br />

322 (a) Mori, T.; Inokuchi, H. Bull. Chem. Soc.<br />

Jpn. 1988, 61, 591. (b) Guionneau, P.;<br />

Kepert, C. J.; Rosseinsky, M.;<br />

Chasseau, D.; Gaultier, J.; Kurmoo, M.;<br />

Hursthouse, M. B.; Day, P. J. Mater.<br />

Chem. 1998, 8, 367. (c) Turner, S. S.;<br />

Day, P.; Gelbrich, T.; Hursthouse, M. B.<br />

J. Solid State Chem. 2001, 159, 385.<br />

323 (a) Ogawa, K.; Kasahara, Y.; Ohtani, Y.;<br />

Harada, J. J. Am. Chem. Soc. 1998,<br />

120, 7107. (b) Nazir, H.; Yıldız, M.;<br />

Yılmaz, H.; Tahir, M. N.; Ülkü, D. J. Mol.<br />

Struct. 2000, 524, 241. (c) Kurahashi, M.<br />

Bull. Chem. Soc. Jpn. 1976, 49, 2927.<br />

324 Tables of Interatomic Distances and<br />

Configuration in Molecules and Ions:<br />

Supplement 1956–1959 (Chemical Society<br />

(London) Special Publication No. 18),<br />

(Ed.: L. E. Sutton), The Chemical Society,<br />

London, 1965.<br />

325 Allen, F. H. Acta Cryst. 1981, B37, 890.<br />

326 Kuchitsu, K.; Fukuyama, T.; Morino, Y.<br />

J. Mol. Struct. 1968, 1, 463.<br />

327 Tanimoto, M.; Kuchitsu, K.; Morino, Y.<br />

Bull. Chem. Soc. Jpn. 1971, 44, 386.<br />

328 v. Doering, W. E.; Pomerantz, M.<br />

Tetrahedron Lett. 1964, 961.<br />

329 Masamune, S. J. Am. Chem. Soc. 1964,<br />

86, 735.<br />

References<br />

330 Trotter, J.; Gibbons, C. S.; Nakatsuka, N.;<br />

Masamune, S. J. Am. Chem. Soc. 1967,<br />

89, 2792.<br />

331 Gibbons, C. S.; Trotter, J. J. Chem. Soc.<br />

(A) 1967, 2027.<br />

332 Irngartinger, H.; Lukas, K. L. Angew.<br />

Chem. Int. Ed. Engl. 1979, 18, 694.<br />

333 Irngartinger, H.; Goldmann, A.;<br />

Schappert, R.; Garner, P.; Dowd, P.<br />

J. Chem. Soc. Chem. Commun. 1981,<br />

455.<br />

334 Irngartinger, H.; Goldmann, A. Angew.<br />

Chem. Int. Ed. Engl. 1982, 21, 775.<br />

335 Daud, P.; Garner, P.; Schappert, R.;<br />

Irngartinger, H.; Goldman, A. J. Org.<br />

Chem. 1982, 47, 4240.<br />

336 Irngartinger, H.; Goldmann, A.;<br />

Schappert, R.; Garner, P.; Go, C. L.;<br />

Dowd, P. J. Chem. Soc. Chem. Commun.<br />

1985, 113.<br />

337 Irngartinger, H.; Jahn, H.; Rodewald, H.;<br />

Paik, Y. H.; Dowd, P. J. Am. Chem. Soc.<br />

1987, 109, 6547.<br />

338 Irngartinger, H.; Goldmann, A.;<br />

Huber-Patz, U.; Garner, P.; Paik, Y. H.;<br />

Dowd, P. Acta Cryst. 1988, C44, 1472.<br />

339 Dowd, P.; Irngartinger, H. Chem. Rev.<br />

1989, 89, 985.<br />

340 Levin, M. D.; Kaszynski, P.; Michl, J.<br />

Chem. Rev. 2000, 100, 169.<br />

341 Lee, V. Ya.; Ichinohe, M.; Sekiguchi, A.<br />

J. Am. Chem. Soc. 2002, 124, 9962.<br />

342 Galasso, V.; Carmichael, I. J. Phys.<br />

Chem. A 2000, 104, 6271.<br />

343 Paddon-Row, M. N.; Houk, K. N.;<br />

Dowd, P.; Garner, P.; Schappert, R.<br />

Tetrahedron Lett. 1981, 22, 4799.<br />

344 Yonezawa, T.; Simizu, K.; Kato, H.<br />

Bull. Chem. Soc. Jpn. 1968, 41, 2336.<br />

345 Closs, G. L.; Larrabee, R. B. Tetrahedron<br />

Lett. 1965, 287.<br />

346 Gleiter, R.; Haider, R.; Bischof, P.;<br />

Zefirov, N. S.; Boganov, A. M.<br />

J. Org. Chem. 1984, 49, 375.<br />

347 Ermer, O.; Lex, J. Angew. Chem. Int. Ed.<br />

Engl. 1987, 26, 447.<br />

348 Gilardi, R.; Maggini, M.; Eaton, P. E.<br />

J. Am. Chem. Soc. 1988, 110, 7232.<br />

349 Alden, R. A.; Kraut, J.; Traylor, T. G.<br />

J. Am. Chem. Soc. 1968, 90, 74.<br />

350 Kaszynski, P.; Michl, J. J. Am. Chem.<br />

Soc. 1988, 110, 5225.<br />

351 Friedli, A. C.; Kaszynski, P.; Michl, J.<br />

Tetrahedron Lett. 1989, 30, 455.<br />

101


102 2 Distorted Saturated <strong>Hydrocarbons</strong><br />

352 Bunz, U.; Polborn, K.; Wagner, H.-U.;<br />

Szeimies, G. Chem. Ber. 1988, 121, 1785.<br />

353 Schäfer, J.; Polborn, K.; Szeimies, G.<br />

Chem. Ber. 1988, 121, 2263.<br />

354 Ermer, O.; Bell, P.; Schäfer, J.;<br />

Szeimies, G. Angew. Chem. Int. Ed. Engl.<br />

1989, 28, 473.<br />

355 Xie, Y.; Schaeffer III, H. F.; Aped, P.;<br />

Chen, K.; Allinger, N. L. Intern. J. Quant.<br />

Chem. 1992, 42, 953.<br />

356 Xie, Y.; Schaeffer III, H. F. Chem. Phys.<br />

Lett. 1989, 161, 516.<br />

357 Huang, M.-J.; Bodor, N. Chem. Phys. Lett.<br />

1992, 190, 25.<br />

358 v. Schleyer, P. R.; Bremer, M. Angew.<br />

Chem. Int. Ed. Engl. 1989, 28, 1226.<br />

359 Galasso, V. Chem. Phys. 1994, 181, 363.<br />

360 Spoormaker, T.; de Bie, M. J. A. Rec.<br />

Trav. Pays Bas 1979, 98, 380.<br />

361 Becher, G.; Lüttke, W.; Schrumpf, G.<br />

Angew. Chem. Int. Ed. Engl. 1973, 12,<br />

339.<br />

362 Kamienska-Trela, K. Org. Magn. Res.<br />

1980, 14, 398.<br />

363 Tanaka, M.; Sekiguchi, A. Angew. Chem.<br />

Int. Ed. 2005, 44, 5821.<br />

364 Mo, Y. Org. Lett. 2006, 8, 535.<br />

365 Tanaka, M.; Sekiguchi, A., unpublished<br />

results.<br />

366 For the synthesis of tetrakis (trimethylsilyl)tetrahedrane:<br />

Maier, G.; Neudert, J.;<br />

Wolf, O.; Pappusch, D.; Sekiguchi, A.;<br />

Tanaka, M.; Matsuo, T. J. Am. Chem. Soc.<br />

2002, 124, 13819.<br />

367 For reviews on tetrahedrane derivatives:<br />

(a) Maier, G. Angew. Chem. Int. Ed. 1988,<br />

27, 309. (b) Matsuo, T.; Sekiguchi, A.<br />

Bull. Chem. Soc. Jpn. 2004, 77, 211.<br />

(c) Sekiguchi, A.; Matsuo, T. Synlett<br />

2006, 2683.<br />

368 Huntley, D. R.; Markopoulos, G.;<br />

Donovan, P. M.; Scott, L. T.;<br />

Hoffmann, R. Angew. Chem. Int. Ed.<br />

2005, 44, 7549.<br />

369 Pascal, R. A., Jr. Eur. J. Org. Chem. 2004,<br />

3763.<br />

370 Song, Q.; Ho, D. M.; Pascal, R. A., Jr.<br />

J. Am. Chem. Soc. 2005, 127, 11246.


3<br />

Distorted Alkenes<br />

3.1<br />

Nonplanar Alkenes<br />

Dieter Lenoir, Paul J. Smith and Joel F. Liebman<br />

3.1.1<br />

Introduction and Context<br />

Sterically-strained alkenes are important species in the study of physical organic<br />

chemistry [1]. Steric strain can result in torsion (twisting) and/or bending of<br />

the double bond [2]. Bending can occur in a syn- or anti-mode fashion as the<br />

two pairs of geminal groups on the double bond crinkle towards each other or<br />

apart. Strain diversely affects alkenes: (a) energy, measured by standard heat<br />

(enthalpy) of formation (e.g. calorimetric measurement of heat of combustion<br />

and hydrogenation); (b) kinetics, e.g. as measured by thermal (Z)/(E)-rotational<br />

barriers; (c) reaction mechanisms and reactive intermediates. While energetics and<br />

kinetics dominate our interest, examples of such special behavior are sprinkled<br />

throughout this chapter.<br />

3.1.2<br />

Bridgehead Alkenes<br />

Bridgehead alkenes (also discussed in Section 3.2) have been widely investigated<br />

[1–3]. For example, consider adamantene 1 [4] shown below. According to Bredt’s<br />

rule, which is the absolute proscription against bridgehead double bonds, this<br />

species should not exist. Highly strained bridgehead alkenes such as 1 cannot be<br />

isolated, since they tend to dimerize spontaneously after their formation. Like some<br />

other bridgehead alkenes, 1 has been trapped after its formation in solution by<br />

cycloaddition with butadiene [5]. Wiseman [6] has developed a rule to predict the<br />

stability of bridgehead alkenes by comparing the stability of these species with that<br />

of the corresponding E-cycloalkene that serves as the substructure thereof. 1 is a<br />

derivative of E-cyclohexene, which does not exist under normal conditions, while<br />

alkene 2 is a derivative of the stable E-cyclooctene and does exist. There is another<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

103


104 3 Distorted Alkenes<br />

general rule developed by Maier and Schleyer [7] to predict whether a cyclic alkene<br />

can be isolated as a monomer: if the calculated strain energy difference of alkene<br />

and the corresponding saturated hydrocarbon (OS, the ‘olefin strain’) exceeds a<br />

value of 21 kcal mol –1 , then the olefin tends to dimerize at room temperature.<br />

Several highly strained alkenes have been isolated in a cryogenic matrix [8a],<br />

e.g. 1 and the related aza [8b] and chloro [8c] derivatives. However, 2-(1-adamantyl)<br />

adamantene cannot be isolated because of fast rearrangement to 3-(1-adamantyl)-<br />

4-protoadamantylidene [8d]. E-cycloheptene is the smallest E-cycloalkene that can<br />

be isolated at –40 °C, rapidly transforming to its Z-isomer at room temperature<br />

[9] through a low barrier biradical Z/E transition state.<br />

3.1.2.1 t-Butyl-substituted Ethylenes<br />

t-Butyl is sterically an extremely bulky alkyl group. Sequential addition of this<br />

group to a C=C double bond leads to a significant increase in strain. The five<br />

alkenes 3–7, as both the E and Z-isomers, have been prepared. The E-isomers<br />

can be prepared by McMurry coupling [10] from the corresponding ketones and<br />

their Z-isomers by photochemical Z/E isomerization and subsequent separation<br />

using column chromatography. The heats of formation have been determined<br />

from the measurement of the corresponding heats of hydrogenation [11]. Strain<br />

energies have been calculated from these values; the highest strain energy value<br />

of 31.5 kcal mol –1 is found for the Z-isomer of 3,4-diethyl-2,2,5,5-tetramethylhex-<br />

3-ene 7 [11].<br />

The gas phase Z/E barriers of these alkenes have been measured using high<br />

temperature, low pressure, shock tubes. The values for these alkenes vary greatly<br />

with a range of ca. 30 kcal mol –1 . The so-called inherent value, the barrier of<br />

ethylene, has been calculated to be 65.9(0.9) kcal mol –1 [11] (Parentheses convey<br />

uncertainties in measured values). Some of their isomers, the geminal 1,1-dit-butyl-2,2-dialkyl<br />

ethylenes, 8 and 9, have been synthesized, and their electrochemical<br />

oxidation potentials leading to radical cations investigated [12a]. 9 is<br />

more strained by 9.2 kcal mol –1 than its regioisomer E-7, an effect of geminal vs<br />

vicinal t-butyl groups in compound 9 [12b].<br />

Strain energies of these highly crowded Z-ethylenes can be correlated with rotational<br />

Z/E barriers [13] using an exponential function, see Figure 3.1 combining<br />

thermodynamic and kinetic data. Other functional forms relating rotational<br />

barriers and strain energies are of lower accuracy [13].<br />

As recently reviewed [14], tetra-t-butylethylene 10 remains unprepared despite<br />

numerous attempts by different research groups. However, many tied-back<br />

derivatives are known [15] such as 11 and its isomer, syn-fenchylidenefenchane.


3.1 Nonplanar Alkenes<br />

These species have crystallographically-determined twist angles of 17.0° and 11.8°,<br />

respectively, and, as their direct equilibration with CH 2 Cl 2 /AlCl 3 shows, their free<br />

energies differ by only 0.2 kcal mol –1 at 250 °C. However, their reactivity with<br />

elemental bromine is markedly different, the former rapidly reacting and the other<br />

very slowly and both yielding ill-defined, but different sets of products.<br />

Figure 3.1 Correlation of strain energy of Z-ethylenes with Z/E barriers,<br />

Ea,rot = 64.024exp(-0.0176SE), R 2 = 0.9986.<br />

105


106 3 Distorted Alkenes<br />

Some ‘monomeric’ carbenes could be dimerized to the desired alkenes.<br />

However, di-t-butylcarbene does not dimerize to the hoped for 10 since the<br />

carbene kinetically prefers intramolecular insertion into the methyl group leading<br />

to 1,1-dimethyl-2-t-butylcycloprop-ane calculated to be of lower thermodynamic<br />

stability than the desired olefin [16].<br />

3.1.2.2 Investigations of t-Butylated Ethylenes and Other Acyclic Alkenes<br />

We now turn our attention to other nonplanar alkenes for which thermochemical<br />

measurements, enthalpies of formation, combustion and/or hydrogenation have<br />

been reported. The most significant example of this is (Z)-1,2-di-t-butylethylene<br />

(2,2,5,5-tetramethylhex-3-ene) 3. As cited in a recent monograph on heats of hydrogenation<br />

[17], there are two reported measurements, both in glacial acetic acid,<br />

as opposed to the recommended [18] medium of heptane or other nonpolar, inert<br />

gas mimicking media. The two values are 36.2(0.2) and 37.7(0.1) kcal mol –1 from<br />

which we derive an average or consensus value of 37.0(0.8) kcal mol –1 . This value<br />

may be compared with those of a variety of strain-free analogs. Perhaps the most<br />

direct analogy is with its own (E)-isomer for which the corresponding two primary<br />

or directly measured values are 26.9(0.1) and 28.1(0.2) kcal mol –1 resulting in a<br />

consensus value of 27.5(0.6) kcal mol –1 . This suggests some 9.5 kcal mol –1 strain<br />

energy associated with the vicinal t-butyl groups. Indeed, the difference in hydrogenation<br />

enthalpies between (Z)- and (E)-alkenes, where no such destabilization<br />

for the (Z)-isomer is expected beyond about one kcal mol –1 .<br />

Two other comparisons may naturally be made. The first is with the 1,2-dineopentylethylenes<br />

12 for which the hydrogenation enthalpies of the (Z) and (E)isomers<br />

are 26.9(0.1) and 26.0(0.1) kcal mol –1 , respectively, with a difference of<br />

0.8(0.2) kcal mol –1 . 12 in both the (Z)- and (E)-forms is essentially unstrained.<br />

The other comparison is with 6, the 1,2-dimethyl derivative of 1,2-di-t-butylethylene.<br />

The relevant hydrogenation enthalpies for the (Z)- and (E)-isomers are<br />

43.7(0.2) and 37.4(0.2) kcal mol –1 , respectively, suggesting that the former is only<br />

6.3(0.3) kcal mol –1 more strained than the latter one. This difference is smaller<br />

than for the aforementioned difference of the parent di-t-butylethylenes. This<br />

does not mean that there is less strain in the dimethyl species. Rather, the hydrogenation<br />

enthalpy is higher for the dimethyl compounds than the corresponding<br />

parent olefins. This reminds us of the importance of carefully defining reference<br />

species and strain energies, when making comparisons even for seemingly highly<br />

related species.<br />

Reactivity considerations accompany discussions of the large strain energy of<br />

(Z)-di-t-butylethylene. For example, despite its considerable destabilization, acidcatalyzed<br />

isomerization to the much more stable (E)-species does not proceed [19].<br />

Both isomers of 3 are readily chlorinated: while the (Z)-isomer cleanly yields the<br />

classical vicinal species, the (E)-isomer results in an isomerized 1,3-dichloride [19,<br />

20]. Photochemically, the isomers of 3 interconvert, also forming 1,1-dimethyl-<br />

2-neopentylcyclopropane [21]. Upon irradiation, tri-t-butylethylene, with strain<br />

presumably interpolating species 3 and 10, results in 1,1-dimethyl-2,3-bis-(t-butyl)<br />

cyclobutane and 1,2-dimethyl-3-bis(t-butyl)methylcyclopropane.


3.1 Nonplanar Alkenes<br />

3.1.2.3 Cyclo and Bicycloalkenes … and on to Polycyclic Analogs<br />

Let us now consider monocyclic species, and start with cycloalkenes [22]. Solution<br />

phase hydrogenation enthalpy measurements have also been useful here for<br />

the understanding of strain and stability of these species. Under comparable<br />

conditions of an inert media, or otherwise correcting for solvent effects, the difference<br />

of hydrogenation enthalpies for (Z)- and the distorted (E)-cyclooctene<br />

13 is about 11.5(0.5) kcal mol –1 ; the latter value is reduced to 9.8(0.4) kcal mol –1<br />

in acetic acid solution, suggestive of enhanced solvation of the strained � bond<br />

in the (E)-species over that of its more stable (Z)-isomer. For the corresponding<br />

cyclononenes, cyclodecenes and cyclododecenes in acetic acid, hydrogenation<br />

of the (E)-isomers is more exothermic than those of the (Z)-isomers by 2.9(0.1),<br />

3.3(0.1) and 0.5 kcal mol –1 , respectively, and the (Z)-isomers are somewhat more<br />

stable than their (E)-counterparts. By contrast, the related acyclic 3-hexene shows<br />

the (E)-isomer to be more stable by 1.5 kcal mol –1 as determined by hydrogenation<br />

in hydrocarbon solution. Summarizing, the additional strain in these higher<br />

cycloalkenes, beyond that of their medium size rings, is minimal.<br />

Introducing additional unsaturation brings us to the corresponding cycloalkadienes<br />

[23], in particular the cycloctadienes. Changing one (Z)- linkage to (E)- in<br />

the 1,3-isomer 14 increases the hydrogenation enthalpy to 15 kcal mol –1 (no experimentally<br />

determined data for the (E,E)-isomer is available). For the 1,5-isomer<br />

15, the hydrogenation enthalpies increase in the order (Z,Z) < (Z,E) < (E,E) with<br />

sequential differences of 12 and 9 kcal mol –1 . The difference between the values for<br />

the (Z)- and (E)-cyclooctene is about 12 kcal mol –1 , suggesting little thermochemical<br />

consequence of the additional unsaturation.<br />

Consider now [22] singly unsaturated bicyclic species. Table 3.1 gives the hydrogenation<br />

enthalpies of the derivatives of bicyclo[m.n.0] alkenes 16 wherein the<br />

double bond spans the 0 bridge, always m = n, and some related polycyclic species.<br />

For comparison, the corresponding data for the acyclic 2,3-dimethyl-2-butene<br />

Table 3.1 Hydrogenation enthalpies (kcal mol –1 ) of some bicycloalkenes and related species.<br />

Compound �H H2 Method and comments [Ref.]<br />

Tetramethylethylene 26.2(0.1) Classical calorimetry<br />

16, m = n = 1 79(10) Gas phase ion, radical, radical ion cycle [24]<br />

16, m = n = 2 43(2) Classical hydrogenation calorimetry of<br />

bicyclo[2.2.0]hexene and -hexane<br />

17, � 1,5 -dehydroquadricyclane 91(5) Gas phase ion, radical, radical ion cycle [25]<br />

18, cubene 88(5) Gas phase ion, radical, radical ion cycle [26]<br />

16, m = n = 3 27(2) Quantum chem calculation for olefin [25, 27]<br />

19, dodecahedrene 64(4) Gas phase ion comparison [28]<br />

107


108 3 Distorted Alkenes<br />

(TME) are also included there. Cubene for which the values are also given in<br />

Table 3.1 is briefly discussed in Chapter 2.3.<br />

Enthalpies of formation of many of the aforementioned olefins of interest are<br />

unavailable because we lack the necessary data for many ancillary species such<br />

as the saturated cubane [29] (see, however, the discussion of the cubane heat of<br />

formation in Chapter 1.2) and dodecahedrane [30]. We thus defer making any<br />

thermochemical comparisons with singly, and a fortiori, the multiply-unsaturated<br />

derivatives of dodecahedrane [31] and diverse hydrogenated fullerenes (presented<br />

in Chapter 5.6) and their derivatives [32], which are fascinating, clearly strained<br />

and worthy of further thermochemical study. For the latter we also need reliable<br />

heat of formation information [33] for the plausibly strain-free bicyclo[4.4.0]dec-<br />

1(9)-ene (16, m = n = 4). Other data are absent: for example, the reader might wish<br />

to use gas phase ion reaction results of the acyclic 2,3-dimethyl-2-butene. This is<br />

not achievable because the stability of the radical anion of the bicyclobutene arises<br />

precisely from its nonplanarity [34] as ‘simple’ alkenes fail to bind the defining<br />

extra electron [35].<br />

3.1.2.4 Adamantylideneadamantane and its Derivatives<br />

While the double bond of adamantylideneadamantane 20 is nearly planar [36]<br />

when R = H, the three derivatives substituted at the trans allylic bridgehead<br />

positions show severe twisting and out-of-plane bending of the double bond.<br />

[37] The calculated strain energy from Allinger’s MM2 method, increases in the<br />

order; 1,1�-dimethyl, 21.6 kcal mol –1 , 1,1�-diethyl, 28.1 kcal mol –1 compared to the<br />

parent alkene. The X-ray structure of the ethyl derivative shows a torsion angle of<br />

12.3° with a pyramidalization angle of 8.9° accompanied by significant elongation<br />

of the C=C double bond and the vinylic bonds. This highly crowded structure of<br />

the ethyl derivative has been used to test, and affirm [37] the validity of Allinger’s<br />

MMP2 method for highly crowded ethylenes. The corresponding 1,1�-diphenyl<br />

derivative of adamantylideneadamantane (singled out to presage phenylated<br />

olefins in the following section) has also been prepared recently and likewise<br />

shows [38] the greatest deformation of the double bond of all species 20. In no<br />

case, however, is there any measured enthalpy of formation or of hydrogenation<br />

to test any thermochemical conclusion.<br />

Adamantylideneadamantanes 20, especially the parent hydrocarbon, have greatly<br />

contributed to our understanding of diverse reactive intermediates. For example,<br />

the unsubstituted olefin forms solid bromonium ion salts [39] determined crystallographically<br />

as the triflate salt [40] while all other derivatives of 20 lead only<br />

to products of addition/elimination when reacting with bromine [41]. An (asymmetric)<br />

chloronium salt is known as well [42]. Parent 20 readily exhibits radical<br />

cation derived chemistry as part of reactions involving excited state uranyl ion<br />

[43a], strong Brønsted acids [43b], and neutral and cationic NO x species [43c, d].<br />

3.1.2.5 t-Butyl-substituted and Cyclic Stilbenes<br />

When both of the vinylic hydrogens in 1,2-di-t-butylethylene are substituted by<br />

phenyl groups or equivalently those of stilbene are substituted by t-butyl groups,


3.1 Nonplanar Alkenes<br />

to form olefins 21, significant geometric change occurs. The increase of strain is<br />

calculated to be about 18.0 kcal mol –1 . The central double bonds do not show the<br />

common distortions like torsion or bending, but the vinylic bonds with the phenyl<br />

groups are rotated into a perpendicular conformation to avoid steric repulsion of<br />

the t-butyl groups [44]. The UV spectrum shows two isolated phenyl and olefinic<br />

chromophores [45]. While the resonance stabilization of the stilbene has presumably<br />

vanished, calorimetric corroboration remains absent.<br />

The thermal Z/E barrier for these isomers is significantly reduced from the<br />

value of 42.8 kcal mol –1 for the parent stilbenes [45]. The Ea of the solid E- and<br />

Z-21 have been determined in the condensed phase [46] in contrast to the gas<br />

phase values measured for ethylenes 3–7 [11]. DSC measurements for the temperature<br />

range of 50–250 °C resulted in a value of Ea = 35.3 (1.0) kcal mol –1 as<br />

determined using an Ozawa-Flynn-Wall analysis [47], in close agreement with<br />

32.0(1.3) kcal mol –1 obtained by NMR in solution. Various physicochemical<br />

properties of stilbenes E- and Z-21 have been studied [48]. Their chemical activity<br />

includes formation of the first stable � complex found for a bromination reaction<br />

[49].<br />

E-1-((2,2-dimethyl-1-tetralinylidene)-2,2-dimethyltetralin, 22, is the most strained<br />

of these species with its central double bond twisted by 36.7° [50]. The Z-isomer<br />

of 22 has evaded isolation. However, its fleeting existence has been documented<br />

by its NMR and UV spectra, when the E-isomer, dissolved in THF was irradiated<br />

with UV (235 nm) in a NMR quartz tube at –70° leading to the Z-isomer of 22.<br />

The Z-isomer rearranges back to the E-isomer at room temperature with a low<br />

thermal barrier of Ea = 21.0 kcal mol –1 [50]. A torsion angle of 73° has been calculated<br />

for the Z-isomer by MMP2 [50], but recent DFT and ab initio calculations<br />

give a smaller value of about 42° [51].<br />

3.1.3<br />

Multiply Unsaturated Bicycloalkenes, Homoaromaticity and Cyclophanes<br />

We now turn to the parent and hydrogenated derivatives of bicyclo[4.4.1]undecapentaene<br />

and bicyclo[5.3.1]undecapentaene [52]. These hydrocarbons are<br />

also known as 1,6-methano[10]annulene 23 and 1,5-methano[10]annulene 24,<br />

respectively, and may be recognized as strained species with distorted aromatic<br />

rings. The hydrogenated counterparts of the former species and those of the<br />

related bicyclo[4.4.2]dodecapentaene 25 series with an ethano bridge, show [53]<br />

an interesting balance between these compounds and the tricyclic species with a<br />

0 bridge, a full C–C bond, spanning the bridgeheads to form corresponding propellane<br />

derivatives (e.g. 26). We recognize the bicyclic species as having severely<br />

twisted double bonds and the possibility of homoaromaticity [54]; the tricyclic<br />

species look quite classical – at least when compared with small-ring propellanes<br />

discussed in Chapter 2.1. To our knowledge, there are no measured enthalpies of<br />

hydrogenation, nor any of formation, for the mostly saturated bicycloundecenes<br />

and dodecenes with but one double bond for which we may disentangle the effect<br />

of twisting, independent of any aromaticity or homoaromaticity.<br />

109


110 3 Distorted Alkenes<br />

For the same reason we will ignore here the literature calorimetric discussion<br />

[55] of the hydrogenation of bicyclo[5.3.1]undeca-1,3,5(11)-triene, also known as<br />

[5]metacyclophane as well as its [6]metacyclophane (a bicyclo[6.3.1]dodecatriene)<br />

homolog (species 27, n = 5 and 6, respectively) (cyclophanes are discussed in<br />

Section 4.2). We merely note that these last species hydrogenate readily, as befits<br />

their distorted benzene derivatives, to form bridgehead monoenes and then the<br />

hydrogenation stops, or more properly slows below calorimetric detectability.


3.1 Nonplanar Alkenes<br />

Model calculations show these monoenes to be ‘hyperstable’ [56], i.e. their hydrogenation<br />

enthalpy is less than that of standard olefinic species and so they are not<br />

really within the purview of the current volume on strained species. Perhaps it<br />

is ‘sour grapes’ to avoid them here because of the lack of data. Interestingly, and<br />

perversely, the hydrogenation of [2.2]paracyclophane 28 to a mixture of bridgehead<br />

dienes has been interpreted in terms of hyperstable olefins [57]. However, while<br />

the totally saturated polycycle has been calorimetrically investigated [58] but the<br />

hydrogenation was not, we are thwarted in this case as well.<br />

3.1.3.1 The Most Distorted Ethylenes and Seemingly Simple Analogs<br />

Contenders for the most distorted olefins are the fulvalene 29 with a distortion<br />

angle of 66° [59] and the hydrindanylidenehydrindane 30 with an angle of 49° [60].<br />

No heat of formation data is known for either compound.<br />

We conclude this chapter by acknowledging the diverse fulvalene derivatives as<br />

examples of distorted olefins (and not just 29 and 30), and note thermochemical<br />

data are essentially absent for this class of compounds. For the parent fulvalene<br />

31, we find a measurement [61] for the enthalpy of hydrogenation of its (Cr-Cr)<br />

bis chromium tricarbonyl derivative to the dihydride, [� 5 -C 5 H 4 Cr(CO) 3 H] 2 32 but<br />

no enthalpy of formation data to even estimate the enthalpy of formation of the<br />

parent hydrocarbon (as opposed to its bimetallic complex). For heptafulvalene 33,<br />

the enthalpy of hydrogenation to form cycloheptylidenecycloheptane 34 has been<br />

111


112 3 Distorted Alkenes<br />

known for 50 years [62] but as with the above hydrogenated cyclophanes, the last<br />

step to form bis(cycloheptyl) has evaded calorimetric determination. So, where are<br />

the synthetic chemists to make the key compounds and calorimetrists to make the<br />

key measurements? Quantitative understanding of the so much of the chemistry<br />

and thermochemistry of strained alkenes is still awaited [63].<br />

Acknowledgment<br />

We thank Dr. C. Wattenbach for his help and technical assistance.<br />

3.2<br />

Small Ring and Cage Structures Involving Nonplanar C=C Bonds<br />

Athanassios Nicolaides<br />

According to introductory textbooks the olefinic carbons of idealized alkenes are<br />

sp 2 hybridized and the overall geometry is planar. At this geometry the lateral<br />

overlap of the two p atomic orbitals of the olefinic carbons is maximal and gives<br />

rise to a strong � bond (Figure 3.2a). Within this model one can think of two ways<br />

to distort the system and introduce strain. One is a torsional distortion around<br />

the C–C � bond, giving rise to torsionally strained olefins, also known as twisted<br />

or anti-Bredt olefins (Figure 3.2b), which are discussed in the preceding section.<br />

The second is by moving the olefinic carbons away from the plane defined by<br />

their four substituents, giving rise to pyramidalized olefins. While twisted olefins<br />

have received considerable attention [64] rather less is known about pyramidalized<br />

olefins [65–68].<br />

Figure 3.2 Planar and distorted geometries of alkenes.<br />

In torsionally strained olefins the loss of � bonding due to twisting is partially<br />

recovered by rehybridization of the olefinic carbons, the consequence of which is<br />

pyramidalization [64c]. On the other hand it is possible to design purely pyramidalized<br />

alkenes without torsionally strained � bonds as in 38 and 40.<br />

3.2.1<br />

Pyramidalized Alkenes<br />

In a strict sense pyramidalization of alkenes is very common and is expected<br />

to take place whenever the two faces of the double bond are not symmetrically


3.2 Small Ring and Cage Structures Involving Nonplanar C=C Bonds<br />

equivalent [66]. However, in the majority of the cases the pyramidalization is quite<br />

small and trivial. Often the term ‘pyramidalized alkene’ implies that pyramidalization<br />

is large enough to affect the physical and chemical properties of the system<br />

in a significant way.<br />

Several ways to quantify pyramidalization have been proposed [68]. The most<br />

widely used is that by Borden [69], which defines the pyramidalization angle �<br />

as the angle between the plane containing one of the doubly-bonded carbons<br />

and its two substituents and the extension of the double bond (Figure 3.2c).<br />

In terms of the bond angles � and �, � can be obtained from the formula:<br />

cos � = –cos �/cos (�/2) [66].<br />

9,9�-Didehydroanthracene (DDA, 35) was the first pyramidalized alkene to be<br />

isolated almost 40 years ago by Weinshenker and Greene [70]. The addition of base<br />

(t-butoxide) to the pyramidalized C=C double bond provided an early demonstration<br />

that pyramidalized alkenes are unusually reactive with nucleophiles. The most<br />

cited pyramidalized polyene is fullerene C 60 [71] and its derivatives, discussed<br />

in Chapter 5. Alkenes with pyramidalized double bonds include cubene [72],<br />

bicyclo[1.1.0]but-1(3)-ene [73] and its 2,4-bridged derivatives [74], dodecahedrene<br />

and related compounds [75] and others which have been reviewed recently in an<br />

authoritative manner [68].<br />

This chapter focuses on tricyclo[3.3.n.0 3,7 ]alk-3(7)-enes 36, a family of pyramidalized<br />

alkenes that has been studied extensively and systematically both experimentally<br />

and theoretically [66, 68].<br />

In a formal sense, bicyclo[3.3.0]oct-1(5)-ene 37 can be thought of as the generator<br />

of the homologous series 36 by connecting carbons 3 and 7 with a bridge of n<br />

methylene groups. The first member of this series is 41 in which carbons 3 and<br />

7 are directly bonded (n = 0). The five-membered rings of 37 are puckered in the<br />

same direction [76, 77], making the two faces of the double bond non-equivalent.<br />

As a result the olefinic carbons are trivially pyramidalized with a computed<br />

pyramidalization angle of � = 5.8° [77]. The small value of � implies that 36 is a<br />

normal tetraalkyl-substituted alkene.<br />

113


114 3 Distorted Alkenes<br />

Table 3.2 Computational data for alkenes 37–41.<br />

Alkene n a)<br />

� b,h)<br />

c,i)<br />

ELUMO d,i)<br />

EHOMO �E e,i)<br />

C=C f,h)<br />

41 0 61.9 –0.50 0.61 10.14 1.380 72.8<br />

40 1 53.7 –2.51 0.65 11.09 1.362 52.3<br />

39 2 42.2 –1.38 0.33 12.54 1.349 37.4<br />

38 3 27.9 –0.79 0.24 13.22 1.342 17.7<br />

37 – 5.8 0 0 14.25 1.337 0<br />

a)<br />

Number of methylene groups in the bridge.<br />

b)<br />

Pyramidalization angle (°) [68].<br />

c)<br />

LUMO energies relative to that of 37 (5.65 eV) [76].<br />

d)<br />

HOMO energies relative to that of 37 (–8.60 eV) [76].<br />

e)<br />

Energy difference LUMO–HOMO (eV) [76].<br />

f)<br />

Bond-length (Å) [73c].<br />

g) –1<br />

Olefin strain energy (kcal mol ) [68].<br />

h)<br />

Computed with the B3LYP/6-31G(d) method.<br />

i)<br />

Computed with the SCF/3-21G method.<br />

OSE g,h)<br />

No special properties are expected to arise for olefins 36 as long as the length of<br />

the methylene bridge is sufficiently long (i.e. for a sufficiently large n). However,<br />

as the bridge becomes shorter it is expected that at some point it will cause considerable<br />

puckering of the bicyclic moiety and pyramidalization of the double<br />

bond. Indeed, the computed data (Table 3.2) show that the pyramidalization<br />

angle increases from about 28° for 38, having a tri-methylene bridge, to about<br />

62° for 41, where carbons 3 and 5 are directly bonded. Thus, homologous series<br />

36 provides a nice way of studying systematically the effect of pyramidalization<br />

on the properties of the C=C double bond as a function of n.<br />

A qualitative model based on simple molecular orbital theory has been proposed<br />

as a guide to understanding the properties of these compounds and of pyramidalized<br />

alkenes in general [66, 76]. The pyramidalization effect on the electronic<br />

properties of the C=C double bond can be understood on the basis of two factors<br />

(Figure 3.3). As pyramidalization (�) increases, the overlap between the p atomic<br />

orbitals forming the � bond decreases and this raises the energy of the bonding �<br />

MO (HOMO) but lowers the energy of the anti-bonding �* (LUMO). Pyramidalization<br />

also causes mixing of the � and � MOs, which in terms of atomic orbitals is a<br />

form of rehybridization. This increases the s character of the atomic orbitals and<br />

therefore energetically stabilizes both the HOMO and the LUMO. Overall, both<br />

factors stabilize the LUMO, but have opposite effects on the HOMO and therefore<br />

they tend to cancel out. Thus, in the case of the LUMO strong stabilization is<br />

expected, but the energy of the HOMO is expected to stay roughly the same.<br />

Indeed computations (Table 3.2) find that the energy of the LUMO in pyramidalized<br />

alkenes 36 is lower with respect to the reference compound 37 and<br />

decreases further as pyramidalization of the alkene increases. On the other hand,


3.2 Small Ring and Cage Structures Involving Nonplanar C=C Bonds<br />

Figure 3.3 Effect of pyramidalization on the relative energies of the HOMO (�) and LUMO (�*)<br />

orbitals of an alkene.<br />

the energy of the HOMO increases, but this is considerably less in magnitude<br />

than the corresponding decrease of the HOMO, in agreement with the qualitative<br />

model. Overall a decrease of the HOMO–LUMO gap is predicted with increased<br />

pyramidalization.<br />

Olefin strain energy (OSE) [76, 78] is a quantitative measure of the strain of the<br />

alkene. With alkene 37 serving as the ‘strain-free’ reference point of homologous<br />

series 36, the computed OSEs show that there is a considerable increase in the<br />

strain of the alkene with pyramidalization. Based on empirical rules and the computational<br />

results (Table 3.2) the most pyramidal alkenes of this series 39–41 are<br />

not predicted to be isolable under normal conditions, while for 38 it is difficult to<br />

make a secure prediction. As the strain of the � bond increases, one expects the<br />

�-bond order to decrease and therefore the C=C bond length to increase. Indeed,<br />

the C=C bond length of 41 is the longest, but it is still sufficiently short to suggest<br />

that the residual � bond is still strong despite the large OSE.<br />

This simple qualitative model rationalizes the computational data shown in<br />

Table 3.2 in a satisfactory way. Additionally, research has sought to confirm this<br />

model by generating the pyramidalized alkenes 37–41 and studying their physicochemical<br />

properties [66, 68].<br />

In general, pyramidalized alkenes, like other strained molecules, require the<br />

development of special synthetic methods for their formation. Apart from the<br />

synthetic challenge, isolation of the most reactive ones is impossible under<br />

normal conditions and special techniques are needed. Some of the most reactive<br />

pyramidalized alkenes have been studied directly by spectroscopy in the gas phase<br />

or after matrix isolation at low temperatures. More often the intermediacy of pyramidalized<br />

alkenes is deduced from the product analysis and/or from chemical<br />

trapping of the alkene [68].<br />

3.2.1.1 Tricyclo[3.3.11.0 3,7 ]undec-3(7)-ene 38<br />

Alkene 38 has been synthesized by the reduction of its dimesylate precursor<br />

38-OMs. Even though 38 is an isolable solid, X-ray crystallographic data for it have<br />

115


116 3 Distorted Alkenes<br />

not been reported. Upon exposure to air it oxidizes readily, forming among other<br />

things its epoxide [79] a reaction characteristic of pyramidalized alkenes [80, 81].<br />

Alkene 38 has been characterized spectroscopically by IR, UV, NMR and electron<br />

transmission (ET) spectroscopy.<br />

As expected of symmetrically substituted alkenes, the C=C bond stretching<br />

vibration of reference compound 37 is forbidden by IR spectroscopy but is observable<br />

by Raman spectroscopy. In contrast, the C=C bond stretch of 38 appears both<br />

in the IR and Raman spectra. In the IR spectrum of 38 it appears as a very weak<br />

band and this can be attributed to the fact that pyramidalization of the double<br />

bond causes the C=C stretching mode to have a transition dipole, which is oriented<br />

perpendicular to the C–C bond. It is noteworthy that the C=C stretching frequency<br />

of 38 is 70 cm –1 lower than that in 37, in qualitative agreement with the weaker<br />

� bond of the former [82, 83].<br />

ET spectroscopy has been employed to measure the vertical electron affinities<br />

(EA) of 37 and 38. Both are negative indicating that in both cases the radical anion<br />

is not bound. However, the important result is that 38 has a greater EA than 37,<br />

compatible with the former having a lower-energy LUMO than the latter [79]. The<br />

well-known property of C 60 to accept electrons forming anions and polyanions<br />

can be attributed to the pyramidalization of its C=C double bonds.<br />

The adiabatic ionization energy (IE) as given by photoelectron spectroscopy is<br />

found to be lower in 38 than in 37 by 0.31 eV. This difference can be related via<br />

Koopmans’ theorem [84] to a higher-energy HOMO in 38 than in 37, in agreement<br />

with the computational data of Table 3.2.<br />

The UV spectrum of 37 does not show a maximum above 200 nm. According<br />

to electron energy loss (EEL) spectroscopy, the � � �* singlet excitation requires<br />

6.54 eV corresponding to photons of 190 nm wavelength. This is very close to the<br />

� � �* transition of tetramethylethylene (6.61 eV, 188 nm) demonstrating that 37<br />

is a rather standard alkene. Olefin 38 exhibits a UV absorption with � max = 217 nm<br />

(5.73 eV). This red shift of 27 nm (or 0.81 eV) is in qualitative agreement with the<br />

smaller HOMO–LUMO gap of 38 as compared to 37 [79].<br />

Protonation of a pyramidalized olefin is expected to relieve olefin strain and<br />

therefore 38 should have a substantially greater proton affinity, PA, than 37. The<br />

experiment was carried out in the gas phase using the kinetic method and established<br />

that the PA of 38 is indeed greater than that of 37. The apparent PA of 38<br />

was measured at 219 kcal mol –1 or 22.5 kcal mol –1 higher than that of 37. On the<br />

other hand, the calculations predicted that the difference in PA between 37 and


Table 3.3 Available spectroscopic data for alkenes 37–40.<br />

Alkene UV a)<br />

3.2 Small Ring and Cage Structures Involving Nonplanar C=C Bonds<br />

IR b)<br />

40 1496 [88]<br />

39 245 [82] 1557 [82]<br />

38 217 [79] 1615 [79] –2.27 [79] 7.80 [79] 208 [85] f)<br />

37 190 [68] 1685 [68] –2.44 [79] 8.11 [79] 196.5 [85]<br />

a)<br />

nm.<br />

b) –1<br />

cm .<br />

c)<br />

Vertical electron affinities (eV).<br />

d)<br />

Adiabatic ionization potentials (eV).<br />

e)<br />

Proton affinities (kcal/mol).<br />

f)<br />

Based on the experimental PA of 37 and the computed difference of 11.7 kcal/mol<br />

between the PAs of 37 and 38.<br />

38 should be only 11.7 kcal mol –1 . Further experimentation and careful analysis<br />

of the experimental data with the help of calculations revealed that under the<br />

experimental conditions, skeletal rearrangements are possible and that what was<br />

measured was not the PA of 38, but rather of a diene isomer of 38. The skeletal<br />

rearrangement seems to take place in the protonated form of 38 and involves a<br />

retrograde vinylcyclopropane rearrangement. This type of rearrangement is endothermic<br />

for alkene 38, but it is known for the lower homologs of 38 [85].<br />

3.2.1.2 Tricyclo[3.3.10.0 3,7 ]dec-3(7)-ene 39 and tricyclo[3.3.9.0 3,7 ]non-3(7)-ene 40<br />

Alkene 39 was first formed by decarboxylation of lactone 39-lac both in solution<br />

and in the gas phase (Figure 3.4) [86]. Evidence for the formation of 39 in solution<br />

was given by trapping the alkene as the Diels–Alder 39-DA. Pyrolysis at 530 °C led<br />

to the isolation of alkene 43, an isomer of 39, which can be formed via a retrograde<br />

vinylcyclopropane rearrangement. From experimental data it was estimated that<br />

the strain in 39 must be at least 21 kcal mol –1 in order for the rearrangement to<br />

be energetically favorable [87].<br />

Figure 3.4 Pyrolysis of lactones 39-lac and 40-lac.<br />

EA c)<br />

IP d)<br />

PA e)<br />

117


118 3 Distorted Alkenes<br />

When the pyrolysis was carried out at lower temperature (410 °C) and the pyrolysate<br />

trapped in an Ar matrix at 10 K, alkene 39 was detected by its IR absorption<br />

at 1557 cm –1 , 58 cm –1 lower than the C=C stretching frequency of 38 [83].<br />

Warming up of the matrix led to the isolation of the dimer of 39, cyclobutane<br />

39-dim, providing more evidence for the formation of 39. The UV spectrum from<br />

the matrix isolated 39, exhibited a � max red-shifted to that of alkene 38.<br />

Lactone 40-lac proved more difficult to pyrolyze than 39-lac. Even at 550 °C only<br />

50% of the lactone reacted, but not by loss of CO 2 but by isomerizing to ketoketene<br />

47, which was identified by IR spectroscopy and by chemical trapping with<br />

methanol. From this remarkable rearrangement it was estimated that the strain in<br />

40 exceeds 39 kcal mol –1 [88]. At temperatures higher than 550 °C decarboxylation<br />

did take place, but only traces of the dimer of 40 (40-dim) were isolated, indicating<br />

that this is not a good route for the formation of 40. The major product of the<br />

pyrolysis was diene 46, presumably formed via alkene 45, which is the product<br />

of the retrograde vinylcyclopropane rearrangement of 40.<br />

Dimesylate 36-OMs (n = 1, 40-OMs) can be easily synthesized from the corresponding<br />

diol 40-OH. However reduction with Na/Hg does not yield alkene 40.<br />

A better precursor for alkene 40 is diiodide 40-I (36-I, n = 1). Treatment of 40-I<br />

in THF at –78 °C with n-butyllithium gave quantitatively the dimer 40-dim [89].<br />

When the reaction was repeated in the presence of 1,3-diphenylisobenzofuran<br />

(1,3-DPIBF, 42), the expected Diels–Alder adduct (40-DA) was isolated. The<br />

reduction can also be carried out at room temperature with Na/Hg again giving<br />

products that imply the intermediate formation of 40. Gas-phase dehalogenation<br />

made possible the matrix isolation of alkene 40 and its IR spectroscopic characterization.<br />

Like its higher homologs, alkene 40 exhibits a C=C stretch as a weak band<br />

in the IR spectrum, with a frequency 61 cm –1 lower than that in the spectrum of<br />

39. Unfortunately, the presence of metal atoms in the matrix prevented the UV<br />

spectrum from being obtained, although indirect experimental evidence for the<br />

existence of a long wavelength absorption in 40 was provided.


3.2 Small Ring and Cage Structures Involving Nonplanar C=C Bonds<br />

3.2.1.3 Tricyclo[3.3.0.0 3,7 ]oct-1(5)-ene 41<br />

Alkene 41 being the consummate member of the homologous series 36 has<br />

attracted considerable attention. Neither dimesylate 41-OMs, n = 0 nor the bistriflate<br />

41-OTf, n = 0 (Tf = CF 3SO 2 – ), give the desired alkene by reduction by Na/Hg,<br />

yielding instead the starting material [68, 90]. In 1996 the first synthesis of 41<br />

was reported [91]. Reduction of diiodide 36-I, n = 0 (41-I) afforded a mixture of<br />

the dimer 41-dim and the diene 41-dn, with the latter thermodynamically more<br />

stable than the former. In the presence of 1,3-DPIBF the Diels–Alder product<br />

41-DA (36-DA, n = 0) was formed. Unfortunately, no spectroscopic information<br />

is available for alkene 41.<br />

Some derivatives of 41 (41� [92], 43 [93] and 44 [94]) have been successfully<br />

generated, but like 41 they have not been characterized spectroscopically. Evidence<br />

for their formation is based on product analysis. The least perturbed homolog of<br />

41 that has been studied is 41�, with two Me groups on the bridge, which can be<br />

formed from its diiodide precursor (41�-I). Evidence for the generation of 41� is the<br />

isolation of the dimer 41�-dm, when 41�-I is reduced in solution, and the isolation of<br />

the expected Diels–Alder products when the reduction is carried out in the presence<br />

of 1,3-DPIBF (42) or of 11,12-dimethylene-9,10-dihydro-9,10-ethanoanthracene<br />

(DDE, 48). Some extra experimental evidence for formation of 41� is the detection<br />

and spectroscopic characterization of its (Ph 3 P) 2 Pt complex 41�-Pt [95].<br />

3.2.1.4 (Ph 3 P) 2 Pt Complexes<br />

Organic reactive intermediates can be stabilized by complexation to organometallic<br />

fragments. A number of such complexes containing highly reactive � bonds,<br />

have been isolated as (Ph 3P) 2Pt complexes. For example, the (Ph 3P) 2Pt complexes<br />

of cycloalkynes that have bent triple bonds: cyclopentyne, cyclohexyne, and cycloheptyne<br />

are known. (Small-ring cycloalkynes are discussed in Chapter 7.) The<br />

(Ph 3P) 2Pt complexes of the pyramidalized alkenes 36, n = 1–3 (36-Pt, n = 1–3) have<br />

been isolated and studied spectroscopically in detail [95]. Recently, the complex<br />

of the 41� (41�-Pt) has also been reported.<br />

119


120 3 Distorted Alkenes<br />

Tetrasubstituted alkenes do not ordinarily form stable complexes with (Ph 3P) 2Pt.<br />

Indeed, the (Ph 3 P) 2 Pt complex of 37, the unbridged generator of alkenes 36, was<br />

not formed [96]. This implies that pyramidalization of the double bond is responsible<br />

for the formation of complexes 36-Pt, n = 1–3 and 41�-Pt. As mentioned<br />

above, the major electronic effect of double bond pyramidalization is to lower<br />

the energy of the �* LUMO, making it a better acceptor of electron density from<br />

the metal fragment [79, 97]. This simple picture is capable of rationalizing most<br />

of the NMR data of Table 3.4. Thus, with increased pyramidalization, 13 C-NMR<br />

resonances move to higher field and 195 Pt ones to lower field although the latter<br />

not in a monotonic way. The trend of the 31 P chemical shifts is not that predicted,<br />

but it is compatible with transfer of electron density from the (Ph 3 P) 2 Pt fragment<br />

to the complexed olefin as judged by other experimental data [96, 98, 99].<br />

Table 3.4 13 C (of the olefinic carbons), 195 Pt and 31 P chemical shifts (ppm) and 1 J Pt–P and 1 J Pt–C<br />

coupling constants (Hz).<br />

41�-Pt a)<br />

40-Pt a)<br />

39-Pt a)<br />

38-Pt a)<br />

� C � Pt � P<br />

–4956.6 30.7 b)<br />

1 JPt–P<br />

2850<br />

1 JPt–C<br />

66.9 –5058.5 30.5 2960 407<br />

74.9 –5105.5 31.1 3115 343<br />

78.8,79.2 c)<br />

–5092.5 32.2 3332 296 d)<br />

(Ph 3 P) 2 PtC 2 H 4 (39.2) –5146.5 34.1 3740 194<br />

a) Benzene-d6 , 298 K.<br />

b) Actual measured value (� 32.4 ppm) adjusted so that the value for the 31 P chemical shift of<br />

(Ph 3 P) 2 PtC 2 H 4 of Ref. [95] (� 35.8 ppm) becomes the same as the one reported in Ref. [96]<br />

(� 34.1 ppm).<br />

c) Toluene-d8 , 229 K.<br />

d) Toluene-d8 , 338 K.


3.2 Small Ring and Cage Structures Involving Nonplanar C=C Bonds<br />

The 1 J Pt–P and 1 J Pt–C coupling constants follow the expected trend. In particular, the<br />

former decreases as the latter increases in a linear fashion, implying that electron<br />

density shifts away from the metal fragment and the Pt–C bonding strengthens as<br />

pyramidalization of the olefin increases. Thus, complex 41�-Pt is expected to have<br />

the largest alkene-Pt(PPh 3) 2 binding energy in this series, an expectation that is in<br />

agreement with computational findings [95]. Despite the stronger C–Pt bonds in<br />

41�-Pt, this complex was found to react with ethanol to give 45. This is interesting<br />

because under the same conditions complexes 36-Pt, n = 1–3 are recrystallized<br />

from ethanol. The alcoholysis of 41�-Pt with ethanol is a rather unusual reaction<br />

[97] and paves the way for studying the reactivity of such complexes with the aim<br />

of using them as synthetic equivalents of reactive pyramidalized olefins.<br />

3.2.2<br />

Conclusions<br />

In conclusion, the above synthetic data for alkenes 36, n = 0–3 are summarized<br />

in Figure 3.5. The general way for the formation of these alkenes is by reduction<br />

of the corresponding diiodide (for the first two members of the homologous<br />

series 36) or dimesylate precursor (for the higher homologs 38 and 39). Alkene<br />

38 is stable enough to be isolated under normal conditions provided oxygen is<br />

excluded [79]. The more pyramidalized alkenes 39–41, cannot be isolated under<br />

the same conditions. Instead, in the absence of any trapping reagent their formal<br />

Figure 3.5 Generation and trapping of tricyclo[3.3.n.0 3,7 ]alk-3(7)-enes (36).<br />

121


122 3 Distorted Alkenes<br />

[2+2] dimers (36-dim) are isolated. The strained 36-dim, n = 0 (41-dim) easily<br />

isomerizes to the diene 41-dn. When the reduction is carried out in the presence<br />

of 1,3-DPIBF 42, the expected Diels–Alder product is isolated. In the presence<br />

of (Ph 3P) 2PtC 2H 4 the corresponding organoplatinum complexes 36-Pt of 38–40<br />

have been isolated. Although, the experimental structure of 38 has not been<br />

determined yet, this alkene has been characterized extensively spectroscopically.<br />

Properties of the alkenes 38–40 are also quite well understood, even though the<br />

UV spectrum of 40 has not been recorded yet. Although, synthetic routes to 41<br />

have been developed, direct spectroscopic observation of 41 is still lacking, and<br />

its (Ph 3P) 2Pt complex (41-Pt) is still not known.<br />

Acknowledgement<br />

This work has been supported significantly by the U.S. National Science Foundation<br />

and also by the Research Promotion Foundation of Cyprus.<br />

3.3<br />

<strong>Strained</strong> Cyclic Allenes and Cumulenes<br />

Richard P. Johnson and Kaleen M. Konrad<br />

3.3.1<br />

Introduction<br />

Cycloalkenes are well known to any student of chemistry but the more exotic<br />

homologous series of cycloallenes 49–54 and cyclobutatrienes 55–60 rarely find<br />

their way into textbooks. Decreasing ring size in these two series leads to a rapid<br />

increase in strain and reactivity caused by deformation of the normally linear<br />

allene or butatriene.<br />

Among milestones in the history of cyclic cumulenes, Ball and Landor first<br />

described the chemistry of 51–53 in 1961 [101], but the synthesis of a derivative<br />

of 1,2-cyclopentadiene 50 awaited the efforts of Balci and co-workers which were<br />

reported in 2002 [102]. Evidence for derivatives of 49 was found in studies on enyne<br />

photochemistry described by Meier and König in 1986 [103]. In the butatriene se-


3.3 <strong>Strained</strong> Cyclic Allenes and Cumulenes<br />

ries, a stable 10-membered ring homolog was described by Moore and Ozretich in<br />

1967 [104]. Synthesis of 57 as a reactive intermediate was reported by Shakespeare<br />

and Johnson in 1990 [105]. The parent hydrocarbon 56 remains unknown, but<br />

heterocyclic analogs have been reported by Wong and co-workers [106, 107].<br />

Essential features of this structural map are thus known. Experiment and theory<br />

support the existence of cyclic allenes in all ring sizes 49 through 54 (and larger),<br />

but structures smaller than an eight-membered ring are not isolable without<br />

special features. The smallest cyclic butatriene 55 is predicted not to be an energy<br />

minimum but instead is a transition state on the C 4H 2 energy surface. There is<br />

evidence for cyclic butatrienes in ring sizes 56 through 60 but only 60 and larger<br />

ring homologs have been isolated.<br />

This chapter presents the structure, syntheses and chemical behavior of four- to<br />

nine-membered ring cyclic allenes and butatrienes, illustrating the status of this<br />

field with selections from recent chemistry. Readers are referred to earlier reviews<br />

for more comprehensive treatments [108–111].<br />

3.3.2<br />

Allene � Bond Deformations and Strain Estimates<br />

In a ring of fewer than ten carbons, the allene group is bent and twisted toward<br />

planarity. Johnson and co-workers recently estimated strain by applying isodesmic<br />

and homodesmic relationships [112]. As one example, allene functional group<br />

strain in 51 may be estimated from the following isodesmic reaction, using<br />

B3LYP/6-311+G(d,p) energies. This equation exchanges strained and unstrained<br />

functional groups.<br />

Figure 3.6 summarizes estimates for allene strain, i.e. strain primarily localized<br />

in that functional group, as well as total molecular strain, and observed chemical<br />

behavior for cyclic allenes. Only modest levels of strain result in kinetic instability<br />

for allenes of eight carbons or fewer; allene 53 and most derivatives dimerize<br />

quickly at ambient temperature [113, 114].<br />

Figure 3.6 Strain estimates and chemical behavior of cyclic allenes.<br />

123


124 3 Distorted Alkenes<br />

Ring constraints alter the barrier to allene chiral inversion. For parent allene 63,<br />

the transition state is best described as planar diradical 64, with one � electron in<br />

plane and three out of plane [115, 116]. Brudzynski and Hudson [117] measured<br />

a torsional barrier for allene of 43 kcal mol –1 ; this is easily reproduced by DFT<br />

calculations [116]. With cyclic allenes, the predicted inversion barrier decreases<br />

in proportion to strain. For 1,2-cyclohexadiene 51, the predicted inversion barrier<br />

65 is 14.1 kcal mol –1 [118]. This interconverts the M and P enantiomers (helicity<br />

nomenclature), consistent with experiments by Balci and Jones that demonstrated<br />

the chirality of 51 [119]. Smaller homolog 50 is predicted to be chiral but with an<br />

inversion barrier of < 1 kcal mol –1 [112], while 1,2-cyclobutadiene 49, should be<br />

a nearly planar diradical [120].<br />

3.3.3<br />

Four- and Five-membered Ring Allenes<br />

Can a remarkable substance such as 1,2-cyclobutadiene 49 exist? In a 1986 paper,<br />

Meier and König reported the photorearrangement of 66 to 68 (Scheme 3.1) and<br />

made the courageous suggestion of 67 as a likely intermediate [103]. Seven years<br />

later, Johnson and co-workers showed this to be a general singlet photorearrangement<br />

[120]. Photoreaction of 69 is typical; irradiation results in the interconversion<br />

of 69 and 71. The most straightforward mechanism is excited state closure to afford<br />

ground state 70, which can open thermally in either direction. MP2, CASSCF and<br />

DFT calculations all support the existence of a 1,2-cyclobutadiene intermediate.<br />

Scheme 3.1 Photorearrangements of enynes.<br />

Balci’s group reported preparation of the first 1,2-cyclopentadiene derivative in<br />

2002. Endo fluoro isomer 72 reacts with methyllithium (Scheme 3.2) to give 73,<br />

which was trapped with furan to give 74 [102]. This achievement followed many<br />

unsuccessful attempts by the same authors [110]. The same group later reported<br />

evidence for 1-phenyl-1,2-cyclopentadiene [121].


Scheme 3.2 Synthesis of a 1,2-cyclopentadiene.<br />

3.3.4<br />

1,2-Cyclohexadienes<br />

3.3 <strong>Strained</strong> Cyclic Allenes and Cumulenes<br />

1,2-Cyclohexadienes are surprisingly common reactive intermediates. The parent<br />

structure 3 is now securely characterized as a chiral allene [119]. Early synthetic<br />

routes to 51 included treatment of 79 with strong base (Scheme 3.3) and reaction<br />

of 77 with methyllithium [101, 108, 122]. This latter reaction presumably proceeds<br />

through carbene 76 and/or the related carbenoid. This was an early example of the<br />

Doering-Moore-Skattebøl rearrangement [123–126]. Schleyer and co-workers have<br />

predicted a barrier of only 0.5 kcal mol –1 for disrotatory ring opening of 76 [127].<br />

Shakespeare and Johnson have reported that 51 is easily generated by treatment<br />

of 78 with fluoride ion [105]. Diels–Alder reaction of enynes is described in a later<br />

section. Werstiuk and co-workers have reported the photoelectron spectrum of 51<br />

which was generated from 75 by retro-cycloaddition [128]. Although there have<br />

been several earlier reports of matrix isolated 51, agreement with the computed<br />

allene vibrational frequency remains uncertain [129–131]; a definitive route to<br />

matrix isolated 51 is needed.<br />

Scheme 3.3 Synthetic routes to 1,2-cyclohexadiene.<br />

Scheme 3.4 summarizes common reactions of 51 and its derivatives. Dimerization<br />

and alkene or diene additions proceed rapidly through diradicals 80 or 81 to<br />

give [2+2] and [2+4] cycloadducts [111]. Caubere reported that nucleophiles such<br />

as enolates add to 3, affording a wide range of products [132]; of course this is<br />

facilitated by strain relief. Christl has presented a detailed review of [2+2] and<br />

[2+4] cycloadditions to 51 and derivatives, summarizing the evidence for stepwise<br />

reactions [111]. Computational studies by Tolbert, Houk and co-workers support<br />

the conclusions of a stepwise mechanism [133].<br />

125


126 3 Distorted Alkenes<br />

Scheme 3.4 Reactions of 1,2-cyclohexadiene.<br />

It has long been recognized that Diels–Alder cycloaddition of enynes with<br />

alkenes should give 1,2-cyclohexadienes. Computational studies by several groups<br />

have yielded activation energies only ca. 6 kcal mol –1 above a conventional Diels–<br />

Alder reaction [134, 135]. Recent examples of this surprising reaction are shown<br />

in Scheme 3.5. Miller and co-workers showed that dienyne 82 undergoes double<br />

addition of fumarate to give 84, presumably through allene 83 [136]. Johnson<br />

and co-workers studied the first intramolecular version of an enyne plus ene<br />

cyclo addition [134]. Flash vapor pyrolysis (FVP) of enyne 85 cleanly afforded 87;<br />

the most straightforward mechanism is cycloaddition to give 86, followed by<br />

cycloreversion. Maas and co-workers have reported preparation of 90 by solution<br />

thermolysis of pyridinium triflate 88 [137].<br />

Many heterocyclic analogs of 1,2-cyclohexadiene 91 have been prepared as<br />

intermediates and these provide routes to complex ring systems [111, 138–141].<br />

Scheme 3.5 Examples of enyne Diels–Alder reactions.


3.3 <strong>Strained</strong> Cyclic Allenes and Cumulenes<br />

Several different strategies have led to isolable derivatives of 51. Longer carbon–<br />

heteroatom bonds and steric protection arising from bulky substituents explain<br />

the isolability of 92 [142] and 93 [143]. Metal complexation as in 94 provides an<br />

additional approach to kinetic stabilization [144, 145].<br />

3.3.5<br />

1,2,4-Cyclohexatrienes<br />

1,2,4-Cyclohexatrienes (‘isobenzenes’) and their benzannelated derivatives are<br />

reactive intermediates that lie at the center of an impressive range of chemistry. For<br />

61, present results support a chiral structure, and a diradical transition state 61a<br />

for inversion, much as described for 51. Replacement of one or more ring carbon<br />

can lead to a large variety of heterocyclic analogs of 1,2,4-cyclohexatrience. As first<br />

noted by Shevlin, these may exist as either a chiral cyclic allene 95 or planar ylide<br />

96 [146–148]. Engels has presented a comprehensive analysis of related structures<br />

containing second and third row elements [118, 149]. Sheridan and co-workers<br />

have provided the clearest confirmation of allenic structure by generating a variety<br />

of heterocyclic 1,2,4-cyclohexatriences in cryogenic matrices [150–152].<br />

Scheme 3.6 summarizes synthetic routes to 1,2,4-cyclohexatrienes. The two<br />

pericyclic reactions have been most widely applied. Christl and co-workers<br />

have reported in great detail on other routes to 1,2,4-cyclohexatriene and its<br />

benzannelated derivative [153, 154]. The Doering-Moore-Skattebøl reaction of 98<br />

and dehydrohalogenation of 100 provide excellent routes to 61.<br />

Common reactions of 1,2,4-cyclohexatrienes are summarized in Scheme 3.7.<br />

The allene is easily trapped with furan. With styrene, both [2+2] and [2+4] products<br />

are isolated, presumably through diradical intermediate 101.<br />

In 1969, Hopf and Musso reported that cis-1,3-hexadiene-5-yne 102 thermally<br />

isomerizes to benzene [155]. This reaction, now known as the Hopf cyclization,<br />

provides explanation for diverse high-temperature chemistry [156]. Experimental<br />

and theoretical results support two general pathways (Scheme 3.8). The<br />

127


128 3 Distorted Alkenes<br />

Scheme 3.6 Common routes to 1,2,4-cyclohexatrienes.<br />

Scheme 3.7 Reactions of 1,2,4-cyclohexatriene.<br />

Scheme 3.8 Hopf cyclization chemistry.


3.3 <strong>Strained</strong> Cyclic Allenes and Cumulenes<br />

Type A mechanism proceeds by electrocyclization to 61, followed by sequential<br />

1,2-hydrogen shifts passing through carbene 104. An alternative Type B route<br />

begins with Brown rearrangement (1,2-shift) to vinylidene 105, followed by C–H<br />

insertion.<br />

Support for the Type A cyclic allene mechanism comes from trapping of 61<br />

with styrene and inhibition of rearrangement in the presence of O 2 [157]. Hopf<br />

and co-workers studied the homologous series 106 [158]. In this case, smaller<br />

rings cyclize readily, while no reaction is observed for the largest homologs. These<br />

results correlated nicely with computations [156, 158].<br />

One chemical process may generate 61 in great abundance. Among the wide<br />

array of intermediates in alkane combustion, Miller and Melius have proposed that<br />

61 may be one stop (Scheme 3.9) on the road to aromatic rings and soot formation.<br />

[159]. Head to tail dimerization of propargyl radical 108, a common species in<br />

flames, might be followed by Brown rearrangement and C–H insertion to give<br />

61. This can aromatize by the route described earlier. Alternatively, we note that<br />

61 should have a low barrier for C–H dissociation to give phenyl radical, another<br />

flame intermediate.<br />

Scheme 3.9 1,2,4-Cyclohexatriene in combustion chemistry.<br />

Photochemical Hopf-type cyclizations are known from the earlier work of Laarhovens<br />

(Scheme 3.10) [160] who showed that 109 undergoes photocyclization,<br />

presumably through cyclic allene 110. Uncertainty remains about the aromatization<br />

steps, since sequential 1,2-H shifts have barriers too high to occur under these<br />

conditions. Lewis and co-workers have studied several chromophores that should<br />

photogenerate cyclic allenes [161, 162]. Experiments with 111 lead to the conclusion<br />

that products arise from intermolecular protonation, rather than hydrogen<br />

shifts in strained allene 112.<br />

Scheme 3.10 Photochemical Hopf cyclizations.<br />

129


130 3 Distorted Alkenes<br />

The enyne + alkyne cycloaddition has increasingly been applied in the synthesis<br />

of polycyclic structures. Computational studies on enyne + yne cycloadditions support<br />

the existence of cyclic allene intermediates [134, 135]. Scheme 3.11 presents<br />

recent examples. In 1994, Danheiser and co-workers reported that ynone + enyne<br />

cycloadditions of 113 proceed either by thermal activation or with AlCl 3 catalysis<br />

[163]. One advantage of this reaction is that it proceeds directly to an aromatic<br />

ring. Johnson and co-workers applied flash vapor pyrolysis to a simpler structure<br />

114 [134]. The product distribution was consistent with competitive electrocyclic<br />

opening and aromatization of the allene intermediate. Ananikov has reported<br />

computational models for this reaction that agree well with experiment [135].<br />

Saa’s group has reported extensive studies on intramolecular enyne + yne cycloadditions.[164].<br />

As one example, solution-phase thermolysis of 115 unexpectedly<br />

afforded a mixture of benzo[b] and benzo[c] fluorenones 116 and 117 [165]. The<br />

authors proposed that the initially formed allene opens to a 10-membered ring,<br />

which can then close in two directions. This novel result suggests that at high temperature<br />

other isonaphthalenes might undergo complex atom scrambling. In the<br />

final example, Danheiser’s group applied a clever intramolecular benzyne + enyne<br />

addition to prepare 119 and a variety of related heterocycles [166].<br />

Scheme 3.11 Dehydro-Diels-Alder routes to 1,2,4-cyclohexatrienes.


3.3 <strong>Strained</strong> Cyclic Allenes and Cumulenes<br />

3.3.5.1 6-Methylene-1,2,4-Cyclohexatrienes and Related Structures<br />

Thermal cyclization of structures 120 affords (Scheme 3.12) an intermediate that<br />

might be characterized as a chiral allene, diradical or zwitterion. These are not<br />

resonance structures since each implies a different geometry and electronic state.<br />

With X = CH 2, this is known as a Myers-Saito cyclization [167, 168]. At present,<br />

there is neither experimental nor computational support for the allene structure<br />

121. Computations with a closed-shell configuration optimize to chiral structure<br />

121 but the wavefunction is unstable, with a lower energy open-shell solution.<br />

Thus, diradical 122 would appear to be the best description for these structures.<br />

and not allene 121 [169].<br />

Scheme 3.12 Myers-Saito type cyclizations.<br />

3.3.6<br />

Seven-membered Ring Allenes<br />

Strain has been estimated at 14 kcal mol –1 in 1,2-cycloheptadiene 52. The principal<br />

routes to 52 (Scheme 3.13) have come from dehydrohalogenation [108, 110, 111].<br />

The carbenoid route to 52 unexpectedly fails, giving 126 as the major product. This<br />

longstanding mystery has now been explained by Schleyer and co-workers who<br />

showed by DFT calculations that the barrier to ring opening of 125 is unusually<br />

large because of a conformational change leading to the transition state [129].<br />

One new route reported to the parent allene is from treatment of �-bromosilane<br />

124 with fluoride [170].<br />

Scheme 3.13 Chemistry of 1,2,-cycloheptadiene.<br />

Selected examples of other novel 1,2-cycloheptadiene derivatives are summarized<br />

in Scheme 3.14. Huisgen has reported a [4+3] cycloaddition strategy that<br />

leads to the isolation of seven-membered ring ketenimines such as 127; this work<br />

has been reviewed [171]. Agosta and co-workers have reported that the dimer of<br />

129 is isolated by photolysis of ketone 128 [172]. This type of radical combination<br />

might offer a more general route to cyclic allenes. A number of sila-substituted<br />

131


132 3 Distorted Alkenes<br />

Scheme 3.14 Examples of 1,2-cycloheptadiene chemistry.<br />

1,2-cycloheptadienes have been reported. In these cases, the longer C–Si and<br />

Si–Si bonds decrease bending and strain in the allene functional group. In the<br />

most recent example, Baines and co-workers reacted tetramesityldisylene with an<br />

alkynylcyclopropane to produce a mixture of allene 133 and other isomers [173].<br />

3.3.6.1 Cycloheptatetraenes<br />

1,2,4,6-Cycloheptatetraene 134 is one of the best studied cyclic allenes, with<br />

sustained interest in related chemistry over nearly four decades. Interest has<br />

focused on the structure of 134 and its relationship with carbene 135. Initially<br />

viewed as highly improbable, the cyclic allene now is securely characterized as<br />

an important intermediate on the C 7 H 6 potential energy surface. Chemistry of<br />

benzannelated derivatives is also well known [174]. The most common routes to<br />

134 are shown in Scheme 3.15.<br />

Nearly four decades ago, Wentrup showed that 139 could undergo ring expansion<br />

[175]; without the presence of a trapping agent, dimer 137 was isolated. In 1982,<br />

Chapman and co-workers generated 134 in an argon matrix by photolysis of 140<br />

[176]. The IR spectrum showed bands at 1824 and 1816 cm –1 which supported<br />

the chiral cyclic allene. This result was a revelation at a time when strained cyclic<br />

allenes were poorly understood; however, the role of carbene 135 remained<br />

uncertain. In 1996, three theoretical studies converged on the currently accepted<br />

view of this energy surface [177–179]. Essential features are summarized in<br />

Figure 3.7. The chiral allene was predicted to be the lowest energy structure on this<br />

portion of the energy surface, in agreement with experiment and earlier computations.<br />

More interestingly, it was shown that the closed shell (135, state symmetry<br />

1 A1) is not an energy minimum. The open shell singlet state 141 ( 1 A 2) was found<br />

to be the most stable planar structure. This is the transition state for allene


Scheme 3.15 Chemistry of cycloheptatetraene.<br />

3.3 <strong>Strained</strong> Cyclic Allenes and Cumulenes<br />

enantiomerization. Matzinger and Bally later subjected 134 and 139 to a complete<br />

characterization by UV/Vis and IR spectroscopy, where the spectra were assigned<br />

based on high-level ab initio calculations [180]. The route from singlet carbene<br />

139 to 134 is a two-step process, passing through 133. This bicyclic intermediate<br />

has not been observed, probably because it sits in a shallow energy well. Effects<br />

of aryl substituents on the interconversion of phenyl carbene, bicycloheptatriene<br />

and cycloheptatetraene were examined by Geise and Hadad [181]. Reaction of<br />

benzene with atomic carbon offers another entry into the C 7 H 6 surface. Shevlin<br />

has provided evidence that this reaction proceeds by insertion into a C–H bond<br />

to generate phenylcarbene [182, 183].<br />

Rapid dimerization precludes observation of 134 under ordinary solution-phase<br />

conditions. However, with the use of Cram’s hemicarcerand [184], Warmuth<br />

and Marvel reported the generation (Scheme 3.16) of this strained cycloallene<br />

Figure 3.7 Energetics of 1,2,4,5-cycloheptatrienes.<br />

133


134 3 Distorted Alkenes<br />

Scheme 3.16 Generation of “incarcerated” cycloheptatetraene.<br />

at room temperature inside the carcerand [185, 186]. This and other reactions<br />

carried out in molecular flasks stabilizing the short-lived guest are discussed in<br />

Chapter 10. Hemicarceplexed phenylcarbene was found to competitively insert in<br />

the C–H bonds of the hemicarcerand, resulting in low yields of 134. The yield was<br />

increased by incarcerating phenyldiazirine in a partially deuterated hemicarcerand,<br />

resulting in a 67% yield of 134. Similar results were found for the photolysis of<br />

p-tolyldiazirine in this molecular container [187]. Small molecules can diffuse into<br />

the hemicarcerand to characterize reactions of the allene. Reaction with HCl gave<br />

142, while O 2 lead to a novel carbon loss.<br />

In spite of this intense scrutiny, we can find no strain estimates reported for 134.<br />

The isodesmic reaction scheme below, calculated at the B3LYP/6-311+G(d,p) +<br />

ZPVE level of theory, predicts 13.8 kcal mol –1 in allene strain. This is very similar<br />

to 1,2-cycloheptadiene in the estimate of allene strain.<br />

Wentrup’s group has reported extensive studies (Scheme 3.17) on the aza<br />

analogs of 134. In this case, carbene 143 and nitrene 146 interconvert through<br />

145 [188, 189].<br />

Scheme 3.17 Azacycloheptatetraene chemistry.<br />

3.3.7<br />

Eight-membered Ring Allenes<br />

Doering-Moore-Skattebøl chemistry (Scheme 3.18) provides the most common<br />

route to 1,2-cyclooctadienes 147. Strain in this ring size has diminished to ca.<br />

5 kcal mol –1 , but the parent structure can only be briefly isolated at ambient temperature<br />

and most examples should be considered as reactive intermediates [113,


3.3 <strong>Strained</strong> Cyclic Allenes and Cumulenes<br />

114, 190]. Price and Johnson found that 1-t-butyl-1-2-cyclooctadiene is isolable and<br />

could even be purified by gas chromatography [191].<br />

More complex 1,2-cyclooctadienes have been utilized as intermediates. For<br />

example, Moore and co-workers designed an alkoxy-Cope rearrangement route to<br />

eight-membered ring allene 148 which provides a versatile triquinane synthesis<br />

[192].<br />

Scheme 3.18 1,2-Cyclooctadiene chemistry.<br />

3.3.8<br />

Polycyclic Allenes<br />

Only a few bicyclic allenes have been reported. Allene 149 (Scheme 3.19) has<br />

been of interest for some years; early reports [193] on its preparation now seem<br />

to be in doubt but Zen and Balci has recently shown that 149 can be successfully<br />

Scheme 3.19 Polycyclic allenes.<br />

135


136 3 Distorted Alkenes<br />

generated by a more reliable carbenoid route [194]. Balci’s group has also reported<br />

on the allene generated from �-pinene [195]. Okazaki and co-workers recently<br />

reported evidence for 3,4-homoadamantadiene 151, the first allene built on such<br />

a complex tricyclic framework [196]. Dehalogenation of 150 led to isolation of<br />

dimers of allene 151. In the presence of diphenylisobenzofuran (DIBF), adduct<br />

152 was isolated. DFT calculations on 151 predicted a pure bending deformation<br />

of the allene, with an angle of 135.9°. Barton and co-workers reported the synthesis<br />

and isolation of 153. In this structure, the allene is linear but twisted to a dihedral<br />

angle of 72.4º [197].<br />

3.3.9<br />

Cyclic Bisallenes<br />

A modest collection of bis-allenes has been reported. Mitchell and Sondheimer<br />

first described the interesting pericyclic cascade of 154 through bis-allene 155<br />

(Scheme 3.20) to dimers of 156 [198]. More recently, Wang has developed this as<br />

a general route to more complex polycyclics [199]. Cyclic bis-allenes can exist as<br />

diastereomers. Dehmlow and co-workers have shown that dibromides such as 157<br />

open stereospecifically, in this case giving the dl bisallene [200]. Barton’s group<br />

reported the synthesis of bis-allenes with silicon in the ring [201]. Kamigata and<br />

co-workers have reported a detailed ab initio study on meso and dl stereoisomers of<br />

7–10-membered ring bisallenes [202]. For 7–9-membered rings, the meso isomers<br />

were predicted to be of slightly lower energy. Strain in this series was estimated<br />

by comparison to deformation in 2,3-pentadiene. In the 9–10-membered ring<br />

allenes, strain was predicted to be < 2 kcal mol –1 per allene unit; these structures<br />

have allenes that are nearly linear. The value increased to 3.6 or 5.9 and 13.7 or<br />

31.0 kcal mol –1 in dl or meso isomers of the eight- and seven-membered rings,<br />

respectively.<br />

Scheme 3.20 Bis-allene chemistry.<br />

3.3.10<br />

Cyclic Butatrienes<br />

Cyclic butatrienes have four contiguous �-bonded carbons. Spectacular examples<br />

of this functional group include 159 (Scheme 3.21), an intermediate in neo-


Scheme 3.21 Cyclization of cyclic cumulenes to diradicals.<br />

3.3 <strong>Strained</strong> Cyclic Allenes and Cumulenes<br />

carzinostatin chemistry [203, 104] and 160, a 10 � aromatic substance first prepared<br />

by Myers and Finney [205, 206]. These structures have minimal strain but still<br />

cyclize readily to diradicals with modest thermal activation.<br />

3.3.10.1 Butatriene � Bond Deformations and Strain Estimates<br />

Isodesmic and homodesmic estimates for butatriene strain and total strain in cyclic<br />

butatrienes are summarized in Figure 3. Strain increases rapidly with decreasing<br />

ring size [112]. In the smallest ring, Mabry and Johnson characterized 1,2,3-cyclobutatriene<br />

55 as a transition state rather than an energy minimum [207]. DFT<br />

optimized structures for 56–60 show that the butatriene group remains close to<br />

planarity. For a comparable ring size, cyclic butatrienes have ca. twice the strain of<br />

cyclic allenes. Yavari and co-workers have studied the conformational properties<br />

of 56–60 using HF and MP2 methods [208].<br />

Figure 3.8 Strain estimates and chemical behavior of cyclic butatrienes.<br />

3.3.10.2 Five- to Nine-membered Ring Cyclic Butatrienes<br />

The parent cumulene 56 remains unreported, in spite of a number of attempts,<br />

presumably because of its 85 kcal mol –1 of estimated strain. However, Wong and<br />

137


138 3 Distorted Alkenes<br />

Scheme 3.22 Generation and trapping of a 1,2,3-cyclopentatriene.<br />

co-workers have reported routes to thia and azacyclopentatrienes (Scheme 3.22)<br />

[106, 107]. Treatment of iodonium triflate 161 with fluoride at ambient temperature<br />

in the presence of various dienes afforded modest yields of substances assigned<br />

as cycloadducts of 162. Very similar chemistry was utilized to trap the aza analog<br />

of 162.<br />

An earlier study by Shakespeare and Johnson had developed this methodology<br />

(Scheme 3.23) for the first synthesis of 1,2,3-cyclohexatriene 57 [105]. This<br />

new benzene isomer was easily trapped. Hickey and Paquette later described the<br />

efficient addition of 57 to cyclopentadiene [209].<br />

Scheme 3.23 1,2,3-Cyclohexatriene chemistry.<br />

Dehydro-Diels–Alder chemistry provides a more novel approach to derivatives<br />

of 57. Johnson and co-workers found that flash pyrolysis of diyne 164 afforded<br />

a high yield of dienyne 166 [134]. Intramolecular cycloaddition, followed by<br />

electrocyclic opening of 117 provide the most logical mechanism. Ring opening<br />

of 165 is expected to be substantially exothermic and it seems unlikely that this<br />

process can be run in reverse. Lu and co-workers have predicted that 1,3-diynes<br />

may add to Si(100) and Ge(100) surfaces to produce layers of reactive 1,2,3-cyclohexatrienes<br />

[210].


3.3 <strong>Strained</strong> Cyclic Allenes and Cumulenes<br />

Cyclic medium-ring 1,2,3-butatrienes are highly reactive (Scheme 3.24).<br />

Szeimies reported the first synthesis of 58 through rearrangement of bicyclobutene<br />

167 and later described Ni catalyzed dimerization to 168 [211, 212]. The<br />

eight-membered ring butatriene 59 prepared from 169 has only a brief lifetime<br />

in solution [213] while the next homolog 60 is isolable [214].<br />

Scheme 3.24 Cyclic butatriene chemistry.<br />

Some of the most spectacular examples of cyclic cumulenes come from the<br />

domain of organometallic chemistry. In 1993, Buchwald and co-workers reported a<br />

serendipitous synthesis of zirconacyclohepta-2,4,5,6-tetraene 170 [215]. The crystal<br />

structure showed a planar cumulene with an unusual zigzag geometry. Rosenthal<br />

and co-workers have reported in great detail on the chemistry of metallacyclopenta-<br />

2,3,4-trienes 171 which can be readily prepared by transfer of Cp 2M to 1,3-diynes<br />

[216, 217]. Internal bond angles in these structures range from 142–150°. The<br />

molecular geometry and computations indicate strong stabilizing interactions<br />

between the metal and in-plane cumulene � bond. Steric stabilization of these<br />

complexes was also suggested.<br />

3.3.11<br />

Conclusions<br />

<strong>Strained</strong> cyclic allenes and cumulenes present an enormous diversity of structure<br />

and a complete range of chemical behavior, from shelf-stable substances through<br />

increasingly fragile reactive intermediates. These simple hydrocarbons and their<br />

derivatives are now recognized as intermediates in a wide array of chemical<br />

reactions. Ease of preparation and high chemical reactivity have encouraged a<br />

growing number of sophisticated applications in synthesis.<br />

139


140 3 Distorted Alkenes<br />

Acknowledgments<br />

Support from the National Science Foundation for our earlier work on strained<br />

cumulenes is gratefully acknowledged. Thanks to Catherine Johnson for typing<br />

a semi-legible manuscript.<br />

References<br />

1 (a) Liebman, J. F.; Greenberg, A. Chem.<br />

Rev., 1976, 76, 311; Greenberg, A.;<br />

Liebman, J. F. <strong>Strained</strong> Organic<br />

Molecules; Academic Press, New York,<br />

1978; (b) Sandström, J. Chem. Rev. 1989,<br />

89, 973, (c) Luef, W.; Keese, R. Top.<br />

Stereochem. 1991, 20, 231.<br />

2 (a) Lenoir, D. Nachr. Chem. Techn: Lab.<br />

1979, 27, 762, (b) Ermer, O. Aspekte von<br />

Kraftfeldrechnungen; W. Baur Verlag,<br />

München, 1981, (c) Wiberg, K. Angew.<br />

Chem. Int. Ed. Engl. 1986, 25, 312,<br />

(d) Hopf, H. Classics in Hydrocarbon<br />

Chemistry; Wiley-VCH, Weinheim, 2000.<br />

3 (a) Mastryukov, V. S.; Boggs, J. E.<br />

Struct. Chem. 2000, 11, 97; (b) Novak, I.<br />

J. Chem. Inf. Model. 2005, 45, 334;<br />

(c) Vazquez, S.; Camps, P. Tetrahedron<br />

2005, 61, 5147.<br />

4 (a) Lenoir, D. Tetrahedron Lett. 1972,<br />

40, 4049, (b) Grant, D.; Rooney, J. J.;<br />

Samman, N. G.; Mc Kervey, M. A.<br />

J. Chem. Soc. Chem. Commun. 1972, 21,<br />

1186.<br />

5 (a) Lenoir, D.; Firl, J. Liebigs Ann.<br />

Chem. 1974, 273, 1467, (b) Lenoir, D.;<br />

Kornrumpf, W.; Fritz, H. P. Chem. Ber.<br />

1983, 116, 2390.<br />

6 (a) Wiseman, J. R. J. Am. Chem. Soc.<br />

1967, 89, 5966, (b) Wiseman, J. R.;<br />

Pletcher, W. A. J. Am. Chem. Soc. 1970,<br />

92, 956.<br />

7 Maier, W. F.; Schleyer, P. v. R.<br />

J. Am. Chem. Soc. 1981, 103, 1891.<br />

8 (a) Conlin, R. T.; Miller, R. D.;<br />

Michl, J. J. Am. Chem. Soc. 1979,<br />

101, 7637; (b) Tae, E. L.; Zhu, Z. O.;<br />

Platz, M. S. J. Phys. Chem. A, 2001,105,<br />

3808; (c) Moss, R. A.; Sauers, R. R.;<br />

Sheridan, R. S.; Tian, J.; Zuev, P. S.<br />

J. Am. Chem. Soc. 2004, 126; 10196,<br />

(d) Okazaki, T.; Isobe, H.; Kitakawa, H.<br />

Bull. Chem. Soc. Japan 1996, 69, 2053.<br />

9 Squillacote, M. E.; DeFellipis, J.; Shu, Q.<br />

J. Am. Chem. Soc. 2005, 127, 15983.<br />

10 (a) Lenoir, D. Synthesis, 1989, 833,<br />

(b) McMurry, J. Chem. Rev. 1989, 89,<br />

1513.<br />

11 Doering, W. v. E.; Roth, W. R.;<br />

Bauer, F.; Breuckmann, R.;<br />

Ebbrecht, T.; Herbold, M.; Schmidt, R.;<br />

Lennartz, H.-W.; Lenoir, D.; Boese, R.<br />

Chem. Ber. 1989, 122, 1263.<br />

12 (a) Lenoir, D.; Dauner, H.; Frank, R. M.<br />

Chem. Ber. 1980, 113, 2636;<br />

(b) Eierdanz, H.; Potthoff, S.; Bolze, R.;<br />

Berndt, A. Angew. Chem. 1984, 96, 513.<br />

13 Lenoir, D.; Liebman, J. F., unpublished<br />

results.<br />

14 Lenoir, D.; Wattenbach, C.;<br />

Liebman, J. F. Struct. Chem. 2006,17,<br />

419.<br />

15 Brooks, P. R.; Bishop, R.; Counter, J. A.;<br />

Tiekink, E. R. T. J. Org. Chem. 1994, 59,<br />

1365.<br />

16 Sulzbach, H. M.; Bolton, E.; Lenoir, D.;<br />

Schleyer, P. v. R.; Schaeffer III, H. F.<br />

J. Am. Chem. Soc. 1996, 118, 9908.<br />

17 Rogers, D. W. Heats of Hydrogenation:<br />

Experimental and Computational<br />

Thermochemistry of Organic Compounds<br />

(World Scientific, New Jersey, 2006).<br />

All unreferenced enthalpies of hydrogenation<br />

are taken implicitly from this<br />

source.<br />

18 Rogers, D. W.; Dagdagan, O. A.;<br />

Allinger, N. L. J. Am. Chem. Soc. 1979,<br />

101, 671.<br />

19 Puterbaugh, W. H.; Newman, M. S.<br />

J. Am. Chem. Soc. 1959, 81, 1611.<br />

20 Fahey, R. C. J. Am. Chem. Soc. 1966, 88,<br />

4681.<br />

21 Kropp, P. J.; Tise, F. J. Am. Chem. Soc.<br />

1981, 103, 7293.<br />

22 Slayden, S. W.; Liebman, J. F. In:<br />

The Chemistry of Functional Groups


Supplement A3: The Chemistry of Doubly-<br />

Bonded Functional Groups (ed. S. Patai),<br />

p. 537; Wiley, Chichester, 1997.<br />

23 Liebman, J. F. In: The Chemistry of<br />

Functional Groups Supplement A2:<br />

The Chemistry of Dienes and Polyenes,<br />

Vol. 1 (ed. Z. Rappoport), p. 67; Wiley,<br />

Chichester, 1997.<br />

24 Chou, P K.; Kass, S. R. J. Am. Chem. Soc.<br />

1991, 113, 697.<br />

25 Hoenigman, R. L.; Kato, S.;<br />

Bierbaum, V. M.; Borden, W. T.<br />

J. Am. Chem. Soc. 2005, 127, 17772.<br />

26 (a) Staneke, P. O; Ingemann, S.;<br />

Eaton, P.; Nibbering, N. M. M.;<br />

Kass, S. R. J. Am. Chem. Soc. 1994,<br />

116, 5445; (b) Hare, M.; Emrick, T.;<br />

Eaton, P. E.; Kass, S. R. J. Am. Chem.<br />

Soc. 1997, 119, 237.<br />

27 Rogers, D. W.; McLafferty, F. J.<br />

J. Phys. Chem. A 2000, 104, 9356.<br />

28 (a) Kiplinger, J. P.; Tollens, F. R.;<br />

Marshall, A. G.; Kobayashi, T.;<br />

Lagerwall, D. R.; Paquette, L. A.;<br />

Bartmess, J. E. J. Am. Chem. Soc.<br />

1989, 111, 6914; (b) Broadus, K. M.;<br />

Kass, S. R.; Osswald, T.; Prinzbach, H.<br />

J. Am. Chem. Soc. 2000, 122, 10964.<br />

29 (a) Diky, V. V.; Frenkel, M.;<br />

Karpushenkava, L. S. Thermochim.<br />

Acta 2003, 408, 115; (b) Roux, M. V.;<br />

Dávalos, J. Z.; Jiménez, P.; Notario, R.;<br />

Castaño, O.; Chickos, J. S.; Hanshaw, W.;<br />

Zhao, H.; Rath, N.; Liebman, J. F.;<br />

Farivar, B. S.; Bashir-Hashemi, A. J. Org.<br />

Chem. 2005, 70, 5461; (c) Liebman, J. F.;<br />

Slayden, S. W. In: The Chemistry of<br />

Cyclobutanes, (ed. Z. Rappoport and<br />

J. F. Liebman), Wiley, Chichester, 133<br />

(2005).<br />

30 (a) Beckhaus, H.-D.; Rüchardt, C.;<br />

Lagerwall, D. R.; Paquette, L. A.;<br />

Wahl, F.; Prinzbach, H. J. Am. Chem.<br />

Soc. 1994, 116, 11775; 1995, 117, 8885.<br />

31 See, for example, the three sequential<br />

papers by Prinzbach and his coworkers,<br />

Sackers, E.; Osswald, T.; Weber, K.;<br />

Keller, M.; Hunkler, D.; Woerth, J.;<br />

Knothe, L.; Prinzbach, H. Chem. Eur.<br />

J. 2006, 12, 6242; Wahl, F.; Weiler, A.;<br />

Landenberger, P.; Sackers, E.; Voss, T.;<br />

Haas, A.; Lieb, M.; Hunkler, D.;<br />

Woerth, J.; Knothe, L.; Prinzbach, H.<br />

Chem. Eur. J. 2006, 12, 6255;<br />

References<br />

Prinzbach, H.; Wahl, F.; Weiler, A.;<br />

Landenberger, P.; Woerth, J.; Scott, L. T.;<br />

Gelmont, M.; Olevano, D.; Sommer, F.;<br />

von Issendorff, B. Chem. Eur. J. 2006, 12,<br />

6268.<br />

32 See, for example C 60 H 36 ,<br />

Pimenova, S. M.; Melkhanova, S. V.;<br />

Kolesov, V. P.; Lobach, A. S.<br />

J. Phys. Chem. B. 2002, 106, 2127;<br />

C 60 F 36 , Papina, T. S.; Kolesov, V. P.;<br />

Lukyanova, V. A.; Boltalina, O. V.;<br />

Lukonin, A. Yu.; Sidorov, L. N.<br />

J. Phys. Chem. B. 2000, 104, 5403.<br />

33 Dallwick, E.; Briner, E. Helv. Chim. Acta<br />

1953, 36, 1166.<br />

34 Hehre, W. J.; Pople, J. A. J. Am. Chem.<br />

Soc. 1975, 97, 6941.<br />

35 Jordan, K. D.; Burrow, P. D. Acc. Chem.<br />

Res. 1978, 11, 341.<br />

36 Swen-Walstra, S. C.; Visser, G. J.<br />

J. Chem. Soc. D 1971, 82.<br />

37 Lenoir, D.; Frank, R. M.; Cordt, F.;<br />

Gieren, A.; Lamm, V. Chem. Ber. 1980,<br />

113, 739.<br />

38 Okazaki, T.; Ogawa, K.; Kitakawa, T.;<br />

Takeuchi, K. J. Org. Chem. 2002, 67,<br />

5981.<br />

39 Strating, J.; Wieringa, J. H.; Wynberg, H.<br />

J. Chem. Soc. D. Chem. Commun. 1969,<br />

907.<br />

40 Brown, R. S. Acc. Chem. Res. 1997, 30,<br />

131.<br />

41 Chiappe, C.; Pomelli, C. S.; Lenoir, D.;<br />

Wattenbach, C. J. Mol. Mod. 2006, 12,<br />

631.<br />

42 (a) Mori, T.; Rathore, R. Chem Commun.<br />

1998, 927, (b) Mori, T.; Rathore, R.;<br />

Lindeman, S. V.; Kochi, J. K. Chem<br />

Commun. 1998, 1238.<br />

43 (a) Kazakov, V. P.; Ostakhov, S. S.;<br />

Alyab’ev, A. S.; Osina, I. O, High Ener.<br />

Chem. 2006, 40, 248; (b) Rathore, R.;<br />

Zhu, C.; Lindeman, S. V.; Kochi, J. K.;<br />

Perkin 2, 2000, 1837; (c) Nelsen, S. F.;<br />

Akaba, R. J. Am. Chem. Soc. 1981, 103,<br />

2096; (d) Bosch, E.; Kochi, J. K. J. Am.<br />

Chem. Soc. 1996, 118, 1319.<br />

44 (a) Gano, J. E.; Kluwe, C.;<br />

Kirschbaum, K.; Pinkerton, A. A.;<br />

Sekher, P.; Skrzypczak-Jankun;<br />

Subramaniam, G.; Lenoir, D. Acta<br />

Cryst. 1997, C53, 1723, (b) Gano, J. E.;<br />

Park, B.-S.; Pinkerton, A. A.; Lenoir, D.<br />

Acta. Cryst. 1991, C47, 162.<br />

141


142 3 Distorted Alkenes<br />

45 Lenoir, D.; Gano, J. E.; McTague, J. A.<br />

Tetrahedron Lett. 1986, 27, 5339.<br />

46 Frank, R., Thesis, Technical University<br />

Munich, 1994.<br />

47 Ozawa, T. Bull. Chem. Soc. Japan, 1965,<br />

38, 188.<br />

48 Gano, J. E.; Park, B.-S.;<br />

Subramaniam, G.; Lenoir, D.; Gleiter, R.<br />

J. Org. Chem. 1991, 56, 4806.<br />

49 (a) Bellucci, G.; Chiappe, C.;<br />

Bianchini, R.; Lenoir, D.; Herges, D.<br />

J. Am. Chem. Soc. 1995, 117, 12001;<br />

(b) Lenoir, D.; Chiappe, C. Chem. Eur. J.<br />

2003, 9, 1037.<br />

50 Gano, J. E.; Park, B.-S.; Pinkerton, A. A.;<br />

Lenoir, D. J. Org. Chem. 1990, 55, 2688.<br />

51 (a) Oelgemoeller, M.; Brem, B.;<br />

Frank, R.; Schneider, S.; Lenoir, D.;<br />

Hertkorn, N.; Origani, P.; Lemmen, P.;<br />

Lex, Y.; Inoue, Y. J. Chem. Soc., Perkin<br />

Trans. 2, 2002, 1760, (b) Oelgemoeller, M.<br />

unpublished results.<br />

52 Roth, W. R.; Bohm, M.; Lenhartz, H.-W.;<br />

Vogel, E. Angew. Chem. 1983, 95, 1011.<br />

53 Roth, W. R.; Klaerner, F. G.; Siepert, G.;<br />

Lennartz, H. W. Chem. Ber. 1992, 125,<br />

217.<br />

54 (a) Williams, R. V. Chem. Rev. 2001,<br />

101, 1185; (b) Caramori, G. F.;<br />

de Oliveira, K. T.; Galembeck, S. E.;<br />

Bultinck, P.; Constantino, M. G.<br />

J. Org. Chem. 2007, 72, 76.<br />

55 van Eis, M. J.; Wijsman, G. W.;<br />

de Wolf, W. H.; Bickelhaupt, F.;<br />

Rogers, D. W.; Kooijman, H.; Spek, A. L.<br />

Eur. J. Chem. 2000, 6, 1537.<br />

56 McEwen, A. B.; Schleyer, P. v. R.;<br />

J. Am. Chem. Soc. 1986, 108, 3951.<br />

57 (a) Lin, S.-T.; Wharry, D. L.; Yates, R. L.;<br />

Garti, N.; Pranata, J.; Siegel, S.;<br />

Skrabal, P. J. Org. Chem. 1994, 59, 7056;<br />

(b) Yang, F.-M.; Lin, S.-T. J. Org. Chem.<br />

1997, 62, 2727; (c) Hopf, H.;<br />

Savinsky, R.; Jones, P. G.; Dix, I.;<br />

Ahrens, B. Liebigs Annalen/Recueil 1997,<br />

1499.<br />

58 Shieh, C.; McNally, D.; Boyd, R. H.<br />

Tetrahedron 1969, 25, 3653.<br />

59 Mollins, E.; Miravitles, C.; Espinosa, E.;<br />

Ballester, M. J. Org. Chem. 2002, 67,<br />

7175.<br />

60 Beck, A.; Gompper, R.; Polborn, K.;<br />

Wagner, H. U. Angew. Chem. Int. Ed.<br />

Engl. 1993, 32, 1352.<br />

61 Vollhardt, K. P. C.; Cammack, J. K.;<br />

Matzger, A. J.; Bauer, A.; Capps, K. B.;<br />

Hoff, C. D. Inorg. Chem. 1999, 38, 2624.<br />

62 Turner, R. B.; Meador, W. R.;<br />

Doering, W. v. E.; Knox, L. H.;<br />

Mayer, J. R.; Wiley, D. W. J. Am. Chem.<br />

Soc. 1957, 79, 4127.<br />

63 We are reminded of the quote given in<br />

the preface of the monograph reference<br />

1a that beckons as it berates “It is not thy<br />

duty to complete the work, but neither<br />

art thou free to desist of it” (from the<br />

Ethics of the Father, the Talmud, 2:21).<br />

64 (a) Keese, R.; Krebs, E.-P. Angew. Chem.,<br />

Int. Ed. 1971, 10, 262. (b) Keese, R.;<br />

Krebs, E.-P. Angew. Chem., Int. Ed.<br />

1972, 11, 518. (c) Warner, P. M. Chem.<br />

Rev. 1989, 89, 1067. (d) Grimme, W.;<br />

Bertsch, A.; Flock, H.; Noack, T.;<br />

Krauthäuser, S. Synlett 1998, 1175.<br />

(e) Roach, P.; Warmuth, R. Angew.<br />

Chem., Int. 2003, 42, 3039.<br />

65 Szeimies, G. Chimia 1981, 35, 243.<br />

66 Borden, W. T. Chem. Rev. 1989, 89, 1095.<br />

67 Mastryukov, V. S.; Chen, K.-H.;<br />

Allinger, N. L. J. Phys. Chem. A 2001,<br />

105, 8562.<br />

68 Vázquez, S.; Camps, P. Tetrahedron 2005,<br />

61, 5147.<br />

69 Volland, W. V.; Davidson, E. R.;<br />

Borden, W. T. J. Am. Chem. Soc. 1979,<br />

101, 533.<br />

70 Weinshenker, N. M.; Greene, F. D.<br />

J. Am. Chem. Soc. 1968, 90, 506.<br />

71 See for example: (a) Kroto, H. W.;<br />

Heath, J. R.; OBrien, S. C.; Curl, R. F.;<br />

Smalley, R. E. Nature (London) 1985, 318,<br />

162. (b) Diederich, F.; Whetten, R. L.<br />

Angew. Chem., Int. Ed. 1991, 30, 678.<br />

(c) Curl, R. F.; Smalley, R. E. Sci. Am.<br />

1991, 54. (d) Kroto, H. W.; Allaf, A. W.;<br />

Balm, S. P. Chem. Rev. 1991, 91, 1213.<br />

72 Lukin, K.; Eaton, P. E. J. Am. Chem. Soc.<br />

1995, 117, 7652.<br />

73 (a) Wagner, H. U.; Szeimies, G.;<br />

Chandrasekhar, J.; Schleyer, P. v. R.;<br />

Pople, J. A.; Binkley, J. S. J. Am. Chem.<br />

Soc. 1978, 100, 1210. (b) Wiberg, K. B.;<br />

Artis, D. R.; Bonneville, G. J. Am. Chem.<br />

Soc. 1991, 113, 7969. (c) Yates, B. F.<br />

J. Organomet. Chem. 2001, 635, 142.<br />

74 (a) Szeimies, G. In Halton, B., Ed.;<br />

Advances in Strain in Organic Chemistry;<br />

JAI: Greenwich, CT, 1992, Vol. 2,


p. 1–55. (b) Sauers, R. R.; Harris, J. S.<br />

J. Org. Chem. 2001, 66, 7951.<br />

75 (a) Prinzbach, H.; Weber, K. Angew.<br />

Chem., Int. Ed. 1994, 33, 2239.<br />

(b) Melder, J.-P.; Weber, K.; Weiler, A.;<br />

Sackers, E.; Fritz, H.; Hunkler, D.;<br />

Prinzbach, H. Res. Chem. Intermed. 1996,<br />

22, 667. (c) Prinzbach, H.; Weiler, A.;<br />

Landerberger, P.; Wahl, F.; Wörth, J.;<br />

Scott, L. T.; Gelmont, M.; Olevano, D.;<br />

Issendorff, B. v. Nature 2000, 407, 60.<br />

(d) Reinbold, J.; Sackers, E.; Obwald, T.;<br />

Weber, K.; Weiler, A.; Voss, T.;<br />

Hunkler, D.; Wörth, J.; Knothe, L.;<br />

Sommer, F.; Morgner, N.; Issendorff, B.<br />

v.; Prinzbach, H. Chem. Eur. J. 2002,<br />

8, 509. (e) Reinbold, J.; Bertau, M.;<br />

Voss, T.; Hunkler, D.; Knothe, L.;<br />

Prinzbach, H.; Neschchadin, D.;<br />

Gescheidt, G.; Mayer, B.; Martin, H.-D.;<br />

Heinze, J.; Prakash, G. K. S.; Olah, G. A.<br />

Helv. Chim. Acta 2001, 84, 1518.<br />

76 Hrovat, D. A.; Borden, W. T.<br />

J. Am. Chem. Soc. 1988, 110, 4710.<br />

77 Camps, P.; Font-Bardia, M.; Méndez, N.;<br />

Pérez, F.; Pujol, X.; Solans, X.;<br />

Vázquez, S.; Vilalta, M. Tetrahedron 1998,<br />

54, 4679.<br />

78 Maier, W. F.; Schleyer, P. v. R.<br />

J. Am. Chem. Soc. 1981, 103, 1891.<br />

79 Smith, J. M.; Hrovat, D. A.;<br />

Borden, W. T.; Allan,M.; Asmis, K. R.;<br />

Bulliard,C.; Haselbach, E.; Meier, U. C.<br />

J. Am. Chem. Soc. 1993, 115, 3816.<br />

80 Wiberg, K. B.; Matturro, M.;<br />

Okarma, P. J.; Jason, M. E. J. Am. Chem.<br />

Soc. 1984, 106, 2194.<br />

81 Bartlett, P. D.; Banavali, R. J. Org. Chem.<br />

1991, 56, 6043.<br />

82 Radziszewski, J. G.; Yin, T.-K.;<br />

Miyake, F.; Renzoni, G. E.;<br />

Hrovat, D. A.; Borden, W. T.; Michl, J.<br />

J. Am. Chem. Soc. 1986, 108, 3544.<br />

83 Hrovat, D. A.; Miyake, F.; Trammell, G.;<br />

Gilbert, K. E.; Mitchell, J.; Clardy, J.;<br />

Borden, W. T. J. Am. Chem. Soc. 1987,<br />

109, 5524.<br />

84 Koopmans, T. Physica 1933, 1, 104.<br />

85 Cleven, C. D.; Hoke, S. H., II;<br />

Cooks, R. G.; Hrovat, D. A.; Smith, J. M.;<br />

Lee, M.-S.; Borden, W. T. J. Am. Chem.<br />

Soc. 1996, 118, 10872.<br />

86 Renzoni, G. E.; Yin, T.-K.; Miyake, F.;<br />

Borden, W. T. Tetrahedron 1986, 42, 1581.<br />

References<br />

87 Yin, T.-K.; Radziszewski, J. G.;<br />

Renzoni, G. E.; Downing, J. W.;<br />

Michl, J.; Borden, W. T. J. Am. Chem.<br />

Soc. 1987, 109, 820.<br />

88 Radziszewski, J. G.; Yin, T.-K.;<br />

Renzoni, G. E.; Hrovat, D. A.;<br />

Borden, W. T.; Michl, J. J. Am. Chem.<br />

Soc. 1993, 115, 1454.<br />

89 Renzoni, G. E.; Yin, T.-K.; Borden, W. T.<br />

J. Am. Chem. Soc. 1986, 108, 7121.<br />

90 Johnson, W. T. G., Ph. D. thesis,<br />

University of Washington, 1999.<br />

91 Camps, P.; Luque, F. J.; Orozco, M.;<br />

Pérez, F.; Vázquez, S. Tetrahedron Lett.<br />

1996, 37, 8605.<br />

92 Camps, P.; Font-Bardia, M.; Pérez, F.;<br />

Solans, X.; Vázquez, S. Angew. Chem.,<br />

Int. Ed. 1995, 34, 912.<br />

93 Camps, P.; Font-Bardia, M.; Méndez, N.;<br />

Pérez, F.; Pujol, X.; Solans, X.;<br />

Vázquez, S.; Vilalta, M. Tetrahedron 1998,<br />

54, 4679.<br />

94 Branan, B. M.; Paquette, L. A.;<br />

Hrovat, D. A.; Borden, W. T. J. Am.<br />

Chem. Soc. 1992, 114, 774.<br />

95 Theophanous, F. A.; Tasiopoulos, A. J.;<br />

Nicolaides, A.; Zhou,X.;<br />

Johnson, W. T. G. Borden, W. T. Org.<br />

Lett. 2006, 8, 3001.<br />

96 Nicolaides, A.; Smith, J. M.; Kumar, A.;<br />

Barnhardt, D. M.; Borden, W. T.<br />

Organometallics 1995, 14, 3475.<br />

97 Morokuma, K.; Borden, W. T. J. Am.<br />

Chem. Soc. 1991, 113, 1912.<br />

98 Pellizer, G.; Graziani, M.; Lenarda, M.;<br />

Heaton, B. T. Polyhedron 1983, 2, 657.<br />

99 Bennet, M. A. Pure Appl. Chem. 1989,<br />

61,1695.<br />

100 (a) Jason, M. E.; McGinnety, J. A. Inorg.<br />

Chem. 1981, 20, 4000. (b) Jason, M. E.;<br />

McGinnety, J. A.; Wiberg, K. B.<br />

J. Am. Chem. Soc. 1974, 96, 6532.<br />

101 Ball, W. J.; Landor, S. R. Proc. Chem. Soc.<br />

1961, 143–144.<br />

102 Algi, F.; Ozen, R.; Balci, M. Tetrahedron<br />

Lett. 2002, 43, 3129–3131.<br />

103 Meier, H.; Koenig, P. Nouv. J. Chim.<br />

1986, 10, 437–438.<br />

104 Moore, W. R.; Ozretich, T. M.<br />

Tetrahedron Lett. 1967, 3205.<br />

105 Shakespeare, W. C.; Johnson, R. P.<br />

J. Am. Chem. Soc. 1990, 112, 8578–8579.<br />

106 Ye, X.-S.; Li, W.-K.; Wong, H. N. C.<br />

J. Am. Chem. Soc. 1996, 118, 2511–2512.<br />

143


144 3 Distorted Alkenes<br />

107 Liu, J. H.; Chan, H. W.; Xue, F.; Wang,<br />

Q. G.; Mak, T. C. W.; Wong, H. N. C.<br />

J. Org. Chem. 1999, 64, 1630–1634.<br />

108 Johnson, R. P. Chem. Rev. 1989, 89,<br />

1111–1124.<br />

109 Johnson, R. P. Advances in Theoretically<br />

Interesting Molecules 1989, 1, 401–436.<br />

110 Balci, M.; Taskesenligil, Y. Advances in<br />

<strong>Strained</strong> and Interesting Organic Molecules<br />

2000, 8, 43–81.<br />

111 Christl, M. Modern Allene Chemistry<br />

2004, 1, 243–357.<br />

112 Daoust, K. J.; Hernandez, S. M.;<br />

Konrad, K. M.; Mackie, I. D.;<br />

Winstanley, J., Jr.; Johnson, R. P.<br />

J. Org. Chem. 2006, 71, 5708–5714.<br />

113 Marquis, E. T.; Gardner, P. D.<br />

Tetrahedron Lett. 1966, 2793–2798.<br />

114 Christl, M.; Rudolph, M.; Peters, E.-M.;<br />

Peters, K.; von Schnering, H. G., Angew.<br />

Chem. Int. Ed. Engl. 1996, 34, 2730–2732.<br />

115 Angus, R. O., Jr.; Schmidt, M. W.;<br />

Johnson, R. P. J. Am. Chem. Soc. 1985,<br />

107, 532–537.<br />

116 Bettinger, H. F.; Schreiner, P. R.;<br />

v. Schleyer, P. R.; Schaefer, III, H. F.<br />

J. Phys. Chem. 1996, 100, 16147–16154.<br />

117 Brudzynski, R. J.; Hudson, B. S.<br />

J. Am. Chem. Soc. 1990, 112, 4963–4965.<br />

118 Engels, B.; Schoneeboom, J. C.;<br />

Munester, A. F.; Groetsch, S.; Christl, M.<br />

J. Am. Chem. Soc. 2002, 124, 287–297.<br />

119 Balci, M.; Jones, W. M. J. Am. Chem. Soc.<br />

1980, 102, 7607–7608.<br />

120 Zheng, M.; DiRico, K. J.;<br />

Kirchhoff, M. M.; Phillips, K. M.;<br />

Cuff, L. M.; Johnson, R. P. J. Am. Chem.<br />

Soc. 1993, 115, 12167–12168.<br />

121 Ceylan, M.; Kinal, A.; Suetbeyaz, Y.;<br />

Balci, M. Turk. J. Chem. 2003, 27,<br />

287–294.<br />

122 Wittig, G.; Fritze, P., Angew. Chem. 1966,<br />

78, 905.<br />

123 v. Doering W. E.; LaFlamme, P. M.<br />

Tetrahedron 1958, 2, 75–79.<br />

124 Moore, W. R.; Ward, H. R. J. Org. Chem.<br />

1960, 25, 2073.<br />

125 Skattebol, L. Tetrahedron Lett. 1961,<br />

167–172.<br />

126 Sydnes, L. K. Chem. Rev. 2003, 103,<br />

1133–1150.<br />

127 Bettinger, H. F.; Schleyer, P. V.;<br />

Schreiner, P.; Schaefer, H. F.<br />

J. Org. Chem. 1997, 62, 9267–9275.<br />

128 Werstiuk, N. H.; Roy, C. D.; Ma, J.<br />

Can. J. Chem. 1996, 74, 1903–1905.<br />

129 Bettinger, H. F.; v. Schleyer, P. R.;<br />

Schreiner, P. R.; Schaefer, III, H. F.<br />

J. Org. Chem. 1997, 62, 9267–9275.<br />

130 Wentrup, C.; Gross, G.; Maquestiau, A.;<br />

Flammang, R. Angew. Chem. 1983, 95,<br />

551.<br />

131 Runge, A.; Sander, W. Tetrahedron Lett.<br />

1986, 27, 5835–5838.<br />

132 Caubere, P. Chem. Rev. 1993, 93,<br />

2317–2334.<br />

133 Nendel, M.; Tolbert, L. M.;<br />

Herring, L. E.; Islam, M. N.; Houk, K. N.<br />

J. Org. Chem. 1999, 64, 976–983.<br />

134 Burrell, R. C.; Daoust, K. J.; Bradley, A. Z.;<br />

DiRico, K. J.; Johnson, R. P. J. Am. Chem.<br />

Soc. 1996, 118, 4218–4219.<br />

135 Ananikov, V. P. J. Phys. Org. Chem. 2001,<br />

14, 109–121.<br />

136 Miller, B.; Ionescu, D. Tetrahedron Lett.<br />

1994, 35, 6615–6618.<br />

137 Herz, H.-G.; Schatz, J.; Maas, G. J. Org.<br />

Chem. 2001, 66, 3176–3181.<br />

138 Christl, M.; Braun, M. Chem. Ber. 1989,<br />

122, 1939–1946.<br />

139 Christl, M.; Braun, M.; Wolz, E.;<br />

Wagner, W. Chem. Ber. 1994, 127,<br />

1137–1142.<br />

140 Ruzziconi, R.; Naruse, Y.; Schlosser, M.<br />

Tetrahedron 1991, 47, 4603–4610.<br />

141 Elliott, R. L.; Nicholson, N. H.;<br />

Peaker, F. E.; Takle, A. K.;<br />

Richardson, C. M.; Tyler, J. W.; White, J.;<br />

Pearson, M. J.; Eggleston, D. S.;<br />

Haltiwanger, R. C. J. Org. Chem. 1997,<br />

62, 4998–5016.<br />

142 Hofmann, M. A.; Bergstrasser, U.;<br />

Reiss, G. J.; Nyulaszi, L.; Regitz, M.<br />

Angew. Chem. Int. Ed. Engl. 2000, 39,<br />

1261–1263.<br />

143 Pang, Y.; Petrich, S. A.; Young, V. G., Jr.;<br />

Gordon, M. S.; Barton, T. J. J. Am. Chem.<br />

Soc. 1993, 115, 2534–2536.<br />

144 Jones, W. M.; Klosin, J. Advances in<br />

Organometallic Chemistry, Vol. 42, 1998,<br />

pp. 147–221.<br />

145 Yen, Y. -S.; Lin, Y.-C.; Huang, S.-L.;<br />

Liu, Y.-H.; Sung, H.-L.; Wang, Y. J. Am.<br />

Chem. Soc. 2005, 127, 18037–18045.<br />

146 Emanuel, C. J.; Shevlin, P. B. J. Am.<br />

Chem. Soc. 1994, 116, 5991–5992.<br />

147 Pan, W.; Shevlin, P. B. J. Am. Chem. Soc.<br />

1997, 119, 5091–5094.


148 McKee, M. L.; Shevlin, P. B.; Zottola, M.<br />

J. Am. Chem. Soc. 2001, 123, 9418–9425.<br />

149 Musch, P. W.; Scheidel, D.;<br />

Engels, B. J. Phys. Chem. A 2003, 107,<br />

11223–11230.<br />

150 Khasanova, T.; Sheridan, R. S. J. Am.<br />

Chem. Soc. 2000, 122, 8585–8586.<br />

151 Nikitina, A.; Sheridan, R. S. J. Am.<br />

Chem. Soc. 2002, 124, 7670–7671.<br />

152 Nikitina, A. F.; Sheridan, R. S. Org. Lett.<br />

2005, 7, 4467–4470.<br />

153 Groetsch, S.; Spuziak, J.; Christl, M.<br />

Tetrahedron 2000, 56, 4163–4171.<br />

154 Christl, M.; Braun, M.; Fischer, H.;<br />

Groetsch, S.; Mueller, G.; Leusser, D.;<br />

Deuerlein, S.; Stalke, D.; Arnone, M.;<br />

Engels, B. Eur. J. Org. Chem. 2006,<br />

5045–5058.<br />

155 Hopf, H.; Musso, H. Angew. Chem. Int.<br />

Ed. Engl. 1969, 8, 680.<br />

156 Zimmermann, G. Eur. J. Org. Chem.<br />

2001, 457–471.<br />

157 Roth, W. R.; Hopf, H.; Horn, C.<br />

Chem. Ber. 1994, 127, 1781–1795.<br />

158 Prall, M.; Kruger, A.; Schreiner, P. R.;<br />

Hopf, H. Chem. Eur. J. 2001, 7,<br />

4386–4394.<br />

159 Miller, J. A.; Klippenstein, S. J. J. Phys.<br />

Chem. A 2003, 107, 7783–7799.<br />

160 Op den Brouw, P. M.; Laarhoven, W. H.<br />

J. Chem. Soc. Perkin 2 1982, 795–799.<br />

161 Sajimon, M. C.; Lewis, F. D. Photochem.<br />

Photobiol. Sci. 2005, 4, 789–791.<br />

162 Lewis, F. D.; Karagiannis, P. C.;<br />

Sajimon, M. C.; Lovejoy, K. S.; Zuo, X.;<br />

Rubin, M.; Gevorgyan, V. Photochem.<br />

Photobiol. Sci. 2006, 5, 369–375.<br />

163 Danheiser, R. L.; Gould, A. E.;<br />

de la Pradilla, R. F.; Helgason, A. L.<br />

J. Org. Chem. 1994, 59, 5514–5515.<br />

164 Rodriguez, D.; Navarro-Vazquez, A.;<br />

Castedo, L.; Dominguez, D.; Saa, C.<br />

J. Org. Chem. 2003, 68, 1938–1946.<br />

165 Rodriguez, D.; Martinez-Esperon, M. F.;<br />

Navarro-Vazquez, A.; Castedo, L.;<br />

Dominguez, D.; Saa, C. J. Org. Chem.<br />

2004, 69, 3842–3848.<br />

166 Hayes, M. E.; Shinokubo, H.;<br />

Danheiser, R. L. Org. Lett. 2005, 7,<br />

3917–3920.<br />

167 Wang, K. K. Modern Allene Chemistry<br />

2004, 2, 1091–1126.<br />

168 Hughes, T. S.; Carpenter, B. K. J. Chem.<br />

Soc., Perkin Trans. 2 1999, 2291–2298.<br />

References<br />

169 Cremeens, M. E.; Hughes, T. S.;<br />

Carpenter, B. K. J. Am. Chem. Soc. 2005,<br />

127, 6652–6661.<br />

170 Sutbeyaz, Y.; Ceylan, M.; Secen, H.<br />

J. Chem. Res., Synop. 1993, 293.<br />

171 Huisgen, R.; Mloston, G.; Langhals, E.;<br />

Oshima, T. Helv. Chim. Acta 2002, 85,<br />

2668–2685.<br />

172 Venugopal, D.; Margaretha, P.;<br />

Agosta, W. C. J. Org. Chem. 1993, 58,<br />

6629–6633.<br />

173 Gottschling, S. E.; Milnes, K. K.;<br />

Jennings, M. C.; Baines, K. M.<br />

Organometallics 2005, 24, 3811–3814.<br />

174 Bonvallet, P. A.; Todd, E. M.; Kim, Y. S.;<br />

McMahon, R. J. J. Org. Chem. 2002, 67,<br />

9031–9042.<br />

175 Wentrup, C.; Wilczek, K. Helv. Chim.<br />

Acta 1970, 53, 1459–1463.<br />

176 West, P. R.; Chapman, O. L.;<br />

LeRoux, J. P. J. Am. Chem. Soc. 1982,<br />

104, 1779–1782.<br />

177 Wong, M. W.; Wentrup, C. J. Org. Chem.<br />

1996, 61, 7022–7029.<br />

178 Schreiner, P. R.; Karney, W. L.;<br />

v. Schleyer, P. R.; Borden, W. T.;<br />

Hamilton, T. P.; Schaefer, III, H. F.<br />

J. Org. Chem. 1996, 61, 7030–7039.<br />

179 Matzinger, S.; Bally, T.; Patterson, E. V.;<br />

McMahon, R. J. J. Am. Chem. Soc. 1996,<br />

118, 1535–1542.<br />

180 Matzinger, S.; Bally, T. J. Phys. Chem. A<br />

2000, 104, 3544–3552.<br />

181 Geise, C. M.; Hadad, C. M. J. Org. Chem.<br />

2002, 67, 2532–2540.<br />

182 Armstrong, B. M.; Zheng, F.;<br />

Shevlin, P. B. J. Am. Chem. Soc. 1998,<br />

120, 6007–6011.<br />

183 Sevin, F.; Sokmen, I.; Duz, B.;<br />

Shevlin, P. B. Tetrahedron Lett. 2003, 44,<br />

3405–3407.<br />

184 Maverick, E.; Cram, D. J. Comprehensive<br />

Supramolecular Chemistry 1996, 2,<br />

367–418.<br />

185 Warmuth, R.; Marvel, M. A. Angew.<br />

Chem. Int. Ed. Engl. 2000, 39,<br />

1117–1119.<br />

186 Warmuth, R. Eur. J. Org. Chem. 2001,<br />

423–437.<br />

187 Kerdelhue, J.-L.; Langenwalter, K. J.;<br />

Warmuth, R. J. Am. Chem. Soc. 2003,<br />

125, 973–986.<br />

188 Kuhn, A.; Vosswinkel, M.; Wentrup, C.<br />

J. Org. Chem. 2002, 67, 9023–9030.<br />

145


146 3 Distorted Alkenes<br />

189 Kvaskoff, D.; Bednarek, P.; George, L.;<br />

Pankajakshan, S.; Wentrup, C. J. Org.<br />

Chem. 2005, 70, 9666–9666.<br />

190 Pietruszka, J.; Koenig, W. A.;<br />

Maelger, H.; Kopf, J. Chem. Ber. 1993,<br />

126, 159–166.<br />

191 Price, J. D.; Johnson, R. P. Tetrahedron<br />

Lett. 1986, 27, 4679–4682.<br />

192 Verma, S. K.; Fleischer, E. B.;<br />

Moore, H. W. J. Org. Chem. 2000, 65,<br />

8564–8573.<br />

193 Balci, M.; Harmandar, M. Tetrahedron<br />

Lett. 1984, 25, 237–240.<br />

194 Ozen, R.; Balci, M. Tetrahedron 2002, 58,<br />

3079–3083.<br />

195 Azizoglu, A.; Oezen, R.; Hoekelek, T.;<br />

Balci, M. J. Org. Chem. 2004, 69,<br />

1202–1206.<br />

196 Ogawa, K.; Okazaki, T.; Kinoshita, T.<br />

J. Org. Chem. 2003, 68, 1579–1581.<br />

197 Petrich, S. A.; Pang, Y.; Young, V. G. Jr.;<br />

Barton, T. J. J. Am. Chem. Soc. 1993, 115,<br />

1591–1593.<br />

198 Mitchell, G. H.; Sondheimer, F.<br />

J. Am. Chem. Soc. 1969, 91, 7520–7521.<br />

199 Wang, K. K.; Liu, B.; Petersen, J. L.<br />

J. Am. Chem. Soc. 1996, 118, 6860–6867.<br />

200 Dehmlow, E. V.; Stiehm, T. Tetrahedron<br />

Lett. 1990, 31, 1841–1844.<br />

201 Lin, J.; Pang, Y.; Young, V. G., Jr.;<br />

Barton, T. J. J. Am. Chem. Soc. 1993, 115,<br />

3794–3795.<br />

202 Shimizu, T.; Kamigata, N.; Ikuta, S.<br />

Theochem 1996, 369, 127–135.<br />

203 Myers, A. G.; Cohen, S. B.; Kwon, B.-M.<br />

J. Am. Chem. Soc. 1994, 116, 1670–1682.<br />

204 Kar, M.; Basak, A. Chem. Rev. 2007, 107,<br />

2861–2890.<br />

205 Myers, A. G.; Finney, N. S. J. Am. Chem.<br />

Soc. 1992, 114, 10986–10987.<br />

206 v. Schleyer, P. R.; Jiao, H.;<br />

Sulzbach, H. M.; Schaefer, III, H. F.<br />

J. Am. Chem. Soc. 1996, 118,<br />

2093–2094.<br />

207 Mabry, J.; Johnson, R. P. J. Am. Chem.<br />

Soc. 2002, 124, 6497–6501.<br />

208 Yavari, I.; Kabiri-Fard, H.; Moradi, S.;<br />

Adib, B.; Matlobi, M. Monatsh. Chem.<br />

2003, 134, 23–35.<br />

209 Hickey, E. R.; Paquette, L. A.<br />

J. Am. Chem. Soc. 1995, 117, 163–176.<br />

210 Lu, X.; Zhu, M.; Wang, W.; Zhang, Q.<br />

J. Phys. Chem. B 2004, 108, 4478–4484.<br />

211 Zoch, H. G.; Szeimies, G.; Roemer, R.;<br />

Germain, G.; Declercq, J. P. Chem. Ber.<br />

1983, 116, 2285–2310.<br />

212 Hashmi, S.; Polborn, K.; Szeimies, G.<br />

Chem. Ber. 1989, 122, 2399–2401.<br />

213 Hernandez, S.; Kirchhoff, M. M.;<br />

Swartz, S. G., Jr.; Johnson, R. P.<br />

Tetrahedron Lett. 1996, 37, 4907–4910.<br />

214 Angus, R. O., Jr.; Johnson, R. P.<br />

J. Org. Chem. 1983, 48, 273–276.<br />

215 Hsu, D. P.; Davis, W. M.;<br />

Buchwald, S. L. J. Am. Chem. Soc. 1993,<br />

115, 10394–10395.<br />

216 Rosenthal, U.; Burlakov, V. V.; Arndt, P.;<br />

Baumann, W.; Spannenberg, A.<br />

Organometallics 2005, 24, 456–471.<br />

217 Rosenthal, U.; Pellny, P.-M.;<br />

Kirchbauer, F. G.; Burlakov, V. V.<br />

Acc. Chem. Res. 2000, 33, 119–129.


4<br />

<strong>Strained</strong> Aromatic Molecules<br />

4.1<br />

Nonstandard Benzenes<br />

Paul J. Smith and Joel F. Liebman<br />

4.1.1<br />

Introduction and Context<br />

It was a tenet of organic chemistry that aromatic species are planar. A lengthy<br />

section of the now classic book [1] on strained organic molecules was devoted to<br />

species with distorted rings, and indeed, the current volume elsewhere discusses<br />

such compounds. Interpolating these past and present monographs on strained<br />

species, recent topical issues of Chemical Reviews [2–4] document that ‘bent and<br />

battered’ rings [5] are increasingly part of the chemical landscape. Terms such as<br />

‘fullerenes’ and ‘single wall nanotubes’ attest to increasing comfort with manifestly<br />

nonplanar aromatic species.<br />

We limit our attention to cases where we can compare the energetics of these<br />

species with essentially planar species. We prefer experimental data over that<br />

derived from calculations. While it was recently found that many ‘popular theoretical<br />

methods predict benzene and other arenes to be nonplanar’ [6], it was also<br />

noted that more sophisticated approaches show this failure while more common,<br />

cheap and conventional approaches do not. As such, this finding should not<br />

threaten the increasingly strong coalition of experimentalists and theorists, nor<br />

prejudice for planar arenes.<br />

We focus on hydrocarbons, preferably gaseous or in some innocuous solvent<br />

to minimize the effects of intermolecular interactions. While relative reactivity<br />

generally runs inverse to stability, we primarily discuss enthalpies of formation<br />

(�H f) and reaction, and not of activation.<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

147


148 4 <strong>Strained</strong> Aromatic Molecules<br />

4.1.2<br />

Alkylated Aromatics<br />

Starting with derivatives of benzene itself, the three isomeric xylenes have essentially<br />

indistinguishable gas phase �H f [7], o-, 19.1(1.0); m-, 17.3(0.7); p-,<br />

18.1(1.0) kJ mol –1 (4.184 kJ � 1 kcal): methyl–methyl repulsions are small. The<br />

three isomeric t-butyltoluenes show the o-isomer to be less stable by 20 kJ mol –1<br />

than its m- and p-isomers [8]. For the more strained di-t-butylbenzenes, condensed<br />

phase studies show that the o-isomer is some 80(10) kJ mol –1 less stable than the<br />

nearly isoenergetic m- and p-isomers [9]. Encouragingly, this difference is very<br />

much the same as for 1,2,4- and 1,3,5-tri-t-butylbenzene, 75(15) kJ mol –1 in the<br />

condensed phase [9a,10].<br />

No experimental data allow comparison of 1- and 2-t-butylnaphthalene and<br />

determination of the destabilization from a t-butyl group in a peri (1,8) position.<br />

Computational chemistry shows the 2-isomer to be ca. 22 kJ mol –1 more stable<br />

than its 1-counter part [11]. The 2-isomer behaves as unstrained hydrocarbon:<br />

the difference of its �H f [12] and that of naphthalene is ca. 110 kJ mol –1 [12], vs.<br />

106 kJ mol –1 for related benzenes.<br />

The presence of two t-butyl groups on naphthalene allows for the possibility of<br />

vicinal (1,2- or 2,3-) and peri- (1,8-) situated groups. Such peri species interconvert<br />

with their Dewar isomers thermally [13a], photochemically [13b], and via 1-electron<br />

reduction [13c]. The aromatic isomer is energetically and kinetically disfavored<br />

by repulsion of the t-butyl groups. Racemization of seriously nonplanar 1,8-dit-butylnaphthalene<br />

requires ca. 95 kJ mol –1 [14].<br />

9-t-butylanthracene 1 is nonplanar [15], readily interconverting with its 9,10-<br />

Dewar isomer 2 [16]. Equilibration studies show 1- and 9-t-butylanthracene are<br />

less stable than the 2-isomer [17]: the former are absent under reaction conditions.<br />

Likewise, the undetected 1- and 4-t-butyl derivatives of phenanthrene 3 must<br />

be more strained than the 2- or 3-isomers [18]. 4-t-butyl-5-methylphenanthrene<br />

(studied as its 1,8-dimethylated derivative) is nonplanar [19] and has a strain<br />

energy (SE) of 142 kJ mol –1 ; by contrast, 4,5-di-t-butylphenanthrene has a SE of<br />

208 kJ mol –1 . The nonplanar [20] 4,5-dimethyl species has a SE of ca. 50 kJ mol –1 ,<br />

defined as the �H f difference of this species and its 2,7-dimethyl isomer [21].<br />

The strain in 4,5-dimethylphenanthrene is also shown by polarography [22a],<br />

calculation-guided spectroscopy [22b] and mass spectrometry [22c]. Were only<br />

some of the above simpler alkylated benzenes, naphthalenes and anthracenes so<br />

thoroughly studied!<br />

4.1.3<br />

Helicenes<br />

As of [4]helicene (benzo[a]phenanthrene, 4), these species are significantly<br />

nonplanar as shown by X-ray crystallography [23a], NMR [23b], photoelectron<br />

[23c] and UV spectroscopy [23d]. (A detailed presentation of helicenes is given in<br />

Section 4.3.) The racemization free energy of the parent [6]helicene 5 and some


4.1 Nonstandard Benzenes<br />

larger counterparts [24] have been measured. However, to determine their strain<br />

energies requires �H f measurements which are absent except for [4]helicene<br />

itself: this species is ca. 17 kJ mol –1 less stable than its tri phenylene isomer 6.<br />

Thus, we cannot deduce how homologous is this series – when is the �H f difference<br />

of [n+1] helicene and [n]helicene constant? We recognize these helicenes<br />

are catacondensed (ortho-fused), and note for the simpler class of homologous<br />

catacondensed species, the acenes, our thermochemical knowledge tantalizingly<br />

ends at n = 4, naphthacene.<br />

We now discuss substituted derivatives. Methyl derivatives of [6]helicene show<br />

positional effects on racemization rates [25]: unsubstituted ~ 3 ~ 4 ~ 13 ~ 14<br />

< 2 ~ 15 � 1 ~ 16. The bis-indeno derivatives of 4,5-diphenylphenanthrene[26a],<br />

1,12-diphenyl[4]helicene and 1,14-diphenyl[5]helicene [26b] show large twist angles<br />

by crystallography. These species arise from mild base-catalyzed rearrangements<br />

of unstrained PhC�CC 6H 4C�CCH 2-derivatives of benzene, naphthalene and<br />

phenanthrene. It is a testimony to the high energy of the C�C triple bond in the<br />

precursors and the resonance stabilization of newly formed benzene rings in the<br />

products that such twisted polynuclear aromatic hydrocarbons are readily formed.<br />

Indeed, the photochemical isomerization of diphenylacetylene to phenanthrene<br />

[27] is ca. 180 kJ mol –1 exothermic using contemporary �H f [28]. This quantity is<br />

large enough to accommodate considerable strain in highly distorted aromatic<br />

products [29].<br />

4.1.4<br />

[n]Circulenes<br />

Two circulenes with known �H f are corannulene (n = 5, 7) and coronene (n = 6,<br />

8). The significantly nonplanar corannulene and planar coronene have gas phase<br />

�H f [30] of 460.6(6.5) kJ mol –1 and 307.5(9.8) kJ mol –1 , respectively. Lacking 4,5<br />

H,H repulsions we take coronene to be unstrained, neglecting SE as found in<br />

phenanthrene. Naively, the �H f of corannulene would be 5/6 that of coronene, but<br />

is more positive by 204 kJ mol –1 , a situation exacerbated by another 40 kJ mol –1 for<br />

planar corannulene [31]. To put this strain and aromaticity difference into context,<br />

the isomeric naphthalene and azulene show the former to be more stable by almost<br />

140 kJ mol –1 : in both cases the totally benzenoid species is more stable by some<br />

12 kJ mol –1 per carbon. On the other hand, the parent (C 60 ) and C 70 -fullerenes<br />

(for which corannulene may be understood as en route to) have �H f of ca. 42 and<br />

39 kJ mol –1 per carbon [32], while ‘gaseous’ benzenoid graphite is much more<br />

stable, with �H f of 6 kJ mol –1 per carbon [33].<br />

While a ‘convenient synthesis … of [7]circulene’ 9 has been reported [34], its<br />

enthalpy of formation is unknown. Interestingly, bowl-shaped corannulene and<br />

saddle-shaped [7]circulene have comparable planarization energies [35]: how<br />

does that relate to their strain energies? We note one synthetic detail [36] for [7]<br />

circulene: dehydrogenation of 1,16-didehydro-2,15-dimethyl[6]helicene 10 with a<br />

‘preformed’ 7-membered ring core did not result in 9 or its 1,2-dihydro derivative,<br />

but instead, the fluorene derivative 11.<br />

149


150 4 <strong>Strained</strong> Aromatic Molecules<br />

4.1.5<br />

Cyclophanes<br />

Cyclophanes with small bridges, discussed in detail in Section 4.2, represent<br />

classic examples [5] of distorted benzene rings, most typically reminiscent of boat<br />

cyclohexane. Commencing with [n]cyclophanes, the quintessential question is how<br />

small a polymethylene chain can be used to affix two carbons of a benzene ring.<br />

Starting with planar orthocyclophanes, perhaps we are jaded but the [1]-derivative<br />

12 looks normal. More commonly known as cycloproparenes discussed in<br />

Section 4.4, the SE of this species was indirectly determined by reaction calorimetry<br />

[37] to be 284 kJ mol –1 . However, orthocyclophanes do not exhibit nonplanar<br />

distortions and so will not be further discussed.<br />

What about [1]metacyclophane 13 and/or [1]paracyclophane 14? The �H f of the<br />

corresponding ‘ring-opened’ species, �,3- and �,4-didehydrotoluene (e.g. 15a/15b),<br />

are 431(13) kJ mol –1 for both isomers[38]. However, should we preclude the<br />

existence of [1]-paracyclophane, perchance recognized as norbornadiene-1,7-diyl<br />

16 the ‘opened’ [2.2.1]propelladiene wherein aromaticity is sacrificed to accommodate<br />

bonding of the methylene group to ring carbons? We remember the high<br />

kinetic stability of [1.1.1]propellane presented in Chapter 2.1 and diverse derivatives<br />

despite exceptional SE relative to ring-opened mono- and bicyclic products<br />

[39]. Regardless, we know of no evidence for [1]paracyclophane, nor its isomer<br />

[1]metacyclophane, nor of their presumably less strained [2] and [3] counterparts.<br />

How small can we go?<br />

[3]Paracyclophane 17 and its meta isomer remain unknown. [4]Metacyclophane<br />

18 remains but a reaction intermediate [40] while its Dewar isomer 19 is isolable<br />

[41], and [4]orthocyclophane is tetralin. [4]Paracyclophanes return us to precariously<br />

isolable species. Discussed in [42], NMR spectra of [4]paracyclophane and<br />

even the crystal structure of some derivatives can be obtained. These species<br />

show both considerable geometric distortion and significant remnants of the<br />

aromaticity of p-disubstituted benzenes. Interconversion of [4]paracyclophane<br />

and its more stable Dewar isomer has been observed in both directions. Two<br />

unsaturated derivatives are particularly interesting: polyene 20, 1,2,3,4-tetradehydro[4]paracyclophane,<br />

has been observed [43] and shown by reaction chemistry,<br />

spectroscopy and model calculations to be energetically preferred over its �-bond<br />

isomer 21, bicyclo[4.2.2]deca-1,3,5,7,9-pentaene, a formally unsaturated analog of<br />

[2]-1,4-cyclooctatetraenophane. There are seemingly no cyclooctatetraenophanes<br />

with such a short bridge, although longer, saturated derivatives are known, as are<br />

longer 1,3- and 1,5-cyclooctatetraenophanes [44]. A less relevant nonhydrocarbon<br />

example [43] has a (NC) 2C=C-CH=CH-C=C(CN) 2 group para-bridging a benzene<br />

ring. Powerfully demonstrating the role of substituents in kinetically stabilizing<br />

small cyclophanes, we have ‘ almost a limiting case of experimentally characterizable<br />

strained paracyclophanes’ [43].<br />

What about [5]cyclophanes? We don’t know �H f and/or reaction enthalpy for [5]<br />

paracyclophane and its Dewar isomer. [5]Metacyclophane is an isolable species [45],<br />

but attempts to determine its �H f were thwarted due to formation of a hyperstable


4.1 Nonstandard Benzenes<br />

olefin during calorimetric hydrogenation [46]. The recently developed alternate<br />

approach to aromaticity, the so-called ‘isomerization method’ [47], contrasts a<br />

methylated putative aromatic (e.g., 22) with its ‘isotoluene’ analog (e.g., 23). The<br />

�H f difference of these isomers (the ISE – isomerization stabilization energy)<br />

from quantum chemical calculations [47] is 139 kJ mol –1 , reduced to 82 kJ mol –1<br />

for [5]paracyclophane, suggesting that most of the aromaticity remains for this<br />

species. We welcome corresponding analysis of other cyclophanes.<br />

For [6]paracyclophane and its derivatives, Dewar isomers were experimentally<br />

shown [48] to be less stable than the benzene form by some 130 kJ mol –1 – however<br />

distorted the benzene ring, aromaticity prevails over strain. [6]metacyclophane is<br />

isolable, but, as with its smaller [5] counterpart, �H f determination by measurement<br />

[46] of its enthalpy of hydrogenation was thwarted.<br />

We now turn to cyclophanes containing ring systems other than benzene.<br />

Recalling earlier discussion of species containing cyclooctatetraene, we jump to<br />

those with a 9,10-bridged central anthracene ring, with –CH 2S(CH 2) n-4SCH 2–<br />

(n = 8, 9, 10, 12). As shown in basic solution by NMR, the methylenedihydroanthracene<br />

isomer with the tautomerized =CHS(CH 2 ) n-4 SCH 2 - bridge is thermodynamically<br />

preferred [49]. Similar isomerism is seen for undistorted ‘acyclic’ anthracene<br />

with two –CH 2 SC 3 H 7 groups, unlike methylanthracene [50]. The all hydrocarbon<br />

parent and 1,4,5,8-tetrasubstituted [6] (9,10)anthracenophanes [51] undergo acidcatalyzed<br />

rearrangement to the corresponding methylenedihydroanthracene. Yet<br />

aromaticity prevails: the Dewar isomers of these anthracenophanes are less stable<br />

than their benzenoid counterparts.<br />

Returning to [1]bridged cyclophanes, stable 1,5- and 1,6-methano[10]annulene<br />

24, 25 were studied by hydrogen calorimetry [52]. The calculated ISE [47] for the<br />

latter is 106 kJ mol –1 vs. 215 kJ mol –1 for naphthalene. No numbers are available<br />

for the former species – because different SEs accompany bridgehead double<br />

bonds in bicyclo[5.3.1] and [4.4.1]undecenes, we hesitate to use the difference for<br />

�H f of the two methanoannulenes, ca. 80 kJ mol –1 . We are even more hesitant to<br />

suggest a related difference of 1,5-methano[10]annulene and its unknown tauto mer<br />

bicyclo[5.3.1]undeca-1,3,5(11),6,9-pentaene 26, a [5]metacyclophane derivative.<br />

This brings us to [m,n]cyclophanes, which we limit to meta 27, metapara 28 or<br />

para 29. As reviewed in [42], [1,1]paracyclophanes are characterized in terms of<br />

structure and activation parameters for isomerization with their Dewar counterparts.<br />

We also note the authors of [45] are optimistic about the eventual formation<br />

and characterization of [1,1]metacyclophane. [1,1]metaparacyclophanes are<br />

likewise unknown – photolysis of diketo derivatives of [2,2]metaparacyclophane do<br />

not produce the desired species by double elimination of CO. Rather, a different<br />

fragmentation process [53] results in m-benzyne, another relatively exotic species<br />

[54]. The relative energies of [1,1]meta, metapara and paracyclophane remain<br />

unknown.<br />

For the larger considerably less strained [2,2]cyclophanes (cf. ref. [5]), �H f of<br />

these species, studied as gases, increase in the same order: meta, 170.5(6.5);<br />

metapara, 218.4(1.6), para, 244.1(2.6) kJ mol –1 . For the corresponding dienes, the<br />

enthalpies of formation are ca. 407 kJ mol –1 [32], unknown, and 493.0(5.0) [55].<br />

151


152 4 <strong>Strained</strong> Aromatic Molecules<br />

We note that the presence of the two double bonds affects changes in �H f : meta,<br />

237 and para, 249 kJ mol –1 . By contrast, twice the difference of orbitally conjugated<br />

[56] (E)-stilbene and 1,2-diphenylethane is 186.4(3.6) kJ mol –1 , from which<br />

it may be deduced that the two double bonds increase the SE of [2,2]meta- and<br />

para cyclophane. Relative reactivity, rather than calorimetry, has been used to derive<br />

the same conclusion for [2,2]metaparacyclophane [57].


4.2 Distorted Cyclophanes<br />

We close this discussion of cyclophanes and this chapter with the conclusion<br />

that [6,6]paracyclophane has negative SE [58] – while this species is therefore<br />

outside the scope of the current chapter because of its lack of strain, this observation<br />

nonetheless intrigues [59].<br />

4.2<br />

Distorted Cyclophanes<br />

Henning Hopf<br />

4.2.1<br />

Introduction<br />

Since Kekulé proposed the benzene ring structure, the image of this hydrocarbon<br />

as a rigid, flat, hexagonal molecular object has burned itself into the minds of<br />

chemists. The ideal of the rigid hexagonal ‘tile’ has not only been reinforced by<br />

the structures of the polycondensed aromatic systems but also by the Hückel<br />

theory of aromaticity which proposes that the ‘aromatic character’ of benzene 30<br />

and its derivatives is most pronounced when the axes of the p-orbitals are parallel<br />

allowing optimal p,p-overlap. Intuitively, one might expect that any disturbance<br />

of this overlap, 31, would reduce the aromaticity. This expectation has been the<br />

main driving force for the attempt to prepare deformed aromatic systems. As<br />

we will see, the planarity of aromatic compounds can be deformed considerably<br />

before there is any disturbance in their chemical and spectroscopic behavior<br />

(Figure 4.1) [60].<br />

In principle, there are many ways to deform a benzene ring and a number of<br />

these distortions are encountered in the different normal modes. To translate this<br />

dynamic behavior into a stable distorted structure the simplest approach uses the<br />

incorporation of a molecular bridge, usually a polymethylene chain, causing the<br />

generation of a cyclophane. For a simple benzene ring as the aromatic moiety to<br />

be spanned, the bridge may be anchored in 1,4- ([n]paracyclophane 32) or in the<br />

Figure 4.1 Bridging the benzene ring.<br />

153


154 4 <strong>Strained</strong> Aromatic Molecules<br />

1,3-position ([n]metacyclophane 33). The benzene ring in ortho-arrangement, 34<br />

is usually planar. Employing the ‘phane concept’ numerous bridged aromatic<br />

compounds has been prepared during the last decades [61], although the first<br />

cyclophanes, Lüttringhaus’ ansa compounds [62], were already prepared in the<br />

early 1940s. A particular attraction is exhibited by cyclophanes possessing a second<br />

aromatic ‘deck’ as in 35, since in these molecules the distance and orientation of<br />

the aromatic ‘layers’ may be adjusted by the lengths of the bridges (m, n in 35).<br />

Because of space limitations this chapter will concentrate on representative<br />

examples derived from 32, 33, and 35, and the main questions to be answered<br />

will be the preparation of these compounds and the relationship between bridge<br />

length, ring distortion, and ‘aromatic character’, the latter being judged mainly be<br />

the NMR properties of the cyclophanes (see Section 4.2.5). Available calorimetry<br />

data for cyclophanes are discussed in Section 4.1.5.<br />

4.2.2<br />

The [n]Cyclophanes<br />

4.2.2.1 [n]Paracyclophanes<br />

Within the class of cyclophanes containing a single aromatic moiety only, this<br />

group has received the highest attention [61, 63]. Calculations by different methods<br />

support the expected trend; the shorter the bridge in 32 the more distorted is<br />

its benzene ring. Table 4.1 summarizes the relationship between bridge length,<br />

distortion angle �, and heat of formation of the respective [n]paracyclophane<br />

according to AM1 calculations [64].<br />

Table 4.1 Relationship between bridge length, distortion angle �, and heat of formation of<br />

respective [n]paracyclophane according to AM1 calculations.<br />

n � [°] �H f (kcal mol –1 ]<br />

14 0.5 –62.3<br />

13 1.3 –54.9<br />

12 1.9 –48.9<br />

11 2.2 –40.5<br />

10 3.4 –29.5<br />

9 9.6 –18.9<br />

8 13.4 –8.7<br />

7 16.4 5.1<br />

6 22.3 26.2<br />

5 28.6 53.8<br />

4 35.6 87.8<br />

3 79.4 117.8


(a) � (23%); (b) Zn/Hg, HOAc, HCl (54%); (c) Br 2 , MeOH, –30 °C, H 2 O (47%);<br />

(d) LiAlH 4 /AlCl 3 (65%)<br />

Scheme 4.1<br />

4.2 Distorted Cyclophanes<br />

To prepare [n]paracyclophanes 32 with n down to about 9 should pose no<br />

particular synthetic problems. The typical methods used to make most of the<br />

smaller [n]paracyclophanes (n � 8) can all be demonstrated by the preparation of<br />

[8]paracyclophane 39. By employing a mixture of the two Hofmann bases 36 and<br />

40 and subjecting it to a 1,6-elimination (Scheme 4.1) [65] the p-xylylene 37 and<br />

its furanoid analog 41 are generated which cycloadd to produce the furanophane<br />

38, that subsequently can be reduced directly to 39 or via the diketone 42.<br />

The gain in aromatic resonance is also made use of when the bis-spiro diene<br />

43 is pyrolyzed in the presence of trapping agents such as 1,3-butadiene (46,<br />

Scheme 4.2). The reaction begins with pyrolysis of a three-membered ring to<br />

provide the diradical 44 � 45 first. Trapping then provides the monoene 47 that is<br />

hydrogenated to the saturated hydrocarbon 39 [66]. An important general approach<br />

(Scheme 4.2) to the smaller [n]paracyclophanes employs spirocyclic ketones such<br />

as 48, which is first converted to a cross-conjugated triene 49 followed by flashvacuum<br />

pyrolysis (FVP) to provide the diradical 50 that recloses to 39 with the<br />

remarkable yield of 70% [63].<br />

In another approach strain-free precursors, like 51, are first built-up and subsequently<br />

subjected to ring-contraction processes as shown in Scheme 4.3 for<br />

4-carboxy[8]paracyclophane 54 [67].<br />

51 was prepared by an acyloin condensation/oxidation protocol from the appropriate<br />

open-chain starting materials, and then ring-contracted by a Wolff rearrangement<br />

of 52 via ketene 53 to the acid 54. After conversion of 54 into 4-keto[8]<br />

paracyclophane the sequence could be repeated, and 3-carboxy[7]paracyclophane<br />

be obtained [68]. The approach beginning with �,��-dichloro-p-xylene 55 and<br />

1,6-hexandithiol 56 illustrates the principle of the uncoupling of the two critical<br />

steps in [n]paracyclophane chemistry: the construction of an intermediate that<br />

155


156 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.2<br />

Scheme 4.3<br />

Scheme 4.4


Scheme 4.5<br />

4.2 Distorted Cyclophanes<br />

already has the correct topology of the target molecule (here [10]paracyclo phane 57)<br />

and the built-up of strain in the last step [69]. Ring contraction is accomplished<br />

by sulfone pyrolysis [70], using the bis sulfone 58, obtained readily from the bis<br />

thioether 57. In the acid 54 the experimentally determined deformation angle �<br />

(see 32) amounts to 9.1° [71], whereas in the lower homolog [68], it has increased<br />

to 16.5°, in excellent agreement with the calculated value (see above).<br />

The parent system of the latter was obtained using the approach via reactive<br />

carbene 60 (Scheme 4.4, Jones route) [72].<br />

Here the spirocyclohexadienone tosylhydrazone 59 was metalated with n-butyl<br />

lithium and the resulting salt thermally decomposed under FVP conditions to 61<br />

[73]. A conceptually interesting route to [n]paracyclophanes involves the conversion<br />

of a heterophane into a benzophane (Tochtermann route) [74]. As shown in<br />

Scheme 4.5 the bridged furan 62 undergoes the expected Diels-Alder addition<br />

with dimethyl acetylene dicarboxylate (DMDA) to the adduct 63, that is converted<br />

via the oxaquadricyclane 64 into the bridged oxepin 65 by a sequence of pericyclic<br />

reactions. When the latter was brominated the dibromide 66 is obtained, which<br />

on debromination/deoxygenation yielded the diester 67. Various derivatives of 67<br />

show deformation angles � around 15° [74].<br />

Moving down to the parent system [6]paracyclophane 71, both the Jones route<br />

(Scheme 4.4) [75] as well as the Tochtermann route to the dimethyl diester were<br />

successful. Apparently, both methods reached their limits here, since they failed in<br />

the preparation of [5]paracyclophane or its derivatives. Several derivatives of these<br />

cyclophanes could be prepared, and their solid state structures be determined: for<br />

[6]paracyclophane-8-carboxylic acid the bending angles � (see 3) amount to 20.3<br />

and 21.1°, respectively [76], and for the dimethyl 8,9-dicarboxylate this angle is<br />

19.5° [77]. The 3,3�-disubstutited bicyclopropenyl derivative 68 can be rearranged<br />

into their Dewar benzene isomers, followed by valence isomerization as shown<br />

in Scheme 4.6. The metal-catalyzed step afforded a mixture of the two isomeric<br />

Dewar benzenes 69 and 70, which after separation were thermally opened to<br />

their aromatic isomers 71 and 72, respectively. On irradiation 71 recloses to 69,<br />

indicating the high strain of this cyclophane [78].<br />

157


158 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.6<br />

A longer, but superior route was developed to prepare [6]paracyclophane<br />

(Scheme 4.7) [79]. The bicyclic ketone 73 was first converted into the propellane<br />

74 by addition of acetylene to the double bond of the substrate 73. Wolff rearrangement<br />

of the �-diazoketone 75, prepared from 74 by azo group transfer,<br />

next furnished the ester 78. The Dewar benzene intermediate of 78 was opened<br />

thermally to the [6]paracyclophane ester 77, which on hydrolysis provided 76 in<br />

10% yield. Alternatively, the parent system was prepared from 78 by first saponifying<br />

it and subsequently oxidizing the formed acid. The 1 H NMR spectrum of<br />

71 does not differ significantly from that of an undistorted aromatic compound<br />

[75, 78].<br />

[5]Paracyclophane 80 could be only generated by irradiation of the 1,4-pentamethylene<br />

Dewar benzene analogue of 69 in THF-d8 at –80 °C [80]. Hydrocarbon<br />

80 decomposes rapidly above 0 °C, but it was identified unequivocally by its UV<br />

Scheme 4.7


4.2 Distorted Cyclophanes<br />

and 1 H NMR spectra [81]. Its aromatic protons absorb as an AA�XX�-multiplet at<br />

7.38 and 7.44 ppm, respectively, whereas the benzylic protons are registered at<br />

2.77 ppm, both shifted downfield from the respective protons of p-xylene. Esters<br />

such as 81 [82] and 82 [83], obtained by photoisomerization of their respective<br />

Dewar benzene isomers have half-lives of several hours at room temperature.<br />

The smallest [n]paracyclophane, identified by trapping experiments, is the<br />

tetra methylene bridged system 84 (Scheme 4.8). Irradiation of the butano-bridged<br />

substrate 83 at –20 °C in THF in the presence of trifluoroacetic acid, provided the<br />

addition products 86 and 87 [84]. If the experiment is performed in methanol the<br />

ether 88 is produced. It is rationalized that the precursor 83 indeed photo isomerizes<br />

to 84, which is protonated to 85, that subsequently yields the trapping products<br />

86–88. In a second trapping experiment the Dewar benzene ester 89 on photolysis<br />

in the presence of ethanol yielded the regioisomeric 1,4-trapping products 91 and<br />

92, for which the [4]paracyclophane ester 90 is the most reasonable precursor.<br />

When 83 was irradiated in a matrix at 77 K, the UV spectrum of 84 could be<br />

recorded [85]. Although computational studies concerning [3]paracyclophane have<br />

been published [86], no preparative efforts have been undertaken.<br />

Scheme 4.8<br />

159


160 4 <strong>Strained</strong> Aromatic Molecules<br />

4.2.2.2 [n]Metacyclophanes<br />

A significant ring strain arises in [n]metacyclophanes only when the bridge<br />

length is seven or smaller [63, 87]. Three basic approaches have been developed<br />

to synthesize [n]metacyclophanes as summarized in Scheme 4.9. The first route<br />

begins with cyclic precursors such as dodecanone 93, i.e. substrates that contain<br />

the future molecular bridge from the very beginning. In case of 93 this is accomplished<br />

via intermediates such as 94 [88]. In the second route the aromatic unit is<br />

present from the start, e.g. as the dithiol 97, and the rest of the synthesis involves<br />

bridging it; a typical example is given by the sulfone pyrolysis of 98 yielding [9]<br />

metacyclophane 99 [89]. In the last approach the metacyclophane is obtained from<br />

its para-isomer by acid-catalyzed isomerisation. All three approaches have been<br />

employed to synthesize e.g. [7]metacyclophane [90–92]. For the synthesis of lower<br />

homologues such as [5]metacyclophane, methods similar to those employed in<br />

the para-series have been developed [93]. Again, the [4]cyclophane represents the<br />

smallest hydrocarbon (generated as an intermediate) in this series also [94].<br />

The parent [n]metacyclophanes have low melting points, so only a few derivatives<br />

have been subjected to X-ray structural analysis. 8,11-dichloro[5]metacyclophane<br />

possesses a pronounced boat-structure, with the functionalised carbon atoms<br />

bent out of the plane by 12° (unbridged) and 27° (bridged side). Its 1 H-NMR<br />

spectrum that displays a ring current equal to that of analogous planar benzene<br />

derivatives [95].<br />

Scheme 4.9


4.2.3<br />

The [m.n]Paracyclophanes<br />

4.2 Distorted Cyclophanes<br />

Among the [m.n]cyclophanes, those in which the two alkano bridges are anchored<br />

in the para-positions of both benzene ‘decks’, have received greatest attention.<br />

We will begin with [0.0]paracyclophane 103 (Scheme 4.10) and end with the [4.4]<br />

homolog for this discussion.<br />

Scheme 4.10<br />

The extremely strained 103 should better be regarded as a valence isomer of<br />

tricyclo[4.2.2.22,5]dodeca-1,3,5,7,9,11-hexaene 104 of which recently the dicyano<br />

derivative 76 has been generated as a reaction intermediate [96]. If one of the<br />

bridges is formally ‘opened’ the para-bridged biphenyl derivatives 106 result [97].<br />

The [1.n]paracyclophane series begins with [1.1]paracyclophane 114 (Scheme 4.11);<br />

its synthesis employs many of the steps that have been discussed above for the<br />

preparation of the smaller [n]paracyclophanes [98].<br />

Tricyclic diketone 107 on photoaddition of acetylene afforded bis-adduct 108,<br />

which was converted to the bis-diazo diketone 109. Wolff rearrangement in<br />

methanol then furnished 110, in which the double Dewar benzene structure is<br />

clearly discernible already. The missing double bonds were introduced by double<br />

Hofmann elimination of diamine 111, obtained from 110 by routine methods.<br />

Irradiation of 112 at 77 K in an EPA matrix then furnished 114 via ‘half-opened’<br />

113. That 114 had been generated was inferred from UV/Vis and NMR spectra.<br />

However, on warming to room temperature, 114 was destroyed within 4 h [98].<br />

On extended irradiation 114 was converted into the interesting hydrocarbon 115.<br />

Calculations revealed that its benzene rings are approximately as distorted as the<br />

one in [5]paracyclophane 80, the bending angle by which the bridgehead carbon<br />

atoms are ‘removed’ from the plane amounting to 23°, compared with the 12.6°<br />

of [2.2]paracyclophane (exp., see below) and 23.5° (calc.) for 80. The intraannular<br />

distance between the atoms of the benzene rings is ca. 236–240 pm, i.e. roughly<br />

100 pm shorter than the inter-layer distance in graphite. These calculations indicate<br />

161


162 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.11<br />

the delocalisation of the electrons despite severe bending of the aromatic moieties.<br />

The calculated strain energy of 114 is in the range of 93–128 kcal/mol depending<br />

on the level of theory.<br />

Although the parent hydrocarbon [1.2]paracyclophane is unknown and presumably<br />

would not differ very much in strain and benzene ring deformation<br />

from the [1.1]compound 114 [99], the [1.n]phanes 116 [100] and the analogous<br />

ketones 117 [101] have been prepared by FVP of the corresponding sulfones. In<br />

the pentamethylene bridged parent hydrocarbon 118 [102] the bridgehead atoms<br />

are not only displaced out of the plane by about 10°, but with only 235 pm the<br />

distance between the substituted carbon atoms at the ‘methylene side’, distinctly<br />

shorter than the analogous distance in [2.2]paracyclophane (278 pm, see below).<br />

This observed intraannular distance agrees remarkably well with the calculated<br />

one for [1.1]paracyclophane 114 (see above) and for different highly functionalized<br />

derivatives of 114 [103].<br />

The by far most thoroughly studied phanes are [2.2]paracyclophane 119<br />

(Scheme 4.12) and its derivatives. Numerous routes to prepare these compounds<br />

have been described, and their chemical and structural properties have been<br />

reviewed [60, 61, 63], that an additional summary is unnecessary here. The basic<br />

strategies to prepare [2.2]paracylophanes are summarized in highly condensed<br />

retrosynthetic form in Scheme 4.12.


Scheme 4.12<br />

4.2 Distorted Cyclophanes<br />

Cleavage of one of the ethano bridges in 119 results in the formation of the<br />

diradical 120 for whose generation either other cyclophanes such as 124 and 125<br />

or open-chain precursors such as the dibromide 126 may be used. All that remains<br />

to be done in the last steps is a ring-contraction, for which the photochemical<br />

or high temperature extrusion of small molecules (e.q. CO, CO 2 or SO 2 ) are the<br />

methods of choice. In the case of 126 only Wurtz coupling is required; however,<br />

since in this process intermolecular coupling competes with intramolecular C-C<br />

bond formation, yields are low [104]. If the other bridge in 120 is also broken, the<br />

diradical 121 results, which is nothing else than p-xylylene or p-quinodimethane<br />

37. For this reactive tetraene numerous derivatives of p-xylene 127 can serve as a<br />

precursor, beginning with p-xylene itself (127, X = Y = H) [105], and ending e.g. with<br />

the Hofmann base derived from �-bromo-p-xylene 127 (X = H, Y = N(CH 3 ) 3 OH)<br />

[106]. Continuing our retro-synthetic deconstruction, we next decompose 37 in<br />

retro-Diels–Alder fashion and arrive at 1,2,4,5-hexatetraene 122 as the diene and<br />

acetylene 123 as the dienophile component. Bisallene 122 is prepared by dimerization<br />

of propargyl bromide 128 and since simple alkynes are not reactive enough<br />

to undergo Diels-Alder additions with it, 123 has to be replaced by an activated<br />

acetylene 129 with various electron-electron-withdrawing substituents [107].<br />

The structure of 119 has been determined several times [108, 109]. Only some<br />

important structural data resulting from the latest study (recorded at 19 K) are<br />

given in Figure 4.2. As expected, extension of the alkano bridges leads to a flattening<br />

of the benzene rings and an increase in the distance between them. In [3.3]<br />

paracyclophane 130 the above deformation angle has been halved to 6.4° (intraannular<br />

distance at the bridgeheads 314 pm and distance between the planes passing<br />

through the remaining aromatic carbon atoms: 320 pm) [110] and [4.4]paracyclophane<br />

131 has flat aromatic rings [111]. In the hybrid case of [2.4]paracyclophane<br />

132 the two benzene rings are inclined at an angle of 21°, the distortion angle<br />

at the bridgeheads is slightly above 5° (at all positions), and the two non-bonded<br />

distances at the bridgeheads amount to 279 pm (ethano side) and 386 pm (butano<br />

163


164 4 <strong>Strained</strong> Aromatic Molecules<br />

Figure 4.2 Some structural parameters of [2.2]paracyclophane.<br />

side) [112]. A fascinating known completely bridged cyclophane called superphane<br />

is mentioned in Chapter 9, its hypothetical hexahydrogenated analogue with a<br />

planar cyclohexane ring is briefly discussed in Section 2.4 while known but insufficiently<br />

studied layered cyclophanes are introduced in Section 1.4.<br />

4.2.4<br />

Distorted Aromatic Rings and ‘Aromatic Character’<br />

The planarity of classical aromatic systems (usually benzene rings) may be<br />

distorted effectively by the introduction of molecular bridges. Using this device<br />

highly bent benzene rings with deformation angles � (see structure 32) up to<br />

25° may be generated readily. As judged from their X-ray structures and NMRproperties<br />

(see Section 5) these strongly bent benzene ring systems are delocalized<br />

to the same extent as benzene and its simple derivatives. On the other hand,<br />

many cyclophanes with short bridges display a reactivity unprecedented for<br />

normal aromatics, for example they undergo Diels-Alder additions readily [113]<br />

or can readily be hydrogenated [114]. For example, it has been calculated that [5]<br />

metacyclophane has a strain energy of 43 kcal/mol [115]; for [2.2]paracyclophane<br />

119 a strain energy of 30.1 kcal/mol has been determined [116]. Strain release<br />

would be effective not only for the products of the reaction, in which the formal<br />

sp 2 hybridization changed to sp 3 , which makes bridge formation easier. As an additional<br />

factor the deformation of the �-electron cloud of the benzene ring could<br />

play a role, since this raises the HOMO and lowers the LUMO energy. Localization<br />

of �-electrons into a cyclohexatriene moiety would increase the violation of<br />

Bredt’s rule even further, and is hence avoided. And it may finally play a role that<br />

it appears to be the �-frame that dominates the delocalization in benzene rather<br />

than the �-system [117].


4.2.5<br />

NMR Characteristics of Cyclophanes<br />

4.2 Distorted Cyclophanes<br />

1 H NMR spectroscopy can furnish detailed information with respect to structures,<br />

conformations and conformational dynamics of this class of compounds<br />

and a number of NMR reviews on cyclophanes is available [118–122]. The most<br />

important feature in the 1 H NMR spectra are the unusual chemical shifts of<br />

protons lying above the plane of the aromatic component of the cyclophane.<br />

Shortly after Pople suggested [123] that protons connected to aromatic rings<br />

are deshielded by a ring current induced by the external magnetic field, Waugh<br />

and Fessenden [124] reported the increased shielding, by 0.7 ppm, of the central<br />

methylene protons in [10]paracyclophane 32 relative to those in cyclohexane. These<br />

protons are held in a position over the center of the ring. Shortening the methylene<br />

bridge decreases the average distance between these protons and the region of<br />

maximum shielding. The most highly shielded proton in [5]paracyclophane 80<br />

has � = 0.01 ppm [81] while in [6]paracyclophane 71 � = –0.62 ppm [125].<br />

In [2.2]paracyclophane 119 the aromatic rings cause mutual shielding of their<br />

aryl protons by –0.62 ppm [�(119) = 6.48 ppm [126], �(1,4-diethylbenzene) =<br />

7.10 ppm)]. This is only a moderate effect because the rings are eclipsed. The<br />

effect in [2.2.2](1,3,5)cyclophane 133 is somewhat larger because the triple<br />

bridging decreases the average distance between the rings [�(133) = 5.71 ppm<br />

[126], �(1,3,5-triethylbenzene) = 6.86 ppm)]. In anti-[2.2]metacyclophane 134,<br />

however, the protons experience strong shielding [�(134) = –2.75, 4.27 ppm [126],<br />

�(1,3-diethyl benzene) = 7.02 ppm)] as they are situated much closer to the region of<br />

maximum shielding of the opposite aromatic ring. Finally, Pascal’s in-cyclophanes<br />

show extreme shielding effects, for example � = –4.03 ppm for H i in 135 [127],<br />

because the geometry of the molecular cage forces the internal proton to point<br />

toward the center of the aromatic ring at a very small distance (172 pm, calc.).<br />

165


166 4 <strong>Strained</strong> Aromatic Molecules<br />

4.3<br />

Helicenes<br />

Ivo Starý and Irena G. Stará<br />

4.3.1<br />

Introduction<br />

Helicenes as unique, inherently chiral three-dimensional aromatics have been<br />

attracting continuous attention for decades. They are usually chemically stable<br />

and soluble in common organic solvents, which makes a difference from many<br />

other large �-conjugated systems. The chemistry of helicenes has been reviewed<br />

several times [128–135]. However, since an enormous amount of progress in their<br />

synthesis and use took place in the early 1990s, the recent period has not been<br />

covered by a comprehensive review. Thus, this chapter is mostly focused on the<br />

important achievements in helicene chemistry within the last two decades.<br />

Helicenes are polycyclic aromatic systems consisting of all-ortho-fused aromatic<br />

rings. The prefix before the name expresses the number of fused cycles as exemplified<br />

by hexahelicene or simply [6]helicene. Provided all these rings are benzenes,<br />

such compounds are called carbohelicenes 136. If one (or more) benzene unit<br />

is formally replaced by a heterocycle, such a skeletal modification leads to heterohelicenes<br />

137. Finally, helicene-like compounds represent the third family of<br />

helicenes, which can differ significantly from fully aromatic parent helicenes but<br />

having a similar molecular shape 138. As helicenes can exist in two enantiomeric<br />

forms regardless of their configurational stability, the handedness of the helix<br />

is specified by adding the (M) (minus) or (P) (plus) prefix (Scheme 4.13). This<br />

chapter deals only with carbohelicenes, which are simply called helicenes within<br />

the following text.<br />

4.3.2<br />

Synthesis of Helicenes<br />

[6]Helicene was the first helicene representative prepared intentionally by Newman<br />

and Lednicer in 1956 [136], who also pioneered resolution of the racemate<br />

employing a �-� donor-acceptor interaction between helicene and optically pure (–)-<br />

or (+)-2-(2,4,5,7-tetranitro-9-fluorenylideneamino-oxy)propionic acid (TAPA).


Scheme 4.13<br />

4.3 Helicenes<br />

The real breakthrough in the synthesis of helicenes came in the late sixties when<br />

the Martin group introduced photodehydrocyclisation of stilbene-type precursors<br />

[137–139] as the first general method for their preparation [140]. It was based on<br />

UV-light induced cis/trans isomerisation of 1,2-diarylethylenes followed by conrotatory<br />

electrocyclisation of the cis isomer to generate a primary dihydroaromatic<br />

product with trans configuration (Scheme 4.14).<br />

Scheme 4.14<br />

Then, in the presence of air and a catalytic amount of iodine, it was immediately<br />

converted to a fully aromatic system [135]. However, photocyclisation leading to<br />

an all-ortho-fused system may often be disfavoured for steric or other reasons and,<br />

therefore, the process may exhibit low regioselectivity. To overcome this problem,<br />

167


168 4 <strong>Strained</strong> Aromatic Molecules<br />

Liu and Katz developed synthetic methodology utilising a bromine substituent<br />

on a benzene ring, which directs photocyclisation away from its ortho position<br />

[141]. A remarkable step forward was made in 1991 when Katz and colleagues<br />

published an updated version of photodehydrocyclisation of stilbene-type precursors<br />

(Scheme 4.15) [142]. They found that iodine in a stoichiometric amount<br />

is superior oxidant and propylene oxide is an effective scavenger of hydrogen<br />

iodide. Since then, this improved methodology has become a standard tool for the<br />

synthesis of helicenes. Numerous examples of successful use of photodehydrocyclisation<br />

in its original or innovated version were published to prepare, inter alia,<br />

[5]helicene [141], [6]helicene [143–145], [7]helicene [143], [8]helicene [145–147],<br />

[9]helicene [145, 146, 148], [10]helicene [145, 148], [11]helicene [145, 148, 149],<br />

[12]helicene [145, 149], [13]helicene [145, 148, 150], and [14]helicene [149]. Despite<br />

the synthetic accessibility of stilbene-type precursors and wide applicability and<br />

simplicity of photodehydrocyclisation, this methodology suffers from several<br />

drawbacks. To prevent photodimerisation, the irradiation must be performed<br />

under high dilution conditions which limits any scale-up [151, 152]. Furthermore,<br />

acetyl, dimethylamino and nitro groups that can perturb or depopulate the reactive<br />

excited state should be avoided [152, 153].<br />

Scheme 4.15<br />

Since the early 1990s, various non photochemical approaches have emerged<br />

to circumvent disadvantages of photodehydrocyclisation. The most important<br />

method eliminating the irradiation was published by the Katz group [152]. It was<br />

based on thermal Diels-Alder reaction of aromatic bisvinylethers with p-benzoquinone<br />

in excess to afford helicenes having terminal quinone rings (Scheme 4.16<br />

[161]). It evolved into a robust and versatile method, which has also been employed<br />

Scheme 4.16


4.3 Helicenes<br />

by other authors [154, 155] to prepare derivatives of [5]helicene [152, 156, 157],<br />

[6]helicene [152, 157], and [7]helicene [158–160]. The importance of the Katz group<br />

contribution basically consists of developing the simple methodology, which for<br />

the first time allows the synthesis of a wide series of functionalised helicenes<br />

on a multigramme scale and in providing highly attractive compounds for other<br />

applications [161].<br />

The simplicity of Katz’ Diels-Alder methodology represents, on the other hand,<br />

its drawback. Variations of functionalities on the terminal (hydro)quinone rings<br />

of a helicene skeleton are rather limited. Therefore, other new methods for the<br />

preparation of helicenes have recently emerged on the stage. Stará and colleagues<br />

have developed a new strategy based on intramolecular [2+2+2] cycloisomerisation<br />

of aromatic triynes catalysed by Ni 0 or Co I complexes (Scheme 4.17 [164])<br />

[162–165]. Using this organometallic approach, various helicene derivatives were<br />

synthesised ranging from penta- to heptacyclic systems. This methodology exhibits<br />

considerable potential due to its modularity, chemo- and regioselectivity and allows<br />

for the formation of three new cycles of the helical scaffold all at once.<br />

Scheme 4.17<br />

Axially chiral biaryls can be viewed as truncated helicenes. Indeed, there is a<br />

group of diverse synthetic methods transforming biaryls to helicenes. Bestmann<br />

and Both [166] and Stará and colleagues [167] described the transformation of 1,1�binaphtyl<br />

derived bisphosphinealkylene or dihydroazepinium salt, respectively,<br />

to [5]helicene (Scheme 4.18 [167]). Gingras and Dubois developed syntheses of<br />

[5]- and [7]helicene based on carbenoid coupling, which started in the case of [7]<br />

helicene from 4,4�-biphenanthrene-3,3�-diol (Scheme 4.19) [168]. Alternatively, they<br />

Scheme 4.18<br />

169


170 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.19<br />

applied McMurry coupling to the synthesis of [5]helicene from 1,1�-binaphthalene-<br />

2,2�-dicarbaldehyde [169]. So far the latest contribution to this field was published<br />

by Collins et al., who employed ring-closing olefin metathesis to construct [5]-,<br />

[6]- and [7]helicene from divinyl biaryls (Scheme 4.20) [170]. The synthesis of<br />

helicenes from biaryls has significant potential but such an approach leads to the<br />

construction of only one new benzene ring in the helicene scaffold.<br />

Scheme 4.20<br />

Photocyclodehydrogenation of stilbenes to prepare helicenes has inspired other<br />

authors to develop its nonphotochemical alternative. Harrowven and colleagues<br />

used homolytic aromatic substitution in the case of diiodo cis,cis-distilbenes, which<br />

under treatment with tributyltin hydride and the VAZO radical initiator provided<br />

functionalised [5]- and [7]helicenes (Scheme 4.21) [171].<br />

Scheme 4.21


4.3.3<br />

Nonracemic Helicenes<br />

4.3 Helicenes<br />

Since Newman’s preparation of optically pure [6]helicene [136], different routes<br />

based on the resolution of racemate have been explored to produce enantiomerically<br />

enriched or pure helicenes. The resolution of helicene racemates by HPLC<br />

on a chiral stationary phase is general and can be employed for analytical as well<br />

as preparative purposes [172–174]. There is only one practical method allowing<br />

for resolution of racemic helicenes on a multigramme scale so far. Katz and colleagues<br />

developed a robust procedure using chromatography of diastereomeric<br />

helicene pairs on silica gel (Scheme 4.22) [175]. This is suitable exclusively for<br />

helicen-1-ols, which are converted to (1S)-camphanates.<br />

Scheme 4.22<br />

Asymmetric synthesis of helicenes and their congeners is envisaged to be<br />

the most straightforward and efficient route to single enantiomers. Various<br />

concepts have emerged demonstrating basic principles rather than generally<br />

useful methodologies. Nevertheless, some of them might be highly promising.<br />

Classical photodehydrocyclisation of stilbene-type precursors can be carried out<br />

in an astonishingly stereoselective fashion. This was well demonstrated by the<br />

pioneering works by Vanest and Martin [176] and Katz and colleagues [177] who<br />

used stereocentre(s) external or internal to the helix to control stereoselectivity<br />

of helicene cyclisations. Carreño et al. developed an asymmetric version of the<br />

Diels–Alder approach providing helicene quinones with excellent optical purities<br />

(Scheme 4.23) [155, 178–180]. In addition to that, the last decade has witnessed<br />

other attempts at asymmetric synthesis of helicenes but stereocontrol observed<br />

has been moderate as published by Stará and colleagues (enantioselective Nicatalysed<br />

[2+2+2] cycloisomerisation of aromatic triynes) [162, 164, 174]. In spite<br />

of the above mentioned achievements, practical asymmetric synthesis of helical<br />

aromatics has so far remained a challenging task.<br />

171


172 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.23<br />

4.3.4<br />

Intriguing Helicene Structures<br />

Regarding the highest helicene homologues so far synthesised, the world record<br />

among carbohelicenes belongs to Martin and Baes, who synthesised [14]helicene<br />

(Scheme 4.24) [149]. In order to prepare much longer helicene structures, the Katz<br />

group explored the ways of polymerising bifunctional helicene units. Placing a<br />

salicylaldehyde functionality at each end of optically pure [6]helicene, reaction with<br />

o-phenylenediamine and nickel(II) salt led to the first polymer with an unbroken<br />

network of double bonds that winds in one direction along a helix [181, 182].<br />

The degree of delocalisation was higher than in optically active [7]helicene-based<br />

cobaltocenium oligomers [183, 184].<br />

Structural diversity of helicene molecules is broad, encompassing helical<br />

metallo cenes by Katz and Pesti (the Fe or Co atom spans the ends of a helicene<br />

backbone, Scheme 4.25) [185], helical metal phthalocyanine derivatives by the<br />

Katz group [186], conjugated helical acetylene-bridged cyclophanes by Fox et al.<br />

[187], 2,2�-bis[6]helicyl by Laarhoven and Veldhuis [188], helicenes containing a<br />

cyclophane unit by Nakazaki et al. [189], Martin and colleagues [190], and S-shaped<br />

[191] or 3-shaped [192] double helicenes by the Laarhoven group, to mention but<br />

a few.<br />

Scheme 4.24 Scheme 4.25


4.3.5<br />

Physicochemical Properties and Applications<br />

4.3 Helicenes<br />

As helicenes have robust, inherently chiral scaffolds, it is not surprising that they<br />

were soon applied by the Martin group to asymmetric synthesis as chiral auxiliaries<br />

(in diastereoselective reduction of �-keto esters [193], ene reaction [194] and<br />

atrolactic synthesis [195]) or chiral reagents (in hydroxyamination [196] or epoxidation<br />

[197] of olefins, Scheme 4.26) with remarkable success. After certain delays,<br />

attention has turned to enantioselective catalysis and so pioneering exploitations of<br />

helicene ligands have recently emerged. Highly stimulating results were obtained<br />

in asymmetric hydrogenation (Reetz et al. [198], Nakano and Yamaguchi [199]),<br />

allylic substitution (Reetz and Sostmann [200], Scheme 4.27) and diorganozinc<br />

addition to aldehydes (Katz et al. [156], Soai et al. [201]).<br />

Scheme 4.26<br />

Scheme 4.27<br />

Regarding self-assembly, one of the most astonishing attributes of helicene<br />

assemblies was described by the Katz group [204] (Scheme 4.28) [130]. Properly<br />

substituted nonracemic helicenes, possessing both electron-rich inside and<br />

electron-deficient outside regions, can aggregate spontaneously to create columnar<br />

structures exhibiting enormous optical rotation values and NLO properties.<br />

173


174 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.28<br />

In Langmuir-Blodgett films, an array of parallel columns can be observed<br />

directly by AFM [202]. Moreover, these columns are further organised into long<br />

micrometre-wide lamellar fibres visible under an optical microscope [203]. The<br />

chiroptical properties of such assemblies are so remarkable that CD spectra could<br />

be measured for a monolayer [204]. The Moore group showed that a [6]helicenecontaining<br />

foldamer provided highly solvent dependent CD spectra [205].<br />

There are remarkable examples of helicene use in molecular recognition. The<br />

[7]helicene-based helicopodand was used by Diederich and colleagues in molecular<br />

recognition of dicarboxylic acids with high diastereoselectivity [206]. Nonracemic<br />

helicene diol was successfully used by Reetz and Sostmann reporting on enantioselective<br />

fluorescence quenching by chiral amines [173]. A helicene derivatising<br />

reagent developed by the Katz group can serve as a remote chirality sensor for chiral<br />

alcohols, amines and phenols when detected by NMR measurements [207].<br />

Helicenes were deposited on metal surfaces and studied by various techniques.<br />

Ernst et al. used near-edge X-ray absorption spectroscopy with linearly polarised<br />

synchrotron radiation (NEXAFS) to study the orientation of (P)-[7]helicene on<br />

a Ni(100) surface under ultrahigh vacuum (UHV) conditions [208]. This group<br />

also studied chirality transfer from (M)-[7]helicene into handed supramolecular<br />

structures on a Cu(111) surface by STM [209] and the orientation and the intramolecular<br />

relaxation due to adsorption of nonracemic [7]helicene on Cu(111)<br />

and Cu(332) surfaces by means of angle-scanned full-hemispherical X-ray photoelectron<br />

diffraction [210].<br />

Various spectral and physicochemical properties of helicenes have been<br />

investigated. The fluorescence spectra, emission lifetimes, quantum yields of<br />

fluorescence and triplet state formation of a series of helicenes were studied by<br />

Vander Donckt and colleagues [211, 212]. They found the photophysical properties<br />

of the helicenes evolved steadily as a function of the number of ortho-fused<br />

benzene rings. Experimental photoelectron spectra of helicenes were analysed<br />

by Obenland and Schmidt to conclude that interaction between the �-orbitals of<br />

overlapping benzene rings is much smaller than in cyclophanes [213]. The comparison<br />

between experimental and calculated VCD spectra allowed Bürgi et al.<br />

to make the unequivocal assignment of the absolute configuration of [7]helicene<br />

[172]. An electrochemical study on helicenes was performed by Laarhoven and Brus<br />

who measured values of polarographic half-wave potentials [214]. Experimental


Scheme 4.29<br />

4.3 Helicenes<br />

estimation of the reduction and oxidation potential of [7]helicene was done by<br />

Rulíšek et al. by measuring current-voltage curves on inert electrodes [215].<br />

Focussing on molecular machinery, there is a fascinating application of a<br />

helicene structure by the Kelly group. In an artificial molecular motor, a helicene<br />

ratchet ensured unidirectional rotary motion fuelled by the chemical energy of<br />

periodic bond making/bond breaking processes (Scheme 4.29) [216].<br />

4.3.6<br />

Theoretical Studies<br />

The unique helicene structure has steadily attracted the attention of theoretical<br />

chemists. Grimme and Peyerimhoff studied the relationship between the structure<br />

and racemisation barrier in the series of helicenes by means of semiempirical<br />

AM-1 and ab initio SCF methods [217]. Similarly, Haufe and colleagues recalculated<br />

barriers to racemisation for helicenes finding an excellent agreement with the<br />

experimental results [218]. They confirmed the fact that the barriers for carbohelicenes<br />

converge to the value of about 45 kcal/mol. Ab initio calculations were<br />

carried out by Schulman and Disch in the series of helicenes and their closest<br />

topological isomers, planar phenacenes [219]. Their comparison revealed only<br />

slight loss of aromatic character in the former molecules.<br />

Most interestingly, the current-voltage characteristics of helicenes were calculated<br />

and their intriguing conductance was foreseen by Treboux et al. [220]. Helicenes<br />

can also be viewed as tiny mechanical objects since their structure resembles a<br />

molecular spring. Hartree–Fock calculations by Lipkowitz and colleagues using<br />

PM3 Hamiltonian revealed that the nanospring stiffness could be modulated by<br />

increasing/decreasing the electron density and length of helicenes [221].<br />

Optical and chiroptical properties of helicenes were studied in detail by computational<br />

methods. The electronic CD spectra were calculated by the Ahlrichs<br />

group for helicenes, exploiting the adiabatic time-dependent DFT method as a<br />

prime tool for chiroptical property investigations. Thus agreement between the<br />

most important spectral features and theory was found [222]. There are also other<br />

theoretical approaches to chiroptical spectroscopy published by Hansen and Bak<br />

[223]. The hyper-Rayleigh scattering second-order NLO responses of helicenes<br />

were investigated by the Botek group employing the time-dependent Hartree-Fock<br />

175


176 4 <strong>Strained</strong> Aromatic Molecules<br />

approach and AM1 semi-empirical Hamiltonian [224]. The results of a series of<br />

DFT and DFT-D (the empirical dispersion energy terms included) calculations<br />

were reported by Rulíšek et al. with the aim to predict the physicochemical properties<br />

(equilibrium structures, stabilisation energies, redox potentials, excitation and<br />

CD spectra, electronic conductivity and elasticity) of elongating helicene structures<br />

[215]. It was shown that many of them are converged on [14]helicene.<br />

4.3.7<br />

Outlook<br />

After five decades of helicene chemistry, interest in these unique aromatic<br />

compounds will certainly continue in the forthcoming years and there are many<br />

good reasons for it. As the synthesis of helicenes has recently witnessed significant<br />

progress, various helical aromatics have become more easily available than<br />

before. Important achievements might be expected, for instance in developing<br />

general asymmetric synthesis of helicenes, which remains still unsolved. However,<br />

applications of helicenes in various branches of chemistry, material science and<br />

nanoscience will be central to further efforts.<br />

4.4<br />

Cycloproparenes<br />

Brian Halton<br />

4.4.1<br />

Introduction<br />

The most highly strained ring-fused aromatics, the cyclopropa- and cyclobutabenzenes,<br />

have provided distinct fields of study for more than 50 years [225, 226].<br />

Cyclopropabenzene 139 made its debut in an 1888 paper by Perkin [227] where<br />

its synthesis was noted as yet to be done; cyclobutabenzene 140 appeared some<br />

20 years later in a 1909 thesis footnote [228]. Each parent has provided classes of<br />

compounds that continue to be the subjects of detailed scrutiny with numerous<br />

reviews covering various aspects of cycloproparene [226, 229–232] and cyclobutarene<br />

[225, 233] chemistry. The present cycloproparene synopsis summarizes<br />

developments in cycloproparene chemistry over the past five years [229].<br />

The first cycloproparene claim [234] made in 1930, was inconclusive [235] and<br />

the 1953 addition [236] of Ph 2 CN 2 to imide 141 did not give 143 as thought but<br />

the sulfonamides 142 [237] as now confirmed from X-ray data [238].


4.4.2<br />

Synthetic Considerations<br />

4.4 Cycloproparenes<br />

That 3H-indazoles, analogs of 143, are precursors to cycloproparenes was<br />

confirmed by dinitrogen loss from 144 that gave the first stable derivative [239],<br />

but the route is not general. Wege found [240] that debromosilylation of 145 led<br />

to quinone 146 as a transient trapped by furan as exo/endo adducts. Attempts to<br />

obtain a kinetically more stable quinone targeted [241] C1 dialkylation via diazopropane<br />

addition to 1,4-naphthoquinone and subsequent deazetation.<br />

The addition gave 147 that was easily oxidized to 148 (37%) (Scheme 4.30)<br />

but irradiation (350 nm) then gave a complex mixture [241]. Ultimately, primary<br />

adduct 147 was diverted to 149 (30%) and this lost nitrogen affording 150 (50%)<br />

and 151 (38%) as expected [232, 239, 242]. Attempts to dideacetylate 149 to hydroquinone<br />

and/or quinone were unsuccessful. Modification of the acetate groups<br />

of 149 gave further six derivatives of which only the 2-acetoxy-7-methoxy affords<br />

a cycloproparene on photolysis (Table 4.2, see p. 179).<br />

The Collis–Wege results [241] indicate limitations in the 3H-indazole route.<br />

Sylvania 350 nm lamps are ideal for excitation of C9 acetoxy compounds absorbing<br />

close to 350 nm (Table 4.2). Substrates that absorb at longer wavelength do not<br />

react because of insufficient energy absorption or enhanced mesomerism to N1<br />

that strengthens the C9a–N1 bond; impact of the C9 substituent on the excited<br />

state also needs consideration [241].<br />

The planarization of annulenes by annelating small rings across strategic sites<br />

has been recognized for some time. Dürr [243] synthesized 152 in 1983 and<br />

demonstrated close planarity by X-ray analysis. Schleyer’s group subsequently<br />

predicted highly aromatic planar all-cis-[10]annulenes from cyclopropa fusions<br />

[244], and Sastry’s group now have extended this to the rim of corannulene 153<br />

(B3LYP/6-31G+G* calculations) where the inversion barrier is reduced by > 50%<br />

if all rim bonds are cyclopropannelated [245]. Now, the first parent non-benzenoid<br />

177


178 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.30<br />

cyclopropannulene has been isolated by Stevenson [246]. Reaction of CH 2 Cl 2 and<br />

cycloC 8 H 7 Br with t-BuOK in HMPA generates anion radical 154, but in THF 155<br />

is formed from I 2 oxidation of the dianion. Unlike the foul smell of 139, 155 has a<br />

sweetly olefinic odor. NMR spectra have the CH 2 protons at � 4.61 (139: � 3.11) and<br />

those of the eight-membered ring at � 3.6–3.7; C1 (� 90.4) is deshielded 72 ppm<br />

relative to 139 and 13 C labeling gives J C1–H 162 Hz (139: 170 Hz [247]; cycloC 3H 4:<br />

167 Hz [248]). Clearly 155 has C1a–C2 and C7–C7a � bonded and a paratropic<br />

ring current. Calculations (B3LPY/6-31G*) confirm this with C1a–C7a 141.7 and<br />

C1a–C2 132.9 pm [246].<br />

The formation of benzdiynes by loss of CO and CO 2 from designed precursors<br />

continues with Sato confirming [249] 157 as a primary product from 156,<br />

and that loss of CO then gives diyne 158. Coupled with low temperature IR<br />

studies, energy diagrams (B3LPY/6-31G*) place the derivatives in the order of<br />

stability 157 > 159 > 160 > 158 with energy differences as shown in Scheme 4.31.<br />

Likewise, benzocyclopropenone 161 is 138 kJ mol –1 more stable than 162, but only<br />

42 kJ mol –1 below ring contracted 163.


4.4 Cycloproparenes<br />

Table 4.2 Photochemical behavior of 4,9-disubstituted 3,3-dimethyl-3H-benz[f]indazoles. a)<br />

R 1<br />

R 2<br />

� max<br />

log �<br />

Outcome b)<br />

Ac Ac 354 (3.49) CPN (50%) + STY (38%)<br />

Me Ac 357 (3.57) CPN (32%) + STY (43%)<br />

H TBDMS c)<br />

390 (3.68) decomposition<br />

Ac TBDMS 368 (3.65) decomposition<br />

Me TBDMS 375 (3.73) decomposition<br />

Ac H 372 (3.67) no reaction<br />

Ac Me 366 (2.99) no reaction<br />

Me H 379 (3.74) no reaction<br />

Me Me 372 (3.71) no reaction<br />

a)<br />

Data reproduced with permission from the Australian Journal of Chemistry:<br />

htpp://www.publish.csiro.au/journals/ajc; see Ref. [17].<br />

b)<br />

CPN = cyclopropanaphthalene; STY = styrene.<br />

c) t<br />

TBDMS = BuMe2Si- 179


180 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.31<br />

Scheme 4.32<br />

The most general route to the cycloproparenes is from use of 1-bromo-2-chlorocyclopropene<br />

in Diels–Alder cycloaddition and subsequent didehydrohalogenation<br />

as illustrated for 164 [229, 250, 251]. Addition of :CCl 2 to cycloC 6H 8 and


Scheme 4.33<br />

4.4 Cycloproparenes<br />

subsequent dehydrochlorination (the Billups synthesis; Scheme 4.32) remains<br />

the method of choice for parents 139 and 165 [229, 252]. Detailed procedures for<br />

preparation of alkylidenes 167 from 165 via anion 166 and silyl-Wittig olefination<br />

(Scheme 4.33) [229, 253] are now formalized into four protocols that gave<br />

~30 new 3,6-dimethoxycyclopropa[b]naphthalenes [254]. Applications are largely<br />

to �-extended, conjugated and cross-conjugated derivatives with ‘push–pull’<br />

character that include dithioles 168/169 [255], trienes 170–172 [256], and ‘spaced’<br />

[257] 173–175; p-HCO-C 6 H 4 -CHO and 166 give 174 with 173 as a trace product.<br />

Not all reactions go to plan [258].<br />

181


182 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.34<br />

Reactive bicyclopropenylidenes–novel C 6 H 4 valence bond isomers–are formed<br />

from anion 166 with bulkily substituted cyclopropenones 176 (Scheme 4.34) [259].<br />

While complex mixtures ensue, those from 176b/176c afford diones 178b/178c<br />

in low yield (~ 8%) via triafulalvenes 177, as established from the mass spectrum<br />

of isolated 177b and its subsequent transformation to 178b in air.<br />

An inverse electron demand route to alkylidenecycloproparenes has yet to receive<br />

the detailed assessment it deserves [260]. Condensation of cation 179 [261] with<br />

anion 180 results in coupling and in situ HCl loss to 181 that co-crystallizes with<br />

water as a hemihydrate.


4.4.3<br />

Chemical Considerations<br />

4.4 Cycloproparenes<br />

Cycloproparene functional group modification can provide access to difficultly<br />

accessible derivatives. Dilithiation of 182, reaction with PhCONMe 2, and condensation<br />

of the ensuing dione with LiC 5H 5 gives � extended 183 that packs in sheets<br />

in the solid state (Scheme 4.35) [255]; by analogy 184 gives 185 that provides 186<br />

and 187. Horner–Wittig reaction of cyclopropaquinone 188 provides 189 and 190<br />

but the reaction fails with lower homolog 191, likely due to its enhanced enedione<br />

character [255]. Bis(dithiole) 190, sensitive to acid and light, decomposes on purification.<br />

In fact, many � extended alkylidenes are only available in poor yields<br />

with their ‘push–pull’ ability for the new materials arena untested.<br />

The HOMO of 139 is located at the bridge and the C3–C4 bond [262] so that<br />

inverse electron demand [2+4] cycloadditions are to the bridge as illustrated by<br />

Scheme 4.35<br />

183


184 4 <strong>Strained</strong> Aromatic Molecules<br />

192/193 [263]. Strain is manifested by these reactions and those that use the<br />

lateral � bond as a two-electron component [229]. Such [3+2] cycloadditions,<br />

usually catalyzed by AgBF 4 or Eu(fod) 3 , involve ionic intermediates and proceed<br />

best in polar media to a wide range of heterocycles as summarized for 165 in<br />

Scheme 4.36 [264]. Notable is the cyclization to 195 via C–S bond formation in<br />

194 in a yield that almost doubles under Eu(fod) 3 catalysis; the alternative C–N<br />

closure is not seen [265]. As for formation of 200, thiotropone adds to 165 but<br />

the initial spirocycle rearranges via [1.7] C shift as shown to give 201 in analogy<br />

to tropone; the yield is but 5% [266].<br />

Exocyclic alkene 181, explicitly prepared for flash vacuum thermolysis (FVT)<br />

study, ejects Me 2 CO and CO 2 as expected and gives 202 (Ar matrix, 20 K) characterized<br />

by its IR spectrum matching that calculated (B3LPY/6-31G*) [260]. The<br />

existence of such strained and bent ethylideneone derivatives of cyclopropabenzene,<br />

first proposed some 17 years ago from bis(diazoketone) photolyses [229,<br />

267] (cf. 163, Scheme 4.31) is now settled.<br />

FVT details for methylidenes 203 and 204 alluded to [268] in 1990 have appeared<br />

[269]; they do not parallel parents 139 and 165 in their behavior [270]. Diphenyl<br />

203 cyclodehydrogenates to 204 and both proceed to a range of C 24H 14 polycyclic<br />

aromatics (Schemes 4.37 and 4.38). FVT of 203 gives [e,k]acephen-206 [271] (7%)<br />

and [a,e]ace-anthrylene 209 (12%). With 204, [e,l]acephenanthrylene 212 (47%) is<br />

obtained via 210; 206 is the minor polycyclic aromatic hydrocarbon (< 4%). Traces<br />

of dimer 215 were also detected. Automerization of the products in analogy to<br />

acephenanthrylene [272] could lead from 206 to 208, and from 212 to 214 [269].


Scheme 4.36<br />

4.4 Cycloproparenes<br />

185


186 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.37<br />

Scheme 4.38<br />

The well used [250, 273] Ag(I) catalyzed dimerization of cycloproparenes fails<br />

[274] with alkylidenes 167 (Scheme 4.39). Conditions effective for 139 and 165<br />

do not apply to 216, the simplest compound tested; ethyne 217 and ethanone<br />

218 (3 : 1) are obtained without 219. While metal ion is involved, none of the aryl<br />

derivatives studied dimerizes due to steric constraints at C1; simpler alkylidene<br />

derivatives such as C1 ethylidene (=CHMe) remain unknown.


Scheme 4.39<br />

4.4.4<br />

Heteroatom Derivatives<br />

4.4 Cycloproparenes<br />

Cycloproparenes carrying a heteroatom other than in the 6–3 fusion are unexceptional<br />

[229], those with the atom in the fused six-membered ring date to 1987 [275],<br />

and those involving N [276], S or Se [277] at C1 have been covered [229]. Recent<br />

efforts are aimed at incorporating group III/IV atoms at C1. Benzoborirenes 220<br />

have been characterized [278]. Computed structures of 220 (R 1 = R 2 = H) and the<br />

cyclopropabenzenyl cation [279] show marked reverse Mills–Nixon deformation<br />

within the aromatic unit due to strong � delocalization over the three-membered<br />

ring and rehybridization of the fusion sites.<br />

Kinetically stabilized metallacyclopropabenzenes are obtained as stable compounds<br />

when the very bulky Dip and Tbt substituents (Scheme 4.40) are present<br />

at C1 [280]. Dibromides 221 transform to 222 that react with o-dibromobenzene<br />

to give not only 223 (34%) [281, 282] and 224 (40%) [283], but also the bisheterocycles<br />

225 [281, 284] and 226 [285], as separable cis/trans-isomers. Structures are<br />

confirmed from X-ray and spectroscopic data (Section 4.4.5) and delineate the<br />

impact of strain in the cycloproparenes.<br />

187


188 4 <strong>Strained</strong> Aromatic Molecules<br />

Scheme 4.40<br />

4.4.5<br />

Physicochemical and Theoretical Considerations<br />

The cycloproparenes are ‘push–pull’ � systems, the direction of which depends on<br />

the nature of the attached substituents [231, 286] as confirmed by 3,6-di methoxynaphthalene<br />

derivatives [254, 287]. Cycloheptatrienylidenes 227–231 with electron<br />

donation from the cycloproparene to the seven-membered ring best illustrate this<br />

[287]. The dipole moments of 227–231 range from 1.4(5) D to 1.8(5) D with those<br />

of parents, 233 and 234 directed away from the cycloproparene core and that of<br />

232 directed towards it (HF/6-31G(d,p) calculations) [288]. That diethers 229 and<br />

231 are each more polar than parents 228 and 230 verify this for 227, 228 and 234.<br />

Furthermore, X-ray analyses of 227 and 229 show the seven-membered rings<br />

markedly distorted from planarity. The cycloheptatrienylidene –CH=CH– moiety<br />

of 227 is bent out of the cycloproparene plane by ~28° and that of more polar 229<br />

by ~45° (Figure 4.3). This represents physical resistance of the seven-membered<br />

ring to developing 8� antiaromatic character enforced by the more powerful<br />

electron donating cycloproparene. In contrast, the 6� 5C cyclopentadienylidenes<br />

235 and 204 are essentially planar [288].<br />

Figure 4.3 Superimposed side perspectives of cycloheptatrienylidenes 227 (dashed) and 229<br />

(solid). (Reproduced from [63] with permission from Elsevier).


4.4 Cycloproparenes<br />

The linearly dependence of cycloproparene 13 C NMR shifts on p-aryl- and p,p�diarylmethylidene<br />

substituents [289] extends to the range of 3,6-dimethoxy derivatives<br />

[254] as shown in Figure 4.4. The correlations apply also to the �-extended<br />

derivatives 168–175.<br />

Cycloproparene charge transfer (CT) complexation is demonstrated by admixture<br />

of dithiole 190 and 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) in<br />

MeCN with an instantaneous green that has a new broad CT absorption in the<br />

540–620 nm range; the complex was too unstable for isolation [255]. A more systematic<br />

study [290] with p-substituted mono- and diaryl derivatives of 167 showed<br />

only the bis(dimethylanilino) of the diaryls to complex, and it did so with, DDQ,<br />

7,7,8,8-tetracyano-1,4-benzoquinodimethane (TCNQ), 2,3,5,6-tetrafluoro-7,7,8,8tetracyano-1,4-benzoquinodimethane<br />

(TCNQF 4 ), tetracyanoethene (TCNE), and<br />

o-chloranil. In contrast, the mono anilino complexes with DDQ only and in<br />

MeCN, while the p-OMe, p-SMe, and p-Ph homologs complex with TCNQF 4 , but<br />

only in CH 2 Cl 2 . HF/6-31G**-derived HOMO-LUMO interactions account for<br />

the complexation as illustrated by 236 and 237. That of the diaryl is via a pendant<br />

aromatic while the essentially planar mono-aryls facilitate face-to-face interaction.<br />

A more detailed study is needed.<br />

189


190 4 <strong>Strained</strong> Aromatic Molecules<br />

+<br />

Figure 4.4 Plots of �p vs �C for (a) 1-(arylmethylidene)-1H-cyclopropa[b]naphthalenes and<br />

(b) 3,6-dimethoxy analogs. (Reproduced from [254[ with permission from the Royal Society of<br />

Chemistry).<br />

Much impetus for cycloproparene study comes from the juxtaposition of aromaticity<br />

and strain with the concepts of aromatic bond localization (Mills–Nixon<br />

effect) and � delocalization central. Noted earlier [245] was the flattening of<br />

cyclopropa-fused conjugated polyenes (B3LPY/6-31G*), and now we know that<br />

the bond localization imposed on benzene by tris-fusion as in 238 [291] and 239<br />

[292] extends to the pyridazines 240 and 241 [293]. Crystallographic analyses of<br />

the essential cycloproparenes have been discussed [229] and current interest lies


4.4 Cycloproparenes<br />

with heteroatom derivatives. The cycloproparenes are fully delocalized aromatics<br />

whose � frames are distorted primarily about the fusion sites [139: C1a–C5a,<br />

133.4(4) pm, C1a–C2, 136.3(3) pm] [294] as expected from strain induced bond<br />

localization (SIBL) [295]. Optimized structures (B3LYP/aug-cc-pVDZ) give negative<br />

nucleus independent chemical shift (NICS) values fully commensurate with aromaticity<br />

[296]. Maksi� identifies positive and negative Mills–Nixon effects with<br />

differing substituents [279, 297] (cf. 220 and cyclopropabenzenyl cations [279]).<br />

Importantly, clear distinction between crystallographic � frame measurements<br />

and strain effects that impact on the � orbitals is vital; cycloproparenes have no<br />

meaningful bond alternation. The diatropic ring current can now be visualized<br />

using the distributed-origin, coupled Hartree–Fock method [298].<br />

Structures of heteroatom derivatives 223–226 have appeared [280, 285]. Table 4.3<br />

shows that a large atom at the 1-position alleviates steric constraints and reduces<br />

bond length deviations; comparative data for 139 and (unknown) biscycloproparene<br />

242 are included. Bridge bonds are lengthened from 133.4 pm in 139 to<br />

between 139.0 pm (223) and 141.5 pm (cis-225), and there is essentially no aro matic<br />

bond length variation in bis-225 and 226. This is in stark contrast to unknown 242<br />

where marked variations are predicted. Despite larger atom sizes, the hetero atom<br />

derivatives retain internal six-membered ring angles � less (111.9–117.3°) and �/�<br />

greater (121.2–124.7°) than the 120° of the ideal sp 2 -hybrid with deviations greatest<br />

in the bis-derivatives. Three-membered ring fusion continues to manifest itself<br />

even with heteroatoms present, but it is the parent hydrocarbons that are forced<br />

to maximize the distortions and provide unique structures.<br />

Cyclopropabenzenyl anion (139A), generated by hydroxide ion deprotonation<br />

in the gas phase and computed at MP2/6-31+G(d)//HF/6-31+G(d), is more<br />

stable [299] than its C2 and C3 isomers by 28 and 42 kJ mol –1 . According to<br />

MP2/6-31+G(d) computation, 4� interaction in the three-membered ring is minimized<br />

when bonds about the bridge sites (C1a–C5a, 139 pm; C1a–C2, 137.9 pm)<br />

are lengthened and C1 pyramidalized (C1–H is bent 50.6° from the cycloproparene<br />

plane) [300]. Homolog 165A comes from fluoride ion deprotonation (Fourier<br />

transform mass spectrometry) [300, 301]. The proton affinities (PA) are 139A:<br />

1614.2; 165A: 1535.5; 243A (unknown) 1607.5 kJ mol –1 , respectively [300]. While<br />

the calculated energies of 165 and 243 differ by only 4 kJ mol –1 , the PAs show the<br />

former to be 72 kJ mol –1 more acidic consistent with resonance structures and<br />

avoidance of 8� antiaromaticity. Notable is the lower acidity of 139 than that of<br />

toluene by ~12.5 ± 13 kJ mol –1 while 165A is more acidic than 2-methylnaphthalene<br />

by 32 ± 13 kJ mol –1 due to increased ring size and reduced importance of the 4� interaction;<br />

165A is an aromatic anion [301]. Because 243A is ~40 kJ mol –1 less stable<br />

191


192 4 <strong>Strained</strong> Aromatic Molecules<br />

Table 4.3 Observed and calculated structure parameters for cyclopropabenzene and its<br />

heteroatom derivatives. a)<br />

139 b)<br />

223 c)<br />

224 c)<br />

cis-225 c)<br />

trans-<br />

225 c,d)<br />

Mol A<br />

trans-<br />

225 c,d)<br />

Mol B<br />

cis-226 c)<br />

trans-<br />

226 c)<br />

242 e)<br />

a 133.4(4) 139.0(4) 139.1(3) 141.5(6) 140.5(11) 139.8(11) 139.4(6) 139.4(6) 134.6<br />

b 136.3(3) 138.8(4)<br />

139.4(4)<br />

c 138.7(4) 138.1(4)<br />

138.3(4)<br />

138.5(3)<br />

138.8(3)<br />

139.2(3)<br />

139.1(2)<br />

140.9(5)<br />

141.4(6)<br />

139.7(6)<br />

139.4(6)<br />

138.8(11) 138.9(11) 138.1(8)<br />

141.4(11) 137.6(11) 140.3(9)<br />

139.4(6)<br />

140.3(6)<br />

139.6(9) 140.5(11) 139.6(9) 139.7(6)<br />

138.9(11) 140.1(11) 139.1(6)<br />

139.2<br />

139.2<br />

d 139.0(5) 140.3(3) 139.7(3) 141.2(6) 1.420(10) 138.0(11) 136.4(9) 138.9(6) 134.6<br />

e 149.8(3) 182.6(2)<br />

182.8(3)<br />

� 124.5(2) 121.3(3)<br />

121.5(2)<br />

� 113.2(2) 117.3(3)<br />

117.3(3)<br />

� 122.4(2) 121.2(3)<br />

121.4(3)<br />

198.2(2)<br />

194.0(2)<br />

121.9(2)<br />

122.0(2)<br />

116.5(2)<br />

116.3(2)<br />

121.6(2)<br />

121.7(2)<br />

– – – – – –<br />

123.2(4)<br />

122.9(3)<br />

113.7(7)<br />

114.3(4)<br />

122.9(4)<br />

123.1(4)<br />

124.7(7)<br />

122.7(4)<br />

113.7(7)<br />

114.8(7)<br />

122.9(4)<br />

122.7(4)<br />

123.0(7)<br />

123.2(8)<br />

113.0(7)<br />

114.6(7)<br />

124.2(7)<br />

121.8(7)<br />

123.1(6)<br />

123.5(6)<br />

113.3(6)<br />

111.9(6)<br />

124.2(6)<br />

122.7(4)<br />

123.8(4)<br />

123.6(4)<br />

112.5(4)<br />

112.3(4)<br />

123.7(4)<br />

124.2(4)<br />

� 52.8(2) 44.7(1) 42.11(8) – – – – – –<br />

a) Data taken from Refs. [280] and [285]; bond lengths in picometres (pm), angles in degrees (°).<br />

b)<br />

X-ray data at –150 °C taken from Ref. [294].<br />

c)<br />

X-ray at –170 °C.<br />

d)<br />

Independent molecules A and B are in the unit cell.<br />

e)<br />

Optimized at B3LYP/6-311G(2d,p).<br />

126.1<br />

107.8<br />

126.1


References<br />

than its acyclic counterparts, it is less able to alleviate the unfavorable 4� interaction<br />

in the three-membered ring. It is best regarded as antiaromatic. Calculated<br />

proton acidities of homologs 244A–246A are 1483.6, 1611.7, and 1531.3 kJ mol –1<br />

respectively [300]. Computational study of mono-anionic dicycloproparenes 247A<br />

and 248A (PA: 1579.0/1605.4 kJ mol –1 ) shows that the second three-membered<br />

ring enhances acidity. The proton affinity of 139A is 35.2 kJ mol –1 above that of<br />

247A but only 8.8 kJ mol –1 above that of 248A [300].<br />

The impact of mesomeric CN groups and inductive F substituents on the<br />

stability of 139A has been assessed [301]. The CN group, able to conjugate to<br />

the C1 anion, increases acidity more at C2(5) than C3(4). Difluorination inductively<br />

raises acidity by 43.1 kJ mol –1 at C2(5) but reduces it by 43.9 kJ mol –1 at<br />

C3(4); replacement of all aromatic ring protons in 139A by F decreases acidity<br />

by 83.7 kJ mol –1 [300]. A range of C1 substituents enhance acidity; Me has least<br />

(< 1 kJ mol –1 ) and SO 2 H most (141 kJ mol –1 ) impact [302], but their impact is less<br />

than on the cyclopropenyl anion.<br />

References<br />

1 Greenberg, A.; Liebman, J. F. <strong>Strained</strong><br />

Organic Molecules; Academic Press, New<br />

York, 1978, in particular, Section 3. H.,<br />

pp. 139–178.<br />

2 Aromaticity a topical issue of Chem. Rev.<br />

101(5) 2001, 1115–1566 (special editor,<br />

Schleyer, P. v. R.).<br />

3 Delocalization – Pi and Sigma a<br />

topical issue of Chem. Rev., 105(10)<br />

2005, 3433–3947 (special editor,<br />

Schleyer, P. v. R.).<br />

4 Designing the Molecular World a<br />

topical issue of Chem. Rev. 106(12)<br />

2006, 4785–5430 (special editors<br />

de Meijere, A.; Hopf, H.).<br />

5 Cram, D. J.; Cram, J. M. Acc. Chem. Res.<br />

1971, 4, 204.<br />

6 Moran, D.; Simmonett, A. C.; Leach III,<br />

F. E.; Allen, W. D.; Schleyer, P. v. R.;<br />

Schaefer III, H. F. J. Am. Chem. Soc.<br />

2006, 128, 9342.<br />

7 Pedley, J. B. Thermochemical Data and<br />

Structures of Organic Compounds, Vol. I.,<br />

TRC Data Series, College Station, Texas<br />

1994. All unreferenced enthalpies of<br />

formation may be assumed to be from<br />

this archival source, and to reiterate,<br />

are assumed to refer to gaseous species<br />

if the phase is unspecified. These gas<br />

phase data were most often obtained by<br />

193


194 4 <strong>Strained</strong> Aromatic Molecules<br />

summing experimental measurements<br />

of the enthalpy of combustion (and<br />

hence of formation) with those of sublimation<br />

or vaporization (and hence of<br />

solid or liquid � gas interconversion).<br />

8 Brown, H. C.; Domash, L. J. Am. Chem.<br />

Soc. 1956, 78, 5384, reporting otherwise<br />

unpublished results by E. J. Prosen.<br />

9 (a) Arnett, E. M.; Sanda, J. C.;<br />

Bollinger, J. M.; Barber, M. J. Am. Chem.<br />

Soc. 1967, 89, 5389; (b) Nesterova, T. N.;<br />

Verevkin, S. P.; Karaseva, S. Y.;<br />

Rozhnov, A. M.; Tsvetkov, V. F. Russ.<br />

J. Phys. Chem. (Engl. Transl.) 1984, 58,<br />

297. Also see, Verevkin, S. P. J. Chem.<br />

Thermodyn. 1998, 30, 1029.<br />

10 Kruerke, U.; Hoogzand, C.; Hubel, W.<br />

Chem. Ber., 1961, 94, 2817.<br />

11 Tasi, G.; Mizukami, F.; Toba, M.;<br />

Niwa, S.; Palinko, I. J. Phys. Chem. A<br />

2000, 104, 1337.<br />

12 The value is the sum of the archival<br />

enthalpy of formation of the liquid<br />

and an estimated enthalpy of vaporization<br />

from the simple approach<br />

in Chickos, J. S.; Hyman, A. S.;<br />

Ladon, L. H.; Liebman, J. F. J. Org.<br />

Chem. 1981, 46, 4294.<br />

13 (a) Miki, S.; Katayama, T.; Yoshida, Z.<br />

Chem. Lett. 1992, 41; (b) Mandella, W. L.;<br />

Franck, R. W. J. Am. Chem. Soc. 1973,<br />

95, 971; (c) Goldberg, I. B.; Crowe, H. R.;<br />

Franck, R. W. J. Am. Chem. Soc. 1976, 98,<br />

7641.<br />

14 (a) Anderson, J. E.; Franck, R. W.;<br />

Mandella, W. L. J. Am. Chem. Soc., 1972<br />

94, 4608; Anderson, J. E.; Franck, R. W.<br />

J. Chem. Soc. Perkin Trans. 2, 1984,<br />

1581; (b) Handal, J.; White, J. G.;<br />

Franck, R. W.; Yuh, Y. H.; Allinger, N. L.<br />

J. Am. Chem. Soc. 1977, 99, 3345.<br />

15 Penner, G. H.; Chang, Y.-C. P.;<br />

Nechala, P.; Froese, R. J. Org. Chem.<br />

1999, 64, 447.<br />

16 (a) Kraljic, I.; Mintas, M.; Klasinc, L.;<br />

Ranogajec, F.; Guesten, H.<br />

Nouv. J. Chim. 1983, 7, 239;<br />

(b) Abdul-Ghani, A. J.; Bashi, N. O. T.;<br />

Maree, S. N. J. Solar Energ. Res. 1987,<br />

5, 53.<br />

17 Fu, P. P.; Harvey, R. G. J. Org. Chem.<br />

1977, 42, 2407.<br />

18 Pozdnyakovich, Y. V. Zhur. Org. Khim<br />

1988, 24, 1076.<br />

19 Grimme, S.; Pischel, I.; Nieger, M.;<br />

Voegtle, F. J. Chem. Soc., Perkin Trans. 2<br />

(Phys. Org. Chem.) 1996, 2771.<br />

20 Armstrong, R. N.; Ammon, H. L.;<br />

Darnow, J. N. Am. Chem. Soc. 1987, 109,<br />

2077.<br />

21 Frisch, M. A.; Barker, C.; Margrave, J. L.;<br />

Newman, M. S. J. Am. Chem. Soc. 1963,<br />

85, 2356. Actually, we should have said<br />

negligibly strained in that the parent<br />

hydrocarbon phenanthrene itself has<br />

been deduced to be strained by 4.9(2.8)<br />

kJ mol –1 , Nagano, Y.; Nakano, M.<br />

J. Chem. Thermodyn. 2003, 35, 1403.<br />

22 (a) Mark, H. B., Jr.; Jezorek, J. R. J. Org.<br />

Chem.1971, 36, 666; (b) Grimme, S.;<br />

Loehmannsroeben, H. G.<br />

J. Phys. Chem. 1992, 96, 7005;<br />

(c) Dougherty, R. C.; Bertorello, H. E.;<br />

Martinez de Bertorello, M. Org. Mass<br />

Spectry. 1971, 5, 1321.<br />

23 (a) Herbstein, F. H.; Schmidt, G. M. J.;<br />

J. Chem. Soc. 1954, 3302;<br />

(b) Haigh, C. W.; Mallion, R. B. Mol.<br />

Phys. 1971, 22, 945; Martin, R. H.;<br />

Defay, N.; Zimmermann, D. Tetra hedron<br />

Lett. 1971, 1871; (c) Schmidt, W. J. Chem.<br />

Phys. 1977, 66, 828; (d) Wynberg, H.;<br />

Nieuwpoort, W. C.; Jonkman, H. T.<br />

Tetrahedron Lett. 1973, 4623.<br />

24 Martin, R. H.; Marchant, M. J.<br />

Tetrahedron Lett. 1972, 3707.<br />

25 Borkent, J. H.; Laarhoven, W. H.<br />

Tetrahedron 1978, 34, 2565.<br />

26 (a) Li, H.; Petersen, J. L.; Wang, K. K.<br />

J. Org. Chem. 2001, 66, 7804;<br />

(b) Zhang, Y.; Petersen, J. L.; Wang, K. K.<br />

Org. Lett. 2007, 9, 1025<br />

27 (a) Templeton, W. J. Chem. Soc. D:<br />

Chem. Commun. 1970, 21, 1412;<br />

(b) Roberts, T. D. J. Chem. Soc. D: Chem.<br />

Commun. 1971, 362.<br />

28 (a) Phenanthrene, Nagano and Nakano,<br />

op. cit. ref. [21] and Steele, W. V.;<br />

Chirico, R. D.; Nguyen, A.;<br />

Hossenlopp, I. A.; Smith, N. K. Am. Inst.<br />

Chem. Eng. Symp. Ser. (AIChE Symp.<br />

Ser.), 1990, 138; (b) Diphenylacetylene,<br />

Davis, H. E.; Allinger, N. L.;<br />

Rogers, D. W. J. Org. Chem. 1985, 50,<br />

3601.<br />

29 However, stability for such strained<br />

species as the bisindenoarenes is clearly<br />

precarious in that a not altogether


different set of precursor molecules<br />

dismembered a benzene ring to form a<br />

nonaromatic, at least, presumably less<br />

aromatic, fulvalene derivative: Yang, Y.;<br />

Petersen, J. L.; Wang, K. K. J. Org. Chem.<br />

2003, 68, 5832.<br />

30 Chickos, J. S.; Webb, P.; Nichols, G.;<br />

Kiyobayashi, T.; Cheng, P.-C.; Scott, L. T.<br />

J. Chem. Thermodyn. 2002, 34, 1195.<br />

31 (a) Scott, L. T.; Hashemi, M. M.;<br />

Bratcher, M. S. J. Am. Chem. Soc.<br />

1992, 114, 1920; (b) Borchardt, A.;<br />

Fuchicello, A.; Kilway, K. V.;<br />

Baldridge, K. K.; Siegel, J. S. J. Am.<br />

Chem. Soc. 1992, 114, 1921.<br />

32 Slayden, S. W.; Liebman, J. F. Chem. Rev.<br />

2001, 101, 1541.<br />

33 This equates gaseous graphite and the<br />

C BF -(C BF ) 3 group increment, Stein, S. E.;<br />

Golden, D. M.; Benson, S. W. J. Phys.<br />

Chem. 1977, 81, 314.<br />

34 Sato, M.; Yamamoto, K.; Sonobe, H.;<br />

Yano, K.; Matsubara, H.; Fujita, H.;<br />

Sugimoto, T.; Yamamoto, K. J. Chem.<br />

Soc., Perkin Trans. 2: Phys. Org. Chem.<br />

1998, 1909.<br />

35 Shen, M.; Ignatyev, I. S.; Xie, Y.;<br />

Schaefer, H. F., III. J. Phys. Chem. 1993,<br />

97, 3212.<br />

36 Yamamoto, K.; Harada, T.; Okamoto, Y.;<br />

Chikamatsu, H.; Nakazaki, M.; Kai, Y.;<br />

Nakao, T.; Tanaka, M.; Harada, S.;<br />

Kasai, N. J. Am. Chem. Soc. 1988, 110,<br />

3578.<br />

37 Billups, W. E.; Chow, W. Y.;<br />

Leavell, K. H.; Lewis, E. S.;<br />

Margrave, J. L.; Sass, R. L.; Shieh, J. J.;<br />

Werness, P. G.; Wood, J. L. J. Am. Chem.<br />

Soc. 1973, 95, 7878.<br />

38 Wenthold, P. G.; Wierschke, S. G.;<br />

Nash, J. J.; Squires, R. R. J. Am. Chem.<br />

Soc., 1994, 116, 7378.<br />

39 Levin, M. D.; Kaszynski, P.; Michl, J.<br />

Chem. Rev. 2000, 100, 169.<br />

40 (a) Kostermans, G. B. M.;<br />

Van Dansik, P.; De Wolf, W. H.;<br />

Bickelhaupt, F. J. Org. Chem. 1988,<br />

53, 4531; (b) Kostermans, G. B. M.;<br />

Van Dansik, P. L.; De Wolf, W. H.;<br />

Bickelhaupt, F. J. Am. Chem. Soc. 1987,<br />

109, 7887.<br />

41 Turkenburg, L. A. M.; Van Straten, J. W.;<br />

De Wolf, W. H.; Bickelhaupt, F. J. Am.<br />

Chem. Soc. 1980, 102, 3256.<br />

References<br />

42 Tsuji, T.; Ohkita, M.; Kawai, H.<br />

Bull. Chem. Soc. Jpn 2002, 75, 415.<br />

43 Tsuji, T.; Nishida, S. O. M.; Osawa, E.<br />

J. Am. Chem. Soc. 1995, 117, 9804.<br />

44 (a) 1,4-, Paquette, L. A.; Trova, M. P.<br />

J. Am. Chem. Soc. 1988, 110, 8197;<br />

(b) 1,3-, Wang, T. Z.; Paquette, L. A.<br />

Tetrahedron Lett. 1988, 29, 41;<br />

(c) 1,5-, Paquette, L. A.; Trova, M. P.;<br />

Luo, J.; Clough, A. E.; Anderson, L. B.<br />

J. Am. Chem. Soc. 1990, 112, 228.<br />

45 Van Eis, M. J.; de Wolf, W. H.; Bickelhaupt,<br />

F.; Boese, R. J. Chem. Soc., Perkin<br />

Trans. 2 (Phys. Org. Chem.) 2000, 793.<br />

46 Van Eis, M. J.; Wijsman, G. W.;<br />

De Wolf, W. H.; Bickelhaupt, F.;<br />

Rogers, D. W.; Kooijman, H.; Spek, A. L.<br />

Chem. Eur. J. 2000, 6, 1537.<br />

47 v. Schleyer, P. R.; Puehlhofer, F. Org.<br />

Lett. 2002, 4, 2873.<br />

48 Dreeskamp, H.; Sarge, S. M.;<br />

Tochtermann, W. Tetrahedron 1995, 51,<br />

3137.<br />

49 Rosenfeld, S.; Shedlow, A. J. Org. Chem.<br />

1991, 56, 2247.<br />

50 Bartmess, J. E.; Griffith, S. S. J. Am.<br />

Chem. Soc. 1990, 112, 2931. This<br />

reference shows using both gas and<br />

solution phase acidity measurements<br />

(a laboratory realization of the above ISE<br />

analysis, see [47]) that the enthalpy of<br />

formation difference of 9-methylanthracene<br />

and 9-methylenedihydroanthracene<br />

is some 20 kJ mol –1 favoring the former.<br />

51 Tobe, Y.; Saiki, S.; Utsumi, N.;<br />

Kusumoto, T.; Ishii, H.; Kakiuchi, K.;<br />

Kobiro, K.; Naemura, K. J. Am. Chem.<br />

Soc. 1996, 118, 9488.<br />

52 Roth, W. R.; Boehm, M.; Lennartz,<br />

H.-W.; Vogel, E. Angew. Chem. 1983,<br />

95, 1011.<br />

53 Sander, W.; Exner, M. J. Chem. Soc.,<br />

Perkin 2 1999, 2285.<br />

54 JFL once asked pioneering fluorine<br />

chemist Harry J. Emeléus how he<br />

devised the syntheses of so many exotic<br />

species. He replied (within the limits<br />

of memory): “When I ask myself how<br />

to make a given compound, I think of<br />

logical syntheses. If the compounds<br />

react as I hope, that’s fine. If they don’t,<br />

the compounds are reactive enough to<br />

produce something else of interest. So,<br />

how can I lose?”<br />

195


196 4 <strong>Strained</strong> Aromatic Molecules<br />

55 de Meijere, A.; Kozhushkov, S. I.;<br />

Rauch, K.; Schill, H.; Verevkin, S. P.;<br />

Kuemmerlin, M.; Beckhaus, H.-D.;<br />

Ruechardt, C.; Yufit, D. S. J. Am. Chem.<br />

Soc. 2003, 125, 15110.<br />

56 (a) Cram, D. J.; Dewhirst, K. C. J. Am.<br />

Chem. Soc. 1959, 81, 5963; (b) Hopf, H.;<br />

Laue, T.; Zander, M. Angew. Chem. Int.<br />

Ed. Engl. 1991, 30, 432.<br />

57 Yamato, T.; Noda, K.; Tsuzuki, H.;<br />

New J. Chem. 2001, 25, 721.<br />

58 Shieh, C.-F.; McNally, D.; Boyd, R. H.<br />

Tetrahedron 1969, 25, 3653.<br />

59 The following story told by Irvin Greenberg<br />

to JFL may be evocative. A talk to<br />

a major professional organization by<br />

Carl Rogers, a premier humanistic psychologist,<br />

and roughly entitled “What<br />

it Means to be a Psychologist” resulted<br />

in a listener’s question “What do you<br />

do when you find a client boring?”, also<br />

admitting concern because it seemed<br />

unprofessional. Rogers acknowledged<br />

having similar feelings and said<br />

“Whenever I find myself bored with a<br />

patient, I ask myself why I am bored.<br />

Attempting to answer this, now that I<br />

have a research problem, I am no longer<br />

bored.” Translating to chemistry, who<br />

would have thought that cyclophanes are<br />

normal looking?<br />

60 The title of this chapter was first used<br />

in a review on his cyclophane work<br />

by Donald Cram, one of the founding<br />

fathers of cyclophane chemistry:<br />

Cram, D. J.; Cram, J. M. Acc. Chem. Res.<br />

1971, 4, 204.<br />

61 Hopf, H. Classics in Hydrocarbon<br />

Chemistry; Wiley-VCH, Weinheim,<br />

2000; R. Gleiter, H. Hopf (eds.) Modern<br />

Cyclophane Chemistry; Wiley-VCH,<br />

Weinheim, 2005.<br />

62 Lüttringhaus, A.; Gralheer, H. Justus<br />

Liebigs Ann. Chem. 1942, 550, 67.<br />

63 Bickelhaupt, F.; de Wolf, W. H.<br />

Advances in Strain in Organic Chemistry<br />

(Halton, B., ed.), JAI Press, London,<br />

1993, Vol. 3, pp. 185; Kane, V. V.;<br />

de Wolf, W. H.; Bickelhaupt, F.<br />

Tetrahedron, 1994, 50, 4575;<br />

Bickelhaupt, F.; de Wolf, W. H. J. Phys.<br />

Org. Chem. 1998, 11, 362.<br />

64 Hopf, H.; Goldberg, N., unpublished<br />

results; for high level ab inbitio<br />

calculations see Rice, J. E.; Lee, T. J.;<br />

Remington, R. B.; Allen, W. D.;<br />

Clabo, D. A., Jr.; Schaefer, III, H. F.<br />

J. Am. Chem. Soc. 1987, 109, 2902;<br />

Lee, T. J.; Rice, J. E.; Allen, W. D.;<br />

Remington, R. B.; Schaefer, III, H. F.<br />

Chem. Phys. 1988, 123, 1; Grimme, S. T.<br />

J. Am. Chem. Soc. 1992, 114, 10542.<br />

65 Cram, D. J.; Knox, G. R. J. Am. Chem.<br />

Soc. 1961, 83, 2204; cf. Cram, D. J.;<br />

Montgomery, C. S.; Knox, G. R. J. Am.<br />

Chem. Soc. 1966, 88, 515; Noble, K. L.;<br />

Hopf, H.; Ernst, L. Chem. Ber. 1984, 117,<br />

455.<br />

66 Tsuji, T.; Nishida, S. Acc. Chem. Res.<br />

1984, 17, 56.<br />

67 Allinger, N. L.; Freiberg, L. A.;<br />

Hermann, R. B.; Miller, M. A.<br />

J. Am. Chem. Soc. 1963, 85, 1171;<br />

cf. Newton, M. G.; Walter, T. J.;<br />

Allinger, N. L. J. Am. Chem. Soc. 1973,<br />

95, 5652.<br />

68 Allinger, N. L.; Walter, T. J. J. Am. Chem.<br />

Soc. 1972, 94, 9267; cf. Allinger, N. L.;<br />

Walter, T. J.; Newton, M. G. J. Am. Chem.<br />

Soc. 1974, 96, 4588–4597.<br />

69 Otsubo, T.; Misumi, S. Synth. Comm.<br />

1978, 8, 285; Higuchi, H.; Kobayashi, E.;<br />

Sakata, Y.; Misumi, S. Tetrahedron, 1986,<br />

42, 1731.<br />

70 Dohm, J.; Vögtle, F. Top. Curr. Chem.<br />

1992, 161, 69; cf. Vögtle, F.; Rossa, L.<br />

Angew. Chem. Int. Ed. Engl. 1979, 18,<br />

274.<br />

71 Newton, M. G.; Walter, T. J.;<br />

Allinger, N. L. J. Am. Chem. Soc. 1973,<br />

95, 5652.<br />

72 Wolfe, A. D.; Kane, V. V.; Levin, R. H.;<br />

Jones, M., Jr. J. Am. Chem. Soc. 1973, 95,<br />

1680.<br />

73 Kane, V. V.; Jones, M., Jr. Org. Synth.<br />

1981, 61, 129.<br />

74 Hunger, J.; Wolff, W.; Tochtermann, W.;<br />

Peters, E. M.; Peters, K.;<br />

v. Schnering, H. G. Chem. Ber. 1986,<br />

119, 2698.<br />

75 Kane, V. V.; Wolf, A. D.; Jones, M., Jr.<br />

J. Am. Chem. Soc. 1974, 96, 2643.<br />

76 Tobe, Y.; Kakiuchi, K.; Odeira, Y.;<br />

Hosaki, T.; Kai, Y.; Kasai, N. J. Am. Chem<br />

Soc. 1983, 105, 1376.<br />

77 Liebe, J.; Wolff, Chr.; Krieger, C.;<br />

Weiss, J.; Tochtermann, W. Chem. Ber.<br />

1985, 118, 4144.


78 Kammula, S. L.; Iroff, L. D.;<br />

Jones, M., Jr.; van Straten, J. W.;<br />

de Wolf, W. H.; Bickelhaupt, F. J. Am.<br />

Chem. Soc. 1977, 99, 5815; cf. Weiss, R.;<br />

Schlierf, C. Angew. Chem. Int. Ed. Engl.<br />

1971, 10, 811; Weiss, R.; Andrae, S.<br />

Angew. Chem. Int. Ed. 1973, 12, 150.<br />

79 Tobe, Y.; Ueda, K.; Kikiuchi, K.;<br />

Odeira, Y.; Kai, Y.; Kasai, N. Tetrahedron,<br />

1986, 42, 1851; cf. Tobe, Y.; Ueda, K.;<br />

Kaneda, T.; Kakiuchi, K.; Odaira, Y.;<br />

Kasai, N. J. Am. Chem Soc. 1987, 109,<br />

1136.<br />

80 Jenneskens, L. W.; Louwen, J. N.;<br />

de Wolf, W. H.; Bickelhaupt, F.<br />

J. Phys. Org. Chem. 1990, 3, 295.<br />

81 Jenneskens, L. W.; de Kanter, F. J. J.;<br />

Kraakman, P. A.; Turkenburg, L. A. M.;<br />

Koolhaas, W. E.; de Wolf, W. H.;<br />

Bickel haupt, F.; Tobe, Y.; Kakiuchi, K.;<br />

Odeira, Y. J. Am. Chem. Soc. 1985, 107,<br />

3716.<br />

82 Tobe, Y.; Kaneda, T.; Kakiuchi, K.;<br />

Odeira, Y. Chem. Lett. 1985, 1301.<br />

83 Kostermans, G. B. M.; de Wolf, W. H.;<br />

Bickelhaupt, F. Tetrahedron Lett. 1986,<br />

27, 1095; cf. Kostermans, G. B. M.;<br />

de Wolf, W. H.; Bickelhaupt, F.<br />

Tetrahedron, 1987, 43, 2955.<br />

84 Kostermans, G. B. M.; Bobeldijk, M.;<br />

de Wolf, W. H.; Bickelhaupt, F.<br />

J. Am. Chem. Soc. 1987, 109, 2471.<br />

85 Tsuji, T.; Nishida, S. J. Am. Chem. Soc.<br />

1988, 110, 2157; cf. Tsuji, T.; Nishida, S.<br />

J. Chem. Soc. Chem. Comun. 1987, 1189.<br />

86 Bockisch, F.; Rayez, J. C.; Liotard, D.;<br />

Duguay, B. Theochem. 1993, 103, 75;<br />

Bockisch, F.; Rayez, J. C.; Liotard, D.;<br />

Duguay, B. J. Comp. Chem. 1992, 13,<br />

1047.<br />

87 Bickelhaupt, F.; de Wolf, W. H. Recl.<br />

Trav. Chim. Pays-Bas, 1988, 107, 459.<br />

88 Marchesini, A.; Bradamante, S.;<br />

Fusco, R.; Pagani, P. Tetrahedron Lett.<br />

1971, 671; Kang, G. J.; Chan, T. H.<br />

J. Org. Chem. 1985, 50, 452;<br />

Lorenzi-Riatsch, A.; Wälchli, R.;<br />

Hesse, M. Helv. Chim. Acta, 1985, 68,<br />

2177; cf. Nelson, P. H.; Nelson, J. T.<br />

Synthesis, 1991, 192.<br />

89 Vögtle, F.; Koo Tze Mew, P. Angew.<br />

Chem. Int. Ed. Engl. 1978, 17, 60;<br />

Higuchi, H.; Misumi, S. Tetrahedron<br />

Lett. 1982, 23, 5571; Higuchi, H.;<br />

References<br />

Kugimiya, M.; Otsubo, T.; Sakata, Y.;<br />

Misumi, S. Tetrahedron Lett. 1983,<br />

24, 2593; Higuchi, H.; Tani, K.;<br />

Otsubo, T.; Sakata, Y.; Misumi, S.<br />

Bull. Chem. Soc. Jpn. 1987, 60, 4027;<br />

Hubert, A. J.; Dale, J. J. Chem. Soc.<br />

1963, 86; Tamao, K.; Kodama, S.;<br />

Nakatsuke, T.; Kiso, Y.; Kumada, M.<br />

J. Am. Chem. Soc. 1975, 97, 4405;<br />

Bates, R. B.; Camou, F. A.; Kane, V. V.;<br />

Mishra, P. K.; Suvannachut, K.;<br />

White, J. J. J. Org. Chem. 1989, 54, 311;<br />

Bates, R. B.; Gangwar, S.; Kane, V. V.;<br />

Suvannachut, K.; Taylor, S. R. J. Org.<br />

Chem. 1991, 56, 1696.<br />

90 Bates, R. B.; Ogle, C. A. J. Org. Chem.<br />

1982, 47, 3949.<br />

91 Noble, K. L.; Hopf, H.; Jones, M., Jr.;<br />

Kammula, S. L. Angew. Chem. Int. Ed.<br />

Engl. 1978, 17, 602.<br />

92 Hirano, S.; Hara, H.; Hiyama, T.;<br />

Fujita, S.; Nozaki, H. Tetrahedron, 1975,<br />

31, 2219.<br />

93 Turkenburg, L. A. M.; Blok, P. M. L.;<br />

de Wolf, W. H.; Bickelhaupt, F.<br />

Tetrahedron Lett. 1981, 22, 3317;<br />

Jenneskens, L. W.; De Kanter, F. J. J.;<br />

Turkenburg, L. A. M.; De Boer, H. J. R.;<br />

de Wolf, W. H.; Bickelhaupt, F.<br />

Tetrahedron, 1984, 21, 4401.<br />

94 Kostermans, G. B. M.; Hogenbirk, M.;<br />

Turkenburg, L. A. M.; de Wolf, W. H.;<br />

Bickelhaupt, F. J. Am. Chem. Soc. 1987,<br />

109, 2855; cf. Kostermans, G. B. M.;<br />

van Dansink, P.; de Wolf, W. H.;<br />

Bickelhaupt, F. J. Org. Chem. 1988, 53,<br />

4531.<br />

95 van Eis, M. J.; de Wolf, W. H.;<br />

Bickelhaupt, F.; Boese, R. J. Chem. Soc.,<br />

Perkin Trans. 2, 2000, 793.<br />

96 Tsuji, T.; Okuyama, M.; Ohkita, M.;<br />

Imai, T.; Suzuki, T. J. Chem. Soc. Chem.<br />

Commun. 1997, 2151.<br />

97 Nakazaki, M.; Yamamoto, K.<br />

Chem. & Ind. 1965, 468; Dehne, H.<br />

Z. Chem. 1969, 9, 308; Dehne, H.;<br />

Zschunke, A. Z. Chem. 1969, 9, 342;<br />

Ishizu, K.; Nemoto, F.; Hasegawa, H.;<br />

Yamamoto, K.; Nakazaki, M. Bull. Chem.<br />

Soc. Jpn. 1973, 46, 140.<br />

98 Tsuji, T.; Ohkita, M.; Konno, T.;<br />

Nishida, S. J. Am. Chem. Soc. 1997, 119,<br />

8425; cf. Tsuji, T.; Ohkita, M.; Nishida, S.<br />

J. Am. Chem. Soc. 1993, 115, 5284.<br />

197


198 4 <strong>Strained</strong> Aromatic Molecules<br />

99 Buchholz, A.; de Meijere, A. Synlett,<br />

1993, 253.<br />

100 Staab, H. A.; Ruland, A.; Kuo-chen, C.<br />

Chem. Ber. 1982, 115, 1755; Staab, H. A.;<br />

Kuo-chen, C.; Ruland, A. Chem. Ber.<br />

1982, 115, 1765.<br />

101 Staab, H. A.; Alt, R. Chem. Ber. 1984,<br />

117, 850.<br />

102 Ruland, A.; Staab, H. A. Chem. Ber. 1978,<br />

111, 2997.<br />

103 Kawai, H.; Suzuki, T.; Ohkita, M.;<br />

Tsuji, T. Angew. Chem. Int. Ed. 1998, 37,<br />

817–819.<br />

104 Cram, D. J.; Steinberg, H. J. J. Am.<br />

Chem. Soc. 1951, 73, 5691.<br />

105 Brown, C. J.; Farthing, A. C. Nature,<br />

1949, 164, 915.<br />

106 Winberg, H. E.; Fawcett, F. S. Org.<br />

Synth. Coll. Vol. V, J. Wiley and Sons,<br />

New York, 1973, 883.<br />

107 Kleinschroth, H.; Hopf, H.; Böhm, I.<br />

Org. Synth. 1981, 60, 41; cf. Hopf, H.;<br />

Lenich, F. T. Chem. Ber. 1974, 107, 1891.<br />

108 Hope, H.; Bernstein, J.; Trueblood, K. N.<br />

Acta Cryst., Section B: Structural<br />

Crystallog. and Cryst. Chem. 1972, 28,<br />

1733.<br />

109 Stalke, D., private communication to<br />

Hopf, H., Feb. 19, 2007.<br />

110 Gantzel, P. K.; Trueblood, K. N. Acta<br />

Crystallogr. 1965, 18, 958.<br />

111 Jones, P. G.; Hopf, H.; Pechlivanidis, Z.;<br />

Boese, R. Z. Krist. 1994, 209, 673.<br />

112 Jones, P. G.; Pechlivanidis, Z.; Hopf, Z.<br />

Z. Naturforsch. 1989, 44b, 860.<br />

113 Murad, A. F.; Kleinschroth, J.; Hopf, H.<br />

Angew. Chem. Int. Ed. Engl. 1980, 19, 389;<br />

Turkenburg, L. A. M.; Blok, P. M. L.;<br />

de Wolf, W. H.; Bickelhaupt, F. Angew.<br />

Chem. Int. Ed. Engl. 1982, 21, 298.<br />

114 Hopf, H.; Savinsky, R.; Jones, P. G.;<br />

Dix, I.; Ahrens, B. Liebigs Ann.Chem.<br />

1997, 1499.<br />

115 Kraakman, P. A.; Valk, J.-M.;<br />

Niederländer, H. A. G.;<br />

Brouwer, D. B. E.; Bickelhaupt, F. M.;<br />

de Wolf, W. H.; Bickelhaupt, F.;<br />

Stam, C. H. J. Am. Chem. Soc. 1990, 112,<br />

6638.<br />

116 de Meijere, A.; Kozhushkov, S. I.;<br />

Rauch, K.; Schill, H.; Verevkin, S. P.;<br />

Kümmerlin, M.; Beckhaus, H.-D.;<br />

Rüchardt, C.; Yufit, D. S. J. Am. Chem.<br />

Soc. 2003, 125, 15110.<br />

117 Shaik, S. S.; Hiberty, P. C.;<br />

Ohanessian, G.; Lefour, J.-M. J. Phys.<br />

Chem. 1988, 92, 5086.<br />

118 Smith, B. H. Bridged Aromatic<br />

Compounds; Academic Press, New York,<br />

1964.<br />

119 Mitchell, R. H. Nuclear magnetic<br />

resonance properties and conformational<br />

behavior of cyclophanes,<br />

in Cyclophanes, Vol. 1, Keehn, P. M.;<br />

Rosenfeld, S. M. (eds.), Academic Press,<br />

New York, 1983, pp. 239.<br />

120 Ernst, L. Prog. NMR Spectrosc. 2000, 37,<br />

47.<br />

121 Ernst, L.; Ibrom, K. NMR spectra of<br />

cyclophanes, in Modern Cyclophane<br />

Chemistry, Gleiter, R.; Hopf, H. (eds.),<br />

Wiley-VCH, Weinheim, 2004, pp. 381.<br />

122 Ernst, L. Annu. Rep. NMR Spectrosc.<br />

2007, 60, 77.<br />

123 Pople, J. A. J. Chem. Phys. 1956, 24,<br />

1111.<br />

124 Waugh, J. S.; Fessenden, R. W. J. Am.<br />

Chem. Soc. 1957, 79, 846.<br />

125 Tobe, Y.; Takahashi, T.; Ishikawa, T.;<br />

Yoshimura, M.; Suwa, M.; Kobiro, K.;<br />

Kakiuchi, K.; Gleiter, R. J. Am. Chem.<br />

Soc. 1990, 112, 8889.<br />

126 Ernst, L.; Boekelheide, V.; Hopf, H.<br />

Magn. Reson. Chem. 1993, 31, 669.<br />

127 Pascal, R. A., Jr.; Grossman, R. B.;<br />

Van Engen, D. J. Am. Chem. Soc. 1987,<br />

109, 6878.<br />

128 Urbano, A. Angew. Chem. Int. Ed. 2003,<br />

42, 3986.<br />

129 Hopf, H. Classics in Hydrocarbon<br />

Chemistry: Syntheses, Concepts,<br />

Perspectives; VCH: Weinheim, 2000,<br />

Chapter 12.2, p. 323.<br />

130 Katz, T. J. Angew. Chem. Int. Ed. 2000,<br />

39, 1921.<br />

131 Oremek, G.; Seiffert, U.; Janecka, A.<br />

Chem.-Ztg. 1987, 111, 69.<br />

132 Vögtle, F. Fascinating Molecules in<br />

Organic Chemistry; Wiley: New York,<br />

1992, p. 156.<br />

133 Meurer, K. P.; Vögtle, F. Top. Curr. Chem.<br />

1985, 127, 1.<br />

134 Laarhoven, W. H.; Prinsen, W. J. C.<br />

Top. Curr. Chem. 1984, 125, 63.<br />

135 Martin, R. H. Anqew. Chem. Int. Ed.<br />

1974, 13, 649.<br />

136 Newman, M. S.; Lednicer, D.<br />

J. Am. Chem. Soc. 1956, 78, 4765.


137 Floyd, A. J.; Dyke, S. F.; Ward, S. E.<br />

Chem. Rev. 1976, 76, 509.<br />

138 Mallory, F. B.; Mallory, C. W. Organic<br />

Reactions, Vol. 30; Wiley & Sons:<br />

New York, 1984, p. 1.<br />

139 Laarhoven, W. H. In: Organic Photochemistry,<br />

Vol. 10, Padwa, A., Ed;<br />

Marcel Dekker: New York, 1989, p. 163.<br />

140 Flammang-Barbieux, M.; Nasielski, J.;<br />

Martin, R. H. Tetrahedron Lett. 1967, 8,<br />

743.<br />

141 Liu, L.; Katz, T. J. Tetrahedron Lett 1991,<br />

32, 6831.<br />

142 Liu, L.; Yang, B.; Katz, T. J.; Poindexter,<br />

M. K. J. Org. Chem. 1991, 56, 3769.<br />

143 Martin, R. H.; Marchant, M.-J.; Baes, M.<br />

Helv. Chim. Acta 1971, 54, 358.<br />

144 Moradpour, A.; Nicoud, J. F.;<br />

Balavoise, G.; Kagan, H.; Tsoucaris, G.<br />

J. Am. Chem. Soc. 1971, 93, 2353.<br />

145 Moradpour, A.; Kagan, H.; Baes, M.;<br />

Morren, G.; Martin, R. H. Tetrahedron<br />

1975, 31, 2139.<br />

146 Martin, R. H.; Flammang-Barbieux, M.;<br />

Cosyn, J. P.; Gelbcke, M. Tetrahedron<br />

Lett. 1968, 9, 3507.<br />

147 Martin, R. H.; Schurter, J. J. Tetrahedron<br />

1972, 28, 1749.<br />

148 Martin, R. H.; Libert, V. J. Chem. Res.<br />

Miniprint 1980, 4, 1940.<br />

149 Martin, R. H.; Baes, M. Tetrahedron<br />

1975, 31, 2135.<br />

150 Martin, R. H.; Morren, G.; Schurter, J. J.<br />

Tetrahedron Lett. 1969, 10, 3683.<br />

151 Kaupp, G. In: Methoden der Organischen<br />

Chemie (Houben-Weyl), Vol. IV/5a;<br />

Müller, E., Ed.; Thieme Verlag, Stuttgart,<br />

1975, p. 326.<br />

152 Willmore, N. D.; Liu, L. B.; Katz, T. J.<br />

Angew. Chem. Int. Ed. 1992, 31, 1093.<br />

153 Blackburn, E. V.; Timmons, C. J. Quart.<br />

Rev. Chem. Soc. 1969, 23, 482.<br />

154 Minuti, L.; Taticchi, A.; Marrocchi, A.;<br />

Gacs-Baitz, E.; Galeazzi, R. Eur. J. Org.<br />

Chem. 1999, 3155.<br />

155 Carreño, M. C.; Garcia-Cerrada, S.;<br />

Urbano, A. J. Am. Chem. Soc. 2001, 123,<br />

7929.<br />

156 Dreher, S. D.; Katz, T. J.; Lam, K. C.;<br />

Rheingold, A. L. J. Org. Chem. 2000, 65,<br />

815.<br />

157 Katz, T. J.; Liu, L. B.; Willmore, N. D.;<br />

Fox, J. M.; Rheingold, A. L.; Shi, S. H.;<br />

Nuckolls, C.; Rickman, B. H. J. Am.<br />

Chem. Soc. 1997, 119, 10054.<br />

References<br />

158 Fox, J. M.; Goldberg, N. R.; Katz, T. J.<br />

J. Org. Chem. 1998, 63, 7456.<br />

159 Paruch, K.; Katz, T. J.; Incarvito, C.;<br />

Lam, K. C.; Rhatigan, B.;<br />

Rheingold, A. L. J. Org. Chem. 2000, 65,<br />

7602.<br />

160 Dreher, S. D.; Weix, D. J.; Katz, T. J.<br />

J. Org. Chem. 1999, 64, 3671.<br />

161 Nuckolls, C.; Katz, T. J.; Katz, G.;<br />

Collings, P. J.; Castellanos, L. J. Am.<br />

Chem. Soc. 1999, 121, 79.<br />

162 Stará, I. G.; Starý, I.; Kollárovi�, A.;<br />

Teplý, F.; Vysko�il, Š.; Šaman, D.:<br />

Tetrahedron Lett. 1999, 40, 1993.<br />

163 Teplý, F.; Stará, I. G.; Starý, I.;<br />

Kollárovi�, A.; Šaman, D.; Rulíšek, L.;<br />

Fiedler, P. J. Am. Chem. Soc. 2002, 124,<br />

9175.<br />

164 Teplý, F.; Stará, I. G.; Starý, I.;<br />

Kollárovi�, A.; Šaman, D.; Fiedler, P.;<br />

Vysko�il, Š. J. Org. Chem. 2003, 68, 5193.<br />

165 Teplý, F.; Stará, I. G.; Starý, I.;<br />

Kollárovi�, A.; Luštinec, D.; Krausová, Z.;<br />

Šaman, D.; Fiedler, P. Eur. J. Org. Chem.<br />

2007, 4244.<br />

166 Bestmann, H. J.; Both, W. Angew. Chem.<br />

1972, 84, 293.<br />

167 Stará, I. G.; Starý, I.; Tichý, M.;<br />

Závada, J.; Hanuš, V. J. Am. Chem. Soc.<br />

1994, 116, 5084.<br />

168 Gingras, M.; Dubois, F. Tetrahedron Lett.<br />

1999, 40, 1309.<br />

169 Dubois, F.; Gingras, M. Tetrahedron Lett.<br />

1998, 39, 5039.<br />

170 Collins, S. K.; Grandbois, A.;<br />

Vachon, M. P.; Côté, J. Angew. Chem. Int.<br />

Ed. 2006, 45, 2923.<br />

171 Harrowven, D. C.; Guy, I. L.; Nanson, L.<br />

Angew. Chem. Int. Ed. 2006, 45, 2242.<br />

172 Bürgi, T.; Urakawa, A.; Behzadi, B.;<br />

Ernst, K. H.; Baiker, A. New J. Chem.<br />

2004, 28, 332.<br />

173 Reetz, M. T.; Sostmann, S. Tetrahedron<br />

2001, 57, 2515.<br />

174 Alexandrová, Z.; Sehnal, P.; Stará, I. G.;<br />

Starý, I.; Šaman, D.; Urquhart, S. G.;<br />

Otero, E. Collect. Czech. Chem. Commun.<br />

2006, 71, 1256.<br />

175 Thongpanchang, T.; Paruch, K.;<br />

Katz, T. J.; Rheingold, A. L.; Lam, K. C.;<br />

Liablesands, L. J. Org. Chem. 2000, 65,<br />

1850.<br />

176 Vanest, J.; Martin, R. H. Recl. Trav. Chim.<br />

Pays-Bas 1979, 98, 113.<br />

199


200 4 <strong>Strained</strong> Aromatic Molecules<br />

177 Katz, T. J.; Sudhakar, A.; Teasley, M. F.;<br />

Gilbert, A. M.; Geiger, W. E.; Robben,<br />

M. P.; Wuensch, M.; Ward, M. D. J. Am.<br />

Chem. Soc. 1993, 115, 3182.<br />

178 Carreño, M. C.; Garcia-Cerrada, S.;<br />

Urbano, A. Chem. Commun. 2002, 1412.<br />

179 Carreño, M. C.; Garcia-Cerrada, S.;<br />

Urbano, A. Chem. Eur. J. 2003, 9, 4118.<br />

180 Carreño, M. C.; Gonzales-Lopez, M.;<br />

Urbano, A. Chem. Commun. 2005, 611.<br />

181 Dai, Y. J.; Katz, T. J.; Nichols, D. A.<br />

Angew. Chem. Int. Ed. 1996, 35, 2109.<br />

182 Dai, Y. J.; Katz, T. J. J. Org. Chem. 1997,<br />

62, 1274.<br />

183 Sudhakar, A.; Katz, T. J.; Yang, B.-W.<br />

J. Am. Chem. Soc. 1986, 108, 2790.<br />

184 Katz, T. J.; Sudhakar, A.; Teasley, M. F.;<br />

Gilbert, A. M.; Geiger, W. E.;<br />

Robben, M. P.; Wuensch, M.;<br />

Ward, M. D. J. Am. Chem. Soc. 1993,<br />

115, 3182.<br />

185 Katz, T. J.; Pesti, J. J. Am. Chem. Soc.<br />

1982, 103, 346.<br />

186 Fox, J. M.; Katz, T. J.; Van Elshocht, S.;<br />

Verbiest, T.; Kauranen, M.; Persoons, A.;<br />

Thongpanchang, T.; Krauss, T.; Brus, L.<br />

J. Am. Chem. Soc. 1999, 121, 3453.<br />

187 Fox, J. M.; Lin, D.; Itagaki, Y.; Fujita, T.<br />

J. Org. Chem. 1998, 63, 2031.<br />

188 Laarhoven, W. H.; Veldhuis, R. G. M.<br />

Tetrahedron 1972, 28, 1823.<br />

189 Nakazaki, M.; Yahamoto, K.; Maeda, M.<br />

Chem. Lett. 1980, 1553.<br />

190 Tribout, J.; Martin, R. H.; Dogle, M.;<br />

Wynberg, H. Tetrahedron Lett. 1972, 13,<br />

2839.<br />

191 Laarhoven, W. H.; Cuppen, T. H. J. M.<br />

Rec. Trav. Chim. Pays-Bas 1973, 92,<br />

553.<br />

192 Laarhoven, W. H.; de Jong, M. H. Rec.<br />

Trav. Chim. Pays-Bas 1973, 92, 651.<br />

193 Ben Hassine, B.; Gorsane, M.; Pecher, J.;<br />

Martin, R. H. Bull. Soc. Chim. Belg. 1985,<br />

94, 597.<br />

194 Ben Hassine, B.; Gorsane, M.; Pecher, J.;<br />

Martin, R. H. Bull. Soc. Chim. Belg. 1987,<br />

96, 801.<br />

195 Ben Hassine, B.; Gorsane, M.;<br />

Geerts-Evrard, F.; Pecher, J.;<br />

Martin, R. H.; Castelet, D. Bull. Soc.<br />

Chim. Belg. 1986, 95, 557.<br />

196 Ben Hassine, B.; Gorsane, M.; Pecher, J.;<br />

Martin, R. H. Bull. Soc. Chim. Belg. 1985,<br />

94, 759.<br />

197 Ben Hassine, B.; Gorsane, M.; Pecher, J.;<br />

Martin, R. H. Bull. Soc. Chim. Belg. 1986,<br />

95, 547.<br />

198 Reetz, M. T.; Beuttenmuller, E. W.;<br />

Goddard, R. Tetrahedron Lett. 1997, 38,<br />

3211.<br />

199 Nakano, D.; Yamaguchi, M. Tetrahedron<br />

Lett. 2003, 44, 4969.<br />

200 Reetz, M. T.; Sostmann, S. J. Organomet.<br />

Chem. 2000, 603, 105.<br />

201 Sato, I.; Yamashima, R.; Kadowaki, K.;<br />

Yamamoto, J.; Shibata, T.; Soai, K.<br />

Angew. Chem. Int. Ed. 2001, 40, 1096.<br />

202 Verbiest, T.; Van Elshocht, S.;<br />

Kauranen, M.; Hellemans, L.;<br />

Snauwaert, J.; Nuckolls, C.; Katz, T. J.;<br />

Persoons, A. Science 1998, 282, 913.<br />

203 Lovinger, A. J.; Nuckolls, C.; Katz, T. J.<br />

J. Am. Chem. Soc. 1998, 120, 264.<br />

204 Nuckolls, C.; Katz, T. J.; Verbiest, T.;<br />

Vanelshocht, S.; Kuball, H. G.;<br />

Kiesewalter, S.; Lovinger, A. J.;<br />

Persoons, A. J. Am. Chem. Soc. 1998,<br />

120, 8656.<br />

205 Stone, M. T.; Fox, J. M.; Moore, J. S. Org.<br />

Lett. 2004, 6, 3317.<br />

206 Owens, L.; Thilgen, C.; Diederich, F.;<br />

Knobler, C. B. Helv. Chim. Acta 1993, 76,<br />

2757.<br />

207 Weix, D. J.; Dreher, S. D.; Katz, T. J.<br />

J. Am. Chem. Soc. 2000, 122, 10027.<br />

208 Ernst, K. H.; Neuber, M.; Grunze, M.;<br />

Ellerbeck, U. J. Am. Chem. Soc. 2001,<br />

123, 493.<br />

209 Fasel, R.; Parschau, M.; Ernst, K. H.<br />

Angew. Chem. Int. Ed. 2003, 42, 5178.<br />

210 Fasel, R.; Cossy, A.; Ernst, K. H.;<br />

Baumberger, F.; Greber, T.; Osterwalder,<br />

J. J. Chem. Phys. 2001, 115, 1020.<br />

211 Vander Donckt, E.; Nasielski, J.;<br />

Greenleaf, J. R.; Birks, J. B. Chem. Phys.<br />

Lett. 1968, 2, 409.<br />

212 Sapir, M.; Vander Donckt, E. Chem. Phys.<br />

Lett. 1975, 36, 108.<br />

213 Obenland, S.; Schmidt, W. J. Am. Chem.<br />

Soc. 1975, 97, 6633.<br />

214 Laarhoven, W. H.; Brus, G. J. M.<br />

J. Chem. Soc. B 1971, 1433.<br />

215 Rulíšek, L.; Exner, O.; Cwiklik, L.;<br />

Jungwirth, P.; Starý, I.; Pospíšil, L.;<br />

Havlas, Z. J. Phys. Chem. C 2007, 111,<br />

14948.<br />

216 Kelly, T. R.; De Silva, H.; Silva, R. A.<br />

Nature 1999, 40 (6749), 150.


217 Grimme, S.; Peyerimhoff, S. D.<br />

Chem. Phys. 1996, 204, 411.<br />

218 Janke, R. H.; Haufe, G.;<br />

Würthwein, E.-U.; Borkent, J. H.<br />

J. Am. Chem. Soc. 1996, 118, 6031.<br />

219 Schulman, J. M.; Disch, R. L.<br />

J. Phys. Chem. A 1999, 103, 6669.<br />

220 Treboux, G.; Lapstun, P.; Wu, Z. H.;<br />

Silverbrook, K. Chem. Phys. Lett. 1999,<br />

301, 493.<br />

221 Jalaie, M.; Weatherhead, S.;<br />

Lipkowitz, K. B.; Robertson, D. Electron.<br />

J. Theor. Chem. 1997, 2, 268.<br />

222 Furche, F.; Ahlrichs, R.; Wachsmann, C.;<br />

Weber, E.; Sobanski, A.; Vögtle, F.;<br />

Grimme, S. J. Am. Chem. Soc. 2000, 122,<br />

1717.<br />

223 Hansen, A. E.; Bak, K. L. Enantiomer<br />

1999, 4, 455.<br />

224 Botek, E.; Spassova, M.; Champagne, B.;<br />

Asselberghs, I.; Persoons, A.; Clays, K.<br />

Chem. Phys Lett. 2005, 412, 274.<br />

225 Klundt, I. L. Chem. Rev. 1970, 70,<br />

471–486.<br />

226 Halton, B. Chem. Rev. 1973, 73, 113–126.<br />

227 Perkin, W. H. J. Chem. Soc. 1888, 1–20.<br />

228 Finklestein, H., PhD thesis, University<br />

of Strassbourg, 1909.<br />

229 Halton, B., Chem. Rev. 2003, 103,<br />

1327–1370.<br />

230 Wegem D., Eur. J. Org. Chem. 2001,<br />

849–862; Müller, P. In: Carbocyclic Three-<br />

Membered Ring Compounds, Vol. E 17 d.<br />

(Ed.: de Meijere, A.), Thieme, Stuttgart,<br />

1997, 2865–2948; Müller, P.<br />

In: Adv. Theor. Int. Mols., Vol. 3. (Ed.:<br />

Thummel, R. P.), JAI Press, Greenwich,<br />

CT, 1995, 37–107; Halton, B., Chem.<br />

Rev. 1989, 89, 1161–1185; Billups, W. E.;<br />

Rodin, W. A.; Haley, M. M. Tetrahedron<br />

1988, 44, 1305–1338; Halton, B.;<br />

Stang, P.J. Acc. Chem. Res. 1987, 20,<br />

443–448; Halton, B., Ind. Eng. Chem.,<br />

Prod. Res. Dev. 1980, 19, 349–364.<br />

231 Halton, B.; Stang, P. J. Synlett 1997,<br />

145–158.<br />

232 Halton, B. In: The Chemistry of the Cyclopropyl<br />

Group, Vol. 2. (Ed.: Rappoport, Z.),<br />

Wiley, Chichester, 1995, 707–772.<br />

233 Cava, M. P. Bull. Soc. Chim. Fr. 1959,<br />

1744–1747; Kündig, E. P.; Pache, S. H.<br />

Science of Synthesis 2003, 2, 153–228;<br />

Mehta, G.; Kotha, S. Tetrahedron 2001,<br />

57, 625–659; Sadana, A. K.; Saini, R. K.;<br />

References<br />

Billups, W. E. Chem. Rev. 2003, 103,<br />

1539–6602; Vollhardt, K. P. C., Acc.<br />

Chem. Res. 1977, 10, 1–8.<br />

234 De, S. C.; Dutt, D. N. J. Indian Chem.<br />

Soc. 1930, 7, 537–544.<br />

235 Halton, B.; Harrison, S. A. R.; Spangler,<br />

C. W. Aust. J. Chem. 1975, 28, 681–685.<br />

236 Mustafa, A.; Kamel, M. J. Am. Chem.<br />

Soc. 1953, 75, 2939–2941.<br />

237 Jones, G. W.; Kerur, D. R.; Yamazaki, T.;<br />

Shechter, H.; Woolhouse, A. D.;<br />

Halton, B. J. Org. Chem. 1974, 39,<br />

492–497.<br />

238 Pinkus, A. G.; Klausmeyer, K. K.;<br />

Feazell, R. P.; Tsuji, J. Acta Cryst., Sect. E,<br />

Str.Rpts. Online 2004, E60, o982-o984.<br />

239 Anet, R.; Anet, F. A. L. J. Am. Chem. Soc.<br />

1964, 86, 525–526.<br />

240 Collis, G. E.; Jayatilaka, D.; Wege, D.<br />

Aust. J. Chem. 1997, 50, 505–513.<br />

241 Collis, G. E.; Wege, D., Aust. J. Chem.<br />

2003, 56, 903–908.<br />

242 Huben, K.; Kuberski, S.; Frankowski, A.;<br />

Gebicki, J.; Streith, J. J. Chem. Soc.,<br />

Chem. Commun. 1995, 315–316.<br />

243 Dürr, H.; Klauck, G.; Peters, K.;<br />

v. Schering, H. G. Angew. Chem., Int. Ed.<br />

Engl. 1983, 22, 332–333; Angew. Chem.<br />

Suppl. 1983, 1347–1362.<br />

244 v. Schleyer, P. R.; Jiao, H.;<br />

Sulzbach, H. M.; Schaefer III, H. F.<br />

J. Am. Chem. Soc. 1996, 118, 2093–2094.<br />

245 Dinadayalane, T. C.; Deepa, S.;<br />

Sastry, G. N. Tetrahedron Lett. 2003, 44,<br />

4527–4529.<br />

246 Kiesewetter, M. K.; Reiter, R. C.;<br />

Stevenson, C. D. J. Am. Chem. Soc. 2005,<br />

127, 1118–1119.<br />

247 Adcock, W.; Gupta, B. D.; Khor, T. C.;<br />

Doddrell, D.; Jordan, D.; Kitching, W.<br />

J. Am. Chem. Soc. 1974, 96, 1595–1597.<br />

248 Günther, H.; Seel, H. Org. Mag. Reson.<br />

1976, 8, 299–300.<br />

249 Sato, T.; Niino, H.; Yabe, Y. J. Am. Chem.<br />

Soc. 2003, 125, 11936–11941.<br />

250 Billups, W. E.; Luo, W.; Wagner, R.;<br />

Hopf, H.; König, B.; Psiorz, M.<br />

Tetrahedron 1999, 55, 10893–10898;<br />

Billups, W. E.; Luo, W.; McCord, D.;<br />

Wagner, R. Pure Appl. Chem. 1996, 68,<br />

275–280.<br />

251 Billups, W. E.; Luo, W.; Harmon, D.;<br />

McCord, D.; Wagner, R. Tetrahedron Lett.<br />

1997, 38, 4533–4534.<br />

201


202 4 <strong>Strained</strong> Aromatic Molecules<br />

252 Billups, W. E.; Blakeney, A. J.;<br />

Chow, W. Y. J. Chem. Soc., Chem.<br />

Commun. 1971, 1461–1462.<br />

253 Halton, B.; Randall, C. J.; Stang, P. J.<br />

J. Am. Chem. Soc. 1984, 106, 6108–6110.<br />

254 Halton, B.; Dixon, G. M. Org. Biomol.<br />

Chem. 2004, 2, 3139–3149.<br />

255 Halton, B.; Jones, C. S. Eur. J. Org.<br />

Chem. 2004, 138–146.<br />

256 Dixon, G. M.; Halton, B. Eur. J. Org.<br />

Chem. 2004, 3707–3713.<br />

257 Halton, B.; Cooney, M. J. Aust. J. Chem.<br />

2006, 59, 118–122.<br />

258 Halton, B.; Boese, R.; Dixon, G. M. Eur.<br />

J. Org. Chem. 2003, 4507–4512.<br />

259 Halton, B.; Cooney, M. J.; Jones, C. S.;<br />

Boese, R.; Bläser, D. Org. Lett. 2004, 6,<br />

4017–4020.<br />

260 Halton, B.; Dixon, G. M.; Jones, C. S.;<br />

Parkin, C. T.; Veedu, R. N.;<br />

Bornemann, H.; Wentrup, C. Org. Lett.<br />

2005, 7, 949–952.<br />

261 Halton, B.; Hügel, H. M.; Kelly, D. P.;<br />

Muller, P.; Bürger, U. J. Chem. Soc.,<br />

Perkin Trans. 2 1976, 258–263.<br />

262 Apeloig, Y.; Arad, D. J. Am. Chem.<br />

Soc. 1986, 108, 3241–3247; Apeloig, Y.;<br />

Arad, D.; Halton, B.; Randall, C. J.<br />

J. Am. Chem. Soc. 1986, 108, 4932–4937;<br />

Brogli, F.; Giovanni, E.; Heilbronner, E.;<br />

Schurter, R. Chem. Ber. 1973, 106,<br />

961–969.<br />

263 Brinker, U. H.; Wüster, H. Tetrahedron<br />

Lett. 1991, 32, 593–596; Halton, B.;<br />

Russell, S. G. G. Aust. J. Chem. 1990, 43,<br />

2099–2105.<br />

264 Saito, K.; Ito, N.; Ando, S. Heterocycles<br />

2002, 56, 59–68; Saito, K.; Ono, K.;<br />

Ito, N.; Tada, N.; Ando, S. Heterocycles<br />

2002, 57, 235–240; Saito, K.; Ono, K.;<br />

Ohkita, M.; Fukaya, M.; Ono, K.;<br />

Kondo, Y. Heterocycles 2003, 60, 773–778;<br />

Saito, K.; Ohzone, Y.; Kondo, Y.; Ono, K.;<br />

Ohkita, M. Heterocycles 2004, 62,<br />

773–778.<br />

265 Saito, K.; Hagari, S.; Ito, N.;<br />

Kuriyama, T.; Ono, K.; Ohkita, M.<br />

Heterocycles 2003, 61, 281–286.<br />

266 Saito, S.; Ando, S.; Kondo, Y. Heterocycles<br />

2000, 53, 2601–2606.<br />

267 Murata, S.; Kobayashi, J.; Kongau, C.;<br />

Miyata, M.; Matsushita, T.; Tomioka, H.<br />

J. Org. Chem. 2000, 65, 6082–6092 and<br />

refs. cited.<br />

268 Halton, B. Pure Appl. Chem. 1990, 62,<br />

541–546.<br />

269 Halton, B. Tetrahedron Lett. 2006, 47,<br />

1077–1079.<br />

270 Wentrup, C.; Wentrup-Byrne, E.;<br />

Müller, P. J. Chem. Soc., Chem.<br />

Commun. 1977, 210–211; Wentrup, C.;<br />

Wentrup-Byrne, E.; Müller, P.; Becker, J.<br />

Tetrahedron Lett. 1979, 4249–4252.<br />

271 Campbell, N.; Wang, H. J. Chem. Soc.<br />

1949, 1513.<br />

272 Scott, L. T.; Roelofs, N. H. J. Am. Chem.<br />

Soc. 1987, 109, 5461–5465.<br />

273 Billups, W. E.; McCord, D. J.;<br />

Maughon, B. R. Tetrahedron Lett. 1994,<br />

35, 4493–4496, J. Am. Chem. Soc. 1994,<br />

116, 8831–8832.<br />

274 Halton, B.; Dixon, G. M.; Forman, G. S.<br />

ARKIVOC 2006, 2006, 38–45.<br />

275 Bambal, R.; Fritz, H.; Rihs, G.;<br />

Tschamber, T.; Streith, J. Angew. Chem.,<br />

Int. Ed. Engl. 1987, 26, 668–669.<br />

276 Yranzo, G. I.; Elguero, J. E.;<br />

Flammang, R.; Wentrup, C. Eur. J. Org.<br />

Chem. 2001, 2209–2220 and refs. cited.<br />

277 Schulz, R.; Schweig, A. Tetrahedron Lett.<br />

1979, 59–62, Tetrahedron Lett. 1984, 25,<br />

2337–2340.<br />

278 Kaiser, R. I.; Bettinger, H. F. Angew.<br />

Chem., Int. Ed. 2002, 41, 2350–2353;<br />

Bettinger, H. J. Am. Chem. Soc. 2006,<br />

128, 2534–2535, Chem. Comm. 2005,<br />

2756–2757.<br />

279 Maksic, Z. B.; Eckert-Maksic, M.;<br />

Pfeifer, K. H. J. Mol. Struct. 1993, 300,<br />

445–453.<br />

280 Sasamori, S.; Tokitoh, N. Organometallics<br />

2006, 25, 3522–3532.<br />

281 Tajima, T.; Hatano, H.; Sasaki, T.;<br />

Sasamori, T.; Takeda, N.; Tokitoh, N.;<br />

Takagi, N.; Nagase, S. J. Organometal.<br />

Chem. 2003, 686, 118–126.<br />

282 Tokitoh, N. Phosphorus, Sulfur Silicon<br />

Relat. Elem. 2001, 168–169, 31–40;<br />

Hatano, K.; Tokitoh, N.; Takagi, N.;<br />

Nagase, S. J. Am. Chem. Soc. 2000, 122,<br />

4829–4830.<br />

283 Tokitoh, N.; Hatano, K.; Sasaki, T.;<br />

Sasamori, T.; Takeda, N.; Takagi, N.;<br />

Nagase, S. Organometallics 2002, 21,<br />

4309–4311.<br />

284 Tajima, T.; Hatano, K.; Sasaki, T.;<br />

Sasamori, T.; Takeda, N.; Tokitoh, N.<br />

Chem. Lett. 2003, 32, 220–221.


285 Tajima, T.; Sasamori, T.; Takeda, N.;<br />

Tokitoh, N.; Yoshida, K.; Nakahara, M.<br />

Organometallics 2006, 25, 230–235.<br />

286 Halton, B.; Buckland, S. J.; Lu, Q.;<br />

Mei, Q.; Stang, P. J. J. Org. Chem. 1988,<br />

53, 2418–2422.<br />

287 Halton, B.; Boese, R.; Dixon, G. M.<br />

Tetrahedron Lett. 2004, 45, 2167–2170.<br />

288 Apeloig, Y.; Boese, R.; Bläser, D.;<br />

Halton, B.; Maulitz, A. H. J. Am. Chem.<br />

Soc. 1998, 120, 10147–10153.<br />

289 Halton, B.; Lu, Q.; Stang, P. J.<br />

J. Org. Chem. 1990, 55, 3056–3060.<br />

290 Halton, B.; Ward, J. M. Aust. J. Chem.<br />

2005, 58, 137–142.<br />

291 Burgi, H.-B.; Baldridge, K. K.;<br />

Hardcastle, K.; Frank, N. L.; Gantzel, P.;<br />

Siegel, J. S.; Ziller, Z. Angew. Chem., Int.<br />

Ed. Engl. 1995, 34, 1454–1456.<br />

292 Frank, N. L.; Baldridge, K. K.;<br />

Gantzel, P.; Siegel, J. S. Tetrahedron Lett.<br />

1995, 36, 4389–4392; Rathore, R.; Lindeman,<br />

S. V.; Kumar, A. J.; Kochi, J. K.<br />

J. Am. Chem. Soc. 1998, 120, 6012–6018.<br />

293 Cohrs, C.; Reuchlein, H.; Musch, P. W.;<br />

Selinka, C.; Walfort, B.; Stalke, D.;<br />

Christl, M. Eur. J. Org. Chem. 2003,<br />

901–906.<br />

294 Neidlein, R.; Christen, D.; Poignée, V.;<br />

Boese, R.; Bläser, D.; Gieren, A.;<br />

Ruiz-Pérez, C.; Hüber, T. Angew. Chem.,<br />

Int. Ed. Engl. 1988, 27, 294–295.<br />

References<br />

295 Stanger, A.; Ben-Mergui, N.; Perl, S. Eur.<br />

J. Org. Chem. 2003, 2709–2712 and refs.<br />

cited.<br />

296 Bachrach, S. M. J. Organometal. Chem.<br />

2002, 643, 39–46.<br />

297 Mo, O.; Yanzek, M.; Eckert-Maksic, M.;<br />

Maksic, Z. B. J. Org. Chem. 1995, 60,<br />

1638–1646.<br />

298 Havenith, R. W. A.; Jenneskens, L. W.;<br />

Fowler, P. W.; Soncini, A. Org. Biomol.<br />

Chem. 2004, 2, 1281–1286; Soncini, A.;<br />

Fowler, P. W.; Jenneskens, L. W.<br />

In: Structure and Bonding (Intermolecular<br />

Forces and Clusters 1), Vol. 115. (Ed.,<br />

Springer GmbH, Berlin, Germany)<br />

2005, 57–79; Havenith, R. W. A.;<br />

Fowler, P. W.; Jenneskens, L. W.<br />

Organic Letters 2006, 8, 1255–1258 and<br />

references cited.<br />

299 Moore, L.; Lubinski, R.; Baschky, M. C.;<br />

Dahlke, G. D.; Hare, M.; Arrowood, T.;<br />

Glasovac, Z.; Eckert-Maksic, M.;<br />

Kass, S. R. J. Org. Chem. 1997, 62,<br />

7390–7396.<br />

300 Eckert-Maksic, M.; Glasovac, Z.<br />

Pure Appl. Chem. 2005, 77,<br />

1835–1850.<br />

301 Antol, I.; Glasovac, Z.; Hare, M. C.;<br />

Eckert-Maksic, M.; Kass, S. R. Int.<br />

J. Mass Spect. 2003, 222, 11–26.<br />

302 Eckert-Maksic, M.; Glasovac, Z. J. Phys.<br />

Org. Chem. 2005, 18, 763–772.<br />

203


5<br />

Fullerenes<br />

5.1<br />

Introduction<br />

Helena Dodziuk<br />

At the beginning of this chapter the inclusion of fullerenes and nanotubes in<br />

this book has to be explained once more. Of course, speaking formally these<br />

compounds are not hydrocarbons, since in the pure state they consist only of<br />

the carbon atoms. However, their extended systems of nonplanar aromatic rings<br />

provide a model for the analysis of the effect of distortions on physicochemical<br />

properties.<br />

The serendipitous discovery of C 60, 1, resembles a detective story [1]. The<br />

molecule is remarkable for its high symmetry that determines the equivalence<br />

of all carbon atoms. However, the possibility of the mere existence of molecules<br />

of this point group was denied by Herzberg [2], later awarded Nobel Prize in<br />

chemistry for his spectroscopic studies. In his seminal monograph he stated:<br />

‘The regular icosahedron and the regular dodecahedron belong to point group<br />

I h. It is not likely that molecules of such a symmetry will ever be found.’ Then C 60<br />

was independently proposed by two groups of theoreticians [3, 4]. However, the<br />

suggestion was totally forgotten. Interestingly, Rohlfing and coworkers observed<br />

a strong signal corresponding to C 60 in their study of small carbon clusters in<br />

1984 [5]. They even reported the whole mass spectrum with the remarkable C 60<br />

signal but, focused on small structures, they limited their discussion exclusively<br />

to ions smaller than 40. The following year, an observation by Curl, Kroto and<br />

Smalley of a strong C 60 peak (accompanied by a smaller C 70 one) in a carbon soot,<br />

produced by discharges in vacuum that should mimic conditions in the cosmos,<br />

marked the beginning of the fullerene era. Intriguingly, the authors, not knowing<br />

the earlier proposals, had trouble assigning the structure and did not know what<br />

shape it represented. Kroto’s hobby interest in graphics and his fascination for<br />

the geodesic domes of the American architect Buckminster Fuller provided clues,<br />

helped clear up the problem and brought the molecule its name. Then, a phone<br />

call to the Mathe matical Department brought the answer: soccer ball.<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

205


206 5 Fullerenes<br />

To begin with, the available amount of fullerene was very small, preventing proof<br />

of the structure. A purification procedure developed by the Krätschmer group<br />

solved the problem of obtaining bulk amounts of C 60 [6], thus enabling further<br />

studies. Four sharp bands in the IR spectrum of the enriched soot were compatible<br />

with the I h symmetry of the molecule. Kroto learned about the purification<br />

procedure by refereeing the paper. After writing a positive review, he purified a<br />

sample and let it be measured by 13 C NMR [1]. The spectrum consisted of only<br />

one line providing the first proof of the equivalence of all 60 carbon atoms, thus<br />

proving the spatial fullerene structure. Interestingly, the first X-ray measurements<br />

yielded the C 60 cage diameter and the distance between the molecules in the crystal<br />

only since the near-spherical shape allowed practically undisturbed rotation of the<br />

molecules at room temperature. As a result, two different C–C bond lengths were<br />

first, although quite inaccurately, determined by NMR technique [7]. Structural<br />

aspects of fullerene chemistry and chirality were recently summarized by Thilgen<br />

and Diederich [8].<br />

The highly symmetrical structure of 1 seems quite artificial but the molecule<br />

has been found in some minerals (http://webmineral.com/data/Fullerite.shtml)<br />

and in the cosmos [9, 10] and its presence in Nature precluded its patenting. A few<br />

C 60 molecules are even produced by burning dinner candles [11].<br />

All fullerenes are composed of 12 cyclopentane rings and a certain number of<br />

six-membered ones. Most stable fullerenes are known to conform to so called<br />

Isolated Pentagon Rule, IPR, stating that their cyclopentane rings do not have<br />

common atoms or bonds. The smallest carbon cage satisfying this condition is<br />

that of 1 and it represents the most frequently encountered fullerene structure.<br />

The next one, discovered together with C 60 , is 2. In addition, numerous smaller<br />

(not complying with IPR) and higher fullerenes have been reported. Contrary to<br />

1 and 2, for which only one isomer is possible, the higher members of the family<br />

can assume several structures of, sometimes different, symmetry. An Atlas of<br />

Fullerenes [12] collects all possible IPR structures and their symmetries from C 20<br />

to C 100 . The Fullerene Gallery at the address: http://cochem2.tutkie.tut.ac.jp:8000/<br />

Fuller/fsl/fsl.html presents many of them including non-IPR ones. Seven possible<br />

IPR isomers of C 80 are shown in Figure 5.1.<br />

C 20 can be considered as the archetypal fullerene, built of fully unsaturated<br />

five-membered rings, since it is a highly unstable fulleroid structure without any<br />

six-membered rings. Its synthesis was accomplished by the Prinzbach group<br />

[13–15].


Figure 5.1 Seven possible isomers of C 80 with their symmetry according to<br />

An Atlas of Fullerenes [12].<br />

Scheme 5.1<br />

5.1 Introduction<br />

As noticed by J.-M. Lehn [16], the synthesis of C 60 represents a unique example<br />

of the covalent self-assembly in which 90 bonds are formed simultaneously. Efforts<br />

of several groups trying to synthesize the molecule in a series of ‘true chemical<br />

reactions’ resulted in rich corannulene chemistry (Scheme 5.1) [17–19].<br />

The interest in fullerenes was triggered by the observation that other molecules<br />

or ions can be hosted inside their cages. Such supramolecular complexes, which<br />

bear the name endohedral fullerene complexes, were thought to allow one to<br />

propose exciting applications. The first suggestions seemed to be too good to<br />

be true. For instance, Stoddart wrote about a fullerene cage with ‘a door’ which<br />

could serve as a drug carrier releasing the content at a desired place [20]. He also<br />

predicted that extending the superconductivity of 1 to room temperature or using<br />

perfluorinated C 60 F 60 as a lubricant would revolutionize our all lives since we<br />

would avoid losses due to electrical resistance or friction. Later we have learned<br />

207


208 5 Fullerenes<br />

that these hopes (discussed in some details in Chapter 2.4) cannot be fulfilled<br />

[21, 22]. Model calculations have shown that production and purification of<br />

much larger fullerenes must be mastered to allow much larger guests of practical<br />

interest to be hosted [23].<br />

Fullerenes are relatively large objects, difficult to study both experimentally<br />

and computationally. Highly symmetrical C 60 is a kind of exception although it<br />

also poses several problems. Other fullerenes are available in small quantities<br />

and their lower symmetry complicates their study. Only the most important<br />

experimental methods, often combined with discussions of theoretical results,<br />

will be presented in this chapter and several exciting aspects of fullerene research<br />

will not be covered. For instance, endohedral fullerene complexes with metal<br />

ions inside forming untypical salts that can be dissociated only by destroying the<br />

cage, will practically not be discussed here. It should be stressed, however, that<br />

the formation of endohedral complexes with these or other guests can stabilize<br />

otherwise unstable non-IPR fullerenes [24].<br />

Fullerene studies are often triggered by practical considerations. However,<br />

these highly symmetrical systems are of critical importance for basic research. As<br />

shown by thought-inspiring experiments in the Zeilinger group on the obtaining<br />

of diffraction patterns of C 60 and C 70 [25, 26], such purely scientific studies could<br />

eventually lead to marketable devices.<br />

Applications of fullerene will be in detail discussed in Section 5.7. It suffices<br />

to mention here that at present only few have reached the market. Interestingly,<br />

the hopes for the applications were high in the late 1980s and early 1990s. Then<br />

carbon nanotubes took the lead and electronic devices built from one nanotube<br />

were thought to revolutionize our whole life. For instance, nanotube transistor<br />

[27] or flat displays (http://news.com.com/Carbon+nanotube+TV+trials+<br />

on+horizon/2100-1041_3-6051476.htm) were expected to be sold soon. However,<br />

they seem to pose numerous technical problems, resulting in new interest in<br />

fullerene applications. In any case, in spite of very few practical applications of<br />

fullerenes and nanotubes at present, these molecules have given a strong boost to<br />

the development of nanotechnology which sooner or later is destined to change<br />

our everyday life.<br />

5.2<br />

Chemistry Influenced by the Nontypical Structure: Modification of [60]Fullerene<br />

Takuma Hara, Takashi Konno, Yosuke Nakamura and Jun Nishimura<br />

5.2.1<br />

Introduction<br />

As is well known, Kroto, Smally, and Curl worked together with their associates<br />

and during their study on interstellar carbon clusters, serendipitously discovered<br />

a soccer ball compound or C 60 which originally was named buckminsterfullerene


5.2 Chemistry Infl uenced by the Nontypical Structure: Modifi cation of [60]Fullerene<br />

Figure 5.2 Number of papers reported on fullerenes and CNTs.<br />

dedicated to genius architect Richard Buckminster Fuller [28]. Since its large-scale<br />

preparation was developed by Krätschmer and Huffman [29], the chemical world<br />

has been caught by the fullerene fever for many years [30]. Only recently, because<br />

of the prospects of their applications, the fever has passed most of its heat to carbon<br />

nanotubes (CNTs), as shown in Figure 5.2. However, the unusual properties of C 60<br />

and other fullerenes continue to make them fascinating objects for study.<br />

During the hot years, many useful C 60 derivatives were obtained. Most modifications<br />

are of three general types: i.e. those with nucleophilic, radical, and<br />

electrophilic reagents. Since C 60 is an electron acceptor itself, most of the modifications<br />

have been carried out by using nucleophilic, electron-donating species<br />

or radicals.<br />

In this chapter we first overview the fullerene reactions and then summarize the<br />

modifications, dividing them into six subclasses according to the addend structures.<br />

Since, at least theoretically, almost all reactions are applicable to all types<br />

of fullerenes, the reactions of C 60 which is the best studied, most easily available<br />

and also interesting for applications, will mainly be discussed here.<br />

Moreover it should be stressed that within the last 15 years the circumstances<br />

around the fullerene chemistry have been well established except for the safety<br />

issues [31]; i.e. the supply of various fullerenes [32], the IUPAC nomenclature<br />

[33], and many useful HPLC columns for their separations [34], etc. A variety<br />

of findings in C 60 chemistry have surprised a number of audiences not only in<br />

organic, theoretical and general chemistry but even in natural philosophy.<br />

5.2.2<br />

General Overview<br />

Fullerenes have a system of conjugated double bonds and receive various additions<br />

at the sites, although these are not regular double bonds. The bonds are classified<br />

into (5,6) junction (single bond) and (6,6) junction (double bond), which mean<br />

the edges connecting a five-membered ring and a six-membered ring, and two<br />

five-membered rings, respectively [35]. These bonds have different properties, as<br />

shown in Figure 5.3. Many reactions do not usually occur at (5,6) junctions, but<br />

at (6,6) junctions to give 1,2-adducts [36].<br />

209


210 5 Fullerenes<br />

Figure 5.3 Two different bonds: (5,6) junction with 1.45 Å of bond length and (6,6) one<br />

with 1.38 Å [9].<br />

A variety of reaction mechanisms have been suggested for the novel electronaccepting,<br />

large �-conjugating fullerenes, which can be roughly classified as<br />

depicted in Figure 5.4.<br />

Figure 5.4 Reactions classified.


5.2 Chemistry Infl uenced by the Nontypical Structure: Modifi cation of [60]Fullerene<br />

Species 3 are generated in many ways by the reactions of organolithium [37]<br />

including acetylides [37d], Grignard reagents with [30] or without copper salt<br />

[37b,c, 38a–c], and typical nucleophiles [39] such as CN – [40]. It was reported<br />

that the negative charge of a typical t-BuC 60 – anion 3 is not fully delocalized but<br />

rather localized at the adjacent carbon (2-position) of the t-Bu addend (1-position)<br />

[37b]. Grignard reagents react with C 60 more slowly than organolithium, and the<br />

products are 1,4-adducts in contrast with the organolithium cases providing 1,2adducts<br />

[37b,c]. Grignard reagents like the phenyl one together in the presence<br />

of CuBr-Me 2S complex react with the fullerene to afford interesting adducts<br />

having five addends at the 2,5,10,21,24-positions, which have been extensively<br />

applied to various fields, not only organic chemistry but also materials science<br />

[38d–f]. Sodium cyanide and fullerene dissolved in DMF-o-dichlorobenzene mixed<br />

solvent generated carbanion 3, which was trapped by various electrophiles like<br />

p-t-butyl(bromomethyl)benzene to give the appropriate 1,2-adducts [40]. Diethyl<br />

bromomalonate gives 3 through the addition of its ester enolate, which rapidly<br />

eliminates bromide anion to end up with the formation of the corresponding<br />

methanofullerene. This reaction is called the Bingel reaction, dedicated to the<br />

inventor [39]. Dianion C 60 2– formed by the two-electron reduction of C60 can<br />

act as a nucleophile, which reacts with, for example, benzyl bromide to afford<br />

1,4-dibenzylfullerene [41].<br />

Since fullerene reacts with radicals so rapidly, it is often called a radical<br />

sponge [42]. Generally, a radical species adds at the fullerene double bonds and<br />

generates a fullerene monoadduct radical species, which accepts another radical<br />

at the 4-position toward the 1-addend position. Such reactivity is explained by<br />

acknowledging that fullerene does not have a fully conjugated �-system but rather<br />

[5]-radialene substructure which becomes the stable conjugated cyclopentadienyl<br />

radical 10 after the addition of five radicals as shown in Scheme 5.2.<br />

Scheme 5.2 Radical addition to the fullerene with [5]radialene substructure.<br />

211


212 5 Fullerenes<br />

The fullerene does not usually react well with Brønsted acids, but does somehow<br />

with sulfuric acid–nitric acid mixture to afford polyhydroxyfullerenes [43]. Lewisacid<br />

catalyzed transformation was reported to proceed through arenium cation<br />

[44].<br />

The double bond of fullerene generally behaves as an electron-accepting site<br />

(7) such as dienophile or dipolarophile. Therefore, C 60 exhibits many pericyclic<br />

reactions; for example, Diels–Alder reaction, 1,3-dipolar cycloaddition (Huisgen<br />

reaction) such as Prato reaction, and so forth.<br />

There are many synthetic data accumulated so far on several reactions. These<br />

reactions proceed with the interaction of fullerene (acceptor) LUMO and reagent<br />

(donor) HOMO (7) or the reverse combination (8). The bisadduct distribution<br />

seems to provide a method that can qualitatively define whether the reagents<br />

behave as acceptors or donors. Several popular reactions like Bingel [45], Prato<br />

[46], and Diels–Alder [47], which give 3-membered, 5-membered, and 6-membered<br />

rings as the junctions between the fullerene and addends, were reported to have<br />

nucleophilic character, while benzyne [48] and nitrene [49] cycloadditions seem<br />

to be electrophilic.<br />

The multi-addition, especially bisaddition, is one of the interesting fields in<br />

fullerene chemistry. Yet even the bisaddition of a symmetric reagent to C 60 can<br />

give at least 8 isomers: cis-1, cis-2, cis-3, e, trans-4, trans-3, trans-2, and trans-1 (see<br />

Figure 5.5 and Table 5.1). The regioselectivities of bisaddition were thoroughly<br />

examined on the three major fullerene modification reactions mentioned above.<br />

Moreover, several C 2 -symmetric regioisomers, cis-3, trans-3, and trans-2, are chiral,<br />

and their absolute configurations have been determined by CD spectroscopy after<br />

optical resolution by using some elaborate chiral HPLC columns [34].<br />

Generally speaking, the regioisomer distribution of bisadducts was qualitatively<br />

correlated with the frontier orbital coefficients of the corresponding monoadducts;<br />

e and trans-3 bisadducts were formed preferentially among eight isomers [45–49].<br />

The isomer ratios in Bingel, Prato, and Diels–Alder bisadditions are depicted in<br />

Figure 5.6a. The reactive species in these reactions, malonate anions, Huisgen<br />

ylides, and dienes, are nucleophiles in their nature, so that similar regioselectivities<br />

are obtained.<br />

The isomer ratios in benzyne and nitrene bisadditions are shown in Figure 5.6b.<br />

The most remarkable difference between these bis-additions and those shown<br />

in Figure 5.6a is the high ratio of cis-1 in the former. Generally, the cis-1 position<br />

is most stressed by the first addition so that this double bond is apt to be more<br />

reactive and to accept the addition of species more than the other sites. However,<br />

the reactions except for benzyne and nitrene additions suffer from steric hindrance<br />

between the first addend moiety and incoming species, and usually result in a<br />

Table 5.1 Number of isomers with symmetrical achiral addends [50].<br />

Number of additions Mono Bis Tris Tetrakis<br />

Number of isomers 1 8 46 262


5.2 Chemistry Infl uenced by the Nontypical Structure: Modifi cation of [60]Fullerene<br />

Figure 5.5 Eight possible isomers of bisadducts.<br />

negligible amount of cis-1 product. Actually benzyne and nitrene can give flat<br />

addends, while other reactions afford more bulky addends.<br />

The second apparent difference of benzyne addition from others is that it has<br />

less total content of the southern hemisphere products. The same tendency is<br />

observed for the nitrene bisaddition, since the methoxycarbonylnitrene is regarded<br />

as an electrophilic species, although there is some subtle difference between the<br />

two electrophilic additions.<br />

For selective modification, especially for multiadditions, the use of tethers<br />

has been proposed. For example, Diederich’s and Hirsch’s groups developed the<br />

selective Bingel addition toward bisadducts and trisadducts [45c, 51]. Nakamura’s<br />

group successfully obtained a chiral bisadduct having a cyclopentane ring at the<br />

junction [52]. Recently a selective Prato reaction was also achieved using the tether<br />

manipulation [53]. Diels–Alder reaction was made selective at the bisaddition<br />

using the same procedure [54].<br />

213


214 5 Fullerenes<br />

Figure 5.6 Distribution of regioisomeric bisadducts.<br />

Another method to make multi-additions selective was developed by Hirsch’s<br />

group, using a reversible template [55]. Using the reversible Diels–Alder reaction of<br />

9,10-dimethylanthracene as a template, they obtained O h -symmetrically modified<br />

hexakisadducts by Bingel addition, although a recent paper reported that similar<br />

selectivities were obtained without using the anthracene derivative [56].<br />

Returning to Figure 5.4, finally the thermal or photochemical electron transfer<br />

reactions from a donor like alkylamine to the fullerene must be mentioned. At<br />

the very beginning of the fullerene chemistry, the addition of alkylamines was<br />

discovered [57]. It is suggested that this reaction involves the formation of fullerene<br />

radical anion and amine radical cation through single electron transfer (SET) and


5.2 Chemistry Infl uenced by the Nontypical Structure: Modifi cation of [60]Fullerene<br />

the radical coupling of both species to afford the corresponding amine adduct<br />

through the proton transfer [57a]. On the other hand, under photoirradiation<br />

an electron-rich ketene silylacetal 11 reacts with C 60 through SET to give an ene<br />

reaction product as shown in Scheme 5.3 [58]. In this reaction, first the fullerene<br />

absorbs light to excite itself to 1 C 60*, which rapidly and efficiently renders intersystem<br />

crossing to 3 C 60*. Then 3 C 60* gets an electron from the acetal to generate<br />

its anion radical and acetal cation radical, which combine with each other finally<br />

giving product 12.<br />

Scheme 5.3 SET under photoirradiation.<br />

Many reactions follow one of the above reaction types. The stability of adducts<br />

is sometimes an important issue, and depends largely on the reversibility of<br />

modification reactions. Since pristine fullerene itself is thermodynamically stable<br />

enough, most fullerene derivatives reproduce fullerene under harsh conditions,<br />

even though the reactions are not reversible. Bingel adducts are decomposed by<br />

electrolysis [59]. Prato adducts give pristine fullerene when they are treated with a<br />

strong dipolarophile such as maleic anhydride in the presence of a catalyst such as<br />

Wilkinson’s complex or copper triflate in o-dichlorobenzene at reflux for 8 to 18 h<br />

[60]. Diels–Alder adducts easily decompose or usually undergo retro-Diels–Alder<br />

reaction at high temperatures or sometimes even at r.t. (see below) [61].<br />

5.2.3<br />

Modification Reactions<br />

Showing typical products with their structures, a variety of modifications are<br />

introduced in this section, focusing on the joint structures between the fullerene<br />

and addends.<br />

5.2.3.1 Reduction and Oxidation<br />

Many investigations related to the reduction have been reported [62], using direct<br />

hydrogenation at high temperatures, electrochemical reduction, BH 3 -THF,<br />

NH=NH, LiBEt 3 H then treated with HBF 4 , Birch reduction, Zn-HCl, Zn(Cu),<br />

hydrozirconation, H 2 Rh(0)-Al 2 O 3 , Pd/C-catalyzed reduction, and so forth.<br />

Theoretically all 30 double bonds ((6,6) junctions) can accept hydrogen atoms<br />

leading to C 60 H 60 . After receiving a certain number of hydrogen atoms, however,<br />

the remaining double bonds are buried inside the C 60 surface too deeply to react or<br />

the products are too much deformed to stay stable. The theoretical studies indicate<br />

that C 60 H 60 with all CH bonds pointing outwards is much less stable than the<br />

215


216 5 Fullerenes<br />

Figure 5.7 Hydrogenation and oxidation of C 60 .<br />

isomers with the bonds pointing inside the C 60 cage [63] (see Section 2.4.1.7). Until<br />

now the structures of C 60 H 2 , C 60 H 4 , C 60 H 6 , and C 60 H 18 have been well examined<br />

and reported (Figure 5.7). Among the series of C 60 H 24 to C 60 H 36 various reducing<br />

agents under various conditions selectively gave C 60 H 36 , which is revealed as a<br />

mixture of isomers. The structure of hydrogenated fullerenes mainly studied by<br />

NMR spectra is discussed in Section 5.6.<br />

The fullerene is easily oxidized, and even its commercial samples usually contain<br />

the oxide. The reactivity to oxidation was anticipated at the very beginning of<br />

C 60 studies. It is oxidized to form fullerene oxide or epoxide 13. The oxidation is<br />

carried out by electrochemical reaction [64a], irradiation with oxygen [64b], direct<br />

oxidation at high temperatures [64c], oxygen with a radical initiator like AIBN<br />

[64d], ozone [64e], dimethyldioxolane [64f], MCPBA [64g], P450 model system<br />

[64h], and so forth.<br />

Recently Tajima and his co-workers reported an intriguing transformation<br />

of the oxide shown in Scheme 5.4 [65], which may be useful in the materials<br />

science.<br />

Scheme 5.4 Transformation of fullerene oxide.


5.2 Chemistry Infl uenced by the Nontypical Structure: Modifi cation of [60]Fullerene<br />

The nitration of C 60 with NO 2 gave mixtures of tetra- and hexanitro derivatives<br />

in the ratio depending on the reaction time [66]. Then a selective synthetic method<br />

toward hexanitro[60]fullerenes was developed [43c]. The nitroallylic moieties in<br />

hexanitro[60]fullerenes were found to be excellent leaving groups for the nucleophilic<br />

substitution by amines, such as anilines, leading to the formation of<br />

hexaanilino[60]fullerenes. Moreover, hexanitro[60]fullerene was reported to be a<br />

better electron acceptor and form nitrofullerene complexes with aniline, o-, m- and<br />

p-phenylenediamine, pyridine, tacrine, triethylamine, 1-aminoanthraquinone,<br />

N,N-dimethylaniline, TTF, and BEDT-TTF (ET) [67].<br />

The nitrated products partly isomerize to the nitrite esters, which, with subsequent<br />

hydrolysis by atmospheric moisture, give nitrofullerols containing 6–8 nitro<br />

and 7–12 hydroxy groups per fullerene [68]. Hydrolysis of polynitrofullerenes in<br />

aqueous NaOH gave the corresponding polyhydroxylated fullerene derivatives or<br />

fullerols in moderate overall yields. Halogenated fullerenes were used as reagents<br />

to prepare fullerols. Halogens were selectively substituted when fullerene bromides<br />

and chlorides were treated with silver trifluoroacetate [69].<br />

5.2.3.2 Alkylation<br />

Here the alkylations of fullerene are summarized. These addends are formed via<br />

intermediates 3, 4, 5, and 9.<br />

Some alkyl- and ethynyl-fullerenes were produced by the reaction with t-butyl-,<br />

9-fluorenyl-, and phenylethynyl-lithium reagents. Typical reaction products are<br />

listed in Figure 5.8 [37c, 38c, 70].<br />

Figure 5.8 Some derivatives obtained from intermediate 3; 15a [37c], b [70a,b], c [70a], d [70c],<br />

e [70e], f [70d], g [70f], h [70f], i [70f], j [37c], k [37c], l [37c], m [37c], n [37c], o [37c], p [38c].<br />

217


218 5 Fullerenes<br />

Grignard reagents react with fullerene to afford some intriguing products as<br />

shown in Figure 5.8 [38c]. It was reported that the reaction is influenced by solvents<br />

and presence of oxygen in the reaction media. Nakamura and his co-workers<br />

discovered so-called Nakamura reaction that, using Grignard reagent combined<br />

with copper reagents and dimethyl sulfide, affords several penta-adducts around<br />

the corannulene subunit in excellent yields [71]. Recently they were successful in<br />

the first synthesis of beltane derivative or 40�-[10]cyclophenacene 16, which can<br />

be regarded as the smallest unit of (5,5)CNT [72].<br />

5.2.3.3 Cycloadditions<br />

Wudl once suggested that some cycloadditions are the most useful in the modifications<br />

of fullerenes [73]. The following four subsections summarize these useful<br />

cycloaddition reactions.<br />

Three-membered Ring-fused C 60 . A three-membered ring as fullerene-addend linkage<br />

is formed through intermediates, 3, 7, and 8. Among the reactions through 3,<br />

Bingel reaction is the most popular [39], which becomes now one of the most<br />

useful fullerene modification reactions (Scheme 5.5) [73, 74]. Using this reaction,<br />

many fascinating and intriguing derivatives have been synthesized as shown in<br />

Figure 5.9. For example, Nakamura, Nishimura, and their co-workers carried out<br />

Bingel reaction at –78 °C to r.t. and for the first time successfully obtained some<br />

catenanes possessing C 60 surface as the ring component [75]. Other interesting<br />

Bingel products are also depicted in Figure 5.9 [36, 76].<br />

Scheme 5.5 Bingel reaction [39].


5.2 Chemistry Infl uenced by the Nontypical Structure: Modifi cation of [60]Fullerene<br />

Figure 5.9 Fascinating Bingel products 18 [75], 19 [36], and 20 [76].<br />

The Bingel products are apt to render electrochemically induced isomerization<br />

by the migration of cyclopropane rings on the C 60 surface, if the controlled<br />

potential electrolysis is not exhaustive [77]. The isomerization was expressed as<br />

‘electrochemical walk on the sphere’.<br />

Many related reactions have been reported. Monoalkyl malonates react with<br />

iodine in the presence of DBU to afford iodomethanofullerenes 21 [78]. BrCH 2CN<br />

and bromoform also give corresponding methanofullerenes by the treatment of<br />

strong base LDA [79]. The reaction with a special enolate ion gives a derivative 22<br />

[80]. Phosphonium [81] and sulfonium [82] ylides afford a variety of methanofullerenes.<br />

Recently Saigo and his co-workers developed fullerene-acid chloride<br />

23 as a building block of the synthetic chemistry via the sulfonium ylide transformation<br />

[83].<br />

219


220 5 Fullerenes<br />

Scheme 5.6 Stepwise and direct formation of fulleroid and/or methanofullerene.<br />

Diazomethane and related diazo compounds react with fullerene double bonds<br />

and form corresponding pyrazolines, which decompose to give fulleroids with<br />

the loss of nitrogen (see Scheme 5.6) [84]. Often fulleroids readily rearrange into<br />

methanofullerenes under heating or photoirradiation [85].<br />

The reactions of diazo compounds with the fullerene do not always provide<br />

pyrazolines. In some cases, the corresponding carbenes are first generated and<br />

then add to fullerene to give methanofullerenes. Recently Akasaka and his coworkers<br />

examined thoroughly whether the reactions with diazoadamantane and<br />

alkyldiazirines pass through carbenes or pyrazolines. They successfully partitioned<br />

reactions after the analysis of methanofullerene and fulleroid contents [86].<br />

As other three-membered ring formations, aziridine framework is used as the<br />

junction between the fullerene and a functional group. RO-CO-N 3 forms aziridine<br />

ring between fullerene and alkoxy- or aryloxy-carbonyl group (see 27) [87]. On the<br />

other hand, RCH 2 N 3 adds to the fullerene to afford azafulleroid 28 [88], one of<br />

which is derived to aza[60]fullerene as its dimer form 29, only accessible hetero-


5.2 Chemistry Infl uenced by the Nontypical Structure: Modifi cation of [60]Fullerene<br />

fullerene. Recently the chemistry of heterofullerene has been developed by Hirsch’s<br />

group [89]. Silylenes generated in situ add to the fullerene to afford corresponding<br />

silacylopropane product 30 [90].<br />

Four-membered Ring-fused C 60 . The most striking four-membered ring between<br />

fullerene and a partner was disclosed as the form of fullerene-dimer 31 by Komatsu<br />

and his co-workers [91]. This dumbbell type structure is often used as a symbol<br />

of chemical achievement. The dumbbell is formed through 3. Naturally a kind of<br />

[2+2] cycloaddition was expected to occur under photoirradiation. Using supramolecular<br />

and photochemical technique, McClenaghan and his co-workers first<br />

successfully obtained corresponding fullerene dimer [92]. Some 1,3-butadienes<br />

[93a], styrenes [93b], and enones [93c] also gave cyclobutane-annulated fullerenes<br />

under photoirradiation through radical species 5. Under thermal conditions<br />

some reactive acetylenes [94a], dimethyleneketene acetals [94b], ketenes [94c],<br />

and allenamides [94d] made cyclobutene-ring-joints for the fullerene. Besides<br />

the examples mentioned above, there are a few ones, mostly from the reaction of<br />

benzyne [47, 95] as shown in Scheme 5.7.<br />

Scheme 5.7 Four membered-ring formation; 31 [91], 32 [94a], and 33 [47].<br />

221


222 5 Fullerenes<br />

Five-membered Ring-fused C 60 . The five-membered rings can be formed by [3+2]<br />

cycloaddition, where three sp 2 atoms come from reagents and two sp 2 atoms derive<br />

from the fullerene double bond at 1,2-positions. Reagents possessing three-atom<br />

active site are 1,3-dipoles or Huisgen ylides, trimethylene methane diradical,<br />

disilylane, and so forth (Scheme 5.8). The reaction path usually is considered<br />

to be through 5 for photoreactions or 7 for pericyclic reactions of electron-rich<br />

ylides. Here in this section only stable products are dealt; unstable pyrazolines<br />

mentioned above are not included. Moreover, some cases treated above are not<br />

touched below such as dioxacyclopentane system 14.<br />

Scheme 5.8 Prato reaction.<br />

Figure 5.10 Fascinating Prato products; 35 [98], 36 [99], and 37 [100].


5.2 Chemistry Infl uenced by the Nontypical Structure: Modifi cation of [60]Fullerene<br />

Figure 5.11 Some five-membered ring-fused fullerenes with reagents or reactive species;<br />

38 [101], 39 [102], 40 [103], 41 [104], 42 [105], 43 [106], 44 [107], 45 [107], and 46 [107].<br />

The most useful convenient method among the kind of modifications is Prato<br />

reaction, one of pericyclic reactions [96]. It is one of convergent syntheses in which<br />

two high molecular weight substrates can be joined by one step. Actually fullerene,<br />

N-alkylglycine, and an aldehyde, such as formaldehyde or a high molecular weight<br />

aldehyde prepared, are dissolved in toluene and the mixture is heated to reflux.<br />

The procedure is just simple and the yield is usually high to excellent. Therefore it<br />

is so popular in materials science field as well as the fullerene chemistry. Recently<br />

this reaction was applied to make CNTs soluble in conventional organic solvents<br />

such as acetone and chloroform [97].<br />

Fascinating Prato products 35–37 [98–100] are depicted in Figure 5.10. Other<br />

five-membered-ring formations are summarized in Figure 5.11 with reagents or<br />

reactive species [101–107].<br />

Six-membered Ring-fused C 60 . There are some stepwise and concerted [4+2] cyclization<br />

reactions toward six-membered ring-fused fullerenes. The most popular one<br />

of the latter category is Diels–Alder reaction. As discussed above, the Diels–Alder<br />

adducts of fullerene are not always stable enough [61], so that unstable dienes such<br />

223


224 5 Fullerenes<br />

Figure 5.12 Some Diels–Alder examples together with a stepwise addition of amine; 47 [108e],<br />

48 [111], 49 [112], and 50 [113].<br />

as o-quinodimethanes are used to prevent the retro-Diels–Alder reaction [108] or<br />

the adducts are further transformed into more stable products [109]. Moreover the<br />

instability of Diels–Alder adducts is applied to the regioselective Bingel reaction<br />

mentioned above [54, 55], as well as the isolation of a higher fullerene such as<br />

C 84 from the mixture with C 76 and C 78 [110]. Nevertheless making precursors for<br />

quinodimethanes, for example, is time consuming so that the Diels–Alder modification<br />

is not applied frequently to the materials science fields. Some Diels–Alder<br />

examples together with a stepwise addition of amine are depicted in Figure 5.12<br />

[108e, 111–113].<br />

5.2.4<br />

Conclusions<br />

In the last one-and-a-half decades, a variety of modification reactions have<br />

been developed using the most easily available C 60 in the chemistry. Yet still<br />

some methods are being searched especially in the transition metal catalyzed<br />

reactions.<br />

Since the application to materials science is now booming, the fullerene properties<br />

have attracted much attention, such as the high coagulation tendency, the<br />

poor solubility in usual organic solvents, the nanometer-scale volume, as well as<br />

the electron-accepting nature. When its electronic properties are important at the<br />

applications, however, the tendency must be kept in mind that C 60 gradually loses<br />

its properties by the modification. For example, its quantum yield of the singlet<br />

oxygen generation under photoirradiation decreases from 1.0 for the pristine<br />

one as a sensitizer to 0.9 for monoadduct, and then to 0.5 for some bisadduct<br />

isomers [114].<br />

And now the isolation of unstable fullerenes seems to be focused on, which<br />

is mentioned briefly above [110]. In the next decade we will see a lot of exotic


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

fullerenes isolated from the soot obtained by a variety of methods. Those will not<br />

only be limited to the higher fullerenes but also found in lower fullerenes, which<br />

will give some impacts in organic chemistry itself at the same time.<br />

5.3<br />

Physicochemical Properties and the Unusual Structure of Fullerenes<br />

5.3.1<br />

Single-crystal X-ray Structures of Fullerenes and Their Derivatives<br />

Olga V. Boltalina, Alexey A. Popov and Steven H. Strauss<br />

5.3.1.1 Introduction<br />

Fullerenes are closed-carbon-cage molecules containing 12 pentagons and (except<br />

for C 20) one or more hexagons of three-connected C atoms that are nominally sp 2<br />

hybridized [115]. All isolated-pentagon-rule (IPR) [116] C 60+x fullerenes have 60<br />

C atoms that are at the junction of two hexagons and one pentagon (HHPJs); the<br />

other x C atoms are at the junctions of three hexagons (triple-hexagon junctions,<br />

or THJs). The curvature of fullerene cages requires that fullerene C atoms cannot<br />

be coplanar with their three nearest neighbors, as shown in Figure 5.13. This is<br />

a significant departure from the usual planar geometry of C(sp 2 ) atoms, and the<br />

nonplanarity introduces steric strain in all fullerenes relative to planar graphite.<br />

The exohedral derivatization of fullerenes is driven, in large measure, by the relief<br />

of this strain as substituents [117] are added to some cage C atoms, changing<br />

the hybridization of those C atoms from (approximately) sp 2 to sp 3 [118]. The<br />

drawings in Figure 5.13, made with the coordinates of DFT-optimized C 60 and<br />

the known D 2 -symmetry isomer of C 76 (which is denoted C 76 -D 2 (1)) [119], show<br />

that HHPJs are more pyramidal with respect to their neighbors than THJs, and<br />

consequently no hollow fullerene derivative with fewer than 38 substituents has<br />

been found to have a substituent on a THJ [120–122]. In other words, HHPJs are<br />

Figure 5.13 One of the 60 hexagon–hexagon–pentagon junctions (HHPJs) in the<br />

DFT-optimized structure of C 60 (top) and one of the 16 triple-hexagon junctions (THJs) in the<br />

DFT-optimized structure of C 76 -D 2 (1) (bottom). Both views are parallel to the plane of the<br />

three darkly-shaded C atoms to which the junction carbon atom is attached. The darkly-shaded<br />

C atoms are connected with dashed lines.<br />

225


226 5 Fullerenes<br />

more reactive with respect to addition reactions than THJs because of the difference<br />

in pyramidalization [117].<br />

There are two excellent reviews of fullerene X-ray structures, one by Neretin<br />

and Slovokhotov [123], which lists nearly 400 fullerene structures, and the more<br />

focused review by Balch [124]. The focus of this chapter is the inherent steric<br />

strain in fullerenes, as measured by �-orbital axis vector (POAV) analysis [125].<br />

Only leading references will be cited. The examples of fullerene derivative X-ray<br />

structures we have chosen for analysis have F and CF 3 substituents because these<br />

structures are more numerous [126, 127] than structures with other substituents.<br />

This will allow a comparison to be made of structures with identical compositions.<br />

X-ray structures of endohedral metallofullerenes (EMFs) will not be discussed,<br />

and space limitations further prevent us from reviewing the many structures of<br />

fullerene cycloadducts [128] and the growing class of norfullerene derivatives [128,<br />

129]. Finally, the DFT-predicted results (PBE functional) in this chapter that are<br />

not explicitly attributed to previous publications were determined for this chapter<br />

using methods previously described [126].<br />

5.3.1.2 Disorder<br />

The first attempts to determine precise C–C distances in crystalline C 60 by X-ray<br />

diffraction revealed that the icosahedral molecule exhibits rotational disorder,<br />

and only the cage radius and intercage distance could be determined [123].<br />

Consequently, as discussed in Section 5.3.3, the first determination of C 60 bond<br />

distances was carried out using solid-state NMR spectroscopy [130], and only later<br />

by X-ray [131] and neutron diffraction [132] (in the latter two cases the fullerene<br />

was disordered even at low temperature). Although the solid-state disorder is<br />

severe, it is not completely isotropic [133]. To varying extents, higher fullerenes and<br />

endohedral fullerenes without exohedral substituents also exhibit disorder of cage<br />

C atoms and hence generally afford only imprecise diffraction-derived positional<br />

parameters [123, 124]. This problem has been overcome in favorable cases by cocrystallization<br />

with other molecules or by lowering the temperature of the crystal<br />

[123, 124, 134, 135]. The most precise structure of C 60 is in C 60 ·Pt(OEP)·2 C 6 H 6 ;<br />

the fullerene is on a general position and the estimated standard deviations (esd)<br />

for the 90 cage C–C distances are 0.003 Å [135]. In most endohedral fullerene<br />

structures, some or all of the endohedral atoms exhibit positional disorder, even<br />

when exohedral substituents or cocrystallization with metalloporphyrins has<br />

minimized or eliminated cage disorder. However, the first completely ordered<br />

EMF derivative was recently reported [136].<br />

5.3.1.3 Nonplanar Steric Strain<br />

The structures, Schlegel diagrams, and IUPAC numbering schemes for C 60 and<br />

C 70 are shown in Figure 5.14. Every fullerene larger than C 60 has more than 20<br />

hexagons (C 70 has 25 hexagons and ten THJs). The nonplanar strain energy is<br />

considerable. It is believed to be ca. 2 000 kJ mol –1 in C 60 (33 kJ mol –1 C-atom –1 ),<br />

which is 80% of the excess energy of the C atoms relative to graphite [125]. The<br />

POAV angle � �� , defined in Figure 5.15 [125], is 101.64° for DFT-optimized C 60


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Figure 5.14 DFT-optimized structures and Schlegel diagrams of C 60 -I h (1) and C 70 -D 5h (1).<br />

The 10 triple hexagon junctions around the equator of C 70 are C2, C5, C12, C20, C30, C40, C50,<br />

C59, C65, and C70.<br />

Figure 5.15 In this version of the �-orbital axis vector approximation, the POAV is defined as<br />

the vector that makes equal angles to the three � bonds at a conjugated C atom. The common<br />

angle is denoted � �� . In C 60 , the � p angle (i.e. � �� – 90°) is 11.64°.<br />

227


228 5 Fullerenes<br />

(the reduced POAV angles, defined as (� �� – 90°) and hereinafter denoted � p,<br />

range from 11.4° to 11.9° for the C 60 molecule in C 60·Pt(OEP)·2 C 6H 6 [135], and<br />

are 8.6°, 10.2°, 11.5°, 11.8°, and 12.0° for C2, C1, C7, C24, and C8, respectively,<br />

in the DFT-optimized structure of C 70-D 5h(1). DFT average values of � p are 10.9°<br />

for C 70, 10.3° for the known D 5d(1) isomer of C 80, and 9.9° for the known D 2(22)<br />

isomer of C 84. It is sensible that the average value of � p decreases as the number<br />

of C atoms in the cage increases because, in a C 60+x fullerene, x C atoms are THJs<br />

and are therefore in a more planar environment than the 60 C atoms that form<br />

the 12 pentagons.<br />

The strain energy in fullerenes varies from isomer to isomer in higher fullerenes,<br />

being lowest for isomers in which the pentagon-induced curvature is distributed<br />

as uniformly as possible over the fullerene surface (i.e. for isomers with pentagons<br />

as far apart as possible) [115]. This is an extension of the isolated-pentagon rule<br />

(IPR) [116]. The lower stability of fullerenes with adjacent pentagons is generally<br />

understood to be due to two factors, (1) resonance destablization and (2) an<br />

increase in steric strain [118]. The larger steric strain in some of the cage C atoms<br />

of a fullerene with adjacent pentagons is clearly seen in Figure 5.16, which shows<br />

the DFT-optimized structure of the hypothetical non-IPR fullerene C 84 -C s (51365)<br />

(an actual derivative of it, the EMF Tb 3 N@C 84 -C s (51365), has been studied by X-ray<br />

diffraction [137]). The two pent-pent junction C atoms have � p angles of 16.1°<br />

(note that the corresponding angle for a C(sp 3 ) atom would be 19.5°). However, the<br />

evidence discussed below, based on DFT-optimized structures [138], indicates that<br />

the overall nonplanar steric strain in C 84 -C s (51365) is not unusually high relative<br />

to many stable IPR fullerenes.<br />

Figure 5.16 The DFT-optimized structure of the hypothetical non-IPR fullerene C 84 -C s (51365).<br />

Note the unusually prominent non-planarity of the cage C atoms that form the pentagon–<br />

pentagon junction at the top of the drawing. The � p angle for these symmetry-related C atoms<br />

is 16.1° (cf. 11.6° for C 60 and 10.5° for the two C atoms the form the pentagon–hexagon junction<br />

at the bottom of the structure.


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

5.3.1.4 Nonplanar Steric Strain Parameters<br />

A measure of steric strain in IPR fullerenes known as � h was described in the<br />

Atlas of Fullerenes [115]. It is ‘the standard deviation … of the hexagon neighbor<br />

index distribution’ (each hexagon in a fullerene has an index, 3, 4, 5, or 6, which<br />

is the number of hexagons that surround it) [115]. For IPR fullerenes up to the 46<br />

isomers of C 90, � h varies from 0.000 (for the ‘least strained’ fullerenes according<br />

to this concept, which include C 60, C 80-I h, and C 80-D 5h) to ca. 2.400, but nearly all<br />

are between 0.000 and 1.000. This follows Raghavachari’s suggestion [139] that<br />

nonplanar steric strain in IPR fullerenes is minimized when ‘the pentagon-induced<br />

curvature is distributed as uniformly as possible over the fullerene surface’<br />

[115]. Some relevant � h values are listed in Table 5.2 [119, 126, 135, 138, 140–151].<br />

For example, � h for C 70 is 0.490. The I h and D 5d isomers of C 80, which differ substantially<br />

in shape and therefore in the distribution of strain energy, are shown in<br />

Figure 5.17. Their � h values are 0.000 and 0.817, respectively. It has proven difficult<br />

to apportion the energy difference between fullerene isomers to nonplanar steric<br />

strain and to � electronic effects that are not due to nonplanarity [115]. What is<br />

known is that, except for C 84 isomers, IPR fullerenes with maximum electronic<br />

stability generally have high steric strain and vice versa [115]. For example, despite<br />

the greater steric strain implied by its high � h value, the D 5d isomer of C 80 has a<br />

DFT energy that is 72 kJ mol –1 lower than for the I h isomer [142].<br />

Several authors have proposed using either (1) the parameter � p (earlier defined<br />

as � �� – 90°) [116] or (2) the fractional s character for a fullerene C atom � atomic<br />

orbital as a measure of the nonplanar strain energy for that C atom, and the sum<br />

over all C atoms as an index of the nonplanar strain energy for the entire fullerene<br />

[115, 152]. Since � h is only defined for IPR fullerenes with no substituents, we<br />

propose that a similar parameter be used to compare overall nonplanar steric<br />

strain in all fullerenes and their derivatives. That dimensionless parameter, �(� p 2 ),<br />

is the standard deviation of the � p 2 distribution for all C(sp 2 ) atoms in the cage.<br />

Figure 5.17 The DFT-optimized structures of two of the seven IPR isomers of C 80 that have<br />

D 5d and I h symmetry. The steric parameter � h is 0.817 for C 80 -D 5d (1) and 0.000 for C 80 -I h (7).<br />

The dimensionless steric parameters �(� p 2 ) is 36.0 for C80 -D 5d (1) and 10.1 for C 80 -I h (7).<br />

229


230 5 Fullerenes<br />

2<br />

Figure 5.18 The correlation of the steric parameters �h and �(�p) for the 24 DFT-optimized IPR<br />

isomers of C84 (the parameters are defined in the text). The line is a linear least-squares fit to<br />

the data. The DFT structure of C84-D2 (1) is shown with its 84 POAV vectors. It is the least stable<br />

of the IPR C84 cages as well as the least uniform in surface curvature. The �p angles for the<br />

vectors labelled a and b are 12.3° and 4.6°, respectively.<br />

To demonstrate that � h and �(� p 2 ) can be used interchangeably for IPR fullerenes,<br />

and therefore that �(� p 2 ) reliably measures nonplanar steric strain, we determined<br />

�(� p 2 ) values of the 24 IPR isomers of C84 and compared them with their � h values<br />

(Figure 5.18). A similar plot for IPR isomers of C 80 (not shown) is also essentially<br />

linear with similar slope and intercept. This suggests that it may be meaningful<br />

to compare �(� p 2 ) values for cages with different numbers of C atoms such as<br />

those listed in Table 5.2.<br />

Table 5.2 Structural parameters and DFT-predicted relative �H f values for fullerenes and<br />

fullerene(X) n derivatives. a)<br />

Compound �H f for isomers,<br />

kJ mol –1<br />

b)<br />

�h [115]<br />

2 c)<br />

�(�p) Average � p ,<br />

deg g)<br />

C 60 0.000 0.0 11.6 DFT<br />

DFT or X-ray<br />

C 60 0.000 2.5 11.6 X-ray [135]<br />

C 70 0.490 25.0 10.9 DFT<br />

C 70 (CF 3 ) 10 -1 49.3 9.0 X-ray [140]<br />

C 74 -D 3h (1) 0.415 23.2 10.5 DFT [119]<br />

(C 74 -D 3h (1))(CF 3 ) 12 35.7 8.4 X-ray [141]


Table 5.2 (continued)<br />

Compound �H f for isomers,<br />

kJ mol –1<br />

C 80 -I h (7) 72 d)<br />

C 80 -C 2v (5) 18 d)<br />

C 80 -D 5d (1) 0 d)<br />

C 84 -D 2 (22) 0.0 e)<br />

C 84 -D 2 (21) 54.0 e)<br />

C 84 -C 2v (7) 88.3 e)<br />

C 84 -D 2 (1) 200.8 e)<br />

C 84 -C s (51365) 258.5 e)<br />

C 84 -D 2 (22) 6–<br />

C 84 -D 2 (21) 6–<br />

C 84 -C 2v (7) 6–<br />

C 84 -C s (51365) 6–<br />

38.5 e)<br />

0.0 e)<br />

48.6 e)<br />

1.8 e)<br />

5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

b)<br />

�h [115]<br />

2 c)<br />

�(�p) Average � p ,<br />

deg g)<br />

DFT or X-ray<br />

0.000 10.1 10.1 DFT [138]<br />

0.365 23.4 10.1 DFT [138]<br />

0.817 36.0 10.3 DFT [138]<br />

0.331 22.8 9.9 DFT [138]<br />

0.331 23.2 9.9 DFT [138]<br />

0.485 28.7 9.9 DFT [138]<br />

0.927 46.0 10.0 DFT [138]<br />

36.1 9.9 DFT [138]<br />

0.331 28.7 9.9 DFT [138]<br />

0.331 27.9 9.9 DFT [138]<br />

0.485 33.1 9.9 DFT [138]<br />

37.5 9.9 DFT [138]<br />

C 60 (CF 3 ) 10 -2 0.1 31.5 9.5 X-ray [143]<br />

C 60 (CF 3 ) 10 -3 4.7 35.9 9.4 X-ray [144]<br />

C 60 (CF 3 ) 10 -4 7.6 36.4 9.5 X-ray [142]<br />

C 60 (CF 3 ) 10 -5 1.2 34.3 9.5 X-ray [126]<br />

C 60 (CF 3 ) 10 -6 0.0 28.7 9.4 X-ray [146]<br />

C 60 (CF 3 ) 12 -1 0.0 f)<br />

C 60 (CF 3 ) 12 -2 39.8 f)<br />

29.7 9.0 X-ray [147]<br />

33.3 8.9 X-ray [148]<br />

C 1 -C 60 (CF 3 ) 18 28.8 7.0 X-ray [150]<br />

C 3v -C 60 F 18 53.9 8.6 X-ray [149]<br />

T-C 60 F 36 3.3 1.7 X-ray [151]<br />

a) All data from this work unless otherwise noted. Reference numbers are shown in square brackets.<br />

b) The �h values are standard deviations of the hexagon neighbor index distributions.<br />

The � h parameter is only defined for IPR fullerenes, not for non-IPR cages (e.g. C 84 -C s (51365))<br />

or for exohedral fullerene derivatives.<br />

c) These dimensionless steric parameters, defined in the text, were calculated for this chapter<br />

using the DFT-optimized coordinates for all of the three-connected carbon atoms in a fullerene<br />

or fullerene derivative.<br />

d) Ref. [142]; these relative energies were calculated for singlet ground states.<br />

e)<br />

Ref. [138].<br />

f)<br />

Ref. [148].<br />

g) 2<br />

This average is over all cage C atoms that are nominally sp hybridized and does not include<br />

sp 3 hybridized C atoms bearing substituents.<br />

231


232 5 Fullerenes<br />

5.3.1.5 Are Non-IPR Fullerenes Sterically Unstable?<br />

2<br />

The relative �Hf and �(�p) values in Table 5.2 reveal an interesting aspect about<br />

the structure of non-IPR C84-Cs(51365). It is predicted to be 58 kJ mol –1 less stable<br />

than the least stable IPR isomer, C84-D2(1), and 258 kJ mol –1 less stable than the<br />

2<br />

most stable IPR isomer, C84-D2(22). However, its �(�p) value is significantly lower<br />

2<br />

than that for C84-D2(1) (it is also lower than �(�p) for two other IPR C84 cages). In<br />

2<br />

addition, its �(�p) value is the same as for IPR C80-D5d(1), the most stable of the<br />

IPR C80 cages [142]. The average values of �p for C84-D2(22) and C84-Cs(51365) are also the same. Based on these results, it does not appear that C84-Cs(51365) has an unusual amount of nonplanar strain energy relative to some of the most<br />

stable IPR fullerenes that are known to exist.The DFT relative energies of several<br />

6–<br />

C84 cages were recently reported [128]. They were calculated to predict the most<br />

stable cages that should be present in isolable iM3N@C84 compounds. The data<br />

2 –<br />

in Table 5.2 show that �(�p) values for the neutral cages and the 6 ions are only<br />

marginally different. Significantly, the C84-Cs (51365) 6– cage is energetically as stable<br />

as the most stable hexaanionic IPR cage, C84-D2 (21) 6– . With the right electron<br />

count, the non-IPR isomer can be as stable as any IPR isomer. This can only be<br />

true if the non-IPR cage is not sterically very unstable. Therefore, it appears that<br />

the instability of neutral C84-Cs(51365) is predominantly electronic in origin. This<br />

may also be true for other non-IPR fullerenes, and the idea that non-IPR fullerenes<br />

are sterically very unstable should be carefully reexamined.<br />

5.3.1.6 Long and Short C(sp 2 )–C(sp 2 ) Bonds in Fullerene Cages<br />

With the exception of three non-IPR EMF X-ray structures, Tb 3N@C 84-C s(51365)<br />

[137]. La@C 72(C 6H 3Cl 2) [153], and Sc 3N@C 68-D 3(6140) [151], all other fullerene<br />

X-ray structures reported to date are of IPR fullerenes. The first X-ray structures<br />

of non-IPR hollow fullerenes were reported in late 2008 [591]. The smallest IPR<br />

fullerene is C 60, and its X-ray structure [135] revealed that it has 30 short bonds<br />

(1.379(3)–1.396(3) Å), which are hexagon-hexagon junctions, and 60 long bonds<br />

(1.440(3)–1.461(3) Å), which are hexagon-pentagon junctions, as shown in<br />

Figure 5.19 (the DFT-optimized distances are 1.399 and 1.453 Å, respectively, and<br />

the NMR-determined distances are 1.40 and 1.45 Å, respectively [130]). The short<br />

hex-hex junctions and long hex-pent junctions are usually referred to as ‘double’<br />

bonds and ‘single’ bonds, respectively, although this is clearly an oversimplification.<br />

These and other results [155, 156] have led to the oft-quoted ‘rule’ that double<br />

bonds in pentagons (DBIPs) are destabilizing in fullerenes and exohedral fullerene<br />

derivatives, although the underlying steric and/or electronic basis for this rule<br />

has not been convincingly explained.<br />

Higher fullerenes have more than two types of cage C–C bonds. The structures<br />

of C 70 ·6 S 8 [157, 158] and C 76 ·6 S 8 [159] were reported in the 1990s, but they did<br />

not have a precision that allowed the C–C bond distances of the various bond<br />

types to be distinguished from one another. For C 70 , this was still true at the end<br />

of 2006 [160]. In order to get a solution to the structure in some cases, noncrystallographic<br />

symmetry had to be imposed (e.g. D 5h symmetry for C 70 [158]) or a<br />

rigid-body DFT-optimized cage has been assumed and used in the refinement (e.g.


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Figure 5.19 The 90 fullerene cage C–C bond distances in the structure of C 60 · Pt(OEP) · 2 C 6 H 6 [135].<br />

The error bars shown are ±3�. The 30 shorter distances (‘double bonds’) are hexagon-hexagon<br />

junctions and the 60 longer distances (‘single bonds’) are hexagon-pentagon junctions.<br />

Figure 5.20 The shortest and longest distance for the eight types of cage C–C bonds in two of<br />

the three C 70 molecules in the X-ray structure of C 9 H 3 Cl 6 N · 3 (C 70 ) · 3 (C 6 H 5 Cl) (±3� error bars)<br />

[163]. Each of the eight diamond-shaped points is the DFT-optimized distance for that C–C<br />

bond type (this work).<br />

233


234 5 Fullerenes<br />

for BaC 74-D 3h [161] and Sc 3N@C 78-D 5h(5) [162]). A precise structure of C 70 was<br />

finally determined in 2007 using crystals of C 9H 3Cl 6N·3(C 70)·3 (C 6H 5Cl), with esd<br />

for cage C–C distances of 0.004 Å for two of the three independent C 70 molecules<br />

[163]. The ranges for the eight types of C–C distances are shown in Figure 5.20<br />

and are compared with the corresponding DFT predicted distances. The addition<br />

of substituents to the fullerene cage can also lead to ordered and precise structures,<br />

albeit at the expense of altering the fullerene cage from its original state as<br />

a pure carbon allotrope. The esd for cage C–C distances can be as low as 0.003<br />

[140], 0.002 [141], or 0.001 Å [147]. For example, the X-ray structure of C 70(CF 3) 10<br />

[140] is shown in Figure 5.21a along with a plot of its 75 C(sp 2 )–C(sp 2 ) distances.<br />

The esd for these distances are 0.003 Å. A plot of the 60 C(sp 2 )–C(sp 2 ) distances<br />

in an isomer of C 60 (CF 3 ) 10 is also shown in Figure 5.21b. These plots show that<br />

the sharp distinction between cage C–C ‘single’ and ‘double’ bonds, which was<br />

meaningful when applied to C 60, is much less distinct for higher fullerenes and<br />

fullerene(X) n derivatives (including C 60X n derivatives).<br />

Figure 5.21 Plots of the 75 (a) and 60 (b) cage C(sp 2 )–C(sp 2 ) distances in the X-ray structures<br />

of one of the isomers of C 70 (CF 3 ) 10 [140] and C 60 (CF 3 ) 10 [143]. The insets are (upper left in<br />

both cases) a 50% probability thermal ellipsoid plot of the molecule and (lower right in both<br />

cases) a Schlegel diagram showing the ribbon of edge-sharing C 6 (CF 3 ) 2 hexagons (each black<br />

circle represents a cage C atom bearing a CF 3 group; the letter ‘m’ denotes the meta-C 6 (CF 3 ) 2<br />

hexagon).


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

There are now nearly 30 X-ray structures of fullerene(CF 3) n compounds (and a<br />

growing number of fullerene(C 2F 5) n structures) [126, 127], and many of these are<br />

among the most precise X-ray structures of any fullerene or fullerene derivative.<br />

For the CF 3 compounds, the value of n varies from 2 to 18. The CF 3 substituent<br />

is sterically bulky (it is larger than an iodine atom) [144]. In most of the structures<br />

with n � 12, the CF 3 groups are distributed one per pentagon (there are only two<br />

exceptions), and CF 3 groups are rarely found on adjacent cage C atoms (two<br />

different exceptions). In nearly every case, the CF 3 groups are found on isolated<br />

para-C 6(CF 3) 2 hexagons and/or on one or more ribbons or loops of edge-sharing<br />

meta- and para-C 6(CF 3) 2 hexagons [126, 127]. In two cases, C 2-p 11 -(C 74-D 3h)<br />

(CF 3) 12 and C 2-p 11 -(C 78-D 3h(5))(CF 3) 12 [141] the X-ray structures were the first<br />

unambiguous proof of the existence of the hollow C 74 -D 3h and C 78 -D 3h (5) cages<br />

(‘p 11 ’ in the formula indicates that the 12 CF 3 groups are arranged on a ribbon<br />

of 11 edge-sharing para-C 6(CF 3) 2 hexagons). The structure of the C 74 compound<br />

was sufficiently precise (esd = 0.002 Å) that the cage C–C distances could be<br />

compared with the DFT-predicted distances (Figure 5.22). For all but three of the<br />

C–C bonds, the X-ray distance matched the DFT distance to within ±3�, validating<br />

the PBE functional used in that study (and in our other publications, including<br />

this chapter) [125].<br />

Figure 5.22 X-ray vs DFT-predicted cage C–C distances for the structure of C 2 -(C 74 -D 3h )(CF 3 ) 12<br />

(the molecule has crystallographic C 2 symmetry) [141]. The error bars shown are ±3�. The<br />

insets are (upper left) a 50% probability thermal ellipsoid plot of the molecule and (lower right)<br />

a Schlegel diagram showing the p 11 [125] ribbon of edge-sharing C 6 (CF 3 ) 2 hexagons (each black<br />

circle represents a cage C atom bearing a CF 3 group).<br />

235


236 5 Fullerenes<br />

5.3.1.7 Steric Strain in C 60(X) n Isomers<br />

One question that has not been addressed very often in the literature is whether<br />

or not different C 60(X) n isomers might have different amounts of nonplanar steric<br />

strain that can affect their relative energies [164–166]. (In this part of our analysis,<br />

we do not consider steric strain induced by proximate X substituents [165] or<br />

potential angle strain in cage C(sp 3 ) atoms). Listed in Table 5.2 are the DFT<br />

relative energies and X-ray derived �(� p 2 ) values for five isomers of C60(CF 3) 10.<br />

Schlegel diagrams of the isomers are shown in Figure 5.23 (the X-ray structure<br />

of isomer 1 has not yet been determined). Not listed in Table 5.2 are the range,<br />

average, and standard deviation for the bond angles for C(sp 2 ) atoms within<br />

the pentagons for each isomer. All three of these angle parameters are virtually<br />

the same for the five isomers as well as for C 60, indicating negligible differences<br />

in angle strain due to C(sp 2 ) atoms having C–C–C angles of ca. 108° [167]. For<br />

example, the sets of {range, average, standard deviation of the average} for these<br />

angles are {107.4–108.6°, 108.0°, and 0.2°} for C 60 , and {106.9–110.8°, 109.0°,<br />

and 0.9°} for C 60 (CF 3 ) 10 -2, and {106.9–110.7°, 109.0°, and 0.9°} for C 60 (CF 3 ) 10 -6<br />

(all three sets from X-ray structures). The relative energies for the five isomers of<br />

C 60(CF 3) 10 range from 0.0 to 7.6 kJ mol –1 and the �(� p 2 ) values range from 28.7<br />

to 36.4. Since these are both small ranges, we conclude that this series of nearly<br />

equienergetic isomers have comparable values of total steric strain (since differences<br />

in angle strain are also negligible) and therefore have similar � electron<br />

energies.<br />

Two isomers of C 60(CF 3) 12 are also listed in Table 5.2 (see Figure 5.23 for Schlegel<br />

diagrams). In this case the relative energies differ by 40 kJ mol –1 , a significant<br />

amount. However, the �(� p 2 ) values are similar, so it would seem that the energy<br />

difference can be attributed to differences in � electron energy. This tentative<br />

conclusion should be tested in the future by a computational study.<br />

Next, compare the two X-ray structures with very different addition patterns,<br />

C 3v -C 60 F 18 and C 1 -C 60 (CF 3 ) 18 (see Table 5.2 and Figure 5.23). In this case, the<br />

�(� p 2 ) values are quite different, 53.9 for C3v -C 60 F 18 but only 28.8 for C 1 -C 60 (CF 3 ) 18 .<br />

This shows that different addition patterns for a given number of substituents<br />

can lead to significantly different amounts of nonplanar steric strain, although<br />

this was not true for the five structurally similar isomers of C 60 (CF 3 ) 10 . The DFT<br />

energies of C 3v -C 60 F 18 and the isomer of C 60 F 18 with the same addition pattern of<br />

C 1 -C 60 (CF 3 ) 18 revealed that C 3v -C 60 F 18 was 290 kJ mol –1 more stable. How much<br />

of that difference is due to differences in steric strain and how much is due to<br />

differences in � electron energy remains to be seen.<br />

Finally, consider the concept that relative planarity for one part of a C 2n fullerene<br />

is sterically destabilizing because, since the surface must be closed, it introduces<br />

‘high curvature in another part of the fullerene surface’ [115, 152, 168]. This is a<br />

valid concept for fullerenes without exohedral substituents, but is it also valid for<br />

fullerene(X) n derivatives? The answer may be ‘No.’ In C 2n fullerenes, all C atoms<br />

are nominally sp 2 hybridized, but in fullerene derivatives there are cage C(sp 3 )<br />

atoms, and these atoms, which do not experience the nonplanar steric strain that<br />

the C(sp 2 ) atoms experience, can contribute to closing the surface of the fullerene.


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Figure 5.23 Schlegel diagrams of some of the C 60 X n derivatives listed in Table 5.2 (X = F, CF 3 ).<br />

Each black circle is a cage carbon atom to which a substituent is attached. For the C 60 (CF 3 ) n<br />

compounds with n = 10 and 12, the ribbons or loops of edge-sharing meta- and para-C 6 (CF 3 ) 2<br />

hexagons are highlighted (the letter ‘m’ denotes the meta-C 6 (CF 3 ) 2 hexagon) and the IUPAC<br />

lowest locants of the cage C(sp 3 ) atoms are indicated.<br />

237


238 5 Fullerenes<br />

In fact, the average � p for the cage C(sp 2 ) atoms in the X-ray structures of C 60,<br />

C 60(CF 3) 10-6, C 60(CF 3) 12-2, C 3v-C 60F 18, C 1-C 60(CF 3) 18, and T-C 60F 36 (Figure 5.23)<br />

are 11.6°, 9.5°, 8.9°, 8.6°, 7.0°, and 1.7°, respectively. The introduction of exohedral<br />

substituents clearly reduces the average nonplanar steric strain in cage C(sp 2 )<br />

atoms, to the point where the 24 remaining C(sp 2 ) atoms in T-C 60F 36 are in nearly<br />

planar environments.<br />

In reviewing the X-ray structures (and some relevant DFT-optimized structures)<br />

of fullerenes and their derivatives, we have raised several questions about how<br />

to determine steric strain in these compounds, especially in exohedral derivatives.<br />

We do not believe that we have answered these questions fully, because our<br />

analysis has been simple. Instead, we have highlighted some aspects of steric<br />

strain in closed-curved-surface carbon compounds that should be studied with<br />

more sophisticated computational methods in the future, especially as more and<br />

more precise X-ray structures become available.<br />

Acknowledgment<br />

We thank Prof. Marilyn Olmstead for her assistance in preparing this chapter, and<br />

we acknowledge the financial support of the Civilian Research and Development<br />

Foundation (RUC2-2830-MO-06 to AAP and SHS) and the U.S. National Science<br />

Foundation (CHE-0707223 to OVB and SHS).<br />

5.3.2<br />

Vibrational and Electronic Spectra<br />

Alexey A. Popov<br />

5.3.2.1 Introduction<br />

Spectroscopic studies of the fullerenes played an important role from the very<br />

beginning of the fullerene era. Historically, when Wolfgang Krätschmer and<br />

coauthors discovered that the soot prepared by the arc-discharge method under<br />

certain conditions was enriched with fullerenes, a characteristic IR spectrum with<br />

four sharp lines emerging over the broad band background was the first proof that<br />

C 60 was present in the sample, because exactly four IR active modes are expected<br />

for the icosahedral C 60 molecule from group theory analysis [169]. Subsequent<br />

discovery of the method of extracting the fullerenes from carbon soot [170] as well<br />

as progress in HPLC separation of fullerene mixtures, boosted numerous studies<br />

of spectroscopic and photophysical properties of these molecules.<br />

In this chapter, a survey of vibrational and electronic spectroscopic studies of<br />

fullerenes is presented. We should note that the number of published papers<br />

dedicated to the given fullerene molecule scales in accordance with the fraction<br />

of this fullerene in the arc-discharge soot. Hence, the ‘archetypical fullerene’<br />

C 60, which constitutes ca 85% of the crude fullerene extract, is by far the best<br />

studied member of the fullerene family, and it is inevitable that any review like<br />

this is to a large extent a review of the numerous studies on C 60. The properties


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

revealed for C 60 are to some extent inherited by all other fullerenes; however, the<br />

high symmetry of the C 60 molecule makes this fullerene very special in terms<br />

of spectroscopic properties, since the majority of its energy levels are degenerated,<br />

and many vibrational and electronic transitions are forbidden by symmetry<br />

selection rules.<br />

5.3.2.2 Vibrational Spectra of Fullerenes<br />

It is natural to expect the vibrational spectra of fullerenes to be quite complex:<br />

even for C 60, the smallest classical fullerene, the number of modes is as large<br />

as 174. However, vibrational representation of the free C 60 molecule spans<br />

2 A g + 3 T 1g + 4 T 2g+ 6 G g + 8 H g + A u + 4 T 1u + 5 T 2u + 6 G u + 7 H u symmetry types<br />

of I h group, and due to the high symmetry, 174 vibrational degrees of freedom<br />

are substantially degenerated and grouped into 46 normal modes, of which only<br />

14 are optically active (A g and H g symmetry types in Raman and T 1u type in IR<br />

spectra). In accordance with the symmetry considerations, four primary IR and<br />

ten Raman lines are observed in the experimental spectra of crystalline C 60 . As the<br />

highest site symmetry of C 60 in the room-temperature lattice is T h (and is further<br />

reduced at lower temperature), the fact that the spectra of solid C 60 samples can<br />

be described in terms of the isolated fullerene molecule with unperturbed I h<br />

symmetry show that crystalline fullerenes can be treated as molecular crystals<br />

with weak intermolecular interactions. The same conclusion can be drawn from<br />

the X-ray data discussed in the former section.<br />

Contrary to the optically-active modes, assignment of the many of 32 inactive<br />

fundamentals of C 60 is still ambiguous (see, for instance, detailed analysis given<br />

in Refs. [171] and [172]). The most detailed, but also the most complicated, information<br />

can be obtained from the second-order IR and Raman spectra [173, 174],<br />

i.e. the spectra of thick films or single crystals, which exhibit multiple bands with<br />

intensities at least an order of magnitude lower than those of the symmetry-allowed<br />

modes (Figure 5.24). The spectra are very rich and can be attributed to the silent<br />

modes as well as overtones and combination modes. The vibrational spectrum<br />

of C 60 was also studied by fluorescence spectroscopy (the measurements of the<br />

matrix-isolated C 60 at helium temperatures are particularly informative) [175, 176],<br />

inelastic neutron scattering (INS) spectroscopy [177, 178], photoluminescence<br />

spectroscopy of the singlet oxygen entrapped in C 60 lattice (SOPL) [179], and high<br />

resolution electron energy loss spectroscopy (HREELS) [180] (see Figure 5.24).<br />

While fluorescence spectroscopy provides information on some optically-forbidden<br />

ungerade modes, INS, SOPL and HREELS in principle provide complete vibrational<br />

density of states (VDOS) irrespective of the symmetry types (however,<br />

selection rules for SOPL spectrum are not yet clear). Unfortunately, the spectral<br />

resolution of these methods is rather low, especially in the high frequency range. In<br />

1998 Heid [178] reported assignment of all vibrational modes in the 250–600 cm –1<br />

range based on the INS study of a large single crystal of C 60.<br />

Besides C 60, detailed vibrational studies by means of IR, Raman, fluorescence,<br />

HREELS, INS and SOPL spectroscopies were reported only for D 5h–C 70. As can be<br />

expected from the lower symmetry of the molecule, the spectra of C 70 are consider-<br />

239


240 5 Fullerenes<br />

Figure 5.24 (a) IR and Raman spectra of C 60 . Raman spectra are shown for two excitation<br />

wavelengths. Note strong resonance enhancement of A g (2) modes in the spectrum excited in<br />

the visible range. (b) ‘Second-order’ IR and Raman spectra of C 60 . (c) Vibrational density of<br />

states in C 60 measured by INS [177], HREELS [181], and SOPL [179] and calculated with DFT<br />

(PBE/TZ2P) [182]. Stretching O-O modes in SOPL spectra of 18 O 2 and 16 O 2 are denoted by<br />

asterisks.


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

ably richer than those of C 60. Due to the much higher number of fundamentals<br />

(46 in C 60 vs 122 in C 70, including 53 Raman active and 31 IR active vibrations),<br />

assignments of C 70 normal modes are less reliable than those for C 60, and some<br />

ambiguities exist even in the assignment of several IR and Raman active modes.<br />

The use of INS or SOPL for C 70 appears less advantageous than for C 60 because<br />

the lower degeneracy of C 70 normal modes and their much higher total number<br />

result in the smeared VDOS in the whole frequency range [183, 184].<br />

The difficulties with the isolation of the isomerically pure samples of higher<br />

fullerenes result in the lack of vibrational spectroscopic data for many of them.<br />

To our knowledge, IR and Raman spectra have been reported for D 2(1)-C 76 [185],<br />

C 2v(2)-C 78 [186], C 2v(3)-C 78 [186], D 2(22)-C 84 [187, 188], and D 2d(23)-C 84 [187, 188];<br />

for D 3 (1)-C 78 and D 2 (2)-C 80 only Raman spectra are available [189]. As anticipated,<br />

complexity of the spectra increases with the number of atoms in the fullerene<br />

molecule and with the decrease of its symmetry. At the same time, vibrational<br />

spectra appear to be very structure sensitive (for instance, IR and Raman spectra<br />

of two C 2v isomers of C 78 are substantially different, Figure 5.25), a property that<br />

can be used for structure determination of the new fullerene isomers, especially<br />

if experimental studies are supported by theoretical calculations.<br />

A systematic study of the Raman spectra of the group of fullerenes were<br />

performed by Eisler et al. [189, 190] in 2000. The authors studied isomerically<br />

pure fullerenes C 60 , C 70 , C 76 (D 2 ), C 78 (D 3 , C 2v , C� 2v ), and C 80 (D 2 ) and isomeric<br />

mixtures of C 82 (two C 2 isomers in 1 : 1 ratio) and C 84 (D 2 and D 2d isomers in 2 : 1<br />

ratio). Room-temperature spectra of all fullerenes were measured with 514, 693,<br />

794, 1064 nm excitation wavelengths (see Figure 5.25 for the spectra obtained<br />

at 693 nm wavelength). To establish the common vibrational phenomena for all<br />

fullerenes, the authors employed Lamb theory of oscillations of isotropic spherical<br />

and spheroidal shells. Two types of vibration, referred to as monopolar and quadrupolar,<br />

were found to be especially characteristic in the analysis of the Raman<br />

spectra of fullerenes. Monopolar modes can be easily recognized as they sustain<br />

medium to strong Raman intensity over all excitation wavelength (the example<br />

is ‘breathing’ A g (1) vibration of C 60 ). Their frequencies scale inversely with the<br />

square root of mass and are shape-insensitive, that is, different isomers with the<br />

same mass (e.g. C 78 ) exhibit nearly the same frequency. Quadrupolar mode (the<br />

example is ‘squashing’ H g (1) vibration of C 60 ) is five-fold degenerate for a sphere<br />

and C 60, but is split into several components for all other fullerenes. Importantly,<br />

the magnitude of splitting correlates with the deviation of the fullerene shape<br />

from the sphere (viz. 15 cm –1 for nearly spherical D 2 –C 84 versus 45 cm –1 for<br />

elongated D 2-C 80), and thus can be used to determine the isomeric structures of<br />

newly isolated fullerenes.<br />

241


242 5 Fullerenes<br />

Figure 5.25 (a) IR spectra of higher fullerenes [185–188]; (b) Raman spectra of higher fullerenes<br />

[185–188] (� ex = 1064 nm for C 70 –C 78 , 514 nm for C 84 ); (c) Raman spectra of some fullerenes,<br />

excitation 693 nm [189]. Asterisks in (c) denote breathing mode.


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

5.3.2.3 The Orbital Picture of Fullerenes: High-energy Electronic Spectra<br />

Though non-planarity of the fullerene molecules results in the partial mixing of �<br />

and � states (see Section 5.3.3.3), high-lying occupied and low-lying unoccupied<br />

molecular orbitals (MOs) of fullerenes are still of the essentially �-nature and may<br />

be adequately described by the theoretical methods developed for the �-systems.<br />

Due to the high symmetry of C 60, it is straightforward to treat this molecule as<br />

a spherical shell whose electronic states can be classified in terms of the orbital<br />

momentum l. In such a ‘superatomic’ model, all levels up to l = 4 are occupied,<br />

while the level with l = 5 is partially filled with 10 electrons (Figure 5.26) [180].<br />

Though very simple, the model gives a reasonable description of the electronic<br />

structure of C 60 which can be correlated with the orbital picture obtained by using<br />

more sophisticated theoretical approaches (Figure 5.26), especially for the occupied<br />

states. Under icosahedral perturbation, the electronic levels are split but still retain<br />

a substantial extent of degeneracy. The majority of C 60 MOs have either t (u,g) , g (u,g)<br />

or h (u,g) symmetry, including five-fold degenerated HOMO (h u ) and three-fold<br />

degenerated LUMO (t 1u ). LDA calculations of the band structure in crystalline<br />

C 60 have shown that the widths of the bands derived from the frontier orbitals do<br />

not exceed 0.5 eV [191], and hence discrete density of states (DOS) characteristic<br />

for C 60 in the gas should be preserved in the solid phase.<br />

Experimentally electronic states of C 60 were probed by a variety of high-energy<br />

spectroscopic techniques (see Refs. [180, 192, 193] for reviews). Valence occupied<br />

states were extensively studied by ultraviolet photoemission spectroscopy (UPS)<br />

[180, 194] providing a direct measure of orbital ionization energies and DOS.<br />

A complementary technique, providing information about density of unoccupied<br />

states (DUOS) is inversed photoemission spectroscopy (IPES) [195], in which<br />

the surplus electron is immersed into unoccupied orbitals. Other methods extensively<br />

used to analyze DUOS of fullerenes are core-level electron energy loss<br />

spectroscopy (EELS) [180, 192, 193] and X-ray absorption spectroscopy (XAS)<br />

[196, 197]. In core-level EELS and XAS, electrons from C1s levels are excited<br />

to un occupied C2p-derived states. Formally this provides a measure of DUOS,<br />

however the energy levels are affected by the interactions with the core hole,<br />

resulting in the possible difference from the IPES spectra (which are free from<br />

the excitonic effect).<br />

Figure 5.27 shows representative UPS, IPES, XAS and EELS spectra of C 60<br />

compared with the DOS and DOUS of the isolated molecule computed at the PBE/<br />

TZ2P level. The spectra in the valence region are characterized by sharp distinct<br />

peaks, which reflect the extensively degenerated MO levels of C 60 . UPS and IPES<br />

spectra show good agreement with theoretically computed DOS and DUOS in the<br />

energy range of � and �* states, respectively, and the features derived from the<br />

highest-energy occupied orbitals can be assigned straightforwardly (Figure 5.26).<br />

The patterns of EELS and XAS spectra are virtually identical and, though they also<br />

resemble IPES spectra, the shift of the peak positions is obvious, pointing to the<br />

importance of the excitonic effect in the fullerene.<br />

UPS spectra measured in the gas and the solid state are very similar proving<br />

weak intermolecular interaction in the latter, however the fine vibronic structure<br />

243


244 5 Fullerenes<br />

Figure 5.26 Upper panel: Electronic energy levels in C 60 as obtained with the model: electron<br />

on a sphere [180] (left), Hückel approximation (middle) and DFT PBE/TZ2P [182] (right).<br />

Note that the energy scales for the three models are different. Middle and lower panels: DOS<br />

and DUOS in C 60 as measured by different methods [180, 194–197] and calculated by DFT [182].<br />

The asterisks in the gas-phase spectra denote CO bands.


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

resolved in the spectrum of the gaseous C 60 is absent in the solid state spectra<br />

[194]. Ionization potential (IP) and electron affinity (EA) of C 60, estimated from<br />

the gas-phase UPS spectra of neutral C 60 and its anion [198], are 7.3 and 2.67 eV,<br />

respectively, which formally gives the energy gap between the HOMO and LUMO<br />

levels of 4.7 eV, while in the solid state the energy difference of the HOMO and<br />

LUMO-derived peaks of UPS and IPEs spectra is reduced to 3.75 eV due to the<br />

screening effect of the surrounding lattice molecules. Significantly, both gas<br />

and solid state values are considerably larger than the band gap � = 1.8–2.2 eV<br />

measured by optical absorption spectroscopy and low-energy EELS. The reason<br />

for such a discrepancy is the different nature of the final states. While the � value<br />

corresponds to the formation of the exciton (hole-electron pair) localized on one<br />

molecule, in the UPS and IPES techniques the final states are ionic. That is, (IP-EA)<br />

difference or the peak-to-peak distance in UPS/IPES spectra is the formation<br />

energy of the hole and the electron localized on two different molecules. The difference<br />

(IP–EA)–� defines the energy of on-site Coulomb interaction (so called<br />

charge correlation energy), Hubbard U, which therefore has the order of 3.0 eV<br />

for the free molecule and 1.6 eV for C 60 in the solid state.<br />

As can be expected on the basis of their common molecular architecture and<br />

all-carbon � systems, DOS and DUOS of C 70 and the higher fullerenes resemble<br />

those of C 60 (Figures 5.26 and 5.27). However, the lower symmetry and larger<br />

number of atoms result in an increase of the number of states with lower degeneracy,<br />

and the spectral patterns are less resolved than those in C 60 exhibiting many<br />

overlapping bands. Meanwhile, the spectra still exhibit individual characteristic<br />

Figure 5.27 UPS and C1s EELS spectra of selected fullerenes [193] (middle panels) compared<br />

to the DOS (left) and DUOS (right) computed at the PBE/TZ2P level [182]. C 84 is a mixture of<br />

D 2 (22) and D 2d (23) isomers in 2 : 1 ratio.<br />

245


246 5 Fullerenes<br />

features, resembling peculiarities of the electronic structure of each fullerene,<br />

which can be best demonstrated by significant difference observed in the DOS<br />

and UPS spectra of two C 2v isomers of C 78 [180, 192, 193].<br />

5.3.2.4 Electronic Excitations. UPS, UV/Vis/NIR Absorption and Fluorescence<br />

Spectroscopy<br />

Excitation spectra of the fullerenes in the near-infrared, visible and near-UV<br />

range are attributed to the �–�* transitions and are very structure sensitive.<br />

The lowest energy transition, which can be ascribed to the excitation through<br />

the HOMO-LUMO gap, is an important fundamental property determining the<br />

kinetic stability of the given fullerene. Only molecules with a considerably large<br />

HOMO-LUMO gap can be isolated from the soot by conventional organic solvents,<br />

while structures with a small gap, albeit they may be formed in considerable<br />

amount and be thermodynamically stable, cannot be extracted from the soot,<br />

presumably because they have polymerized. C74, D3h(5)-C78 or Td-C76 are examples<br />

of such ‘insoluble’ fullerenes, whose formation in the arc-discharge synthesis was<br />

proven by electrochemical or chemical transformation to the soluble forms [199].<br />

Excitation spectra of fullerenes have been studied by many techniques, including,<br />

but not limited to, low-energy EELS, core-level XPS (satellite structure of C1s peak),<br />

UPS of the anionic species, and optical absorption and fluorescence spectroscopy.<br />

Description of the results obtained by EELS and XPS are beyond the scope of this<br />

review (for a review, see Refs. [180, 192, 193]), and here we mainly focus on UPS<br />

and optical spectroscopy of fullerenes.<br />

The UPS of the gas-phase anions is an important information source for two<br />

kinds of fundamental properties of fullerenes. First, the lowest energy band in<br />

the UPS spectrum corresponds to the detachment of the electron from the singleoccupied<br />

orbital of the anion and therefore corresponds exactly to the electron<br />

affinity. Secondly, photodetachment of the electron from the doubly-occupied<br />

orbital results in singlet or triplet excited state of the neutral molecule. Thus, UPS<br />

spectrum of the anion provides information on the excited states of the fullerene<br />

and is especially useful in the determination of the lowest energy excitations. As<br />

UPS measurements of the gas-phase anions require very small amounts of the<br />

sample, while the mass-selection technique allows the mixtures of fullerenes to<br />

be studied without their preliminary separation, it is not surprising that the first<br />

study of the anions of C60 and C70 fullerenes by UPS was done in 1987 [200],<br />

before the method of their bulk-production was discovered. Likewise, UPS is the<br />

only method of choice for the studies of the electronic structure of exotic small<br />

fullerenes C20 –C50 (Figure 5.28) [201, 202]. Thus, C32 has been shown to have a<br />

large HOMO-LUMO gap and is therefore expected to be stable [201]. Vibrationally<br />

resolved UPS spectra of C – 60 [198], and C– 70 [203], cooled in the ion trap are shown<br />

in Figure 5.28. The second and the third lowest energy bands in the spectrum of<br />

C –<br />

60 are attributed to the first triplet and singlet excited states of C60 . The energy<br />

of the former is estimated as 1.62 ± 0.01 eV, while the multiplet splitting (i.e. the<br />

energy gap between triplet and singlet states derived from the same excitation)<br />

is 0.2–0.3 eV.


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Figure 5.28 UPS spectra of anions of some fullerenes; vibrationally resolved UPS spectra of C –<br />

60<br />

[198], and C –<br />

70<br />

[203]. ‘A. D.’ in the spectrum of C–<br />

60<br />

marks an auto-detachment band.<br />

Though UPS remains an invaluable source of information on the structures<br />

whose isolation in bulk amounts is not possible, the method requires rather<br />

complex equipment, while the information it gives is actually limited to the lowest-energy<br />

excitation, because at the higher energies the density of excited states<br />

of fullerenes is usually very high. For this reason, the UV/Vis/NIR absorption<br />

spectroscopy is much more advantageous now, when at least > 1 mg amounts<br />

of the fullerenes have become available. Simple relatively inexpensive instrumentation<br />

and sampling techniques make this method accessible to virtually<br />

all researchers worldwide, and the richness and uniqueness of �–�* excitation<br />

spectrum for each fullerene has made it almost a standard now that the characterization<br />

of any new fullerene cannot be complete without the UV/Vis/NIR<br />

absorption spectrum.<br />

Room-temperature UV/Vis absorption spectrum of C 60 in n-hexane is shown in<br />

Figure 5.29. Three regions with the intensities covering three orders of magnitude<br />

can be roughly distinguished in the spectrum: low-intensity highly structural<br />

bands in the visible range (450–700 nm), several more intense sharp peaks<br />

around 380–420 nm, and finally three very strong broad absorptions at 328 nm<br />

(log(� max ) = 5.20), 256 nm (5.24) and 211 nm (5.20) in the UV range [204]. As only<br />

the excitation to the singlet states of T 1u symmetry are optically allowed for C 60 ,<br />

its HOMO � LUMO (h u � t 1u ) transition is optically forbidden, giving rise to<br />

the excited states of T 1g , T 2g , G g , and H g symmetry. The lowest energy 1 T 1u state<br />

can be attributed to the sharp feature at 408 nm (3.04 eV) in the experimental<br />

247


248 5 Fullerenes<br />

Figure 5.29 Left panel: Room-temperature UV/Vis absorption spectrum of C 60 in n-hexane.<br />

Insets show low temperature ‘absorption’ spectra: resonant two-photon ionization spectrum<br />

in ultrasonic beam [210] (R2PI); fluorescence-excitation spectrum in Ne matrix at 4 K [175],<br />

absorption spectrum in helium droplets at T~0.4 K [207]. Vertical bars show excitation energies<br />

and oscillator strengths predicted by TD-DFT [205]. Right panel: Low temperature fluorescence<br />

spectra of C 60 in different matrices [175, 208].<br />

spectrum. Interpretation of the absorption spectrum of C 60 in the UV range was<br />

given in 1992 by Leach [204]. A new assignment for some bands was proposed by<br />

Bauernschmitt [205] in 1998 with the use of more refined time-dependent (TD)<br />

DFT calculations (see Figure 5.29 for the comparison of TD-DFT computed and<br />

experimental spectra of C 60 ).<br />

Though HOMO � LUMO transitions in C 60 are symmetry forbidden, they<br />

may be activated through a Herzberg–Teller (HT) coupling to the appropriate<br />

molecular vibrations. In due turn, each HT-activated transition is an origin for<br />

the Franck–Condon (FC) and Jahn–Teller (JT) progressions via coupling to the A g<br />

and H g modes, respectively. Altogether, these multiple vibronic excitations form<br />

a complex pattern observed in the experimental spectrum of C 60 in the visible<br />

range. There have been numerous efforts, both theoretical and experimental, to<br />

interpret the low-energy excitation spectrum of C 60 (see Ref. [176] for a detailed<br />

review). Vibrationally resolved data on cold C 60 molecules were obtained by twophoton<br />

ionization spectroscopy in supersonic beam [206], absorption spectra<br />

of C 60 in He droplets (T � 0.4 K) [207], or by measuring the fluorescence and<br />

fluorescence excitation spectra in various matrices or single-crystal C 60 at low<br />

temperatures (T = 1.2–5 K) [175, 176, 208]. The survey of experimental spectra<br />

is presented in Figure 5.29. Semiempirical (CNDO/S, QCFF/PI) and TD-DFT<br />

calculations predict that three lowest energy excited states, 1 T 1g , 1 T 2g and 1 G g at<br />

ca 1.7–2.2 eV are quasi-degenerated within 0.05 eV. Excitations to 1 T 1g state are<br />

activated by A u , H u , and T 1u vibrational modes, to 1 T 2g state – by G u and H u modes,<br />

and to 1 G g state – by T 2u , G u , and H u modes, and therefore vibrationally resolved


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

excitation and especially fluorescence spectra provide additional information on<br />

the silent fundamentals of C 60. Detailed analysis of the vibronic structure based<br />

on the CNDO/S calculations was reported by Negri [175, 209]. It was shown that<br />

the lowest excited state is 1 T 1g with the energy of 1.93–1.94 eV depending on the<br />

matrix and temperature, while 1 T 2g and 1 G g states are ca 10 and 50 cm –1 higher<br />

in energy.<br />

The survey of available absorption spectra for the major isomers of the higher<br />

fullerenes C 70-C 84 is given in Figure 5.30. It can be seen that the spectra are highly<br />

structural and specific for each fullerene. Significant variation of the spectra with<br />

the isomeric structures of the same molecular size observed in C 78 [186, 211], C 80<br />

[212, 213] and C 84 [212, 213] is especially remarkable as it shows that UV/Vis/<br />

NIR absorption spectroscopy may be used to distinguish the fullerene isomers.<br />

In 1998 Bauernschmitt [205] reported that TD-DFT calculations gave very good<br />

agreement with the experimental absorption spectra and thus can be used to<br />

determine the isomeric structures of newly isolated fullerenes when assignment<br />

based on 13 C NMR is ambiguous.<br />

IR and Raman spectra of isomerically pure C 2 (3)-C 82 fullerene have been recently<br />

reported [590].<br />

Figure 5.30 Room-temperature UV/Vis/NIR absorption spectra of isomerically pure fullerenes<br />

[186, 205, 211–213] (for the isomers of a given C 2n , the spectra are given in the order of<br />

their yield; the ratio of C 80 -D 2 (2) to C 80 -D 2d (1) is ca 30 : 1 [213]; for C 84 isomers, the ratio of<br />

D 2(22) : D 2d(23) : C 2(11) : C s(16) : C s(14) : D 2d(4) : D 2(5) is 1 : 0.5 : 0.2 : 0.1 : 0.1 : 0.05 : 0.03 [214]).<br />

Enhanced spectra in the low-energy range are given as additional curves.<br />

249


250 5 Fullerenes<br />

5.3.3<br />

Nuclear Magnetic Resonance<br />

Toni Shiroka<br />

5.3.3.1 Introduction<br />

Nuclear magnetic resonance (NMR) represents the selective absorption of electromagnetic<br />

radiation by nuclei with nonzero spin placed in an external magnetic<br />

field. The dependence of its main parameters (such as the resonance frequency<br />

� and width ��, the spin–lattice T 1 and spin–spin T 2 relaxation times) from the<br />

nucleus surroundings, makes it a highly sensitive probe to the immediate environment<br />

and, therefore, an extremely versatile tool of spectroscopic investigation.<br />

The magnetic resonance spectra can broadly be classified into two categories:<br />

high-resolution solution NMR and solid-state NMR [215, 216]. In the first case,<br />

most appropriate for structural analysis of single molecules, the material is<br />

dissolved in suitable solvents, where the interactions that cause the broadening of<br />

spectral lines are generally averaged to zero as a result of the random molecular<br />

motion. The relative differences in resonance frequency with respect to a reference<br />

nucleus, known as chemical (orbital) shifts, reflect the different electronic screening<br />

of the external field in various chemical environments.<br />

The same principles are also applicable to solid-state NMR. However, the restricted<br />

molecular motion in solids can result in very broad spectra. For metallic<br />

samples moreover, there is an additional contribution to the frequency shift,<br />

arising from the spin polarization of the conduction electrons, known as Knight<br />

(spin) shift [217]. The latter implicitly includes a contact hyperfine interaction<br />

term, due to electrons in s orbitals, and a dipolar contribution arising from non-s<br />

electron spins, respectively related with the isotropic and anisotropic part of the<br />

shift. Advanced techniques, including magic-angle spinning, cross polarization,<br />

spectral editing, etc., allow nevertheless to recover a wealth of information even<br />

from the overcrowded solid-state spectra, thus successfully complementing or<br />

supporting other data.<br />

Two other techniques, closely related to NMR, are the nuclear quadrupole<br />

resonance (NQR) [218] and the muon spin rotation (µSR) [219, 220], both of<br />

which have also contributed to our knowledge of doped fullerenes. NQR relies<br />

on the fact that in many solids, the spectra of nonspherical nuclei (spin I > 1/2)<br />

are dominated by quadrupole effects, arising from the interaction of the nuclear<br />

electric quadrupole moment with the nonspherically symmetrical field gradient<br />

generated by the surrounding electrons. The NQR spectra can be collected even<br />

in zero magnetic field and often provide decisive information about structural<br />

properties and dynamics. µSR on the other hand makes use of 100% polarized<br />

muons, elementary particles similar to electrons, but positively charged, hence<br />

chemically behaving as light isotopes of hydrogen. Once implanted in matter,<br />

muons precess in the local magnetic field and information on both precession<br />

and relaxation is delivered by the decay positrons, emitted preferentially along<br />

the muon spin direction. The main µSR advantages with respect to conventional


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

NMR consist in the higher sensitivity, the complementary (interstitial) ‘point of<br />

view’, and the possibility to perform measurements even in the absence of an<br />

applied magnetic field.<br />

5.3.3.2 NMR of Fullerenes<br />

To illustrate the NMR contribution in the study of fullerene properties we resort<br />

to some selected representative examples which, however, are not intended to be<br />

either definitive, or exhaustive. For a detailed account of the field the interested<br />

reader should refer to the many excellent reviews [221–224].<br />

Depending on the unit being investigated: individual molecules, or ensembles<br />

of them, the reported examples are conveniently grouped accordingly, roughly<br />

corresponding to solution NMR and solid-state NMR experiments, respectively.<br />

Individual Fullerene Property Studies. The first data to support the highly symmetric<br />

structure of C 60 were due to NMR [225]. The exact equivalence of all the carbon<br />

atoms, arising from the icosahedral I h symmetry, implies a single narrow line for<br />

the 13 C solution NMR spectrum. This is indeed the case, as shown in Figure 5.31a,<br />

where a remarkable number of atoms gives rise to a single line resonance, an<br />

Figure 5.31 (a) 13 C NMR spectra of pure C 60 , (b) of a mixture of C 60 with C 70 (b), and<br />

(c) of pure C 70 . Notice the presence of five different lines in the symmetry reduced C 70 ,<br />

whose intensity ratio 10 : 10 : 20 : 20 : 10 reflects the five sets of nonequivalent carbons.<br />

(Reprinted with permission from [225]).<br />

251


252 5 Fullerenes<br />

unprecedented feature for such a large molecule. Despite its simplicity, this early<br />

spectrum is even more revealing. Indeed, the observed chemical shift, 142.7 ppm<br />

from a tetramethyl silane (TMS) reference, is significantly lower than the peaks<br />

for the corresponding positions in aromatic compounds such as benzene<br />

(120 ppm) or naphthalene (133.7 ppm). On the other hand, it is well known that<br />

the strain-induced hybridization produces a downfield shift, as seen for example<br />

in the 13 C peaks for the bridgehead carbons in indane C 9H 10 (143.9 ppm) and<br />

benzocyclobutene C 8H 6 (146.3 ppm). The observation of a similar shift also in<br />

C 60 represented a beautiful demonstration of its curved geometry. This first,<br />

simple NMR experiment contributed decisively to the elucidation of the fullerene<br />

structure, consisting of pentagonal and hexagonal carbon rings.<br />

Even more interesting is the case of C 70, where the icosahedral symmetry is reduced<br />

to a lower D 5h symmetry. Accordingly, its solution NMR spectrum consists<br />

of five different lines [225], reflecting the five nonequivalent carbon atom sites, as<br />

shown in Figure 5.31c. More complex, two-dimensional NMR experiments, making<br />

use of double-quantum coherence in 13 C enriched samples, were able to examine<br />

also the nature of the bonding and to determine the bonding topology [226].<br />

The observation of an isotropic Knight shift, K iso , in the NMR spectra of conducting<br />

fullerenes provides another, independent confirmation of the curved geometry<br />

of the single fullerene units [227–230]. Indeed, since the probability density of<br />

p-electrons at the nucleus site vanishes, they cannot contribute to the isotropic<br />

Knight shift. In the case of a perfect sp 2 hybridization, where neither the sp 2 orbitals,<br />

nor the residual (non hybridized) p z orbitals overlap with the nucleus, K iso should<br />

be exactly zero. In the presence of curvature, though, this simple picture will<br />

change. On curved carbon surfaces, the normally planar sp 2 orbitals (orthogonal<br />

to the axis defined by the p z orbital) must tilt somewhat to match the neighboring<br />

carbons (see Figure 5.32a). This distortion is achieved by a small admixture of<br />

the 2s wave function into the 2p z orbital, producing a state whose hybridization is<br />

intermediate between the planar sp 2 and the tetrahedral sp 3 . The same process can<br />

be described also as a small �–� hybridization, if the overlap of neighboring carbon<br />

atom orbitals is considered. The rehybridization process on curved surfaces will<br />

therefore always retain a small 2s character, resulting in a nonzero direct Fermi<br />

contact interaction and hence in a small isotropic Knight shift, whose magnitude<br />

depends on the degree of curvature. NMR measurements in a metallic compound,<br />

such as Rb 3 C 60 [228], fully confirm this picture by providing an isotropic Knight<br />

shift value of ~44 ppm, less than half the anisotropic shift value. Much smaller<br />

shifts, ~10 ppm, inversely proportional to the tube diameter, are foreseen for the<br />

conducting nanotubes, where the curvature is present only along one direction<br />

[231] (carbon nanotubes will be discussed in Chapter 6).<br />

Quantitative predictions about the degree of orbital mixing, based on the<br />

so-called ‘orbital following’ assumption, initially assigned a ~9% 2s character to<br />

the fullerene molecule, with a pyramidalization angle of ~12% [232, 233]. The<br />

resulting hyperfine coupling, together with the measured susceptibility values,<br />

predicted nevertheless an isotropic Knight shift value (K = A · �) 10 times higher<br />

than that actually measured! The solution to this puzzle came again from NMR


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Figure 5.32 Two possible carbon–carbon bonding schemes in C 60 , depicting one of the three<br />

2sp 2 orbitals, together with the non hybridized 2p z orbital. (a) Diagram showing the ‘orbital<br />

following’ bonding scheme, with the hybridized orbitals inclined with respect to the radially<br />

oriented 2p z orbitals. (b) In the more realistic ‘bent’ � bonding, the hybridized orbitals remain<br />

nearly perpendicular to the respective 2p z orbitals, resulting in a smaller 2s admixture, thus<br />

accounting for the anomalously small isotropic Knight shift observed.<br />

experiments. From the 13 C coupling constants in a 2D experiment, Hawkins and<br />

co-workers [234] could estimate a fractional rehybridization of 3% in neutral C 60<br />

and, more importantly, established that the correct carbon bonding model is that of<br />

‘bent’ �-bonding and not the stiff orbital following (see Figure 5.32b). This model,<br />

which is the currently accepted one [224], resulted also in the correct K iso values.<br />

Because it has a nearly spherical shape, at room temperature C 60 maintains a<br />

high reorientation rate, even in the solid state (vide infra discussed also in the<br />

preceding sections). The resulting static and dynamic disorder makes diffraction<br />

investigations problematic, giving access only to the average charge distribution<br />

and precluding the details of the internal molecular structure. Carr-Purcell solidstate<br />

NMR measurements of 13 C– 13 C magnetic dipolar coupling, on the other<br />

hand, provided the first experimental determination of bond lengths in C 60 [235].<br />

This particular pulse sequence can selectively remove the broadening effects due<br />

to chemical shift anisotropy (differences in local magnetic field shielding in a<br />

powder sample), but retains the dipole–dipole coupling.<br />

Since the dipolar interaction strength depends on the inverse cube of the C–C<br />

distance, the method not only can distinguish between the 1.40 and 1.45 Å bond<br />

lengths, but can assess them with better than 1% precision! Only at a later time,<br />

as described in Section 5.3.1, could low-temperature X-ray studies determine the<br />

bond lengths in C 60 even more accurately.<br />

Figure 5.33 shows both the measured and the simulated powder spectra at<br />

a temperature sufficiently low, so that molecular rotation does not average the<br />

~4 kHz dipolar coupling between directly bonded carbons. The very large central<br />

peak (cropped for clarity) arises from 13 C nuclei with no 13 C nearest neighbors,<br />

and is flanked by two weak doublets with different intensity and frequency, arising<br />

from the directly bonded 13 C atoms. As expected, the inner doublet, due to 60<br />

pentagon bonds, is twice as intense as the external one, due to 30 bonds between<br />

pentagons. This fully confirms the truncated icosahedron geometry, since an early<br />

proposed structure, consisting in a truncated dodecahedron, yields a rather poor<br />

agreement with the experimental data (Figure 5.33, panel c).<br />

253


254 5 Fullerenes<br />

Figure 5.33 (a) Fourier transform of the 13 C signal obtained in a Carr–Purcell sequence on<br />

13 C enriched fullerene sample at 77 K. The intense center line has been cropped for clarity<br />

(see text). (b) Simulation of two Pake doublets with carbon–carbon bond lengths of 1.45 and<br />

1.40 Å; (c) as in (b), but with bond lengths of 1.451 and 1.345 Å. (Reprinted with permission<br />

from [235]).<br />

To conclude this section on individual fullerene properties, we present an<br />

ingenious way to follow the fullerene reactions by monitoring the NMR signal<br />

of their endohedrally labeled versions [236]. Although 13 C NMR spectra of C 60<br />

and C 70 are simple (see above), the spectra of the reaction products are rather<br />

complex, since the attachment of groups to the fullerene skeleton reduces the<br />

initially high symmetry. The study of these spectra is not only time consuming<br />

but, especially when many products are involved, it presents serious difficulties<br />

in assigning the carbon peaks. The 1 H NMR signal of the ubiquitous protons can<br />

in principle help. However, the tendency of solvents to be trapped in the fullerene<br />

lattice, and the hydrogen presence also in the reagents and byproducts makes it<br />

unpractical. On the other hand, (as discussed also in Section 5.6 on hydrogenated<br />

fullerenes) the use of endohedral fullerene complexes of noble gases, as e.g. 3 He,<br />

does not have any of the previous drawbacks. Each helium-labeled fullerene gives<br />

a single sharp peak with a typical line width below 1 Hz and a relatively large<br />

negative chemical shift (–6.3 ppm and –28.8 ppm from a 3 He reference in C 60<br />

and C 70 respectively). The diamagnetic chemical shifts of the centrally located<br />

helium nucleus, not only prove the substantial aromatic ring currents present<br />

in all the fullerenes, but their modifications can be used as a reaction indicator.<br />

Indeed, the conjecture that alterations of the �-bonding structure of the fullerene<br />

through reaction would produce substantial shifts in the 3 He peak, has been<br />

successfully verified experimentally. A simple addition reaction, with just one of<br />

the 30 double bonds being modified, produced a strikingly large –3 ppm shift of<br />

the helium signal. Considering also that in this case the peak areas are accurate<br />

indicators of the relative amounts, and that to date no two different products have<br />

been found to have the same helium shift, the potentialities of the method as a<br />

reaction monitoring tool are clear.


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Collective Fullerene Property Studies. While solution NMR measurements can<br />

reveal properties of single molecules, solid-state NMR is an excellent probe of the<br />

collective behavior of fullerenes. This, together with the fact that both pure and<br />

doped fullerenes are available mainly as powders, explains the extremely vast and<br />

diversified range of experiments belonging to this class.<br />

The investigation of C 60 rotational dynamics in the solid state, represents<br />

perhaps one of the best known examples of this kind of studies. Figure 5.34 shows<br />

the evolution of 13 C NMR line shape of fullerene, as the temperature is lowered<br />

[237, 238]. Surprisingly, at room temperature, instead of a broad and featureless<br />

spectrum, typical of a powder sample, C 60 shows an extremely narrow line, very<br />

similar to those observed in a solution. This extreme ‘motional narrowing’ effect<br />

implies a rapid and isotropic molecular reorientation which, successive relaxation<br />

rate studies have shown to be even more effective than in the liquid phase. At<br />

the lowest temperature the molecules are stationary but randomly oriented, and<br />

Figure 5.34 13 C NMR spectra of solid fullerene taken at different temperatures. The surprisingly<br />

narrow line at room temperature gradually broadens and changes its shape, as the isotropic<br />

rotational motion slows down. (Reprinted with permission from [237]).<br />

255


256 5 Fullerenes<br />

therefore subject to different magnetic field shieldings (technically, to a chemical<br />

shift anisotropy). A fit of the powder pattern yields an asymmetric chemical shift<br />

tensor (220, 186, 25) [237], quite similar to those of aromatic carbon compounds.<br />

The coexistence, at intermediate temperatures, of sharp features superimposed<br />

on a broad spectrum indicates the simultaneous presence of two different phases:<br />

a fast rotating phase with another one where the rotations are blocked. To find<br />

more information about them one can resort to the spin–lattice NMR relaxation<br />

rates. The motional dynamics affects the intensity of magnetic field fluctuations<br />

which, in turn, will be reflected in the measured relaxation rates (1/T 1).<br />

The numerical analysis of room temperature data provides an astonishingly fast<br />

rotational correlation time, � = 9.2 ps, only three times longer than that measured<br />

in the gas phase! Some fascinating, recent attempts towards applications of this<br />

finding, although somewhat speculative, include the use of C 60 as a molecular<br />

ball bearing for nanoscale machine design [239].<br />

Temperature dependent X-ray and neutron diffraction studies [240] hinted<br />

at a structural fcc–sc phase transition at 249 K, but little was known about the<br />

details. An NMR examination of the transition [241, 242], could find a sharp drop<br />

in the T 1 relaxation times, despite no appreciable change in the spectrum shape.<br />

These measurements suggested a high-temperature phase, characterized by free<br />

rotation (‘rotator’ phase), and a low-temperature (‘ratchet’) phase, characterized<br />

by rotational jump diffusion among symmetry-equivalent orientations. Due to<br />

the hampered rotations, the ‘ratchet’ phase shows also much higher relaxation<br />

rates (i.e. shorter relaxation times), and its slower dynamics seems to arise from<br />

an orientational ordering, such that electron-rich interpentagon bonds are facing<br />

electron-poor pentagon centers of neighboring C 60 molecules. This also explains<br />

the lowered structural symmetry (sc instead of fcc) of this phase.<br />

The strong tendency toward polymerization of fullerene molecules into one-,<br />

two-, and even three-dimensionally connected polymers, either by photoexcitation,<br />

or by high-temperature and high-pressure treatments, has attracted a continuous<br />

scientific interest. From the early studies of relatively simple monodimensional<br />

chains in AC 60 (A = K, Rb, Cs) [243], to RbC 60 dimers [244], up to the most recent<br />

elucidation of a particularly complex polymerization in Li 4 C 60 [245], NMR has<br />

always played a key role. This class of experiments has in common the successful<br />

use of magic angle spinning (MAS) [246] to simplify the extremely broad powder<br />

pattern arising from carbon polymerization [247]. The sample spinning produces<br />

an effect very similar to the motional narrowing, by suppressing all the anisotropic<br />

contributions from the relevant interactions.<br />

Figure 5.35 shows the 13 C NMR MAS spectrum of Li 4 C 60 , together with its<br />

structure, consisting in a 2D planar polymerization pattern, characterized by the<br />

coexistence of single and ‘double’ C–C bonds propagating along two orthogonal<br />

directions. The spectrum is broadly divided into two regions, whose frequency<br />

domains are commonly attributed to sp 2 (~150 ppm) and to sp 3 (~60 ppm) hybridized<br />

carbon atoms. The fullerene distortion induced by the polymerization<br />

removes the equivalence of the sp 2 atoms, giving rise to a multitude of isotropic<br />

chemical shifts. The two distinct peaks in the sp 3 region, instead, are reminiscent


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Figure 5.35 Room temperature 13 C MAS (8 kHz) spectrum of Li 4 C 60 showing the different<br />

intensities for sp 2 and sp 3 carbon atoms. The two peaks in the sp 3 region arise from two<br />

different types of C 60 polymerization along a and b directions (see structure in inset).<br />

(Reprinted with permission from [245]).<br />

of the two different kinds of covalent bonds, as confirmed both by the dissimilar<br />

shifts and by the 4 : 2 integrated intensity ratio. Despite the abundance of bonds,<br />

conventional NMR measurements (not shown) did not find any trace of Knight<br />

shifts (either isotropic, or anisotropic), thus demonstrating the insulating character<br />

of the polymer.<br />

We dedicate the last example to the A 3 C 60 (A = K, Rb, Cs) fulleride superconductors,<br />

because their NMR investigation represents also a remarkable example<br />

of electronic property studies [224, 248, 249].<br />

Contrary to endohedral fullerene complexes involving He discussed above,<br />

the solid state fullerides consist of empty fullerenes, with the alkali ions residing<br />

in the voids of the fullerene lattice. Since the superconductivity in fullerides is<br />

widely accepted to be of phononic nature, the most important NMR method in<br />

their studies has naturally been the measurement of the spin-lattice relaxation<br />

times T 1 . Differently from the insulating undoped C 60 , where the main relaxation<br />

mechanisms are the molecular reorientation and the chemical shift anisotropy<br />

(see above), in the metallic A 3 C 60 family the most important mechanism is the<br />

so-called Korringa relaxation [250]. It stems from the significant electron-nucleus<br />

hyperfine interaction in metals which, therefore, provides also the dominant relaxation<br />

channel, with the nuclei being restored to their ground state level through<br />

spin-flip exchanges with electrons. Since the only electronic states available for the<br />

spin-flip transitions are those within kT of the Fermi energy level, the resulting<br />

relaxation rate is proportional to temperature, i.e. [216]:<br />

257


258 5 Fullerenes<br />

1<br />

T<br />

1<br />

� kT<br />

= ⋅ A N( EF)<br />

�<br />

2 2<br />

, (5.1)<br />

with A the hyperfine coupling constant, and N(E F) the electronic density of states at<br />

Fermi level. Taking into account the expression for the (temperature independent)<br />

Knight shift, K = A �, it can easily be shown that (Korringa relation):<br />

1 2<br />

n<br />

2<br />

2<br />

� �e<br />

1<br />

T T = ⋅ ⋅ , (5.2)<br />

4 � k � K<br />

with � e and � n the gyromagnetic ratios of the electron and the nucleus. Since the<br />

spin-lattice relaxation in alkali fullerides is dominated by the dipolar, rather than<br />

the contact hyperfine interaction [251], the Korringa relation should include a<br />

scaling factor of ~3 [229]. However, provided powder averages are considered [252],<br />

the above relation is still valid.<br />

In the normal metallic state, we would therefore expect a temperature independent<br />

1/(T 1 T) factor and, to a very good extent, this is what one really observes in<br />

many alkali-metal fullerides [228]. More interesting, but also more challenging,<br />

were the experiments performed in the superconducting state. Here, the gradual<br />

formation of the Cooper pairs implies an exponential increase of the relaxation<br />

times. This behavior not only was observed, but it could be also used to evaluate<br />

the low-temperature superconducting gap � 0 [228]. The final, and for a long time<br />

puzzling, issue was the absence of a Hebel–Slichter peak [253] in the transition<br />

to the superconducting state. This peak, which appears as a slight upsurge in the<br />

1/(T 1 T) data vs. temperature, is due both to the electronic state ‘pile up’ during<br />

the opening of the superconducting gap and to an electronic coherence factor.<br />

Since it is universally regarded as a strong indicator of BCS superconductivity, its<br />

first observation in Rb 3 C 60 [254] (see Figure 5.36) using the muon-spin relaxation<br />

technique (µSR), and, successively, also in other compounds [255], definitely<br />

Figure 5.36 Muon -spin relaxation (µSR) measurement of 1/(T 1 T) in Rb 3 C 60 at 1.5 T.<br />

The solid line is a fit to the Hebel–Slichter theory, with a broadened density of states.<br />

The almost constant 1/(T 1 T) value above the critical temperature corresponds to<br />

normal metallic behavior. (Reprinted with permission from [254]).


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

confirmed the BCS nature of alkali-fulleride superconductors. Subsequent NMR<br />

measurements [256], performed at relatively low fields, could reproduce these<br />

results and explain the absence of a peak in early data by its suppression by the<br />

high field.<br />

5.3.3.3 Concluding Remarks<br />

Magnetic resonance spectroscopy, both in its solution- and solid-state NMR<br />

forms, has proved an extremely useful tool for the investigation of the structural,<br />

dynamical and electronic properties of fullerenes, which rivals in importance and<br />

successfully complements techniques such as X-ray and neutron diffraction. Even<br />

simple, one-dimensional NMR experiments, carried out on static samples, can<br />

provide useful and often decisive information, otherwise difficult or impossible<br />

to obtain by other methods.<br />

The elucidation of fullerene structure, the study of its rotational dynamics in the<br />

solid state, the investigation of the different electronic properties of doped fullerenes,<br />

and the study of carbon polymerization are just a few of the most important<br />

contributions of NMR to the fullerene field. Even though current research interests<br />

have gradually shifted to nanotubes and similar classes of materials, the investigative<br />

role of NMR remains central in our continuously advancing understanding<br />

of these unusual carbon forms.<br />

5.3.4<br />

Electrochemistry<br />

Renata Bilewicz and Kazimierz Chmurski<br />

5.3.4.1 Electronic Properties of Fullerenes<br />

Fullerenes and their derivatives are unique in the richness of their redox behavior<br />

and electrochemical reactions, and every year new publications including reviews<br />

and monographs appear highlighting various electrochemical aspects of these<br />

fascinating molecules [257–262]. Since early research on the electrochemistry of<br />

C60 to C84 , their isomers and their derivatives has been summarized in excellent<br />

chapters of Echegoyen et al. [261, 263], in the following chapter we focus mainly<br />

on more recent work devoted to the relationship between the fullerene structure<br />

and their electrode behavior.<br />

Fullerenes exhibit a triply degenerated LUMO (t1u ) of low energy and, therefore,<br />

behave as an electronegative molecule reversibly accepting up to 6 e – [264–265].<br />

Electron affinities, EA, and ionization potentials, IP, calculated for C60 to C84 [263]<br />

are shown in Table 5.3.<br />

In early reports, only two or three reduction steps were recorded for C60 using<br />

cyclic voltammetry [264], however, in mixed solvents and lowered temperatures<br />

six reversible 1 e – reductions can be easily resolved (Figure 5.37) [266]. For C70 ,<br />

3–<br />

all six waves are seen at room temperature and starting from C70 the reduction is<br />

easier than that of corresponding C60 anions. This was explained by Cox in terms<br />

of larger size of C70 and charge separation delocalization model [267].<br />

259


260 5 Fullerenes<br />

Table 5.3 Estimated electron affinities and ionization potentials of selected fullerenes [263].<br />

(isomer) EA (eV) IP (eV)<br />

C 60 2.7 7.8<br />

C 70 2.8 7.3<br />

C 76 (D 2 ) 3.2 6.7<br />

C 78 (C 2v ) 3.4 6.8<br />

C 82 (C 2 ) 3.5 6.6<br />

C 84 (D 2 ) 3.5 7.0<br />

C 84 (D 2d ) 3.3 7.0<br />

Figure 5.37 Reduction of C 60 in CH 3 CN/toluene at –10 °C using cyclic voltammetry at a<br />

0.1 V s –1 scan rate and differential pulse voltammetry (0.050 V pulse width, 0.3 s period,<br />

0.025 V s –1 ) upper and lower curves, respectively. (Adapted from [266]).<br />

The interactions of solvent molecules with fullerenes and fullerides are significant<br />

mainly due to the size of the fullerene molecule compared to those of the<br />

solvent, hence, large solvent–fullerene interaction surface [268–274]. The surface<br />

exposes a strained �-orbital system consisting of sp 2 orbitals with enhanced s<br />

n–/(n + 1)–<br />

character. The formal redox potentials of the C60 couples become more<br />

solvent–dependent as the charge increases due to stronger electrostatic interactions<br />

with solvent dipoles and the hydrogen bonding interactions with the fullerides of<br />

increasing basicity (Table 5.4).


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

n–/(n + 1)–<br />

Table 5.4 Formal reduction potentials of C60 couples referenced to the reduction<br />

potential of the Me10Fc +/0 couple in 0.1 M tetra-n-butyl ammoniumperchlorate (TBAClO4 ) [273].<br />

0/1–<br />

Solvent C60 1–/2–<br />

C60 2–/3–<br />

C60 3–/4–<br />

C60 acetonitrile –735 –1225 –1685<br />

aniline b)<br />

–396 c)<br />

–693 c)<br />

–1158 c)<br />

–1626 d)<br />

benzonitrile –397 –817 –1297 –1807<br />

benzyl alcohol b)<br />

bromobenzene b)<br />

–443 c)<br />

–587 d)<br />

–817 c)<br />

chlorobenzene –573 –953 –1438<br />

chloroform b)<br />

1,2-dichlorobenzene b)<br />

–554 d)<br />

–535 c)<br />

–908 d)<br />

–907 c)<br />

–1360 c)<br />

1,2-dichloroethane –448 –848 –1298<br />

–1841 d)<br />

dichloromethane –468 –858 –1308 –1758<br />

N,N-dimethylaniline b)<br />

–547 c)<br />

dimethylformamide –312 –772 –1362 –1902<br />

N-methylaniline b)<br />

–442 d)<br />

nitrobenzene –406<br />

–782 d)<br />

pyridine –343 –763 –1283 –1813<br />

tetrahydrofuran –473 –1063 –1633 –2133<br />

a)<br />

From Ref. [268]. Data recorrected to the Me10Fc +/0 couple.<br />

b)<br />

Determined in the present study.<br />

c)<br />

CV data.<br />

d)<br />

DPV data.<br />

Large interactions of the charged fullerides with the solvent are explained in<br />

terms of solvent dependent Jahn–Teller distortions [273, 275] and large quadrupole<br />

moments [276]. �-stacking interactions between C 60 and aromatic solvents<br />

result in a substantial contribution of the polarizability correction term to the first<br />

reduction. Formation of ion pairs between the more electron-rich fullerides and<br />

the cations of the supporting electrolyte add to the complex solvent/solute interactions<br />

influencing also the values of reduction potentials. Interactions with cations<br />

are especially important in halogenated and nitrile solvents for which slowing<br />

diffusion rates, slower kinetics and changes of reduction potentials were reported<br />

261


262 5 Fullerenes<br />

[274, 277–279]. Dimerization of C – 60 may occur in more concentrated solutions,<br />

+ +<br />

and in the presence of NBu4 cations, the salts of [NBu4 ][C60][C – 60] stoichiometry<br />

can be formed [277, 280]. Fairly large solvent/solute interactions observed were<br />

shown to arise from contributions of different, relatively weak interactions [273<br />

and refs therein]. Attempts to compensate for the effects of counterions included<br />

the use of microelectrodes instead of conventional size electrodes which allowed<br />

one to work without added supporting electrolyte [281]. This useful alternative<br />

approach should be used with caution since unknown residual impurities in the<br />

solution may play the role of counter ions, and affect the experimentally obtained<br />

values of potentials. In the analysis of solvent and counter ions effects on the<br />

electroreduction potential of fullerenes, the quality of the reference electrode or<br />

of the internal redox standard are of major importance. Decamethylferrocenium/<br />

decamethylferrocene couple was proposed as favorable internal standard for the<br />

electrochemical measurements compared to the ferrocene (Fc/Fc + ) system due to<br />

its negligible solvent dependence and interactions with counterions [273].<br />

Low value of HOMO of C60 and high value of IP (Table 5.3) indicate that fullerene<br />

should not be prone to oxidation [263]. When a low-nucleophilicity electrolyte<br />

(Bu4NAsF6 ) was used, three 1 e – oxidation peaks were reported [282]. Contrary to<br />

stable multianions formed upon reduction [266], the cationic species produced<br />

0/+1<br />

upon electrooxidation of C60 or C70 are rather unstable [283]. Reversible C60 oxidation at 1.26 V vs. Fc/Fc + was found in 1,1,2,2-tetrachloroethane (TCE) as<br />

0/+1<br />

solvent [284]. For the analogous C70 process the potential is 1.21 V. The addition<br />

of CF3SO3H to the TCE solutions increases longevity of the cation radicals. C + 60<br />

radicals formed by constant potential electrolysis in a thin layer electrochemical<br />

cell are stable for 3–4 h at 233 K. The characteristic IR bands appear at 10 170 cm –1 ,<br />

11 820 and less intense at 8950 cm –1 . Upon transfer to the EPR tubes, the X-band<br />

EPR spectra of frozen solutions of C + 60 contain one rhombic shaped signal at 133 K<br />

consistent with the cation radical having rhombic symmetry (hence distorted from<br />

the Ih symmetry of the neutral species) [285]. The electrochemical data obtained<br />

upon long term oxidation of C70 are more complicated and indicated instability<br />

of the cation radical even at low temperatures and in the TCE/CF3SO3H medium<br />

in comparison to those in C60 .<br />

With increasing fullerene size and the number of carbon atoms in the cage,<br />

fullerenes become more readily reduced. Half-wave reduction and oxidation potentials<br />

of higher fullerenes (C70 , C76 , C78 , C84 ) and of their isomeric forms were<br />

discussed by Echegoyen et al. [263] (see Chapter 1 and Tables 5.3 to 5.6). The higher<br />

fullerenes are easier to oxidize, e.g. C + 76 is formed in a reversible process at 0.8 V<br />

and isomers of C84 are oxidized at ca. 0.9 V [284, 286, 287].<br />

The low HOMO to LUMO gap (2.34 eV) resulting from unusual hybridization<br />

and the extended conjugated �-electronic structure leads to special electronic<br />

and electrochemical properties of C60 . Both oxidation and reduction peaks can be<br />

recorded for it at 1.27 V and –1.06 V vs. Fc/Fc + reference potential, respectively<br />

[282]. The difference between these values agrees well with the optical gap. The<br />

energy gap decreases as already mentioned with the increasing number of carbon<br />

atoms in the cage.


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Figure 5.38 Potassium ion induced switching of intra- to intermolecular electron transfer<br />

(forward reaction) and 18-crown-6 induced reversible switching of the inter- to intramolecular<br />

process (backward reaction) [289].<br />

Fullerenes act as moderate electron acceptors and form weak charge transfer<br />

complexes with common electron donors both inorganic and organic e.g. TTF<br />

[259, 288]. In the donor–acceptor complexes fullerenes accelerate forward electron<br />

transfer and slow backward electron transfer resulting in the formation of longlived<br />

charge separated states. Porphyrin and fullerene entities show weak �–�<br />

type interactions which can be strengthened by covalent linking the porphyrin<br />

to a crown ether in such a way that both functionalities contribute to the stability<br />

of the complex [288]. Due to two modes of binding depending on the location of<br />

the crown ether either intra or intermolecular electron transfer can be realized.<br />

Interestingly, shifting of the electron-transfer pathway from intra to intermolecular<br />

route was achieved by complexing potassium ions in the crown ether cavity<br />

(Figure 5.38), thus potassium ion induced a reversible switching phenomenon<br />

in this system. With strong electron donors such as alkali metals, radical anion<br />

salts are generated in the solution and often reveal interesting superconducting<br />

properties [260, 289].<br />

5.3.4.2 Electrochemical Properties of Soluble Fullerene Derivatives<br />

In order to tune the electronic properties of fullerenes for specific electronic and<br />

photophysical devices additional units are introduced into the fullerene structure.<br />

Schematic representation of such building blocks linked to the C 60 moiety with<br />

spacer units is shown in Figure 5.39.<br />

Functionalization of fullerene in most cases reduces its electron affinity due<br />

to disruption of the conjugated system. The reduction potentials of fullerene<br />

derivatives classified conveniently as: singly bonded functionalized derivatives,<br />

cyclopropanated and cycloaddition products are collected in [263] while those<br />

of several C 60 -acceptor dyads are listed in Table 5.4 of Ref. [259]. Saturation of a<br />

double bond of the fullerene structure leads to a higher LUMO energy [292, 293].<br />

The loss of conjugation may be, however, compensated by the introduction of<br />

electron withdrawing groups directly connected to the C 60 cage. Such functionalization<br />

may result in even stronger electron accepting molecule than the parent<br />

263


264 5 Fullerenes<br />

Figure 5.39 Schematic representation of dyads and triads involving the C 60 electrophore:<br />

(a) C 60 donor, (b) C 60 acceptor, (c) C 60 donor 1 -donor 2 (monoadduct with respect to C 60) and<br />

(d) C 60 donor-acceptor (bisadduct with respect to C 60 ) [291].<br />

one. (Fullerene functionalization is discussed in Section 5.2.) The reduction<br />

potentials of [1,2]dicyanofullerenes [294] or dicyano-[6,6]methano fullerene [295]<br />

are ca. 0.1 V more positive than that of the parent fullerene. The electron affinity<br />

of fluorofullerenes increases by ca. 0.05 eV per each fluorine atom added, thus,<br />

highly fluorinated fullerenes become exceptionally strong electron acceptors<br />

[296, 297]. Other strategies to improve electron acceptor properties of a fullerene<br />

derivative include insertion of electron-withdrawing substituents connected to<br />

the fullerene cage through a methano bridge, and formation of pyrro lidinium<br />

salts [288, 298].<br />

The electroreduction potential depends on whether the first reduction is<br />

fullerene or addend based. Efficient electron acceptor organofullerenes are those<br />

exhibiting the periconjugative effect. In case of quinone type methanofullerenes,<br />

the intramolecular electronic interaction can occur between p z –� orbitals of the<br />

olefinic carbons of the quinone moiety and the adjacent carbon atoms of C 60<br />

separated by a spiro carbon atom. The molecule with more extended conjugation<br />

has better acceptor properties [288]. Five reversible electroreduction waves were<br />

observed for fullerene linked to the malonate group containing flexible long alkyl<br />

chains to two biphenyl-phenyl units [299]. For some of the derivatives, following<br />

first electron transfer a hemolytic cleavage of one of the bonds connecting the<br />

addend to C 60 may take place leading to irreversible electrochemical behavior<br />

[300]. The voltammetric experiment may even show reversible behavior of the<br />

compound, but the constant potential electrolysis (CPE) performed at a longer<br />

time-scale, often reveals the presence of several side products of chemical reactions<br />

in the reaction mixture which can be identified by HPLC and MALDI-TOF<br />

(Figure 5.40) [288]. In the case of CPE of methanofullerenes with nitrophenyl<br />

groups shown in Figure 5.40 after reoxidation, small amounts of C 60 , bisadducts<br />

and ca. 40% of starting material were found.<br />

Voltammetry performed during CPE may lead to valuable information concerning<br />

the mechanism of reactions and is often very helpful for understanding of the


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Figure 5.40 Cyclic voltammetry of methanofullerenes terminated with an ESR active nitrophenyl<br />

group [288] in THF vs. Fc.<br />

pathways involved in the process. For example, in case of bis(ethoxycarbonyl) derivative<br />

of methano C 70 fullerene, Echegoyen and coworkers [301] obtained evidence<br />

of a new stable intermediate exhibiting reversible electrochemical behavior. The<br />

malonate group with its electron withdrawing affinity allowed stabilization of the<br />

negative charge in the adduct and then formation of an intercage bond between<br />

two fullerene core radicals. Dimerization was found here to inhibit the expected<br />

electrochemical retro-cylopropanation reaction. The proposed mechanism of the<br />

retro-Bingel reaction taking place during reductive electrochemistry is presented<br />

in Figure 5.41.<br />

Addition of K + ion to the solution can cause a positive shift of reduction potential<br />

by ca. 90 mV as shown for a C 60 -dibenzo-18-crown6-conjugate (Figure 5.42) [302,<br />

303]. This shift can be explained by the electrostatic effect of K + bound in the<br />

cavities of the crown ether which decreases the influence of electronic interaction<br />

of the C 60 moiety with the substituent. However, the reduction potential is still<br />

more negative than that of pristine C 60 , showing that raising the LUMO level by<br />

saturation of few double bonds of the fullerene cage is the main factor determining<br />

the value of the reduction potential in this case.<br />

Finding new and improved C 60 -based acceptors, that would show less negative<br />

reduction potentials but at the same time, high stabilities of the anionic radicals<br />

formed, is a challenging goal important for designing fullerene based photovoltaic<br />

devices and artificial photosynthetic systems.<br />

265


266 5 Fullerenes<br />

Figure 5.41 Proposed mechanism for the formation of dimeric C 70 fullerene derivatives from<br />

the electroreduced form (radical position and connecting bond between the C 70 units are<br />

chosen arbitrarily with respect to possible regioisomers) [301].<br />

Figure 5.42 Representative examples of fullerene intramolecular complexes with better<br />

reduction potentials than pristine C 60 fullerene. (Adapted from [288]).


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Effect of Addition Pattern on the Electrochemical Properties of Fullerenes. The addition<br />

pattern is an important factor influencing the value of E 1/2 potential of the C 60/C – 60<br />

step of the exohedral fullerene derivatives. For 18 C 60(CF 3) n derivatives (which<br />

had been discussed in Section 5.3) this potential has been shown to be a linear<br />

function of the DFT-predicted E(LUMO) values [304].<br />

The reduction potential can be changed to a great extent by altering the addition<br />

pattern of C 60(X) n and C 70(X) n as shown for C 60(CF 3) n, C 60(C(CO 2Et) 2) 3 [305] and<br />

C 70(Ph) i [306]. The effect is especially strong when X is large enough so that 1,4<br />

additions are favored over 1,2 ones. Moving e.g. one CF 3 group by only two cage<br />

carbon atoms on its own pentagon, would change the reduction potential by 0.45 V.<br />

The effect of particular substituent on shifting E(LUMO) values is a function of<br />

the substituent ability to withdraw electrons from the fullerene cage [307, 308].<br />

As shown, it depends on the addition patterns of the compounds in question.<br />

Different addition patterns result in different numbers of non-terminal double<br />

bonds in pentagons (defined as short pent-hex junctions that have two C(sp 2 )<br />

nearest neighbors). When the LUMO fragments associated with them overlap,<br />

E(LUMO) becomes very low and the E 1/2 values are relatively high. This shows,<br />

that the addition pattern of a fullerene derivative may be even a more important<br />

factor than the number of subtituents for the electron acceptor efficiency of the<br />

molecule (Figure 5.43) [309].<br />

Figure 5.43 E 1/2 values of C 70 (CF 3 ) n derivatives from electrochemical and DFT study [309].<br />

Addition patterns leading to non-terminal double bonds in pentagons and<br />

extensive LUMO delocalization allow one to obtain controlled electron–acceptor<br />

properties.<br />

Organofullerenes with Electron Donor Moieties. Due to the degree of electron delocalization<br />

within the �-system and large size, the molecules under discussion<br />

are convenient building blocks for the construction of covalent donor–acceptor<br />

267


268 5 Fullerenes<br />

dyads, triads and more complicated systems in molecular devices for electronics,<br />

molecular switches, and photoconductors. Absorption of fullerenes in the visible<br />

spectrum region and their ability of rapid photoinduced charge separation make<br />

fullerene derivatives fascinating materials for photovoltaic applications [310].<br />

The intramolecular transfer dynamics in molecules with fullerenes linked to an<br />

electroactive or photoactive species is function of excited state of the antenna<br />

molecule, donor–acceptor distance and the solvent used [311].<br />

The energy levels of the charge-separated states (� GRIP) can be evaluated using<br />

the Weller type approach utilizing the redox potentials, center-to-center distance,<br />

and dielectric permittivity of the solvent [312]:<br />

� = E − E + �G<br />

,<br />

GRIP ox red s<br />

where �G s = –e 2 /4 � � 0 � R R Ct–Ct and � 0 and � R refer to vacuum and solvent permittivity.<br />

By comparing these energy levels with the energy levels of the excited states,<br />

the driving forces are evaluated. Electrochemical gap as small as 1 eV has been<br />

reported for some fullerene–ferrocene dyads [259]. Ferrocene (Fc, E 1/2 = 0.50 V<br />

vs. SCE), N,N�-dimethylaniline (E 1/2 = 0.81 V vs. SCE), N-methylphenothiazine<br />

(E 1/2 = 0.70 V vs. SCE) and tetrathiafulvalene (TTF, E 1/2 = 0.37 V vs. SCE) are<br />

common donor components of these dyads [313]. This type of hybrids were<br />

often classified as electroactive but, due to the insignificant VIS absorption of<br />

the added group, photoinactive. The electrochemical data for the TTF–C 60 dyads<br />

are reviewed by Bendikov et al. [310]. Due to very different structures such dyads<br />

exhibit different solvation effects and dependence of the electrochemical gap on<br />

the solvent used [314]. A thermodynamically stable supramolecular donor–acceptor<br />

system with C 60 and TTF assembled through guanidinium–carboxylate ion pair<br />

stabilized by two hydrogen bonds has been proposed [315]. Multiple substitution<br />

of the fullerene core leading from dyads (Figure 5.44) to triads [315] tetrads and<br />

quintads [316] enlarged the HOMO–LUMO gap and lead to a decrease of the C 60<br />

reduction potential.<br />

Because of the moderate fullerene absorption in the VIS region, functionalization<br />

of C 60 with a chromophoric addend improves the light harvesting efficiency<br />

of the fullerene dyad. In such a system fullerenes operate as electron or energy<br />

acceptor moieties. Donors such as metalloporphyrins, zinc phthalocyanines,<br />

ruthenium (II) polypyridyl complexes with strong metal-to-ligand charge transfer<br />

(MLCT) are powerful units in the excited state. The radical pair for the latter case<br />

involves oxidized complex and anionic fullerene radical with a lifetime of 304 ns in<br />

deoxygenated CH 2 Cl 2 . However, in the presence of traces of oxygen the fullerene<br />

triplet excited site is quenched to produce cytotoxic singlet oxygen species [311].<br />

In the C 60 -ferrocene conjugates the intimate contacts between the donor and<br />

acceptor result in large ground-state interactions suggesting a substantial charge<br />

density shift from donor to acceptor. In the excited states, the processes in these<br />

conjugates are dominated by rapid charge separation reactions (0.8 ps) to yield<br />

metastable ion pairs with the radical pair lifetime of ca. 30 ps. As discussed by<br />

Guldi and coworkers [317], no prominent charge-transfer features were observed


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Figure 5.44 Cyclic voltammogram of TTF-C 60 diad [R=S(CH 2CH 2O) 4Me] recorded in<br />

1,2-dichlorobenzene. (Adapted from [310]).<br />

for the bucky ruthenocene conjugates and only an intrinsically faster excitedstate<br />

deactivation (ca. 200 ps) evolved. The authors attribute this difference to the<br />

unfavorably shifted oxidation potential of about 0.61 V in ruthenocene. For the<br />

fullerene derivative –�G ET is 0.35 eV, while for ruthenocene conjugate it is –0.26 eV<br />

rendering the charge separation thermodynamically unfeasible. Modulation of the<br />

electronic coupling between C 60 acceptor and various donors for the construction of<br />

new optoelectronic devices requires also searching for new type of connecting units<br />

between the donor and acceptor. Longer-distance charge separation was achieved<br />

for oligothiophene–fullerene dyads (nT–C 60 ) and triads 8T–4T–C 60 [318].<br />

Making use of coordination of fullerene derivatives to transition metal ions is<br />

in general a promising approach. Fullerene coordination ligands in which one<br />

bipirydine or terpyridine unit was directly attached to the nitrogen atom of a fulleropyrrolidine<br />

ensure a linear communication pathway between fullerene and<br />

coordinated ruthenium (II). This formed a linear rod-like donor–acceptor system<br />

(Figure 5.45) [319].<br />

The reduction of the fullerene unit in the complex is easier than in the free<br />

ligand while Ru-based reductions are shifted by 50 mV more negative relative to<br />

the simple [Ru(bpy) 3 ]Cl 2 complex. These electrochemical results indicate electronic<br />

coupling existing between the fullerene core and the ruthenium ion center and are<br />

in line with the MLCT transition band, which is subjected to a red shift for the dyad.<br />

Ferrocene–porphyrin–fullerene constructs with crown ether appended porphyrins<br />

were self-assembled through host–guest interactions with alkylammonium cation<br />

functionalized fullerenes. Interactions between the porphyrin and fullerene entities<br />

were followed by voltammetry. The crown ether–alkylammonium cation complexation<br />

binding strategy lead to well-defined stable triads with charge-recombination<br />

in two steps and direct charge recombination from the porphyrin cation radical.<br />

This system fulfilled the condition of long-lived charge-separated state.<br />

269


270 5 Fullerenes<br />

Figure 5.45 Cyclic voltammograms of [Ru(bpy) 3 ]Cl 2 (black), bpy-C 60 (red) and complex (blue) in<br />

CH 2 Cl 2 with 0.1 M TBA[PF 6 ] (tetrabutylammonium hexafluorophosphate) [319].<br />

An interesting design strategy is based on dendritic molecules appended with<br />

multiple zinc porphyrin units, to trap pyridine compounds carrying multiple<br />

(n = 1–3) fullerene units (Figure 5.46) [320]. The complexes presented in the<br />

scheme have a photoactive layer of spatially segregated donor and acceptor arrays.<br />

Voltammetry experiments revealed electron donation from ligating pyridine to<br />

the zinc porphyrin units. Photoinduced electron transfer was confirmed by flash<br />

photolysis measurements. High ratio of charge-separation rate constant to charge–<br />

recombination rate constant (3400) was obtained.<br />

5.3.4.3 Electrocatalytic Activity of Fullerenes<br />

Early attempts to use fullerenes as catalysts involved reduction of organic halides<br />

[321]. Regeneration of the oxidized form of C60 catalyst at the electrode surface<br />

was responsible for the electrocatalytic current. Electrocatalytic reduction of<br />

dihalogenated alkanes was investigated using C60 and C70 in solution and using<br />

2–/3– 1–/2–<br />

coated electrodes [322–324]. C60 and C60 were the catalytically active couples.<br />

In Figure 5.47, the cyclic voltammograms reveal the catalytic effects observed for<br />

2–/3–<br />

dihaloethanes in the range of potentials corresponding to the C60 electrode<br />

process. The peaks appearing at positive potentials are due to the oxidation of<br />

halogenide released in the catalytic reduction process.<br />

Anions of higher fullerenes C76 , C78 , C84 are relatively more stable with respect<br />

to derivatization and can serve as better electrocatalysts than anions of C60 [323].<br />

The rate constants for the pseudo-first-order conditions can be calculated from<br />

the rotating disk electrode measurements [325]. Alkanes and monoiodoalkanes<br />

were products of the C60 reaction with �,�-diiodoalkanes except for reaction<br />

3–<br />

with 1,3-diiodopropane and 1,5-diiodopentane where C60 adduct formation was<br />

reported [322]. For each fullerene anion kcat increased in the order Cl < Br < I in


5.3 Physicochemical Properties and the Unusual Structure of Fullerenes<br />

Figure 5.46 Molecular structure of zinc complex of multiporphyrin dendrimer DP 24 and schematic<br />

representation of the complexation of DP 24 with fullerene appended bipyridine ligand Py 2 F 3 .<br />

(Adapted from [320]).<br />

271


272 5 Fullerenes<br />

Figure 5.47 Cyclic voltammograms for 0.0001 M C 70 (curve 1), 0.0001 M C 70 and 0.4 M<br />

1,2-dichloroetane (curve 2), 0.0001 mM C 70 and 0.4 M 1,2-dibromoethane (curve 3), as well as<br />

0.0001 M C 70 and 0.0014 M 1,2-diiodoethane (curve 4) in 0.00001 M (TBA)PF 6 in benzonitrile,<br />

at Pt electrode. Potential scan rate is 0.1 V s –1 .<br />

agreement with a dissociative mechanism of reduction [326] and increased in the<br />

order C84 < C78 < C76 < C70 < C60 for each 1,2-dihaloethane [327]. Although,<br />

lower rate constants were obtained for the higher fullerenes, less negative potentials<br />

and the lack of alkyl adduct formation were of advantage.<br />

Electrochemically generated anions of C60 were used also in deprotonation of<br />

the organic acids. A weekly basic monoanion of fullerene deprotonates a relatively<br />

strong acid e.g. ethyl nitroacetate. C – 60 anion catalyzed nitroacetate reaction with<br />

ethyl acrylate and acetonitrile to form double addition products [328].<br />

5.3.4.4 Conclusions and Outlook<br />

Fullerene based films, discussed in the next section, attract considerable interest as<br />

materials possessing the unique properties of fullerenes, but at the same time more<br />

suitable for practical applications in nanotechnological, electrochemical sensing<br />

and photophysical devices. Since C 60 molecules exhibit electron-mediating properties,<br />

upon immobilization in mono- and multilayer films on electrodes they can efficiently<br />

promote reduction and oxidation processes. In these exciting applications,<br />

albeit outside the scope of this chapter, their poor solubility in many solvents and<br />

water becomes an advantage, contrary to the case of diffusion charge mediation.<br />

In general, more uniform coverage of the electrode substrates and higher reversibility<br />

of the electrode processes were reported for C 60 embedded in different type<br />

of matrices than for the polycrystalline C 60 films deposited on electrodes [262]. C 60<br />

modified electrodes can be easily prepared by placing a C 60 drop onto the electrode<br />

which is then overlaid with nafion protecting coat, as shown first by Compton et al.<br />

[280]. Two to four quasireversible 1 e – reductions were usually observed in aqueous<br />

or mixed solvent and electrocatalytic application of surface immobilized C 60<br />

molecules were demonstrated e.g. for oxygen and nitrobenzene reductions [262].


5.4 Fullerene Aggregates<br />

Figure 5.48 Proposed mechanism of cytochrome c immobilization, and electrochemical<br />

reduction by C 60 -Pd polymer film. (Adapted from [331]).<br />

Several attempts concern biocatalytic systems important for the field of bioelectronics<br />

[329], e.g. C 60 molecules assembled in monolayers were used to mediate<br />

electrons between enzymes and the electrodes [330]. Two different fullerenemodified<br />

electrodes were used: electrical contacting an enzyme and cytochrome<br />

c with the electrode (Figure 5.48) [331].<br />

Fast development of applications of the C 60 films as mediating units in sensing<br />

and catalytic devices may be expected in the near future.<br />

5.4<br />

Fullerene Aggregates<br />

Tommi Vuorinen<br />

Many extraordinary properties of fullerenes have alerted researchers to find and<br />

study the properties of aggregated fullerene structures. The fullerenes have also<br />

drawn much interest as construction materials for thin films and other ordered<br />

solid nanostructures. One can prepare photovoltaic devices which use photogenerated<br />

charge separation as a result of the excellent ability of fullerenes to<br />

accept electrons from electron-donating compounds. A well-known example is<br />

the plastic solar cell with bulk hetero junction structure, where an electronegative<br />

fullerene derivative is incorporated into conjugated electron-rich polymer [332].<br />

Many interesting features of fullerene and fullerene derivative films have been<br />

discovered, for example, superconductivity [333, 334], unique redox properties<br />

[335], photo-electrochemical response [336, 337] among others. The most common<br />

fullerenes studied in all respects are C 60 and C 70 , so this chapter will concentrate<br />

on the research based on these two carbon balls.<br />

273


274 5 Fullerenes<br />

5.4.1<br />

Film Preparation Methods<br />

Numerous supramolecular assemblies containing fullerene as a functional moiety<br />

have been synthesized [338–340]. In order to use the functions of the supramolecular<br />

structures created, the assemblies should in most cases be bound as solid<br />

well-ordered films [340]. With ordered nanostructured films one may be able to<br />

scale the functions taking place at the molecular level up to macroscopic level.<br />

This may happen when billions of supramolecular assemblies work coherently<br />

in an ordered film and the sum of all actions at the nanoscale is seen as the<br />

resulting macroscopic function. For this reason thin fullerene films are also in<br />

focus when, for example, molecular electronics are developed. A variety of film<br />

preparation methods have been used in fullerene thin film fabrication and will<br />

be briefly reviewed here.<br />

Drop casting and spin-coating are based on solvent evaporation and precipitation<br />

of the solute. In general, the film morphology can be varied by changing<br />

the solvent evaporation rate. The type of the solvent can also influence how the<br />

precipitate will crystallize. The drop cast fullerene films are used, for example,<br />

as starting material for electrochemical polymerization of fullerene copolymers<br />

[341]. Compared with drop casting, spin-coating can provide better control over<br />

the film formation process. In spin-coating the film is formed by precipitation<br />

onto a substrate which is rotated while evaporation takes place. Thus in addition<br />

to the solvent evaporation the film morphology is affected by the centrifugal<br />

forces caused by the substrate rotation. The film morphology and thickness can<br />

be varied, for example, by changing the rotation speed, the solvent, and the concentration<br />

of active material in the solution. Spin-coating is usually used when<br />

a fullerene derivative is mixed with a conjugated polymer. This way one can get<br />

a bulk hetero junction between the electron-donating conjugated polymer and<br />

electron-accepting fullerene. The bulk hetero junction structure has been proved<br />

to act as an efficient organic photodiode when placed between two electrodes with<br />

different work functions [332].<br />

The Langmuir–Blodgett (LB) and Langmuir–Schäffer (LS) methods are sophisticated<br />

techniques for the preparation of well-ordered one molecule thick films from<br />

surface active molecules [342, 343]. The first step in both techniques is formation<br />

of the Langmuir film that is a stable molecular monolayer at the gas–liquid<br />

interface. Water is definitely the most commonly used liquid as the subphase for<br />

monolayer formation. Water can easily provide very strong anisotropic interaction<br />

between the molecules and the interface. In practice, the subphase is either pure<br />

water or water including a small amount of some salt. Naturally, it is easiest to use<br />

ambient air as the gas phase. The need of anisotropic interaction sets a prerequisite<br />

for the molecules which are used for the monolayer formation. The molecules<br />

should be amphiphilic, having one part of the molecule hydrophilic, i.e. soluble<br />

in water, and the other part hydrophobic. The strong attraction to water forces<br />

molecules to be as close to the water surface as possible. On the other hand the<br />

water insoluble part stops the molecules from dissolving in the subphase. This


5.4 Fullerene Aggregates<br />

kind of anisotropic interaction makes it possible to form stable monomolecular<br />

films at the air–water interface. The attraction to water prevents the molecules at<br />

the interface from forming three-dimensional crystals.<br />

The Langmuir film preparation starts by dissolving the surfactants in a waterimmiscible<br />

volatile solvent, such as chloroform or hexane. The solution is then<br />

carefully spread onto the clean water surface so that the total area of all the<br />

molecules at the air–water interface is much smaller than the available area of the<br />

water surface. One can get a simplified, but descriptive, picture of the phase transitions<br />

of the molecular film at the interface when one considers two-dimensional<br />

phases analogous to the three-dimensional ones. Thus in the beginning, when<br />

the molecules have space to move freely in two dimensions they behave like a<br />

two-dimensional gas (a in Figure 5.49). In this phase the surfactants have very<br />

small, if any, effect on the surface tension of water.<br />

As three-dimensional gas, the two-dimensional gas can also be condensed by<br />

increasing the pressure. In order to increase the two-dimensional pressure of the<br />

molecules at the interface the surface area should be decreased. Usually, the LB<br />

system consists of a trough with computer controlled barriers. By moving these<br />

barriers the area of the water surface can be controlled. When the surface area is<br />

decreased, the free space around the molecules is reduced and molecules start<br />

to influence each other. At some point of the compression, when the surfactants<br />

are close enough to each other, they start to decrease the surface tension of water.<br />

The observed decrease in the surface tension is referred to as the surface pressure<br />

(�) which is the difference in the surface tensions of the pure water surface (�*)<br />

and the water surface with the surfactants (�): � = �* – �. The surface pressure is<br />

plotted as a function of available area per molecule, i.e. mean molecular area. The<br />

resulted graph is referred to as surface pressure–area isotherm. For an example,<br />

an isotherm of an imaginary surfactant is presented in Figure 5.49. The phase<br />

transitions cause changes in the gradient of the area–surface pressure isotherm.<br />

The most common method of obtaining information about the phase transitions in<br />

the monolayer is to measure the surface pressure online with the Wilhelmy method<br />

while the molecules are compressed; the more densely the molecules are packed<br />

Figure 5.49 Schematic illustration of a surface pressure–area isotherm of a surfactant.<br />

The monolayer phases are indicated and illustrated: (a) two-dimensional gas,<br />

(b) two-dimensional liquid, (c) two-dimensional solid, and (d) collapse of the monolayer.<br />

275


276 5 Fullerenes<br />

in the monolayer, the more they lower the surface tension of water. With decreasing<br />

surface tension the surface pressure increases. When the gaseous molecules<br />

are compressed enough a two-dimensional liquid, or expanded condensed phase<br />

(b in Figure 5.49), is formed. Further compression results in a faster rising in the<br />

surface pressure when compared with that observed for the gas phase.<br />

In order to transfer the monolayer from the interface onto a solid substrate, the<br />

monolayer should be stable. An essential requirement to get a stable monolayer<br />

is that the surfactants are able to form a solid phase at the ambient temperature.<br />

When the monolayer has reached the two-dimensional solid phase, or condensed<br />

phase (c in Figure 5.49), the available surface area per molecule corresponds the<br />

molecular cross section which is approximately 0.2 nm 2 for saturated fatty acids.<br />

Usually, the first Langmuir–Blodgett deposition takes place by passing the hydrophilic<br />

substrate through the monolayer at the interface from water to air, so<br />

that the water soluble ends are attached to the substrate. For the layers following<br />

the first deposition, depending on the dipping direction, either from air to water<br />

or from water to air, the molecules in the monolayer may have different orientation<br />

in respect to the substrate. Concurrently with passing the substrate through<br />

the monolayer the surface pressure is kept constant by moving the barriers with<br />

proper speed. The transfer ratio is the area the barriers moved during deposition<br />

divided by the substrate area passed though the film. The transfer ratio describes<br />

the quality of the film deposition, and for the ideal case it is unity. The Langmuir–<br />

Schäffer method is sometimes referred to as horizontal lifting. In the LS method<br />

the substrate is kept parallel to the water surface when the substrate is brought<br />

in contact with the Langmuir film. The Langmuir monolayer adheres onto the<br />

substrate and is lifted with the substrate from the interface.<br />

It might seem that the LB technique is strictly limited by the requirement of<br />

amphiphilicity of the molecules but actually it is possible to prepare, for example,<br />

films where totally hydrophobic compounds are mixed in a surfactant matrix.<br />

Formation of the Langmuir and Langmuir–Blodgett films of pristine fullerenes<br />

is somewhat controversial, but this will be discussed later in the text.<br />

Physical vapor deposition in high vacuum is a very well known technique to<br />

prepare thin metal, semiconductor, and organic films. Among other organic<br />

materials, also fullerene thin films have been prepared by thermal vapor deposition<br />

in vacuum [344–347]. The fullerene vaporized is produced in the high vacuum<br />

by heating electrically the resistive source at temperature of 300 to 550 �C. With<br />

organic materials the temperature control of the source has high importance in<br />

order to avoid destruction of the material.<br />

Nanostructured fullerene films can be fabricated by the method referred to as<br />

electrodeposition [348, 349]. In this method one uses the ability of C 60 and C 70 to<br />

form clusters in solution of mixed solvents at room temperature. A usual diameter<br />

for such a fullerene cluster is 100–300 nm. The driving force for cluster formation<br />

is the three-dimensional hydrophobic interaction between fullerene units in the<br />

solution of polar and non-polar solvents. One method of preparing clusters is to<br />

inject a toluene solution of fullerene into acetonitrile. In the final cluster suspension<br />

the ratio of mixed solvents is between 3 : 1 and 9 : 1 of acetonitrile–toluene. The


5.4 Fullerene Aggregates<br />

fullerene cluster formation process and the stability of the clusters are affected by<br />

the ratio between the solvents and by the fullerene concentration in the original<br />

toluene solution. Especially, C 60 can precipitate, for example, when the fullerene<br />

concentration exceeds 50 µM and when volume fraction of acetonitrile is more<br />

than half [350]. The cluster formation is confirmed by solvatochromic changes<br />

when acetonitrile is added to the toluene solution of fullerene. The color changes<br />

from the purple of C 60 dissolved in toluene to yellow-green of the C 60 cluster<br />

solution. By electrodeposition one can transfer these fullerene aggregates onto a<br />

conductive solid substrate, for example indium-tin-oxide (ITO) coated glass. The<br />

electrodeposition takes place in a two-electrode cell where one of the electrodes<br />

is the substrate. The spacing between the electrodes is few millimeters and a DC<br />

voltage from 20 to 400 V is applied between the substrate and the other electrode.<br />

Under the influence of the DC electric field, the fullerene clusters can become<br />

charged. The charged clusters are transferred by the electric field onto the substrate<br />

with positive potential. Pristine fullerenes in solution could not be charged by<br />

the electric field in a similar way as the clusters, thus fullerene aggregation must<br />

first take place before electrodeposition. By this method relatively thick fullerene<br />

films can be deposited easily and fast. For example, when the cell has been filled<br />

with approximately 0.1 mM fullerene cluster solution and electric field is applied<br />

for one minute, the solution becomes colorless and a brown film forms onto the<br />

substrate. With these conditions the resulted film has a thickness of approximately<br />

1 µm which corresponds to surface concentration of 0.2 µmol cm –2 of fullerene<br />

molecules [348]. The film thickness can be controlled by varying the fullerene<br />

cluster concentration, the deposition time, and applied DC voltage. In order to<br />

increase the film thickness one should use higher cluster concentration, longer<br />

deposition time, and/or apply higher DC voltage.<br />

5.4.2<br />

Fullerene Film Properties<br />

Pristine C 60 and C 70 consist solely of carbon atoms and thus are purely hydrophobic.<br />

The formation of a stable fullerene Langmuir monolayer at the air–water<br />

interface is somewhat controversial. There are reports about successfully formed<br />

monolayers which have been transferred onto the solid substrate from the interface<br />

[351, 352]. When fullerene forms a real monolayer on the water surface the limiting<br />

area, i.e. the molecular area obtained by extrapolating the isotherm to zero pressure<br />

from its steepest part, is approximately 1 nm 2 . Most of the results indicate that<br />

fullerenes tend to form an aggregated multilayer structure rather than a stable<br />

monolayer [353]. The mean molecular area obtained decreases dramatically when<br />

the fullerenes do not form a monolayer at the air–water interface. Usually, the<br />

molecular area is less than half of that of the perfect monolayer. When the surface<br />

pressure–area isotherm results in a molecular area lower than one could expect<br />

from the physical size of the molecule, the conclusion is that the molecules do<br />

not form a perfect monolayer but they are placed on top of each other as a double<br />

layer or a multilayer. The Langmuir monolayer measurements have shown that a<br />

277


278 5 Fullerenes<br />

stable fullerene monolayer can be formed on the water surface in very narrow range<br />

of experimental conditions [351, 352]. For example, with concentrations higher<br />

than 10 –4 M, a perfect monolayer cannot be obtained [351]. The simulations of the<br />

fullerene monolayers on the water surface help to understand the reason for the<br />

observed irregular monolayer formation [354]. The pristine fullerene does not have<br />

anisotropic interaction with the water to keep a stable monolayer structure.<br />

For Langmuir film preparation a dilute fullerene solution is spread onto the<br />

water surface. After solvent evaporation fullerenes form a two-dimensional gas<br />

phase. The single molecule can move away from this phase in two ways: it may<br />

attach itself to other fullerenes and form a monolayer raft at the interface, or it may<br />

escape from the water surface onto the formed raft. A stable monolayer is formed<br />

if every fullerene in the raft has six neighboring fullerenes. Six fullerene–fullerene<br />

bonds keep the surrounded fullerene tightly in place on the water surface [354].<br />

Naturally, the raft is not infinite but it has boundaries where there are also fullerenes<br />

which have less than three neighboring fullerenes. The fullerenes on the edge<br />

can easily be promoted on top of the fullerene raft. For example, if a fullerene has<br />

only two neighboring fullerenes no fullerene–fullerene bonds have to be broken<br />

when this fullerene turns onto the fullerene raft. When conditions are such that<br />

this promotion takes place it is impossible to obtain a fullerene monolayer at the<br />

air–water interface. On the water surface the monolayers of pristine fullerene are<br />

extremely rigid and incompressible having high collapse pressure [351].<br />

In order to overcome the problematic monolayer formation of pristine fullerenes,<br />

caused by the fullerene–fullerene and fullerene–water interactions, one can use<br />

two different strategies. The first is to use some matrix compound which prevents<br />

direct contact, and thus interaction, between fullerenes. The pristine fullerene<br />

can be mixed with long chain alcohols or fatty acids [351]. The surfactants and<br />

fullerenes are supposed to spread homogenously onto the water surface and<br />

thus fullerenes tend to aggregate less than when only pristine fullerene is spread<br />

onto the water surface. In addition, some special matrices with cavities where<br />

fullerene can fit in, such as azocrowns [355], calixarenes [356], and resorcarenes<br />

[357] have been used to enclose fullerenes in the LB films. These cavitants offer a<br />

hydrophobic cavity where fullerene can be incorporated and so isolated from the<br />

water surface when a one-to-one mixture of the cavitant and fullerene is spread<br />

on the water. The cavitant has a hydrophilic part which enables formation of<br />

stable Langmuir monolayer at the air–water interface. When any matrix is used<br />

the resulting fullerene film does not solely consist of fullerene. A rather exciting<br />

way to make a uniform fullerene LB film is to use specially selected matrix<br />

molecules which can be selectively dissolved away from the film deposited onto<br />

the solid substrate [358]. One can enhance miscibility of fullerene in the matrix<br />

by replacing fatty acids with surfactant with, for example, a fulvalene moiety. The<br />

electron-accepting fullerene and the electron-donating matrix molecule may have<br />

ground state interactions which can increase mixing compared with that of fatty<br />

acids and fullerene [358].<br />

The second option to decrease the tendency of fullerenes to form multilayer<br />

stacks at the air–water interface is to use fullerenes substituted with some hydro-


5.4 Fullerene Aggregates<br />

philic group in order to increase surface–molecule interaction. The amphiphilic<br />

fullerene derivatives are capable of forming the Langmuir monolayers themselves<br />

[337, 359–361]. An amphiphilic fullerene can be also mixed with fatty acid matrix<br />

to ease film formation and deposition [336]. The hydrophilic adducts on fullerenes<br />

will keep the molecules as a monolayer at the air–water interface even during compression.<br />

In other words, the interaction between the water surface and the adducts<br />

will defeat the fullerene–fullerene attraction, which usually causes formation of<br />

fullerene multilayers at the interface during compression. Addition of hydrophilic<br />

side group to the fullerene core does not cancel attraction between the fullerenes.<br />

The attraction between the fullerene cores causes irreversible clustering, which<br />

is seen as unequal isotherms for the first compression and for the subsequent<br />

expansion of the molecular film [362]. This is observed when the monolayer of<br />

the hydrophilic fullerene derivative on the water surface is compressed until it<br />

collapses and after that the surface area is increased. The expansion isotherm<br />

resembles the compression isotherm observed for pristine fullerene forming a<br />

multilayer on the water surface. This means that after the monolayer collapse the<br />

fullerene derivatives are no longer spread uniformly onto the water surface but<br />

the fullerene aggregates are held together by the attraction between the fullerene<br />

cores. In order to improve the spreading behavior one can combine the shielding<br />

effect of matrix and the properties of hydrophilic adducts by chemical engineering<br />

[362]. Fullerene derivatives with several long hydrocarbon chains and hydrophilic<br />

groups as adducts have been designed and the LB properties of these molecules<br />

have been tested [363]. Already four alkyl chains added onto the fullerene core<br />

reduced the attraction between the fullerene cores at the air–water interface. This<br />

is seen as better reversibility in isotherms measured with successive compression/expansion<br />

cycles. Still the alkyl chains do not provide proper shielding of<br />

fullerene cores but, for example, the film absorption spectra show broadening<br />

as a result of fullerene aggregation. By placing the fullerene core in the middle<br />

of an amphiphilic dendritic structure, one can achieve a fullerene derivative<br />

with excellent spreading properties on the water surface. The dendritic adducts<br />

make it possible to prepare films where fullerenes are isolated from each other<br />

[364, 365]. The LB monolayers of such fullerene derivatives have shown similar<br />

absorption spectra to those measured for dilute dichloromethane solutions of the<br />

same compounds. Similar absorption spectra for film and solution indicate that<br />

the fullerene cores are isolated even in the LB film. Films consisting of isolated<br />

fullerenes are interesting in point of view of nonlinear optics, especially, optical<br />

limiting applications [362].<br />

Electronegative fullerene can be also modified by linking it covalently with an<br />

electron donating chromophore, such as, porphyrin [338, 366], phthalocyanine<br />

[367], or phytochlorin [368]. The resulting compound is capable of the photoinduced<br />

intramolecular electron transfer from the donor moiety to the fullerene<br />

moiety [366]. The donor–fullerene dyads can be synthesized to have amphiphilic<br />

character and thus they are suitable for the LB film preparation [369–372]. By the<br />

LB technique the electron donor–acceptor dyads can be deposited as monolayers<br />

with specified orientation of all the molecules in the film. In other words, in the<br />

279


280 5 Fullerenes<br />

dyad monolayer the electron transfer takes place perpendicular to the substrate<br />

and in the same direction in all dyad molecules coherently [369, 371]. When the<br />

vectorial photoinduced intramolecular electron transfer takes place from donor<br />

to fullerene in billions of dyads simultaneously, a potential of several millivolts<br />

is produced [371].<br />

Fullerenes are electrochemically interesting. In solutions, pristine C60 and C70 6– 6–<br />

are able to reversibly accept up to six electrons resulting in C60 and C70 anions,<br />

respectively [373]. Fullerenes electrochemistry is discussed in more detail in the<br />

former section. In contrast, the oxidation of fullerene is more difficult and is<br />

usually limited to the formation of dication as the highest oxidation state [374].<br />

Similar step-wise reduction is also seen in fullerene films, but the behavior has<br />

been found to be more complicated than for dissolved molecules [335]. The voltammograms<br />

for drop-cast fullerene films showed larger peak splitting between<br />

the measured reduction and re-oxidation peaks than for dissolved fullerenes. The<br />

observed hysteresis for the films suggests large structural changes in the film [335,<br />

375]. Upon reduction, most of fullerenes are reduced to mono-anions. The charge<br />

in the film is balanced by diffusion of cations from the supporting electrolyte.<br />

The bigger the diffused cations are in size, the larger are the structural changes<br />

seen in the film. The changes are reversible when fullerenes are re-oxidized back<br />

to the neutral form.<br />

With the electrodeposition one can easily prepare micrometer thick fullerene<br />

films on optically transparent electrodes. Absorption spectra for electrodeposited<br />

C60 films on three different substrates are presented in Figure 5.50. The inset<br />

Figure 5.50 Absorption spectra of C 60 clusters deposited on different substrates (a) OTE/(C 60 ) n ,<br />

(b)OTE/TiO 2 /(C 60 ) n , and (c) OTE/SnO 2 /(C 60 ) n film. The absorption spectra of native electrodes,<br />

(d) OTE/TiO 2 and (e) OTE/SnO 2 , recorded prior to electrodeposition are also shown. The inset<br />

shows solution absorption spectra of (f) 19 µM C 60 in toluene and (g) same amount of C 60 in<br />

the form of clusters in 1:3 (v/v) toluene–acetonitrile. (Reproduced with permission from J. Phys.<br />

Chem. B 2000, 104, 4014–4017; Copyright 2000 America Chemical Society).


5.4 Fullerene Aggregates<br />

in Figure 5.50 shows the difference in the absorption spectra of the C 60 toluene<br />

solution and the cluster suspension. The characteristic absorption band of C 60 at<br />

approximately 350 nm is red-shifted and broadened when clusters are formed.<br />

In addition, the very weak absorption band of C 60 monomer in the 500–600 nm<br />

region gets stronger and shifts slightly to shorter wavelengths when aggregation<br />

takes place [350]. Similar spectral shifts and broadening of the absorption bands,<br />

compared with the fullerene solution spectrum, can be seen as an indication of<br />

fullerene–fullerene interaction in films prepared by various methods.<br />

The fullerene films electrodeposited onto the transparent electrode coated with<br />

tin oxide nanoparticles show photocurrent generation in electrochemical photocurrent<br />

measurements. In the experiments, the fullerene film is immersed in an<br />

acetonitrile electrolyte with iodine/iodide redox-pair [348]. The current–voltage<br />

characteristics for this system established a photodiode behavior. The photocurrent<br />

generation in this system is most likely due to photogalvanic type of behavior. This<br />

involves electron injection from the iodine/iodide pair to photoexcited fullerene.<br />

The generated fullerene anion transfers electron to SnO 2 nanoparticle.<br />

For nonlinear optics, fullerenes and fullerene derivatives have shown good<br />

responses, for example, as optical limiting materials. Usually, for this purpose<br />

fullerenes are dispersed into a solid polymer [376] or phosphate glass [377] matrix.<br />

Also, sol-gels of fullerene derivatives can be prepared [378]. By this manner one<br />

may have films with high concentration of isolated fullerenes. The optical power<br />

limiting is a nonlinear optical phenomenon where the optical absorption increases<br />

while incident light intensity increases [379]. As seen in the inset in Figure 5.50,<br />

isolated C 60 molecules in solution have low steady-state absorption in the visible<br />

range. C 60 at its excited singlet or triplet state has considerably higher molar absorption<br />

coefficient in the 300–700 nm range than at the ground state [380]. The<br />

absorption properties result in reverse saturable absorption characteristics for<br />

fullerene, and thus make it a suitable material for optical limiting applications.<br />

Since reversible changes in optical transparency are based on fullerene excited<br />

state absorption, a requirement for an efficient operation is the long lifetime of the<br />

excited state. Actually, fullerene aggregation causes self-quenching of the excited<br />

states and thus restricts the optical limiting time to sub-nanosecond level. The<br />

desire for longer working times directs the interest towards solid films where<br />

fullerenes are not aggregated but are isolated from each other.<br />

High temperature superconductivity has been observed for alkali metal fullerides<br />

[334]. The critical temperature of 33 K can be observed for alkali fulleride<br />

salts with stoichiometries A 3 C 60 , discussed in Section 5.3.3.2. The salt is formed<br />

by refluxing fullerene and alkali metal under an oxygen free atmosphere, and<br />

electronegative fullerenes oxidize alkali metal atoms. The superconductive<br />

phase of the resulted salt adopts, in general, face-centered-cubic or primitive<br />

cubic structure. Essentially complete electron transfer from electropositive alkali<br />

metal to electronegative fullerene fills the fullerene conduction band, the lowest<br />

unoccupied molecular orbital, halfway. Fullerene films with superconductive<br />

character have been prepared also by doping fullerene LB films with potassium<br />

[335]. After a fullerene multilayer LB film has been deposited it has been doped<br />

281


282 5 Fullerenes<br />

with potassium in high vacuum at elevated temperature. The potassium doped<br />

film had superconducting transition at approximately 8 K.<br />

5.4.3<br />

Conclusions<br />

A vast amount of research has been reported in scientific journals, and review<br />

articles in journals and books, about fullerene films, their functions, and future<br />

prospects. The fullerenes have not yet been – and probably never will be – used as<br />

tiny bearings for nanosized machines, although it was seen as a future application<br />

a long time ago. But the fullerenes have shown to be excellent material for organic<br />

solar cells and other applications related to producing and storing electrical energy.<br />

In addition, some applications can be found from the area of optical devices. Thus<br />

far no commercial application of fullerene has been introduced. There have been,<br />

and there are, many potential trials to find a fullerene application useful for the<br />

human race. The enthusiastic research on fullerenes has opened already markets<br />

for mass production of these tiny and fascinating carbon balls.<br />

5.5<br />

Endohedral Fullerenes with Neutral Atoms and Molecules<br />

Sho-ichi Iwamatsu<br />

5.5.1<br />

Introduction<br />

The inside cavity of a fullerene (3.5 Å in diameter for C 60 ) is large enough to<br />

enclose an atom or a very small molecule. This section describes the endohedral<br />

fullerenes enclosing nonmetal substrates inside the cages. Unlike metallofullerenes,<br />

in which an electron transfer takes place from the metal to the fullerene<br />

cage, these guests are bound to the cage only by weak van der Waals’ forces and<br />

by the structural impossibility of escaping. Consequently, there are few changes<br />

in properties of the host fullerenes. Such inert guests have served as valuable<br />

probes for exploring both the interior and exterior of fullerenes. Also, fullerene<br />

containers can keep highly reactive chemical species that are unable to exist in<br />

an ordinary system. Synthetic procedures and properties of the products will be<br />

presented along with theoretical studies.<br />

5.5.2<br />

Preparation<br />

5.5.2.1 Direct Approach Using an Existing Fullerene<br />

Since He@C 60 (C 60 with a helium atom inside) was discovered in a fullerene soot<br />

[381], continuous efforts have been paid to the syntheses of such gas complexes


5.5 Endohedral Fullerenes with Neutral Atoms and Molecules<br />

Scheme 5.9 Direct approach to the endohedral nonmetal complexes of fullerenes.<br />

with high yields. The first successful methodology was the forced penetration of<br />

a guest into an existing fullerene. This is different from the case of metallofullerene<br />

in which a metal fragment is captured in a cage during the formation of a<br />

fullerene framework. (a) High-pressure and high-temperature encapsulations;<br />

(b) ion beam implantations; and (c) nuclear reactions have been developed to<br />

achieve such incorporations (Scheme 5.9).<br />

The high-pressure and high-temperature incorporations are generally carried<br />

out by heating a pristine fullerene at 650 °C under 300 MPa of a desired gas using<br />

a hydrostatic high-pressure vessel [382, 383]. A series of noble gas complexes, from<br />

He@C 60 to Xe@C 60 , have been synthesized by this method [383–389]. Fractions of<br />

the endohedral complexes in recovered fullerenes were 0.03–0.1%. Trace amounts<br />

of complexes containing two helium or neon atoms inside (He 2 @C n and Ne 2 @C n ,<br />

n = 60, 70) were also detected by 3 He NMR and/or mass spectrometry [382, 389].<br />

Chemical bonding between two atoms confined in a fullerene cage have been<br />

investigated in theory [390]. Similarly, formation of some diatomic molecule<br />

complexes with diatomic molecule quests, such as 13 CO@C 60 , N 2 @C 60 , and N 2 @<br />

C 70 , was confirmed by mass spectrometry [391]. Fractions of these products were<br />

estimated to be in the range of 0.02–0.05%. The reaction is applicable to higher<br />

or mixed fullerenes.<br />

Ion beam implantations have been mainly employed for the syntheses of group<br />

V atomic complexes, such as N@C 60 [382, 392]. Commonly, low energetic ions<br />

are provided to a continuously growing fullerene film. The amounts of fullerenes<br />

recovered were 10–20%, because the fullerenes employed decomposed. Fractions<br />

of the desired endohedral complexes in these portions were in the order of 0.01%.<br />

At present, successful examples are limited to N@C 60 , N@C 70 , P@C 60 , and<br />

N 2 @C 60 [393–396]. It has been suggested that an atomic complex with a higher<br />

fullerene cage would be unstable because decreased curvature of the cage might<br />

increase its reactivity toward a substrate inside [392, 394]. Also, the incorporation<br />

of arsenic has been attempted without success [392, 394].<br />

Though the smallest 1 H@C 60 has not been synthesized so far by the above<br />

mentioned methods, its isotopic and radioactive homolog, 3 H@C 60 ( 3 H = T,<br />

tritium), has been synthesized by the nuclear reaction of 6 Li with C 60 or neutron<br />

irradiation on the 3 He atom in 3 He@C 60 [17, 382, 383]. Counting a soluble portion<br />

in a scintillation counter showed the presence of tritium. Other radioactive<br />

complexes, such as 41 Ar, 85 Kr, and 133 Xe@C n (n = 60 or 70), were synthesized by<br />

neutron irradiations or ion implantation [398–400].<br />

283


284 5 Fullerenes<br />

Table 5.5 Separation factors in HPLC and 13 C NMR chemical shifts (in ppm) of X@C 60 .<br />

Separation<br />

factor<br />

in HPLC a)<br />

X@C 60<br />

13 C NMR, � �� b)<br />

Refs.<br />

In all cases, fractions of the desired complexes in recovered fullerenes are less<br />

than 0.5%. Despite such extremely low content, Ar@C 60 [384, 385], Kr@C 60 [386],<br />

Xe@C 60 [388], N@C 60 [393, 395], and N 2 @C 60 [396] were enriched to the analytical<br />

level by high performance liquid chromatography (HPLC). Multistage purification<br />

is essential because of poor separation factors (see retention times relative to that<br />

of C 60 collected in Table 5.5) in addition to low fractions.<br />

5.5.2.2 Molecular Surgery Approach via an Open-cage Fullerene<br />

Since the carbon–carbon bond of C 60 was cleaved to give a hole-opened, so-called<br />

an open-cage fullerene 51 (Scheme 5.10) [401], an alternative strategy has been<br />

proposed as a rational approach to obtain the desired endohedral complex at a<br />

high incorporation level [402]. It consists of the following three stages: (1) chemical<br />

cage scission to make an opening, (2) insertion of a guest through the opening,<br />

and then (3) closing the cage with the guest inside. He@51 prepared from He@<br />

C 60 revealed that an opening formed by one bond scission is too small to permit<br />

passage even of the smallest helium atom. Then, larger openings in 52–56 were<br />

created by regioselective and multiple bond scissions (Scheme 5.10) [403–413].<br />

Among successful molecular incorporations into these derivatives (described in<br />

Section 5.5.2.3), a pure and macroscopic quantity of H 2 @C 60 was recently synthesized<br />

from C 60 through 53 (Scheme 5.11) [406–408]. First, the 100% encapsulation<br />

of a hydrogen molecule into 53 was achieved at 80 MPa of H 2 and at 200 °C [406].<br />

Then, closure of the opening in H 2 @53 was carried out with a trapped H 2 in<br />

four steps to afford H 2 @C 60 [407, 408]. Although slight escape of the trapped H 2<br />

had taken place during the closure, a rich fraction of H 2 @C 60 , higher than 90%,<br />

permitted its complete separation by using the repeated HPLC system, despite<br />

an extremely low separation factor (Table 5.5).<br />

C 60<br />

Ar@C 60 1.04–1.05 143.4 143.2 c)<br />

Kr@C 60 1.09 143.6 143.2 d)<br />

Xe@C 60 1.08 144.5 143.5 c)<br />

+0.2 [384, 385]<br />

+0.4 [386]<br />

+1.0 [388]<br />

N@C 60 1.06–1.08 – – – [393, 395]<br />

N 2 @C 60 1.10–1.13 – – – [396]<br />

H 2 @C 60 1.01 142.84 142.76 e)<br />

a)<br />

Retention time relative to that of C60 .<br />

b)<br />

������(X@C60 )–�(C60 ).<br />

c)<br />

in C6D6. d)<br />

in C6D6/C6H6. e)<br />

in o-dichlorobenzene-d4.<br />

+0.08 [407, 408]


5.5 Endohedral Fullerenes with Neutral Atoms and Molecules<br />

Scheme 5.10 Molecular structures of open-cage C 60 derivatives 51–56.<br />

Scheme 5.11 Synthesis of H 2 @C 60 following the molecular surgery approach.<br />

5.5.2.3 Open-cage Fullerenes, Reversible Molecular Incorporations and Ejections<br />

Though successful restoration of the opening in open-cage fullerenes has been<br />

limited to H 2 @53, several molecular encapsulations have been performed using<br />

52–56. The first successful examples were incorporations of helium and hydrogen<br />

gas (H 2 ) into 52 [404]. Encapsulation of helium was achieved by heating solid 52<br />

at 305 °C under 48 MPa of helium (Table 5.6). The fraction of He@52 was 1.5%.<br />

Compound 53 has a larger opening than that of 52, and the same encapsulation<br />

could be done below 2 MPa [405]. Despite easy encapsulation, however, the fraction<br />

of He@53 was nearly the same as that of He@52. This is because molecular incorporation<br />

is a reversible and equilibrium process, and a larger opening allows<br />

285


286 5 Fullerenes<br />

Table 5.6 Endohedral He and H 2 complexes of open-cage C 60 derivatives: reaction conditions,<br />

fractions, energy barriers to escape (kcal mol –1 ), and NMR chemical shifts of the guests<br />

(in ppm).<br />

Condition Fraction<br />

(%)<br />

Barrier to escape<br />

exp. (theory)<br />

3 He/ 1 H<br />

NMR a)<br />

Refs.<br />

3 He@C60 – – – – –6.40 [38]<br />

3 He@2 48 MPa, ~305 °C 1.5 24.6 (24.3) –10.10 [404]<br />

3 He@3 2 MPa, 80 °C 1.5 22.8 (18.6) –11.86 [405]<br />

H 2 @C 60 – – – – –1.44 [408]<br />

H 2 @2 10 MPa, 400 °C 5 – (40.0) –5.43 [404]<br />

H 2 @3 81 MPa, 200 °C 100 34.3 (28.7) –7.25 [406]<br />

H 2 @4 13.5 MPa, 100 °C 83 21.7 (19.8) –7.34 [410]<br />

a) 3 He signals NMR relative to dissolved 3 He gas, 1 H signals NMR relative to TMS.<br />

not only for encapsulation but also rapid escape. Actually, the trapped helium<br />

atom can be released from the cage by heating. The activation energies for the<br />

helium escape from He@52 and He@53 were estimated to be 24.6 kcal mol –1<br />

and 22.8 kcal mol –1 , respectively, in accordance with the size of the openings in<br />

52 and 53 (Table 5.6) [404, 405]. Easy escape also suggests that there is negligible<br />

binding energy between the trapped helium and the fullerene cage.<br />

Encapsulation of molecular hydrogen has been achieved using 52–54 [404,<br />

406, 409, 410]. H 2 @52 was obtained in 5% fraction by heating solid 52 at 400 °C<br />

under 10 MPa of H 2 (Table 5.5) [404]. Compound 53 allowed for 100% H 2 encapsulation<br />

at 80 MPa of H 2 and at 200 °C as mentioned earlier [406]. Compound<br />

54 has a larger opening than those of 52 and 53, and allowed entry of H 2 under<br />

mild conditions [410]. The trapped H 2 can be released by heating, as in the case<br />

of helium complexes. Energy barriers of the H 2 escape from H 2 @53 and H 2 @54<br />

were estimated to be 34.3 kcal mol –1 and 21.7 kcal mol –1 , respectively (Table 5.6)<br />

[406, 410]. Recently, an useful index was reported for the evaluation of effective<br />

areas of these openings bearing different shapes and functional groups [409].<br />

Among the open-cage derivatives of fullerenes, compound 55 has the largest<br />

opening at present. Although endohedral fullerene complexes bearing a di- or<br />

tri-atomic molecule inside are still limited [391, 396, 407, 408], compound 55<br />

can hold one water or a carbon monoxide molecule inside the cage [411–413].<br />

Unlike other entries, a water molecule is incorporated into 55 spontaneously<br />

[411]. H 2 O@55 was obtained as a mixture with the empty 55 in the product. In<br />

solution, the trapped water molecule in H 2 O@55 is in rapid equilibrium with<br />

residual water in the solvent. The fraction of H 2 O@55 was about 75% in commercial<br />

CDCl 3 , but decreased with decline of the water content in the solvent.


5.5 Endohedral Fullerenes with Neutral Atoms and Molecules<br />

Also, in the 1 H NMR spectrum the signal of H 2O in H 2O@55 disappeared rapidly<br />

after treatment with D 2O. The inclusion property can be refined by the size and<br />

structure of the opening. Compound 56 has a similar but smaller opening than<br />

that in 55. The fraction of H 2O@56 was less than 10% under conditions identical<br />

to those in which H 2O@55 was obtained, but could be improved to 85% by<br />

refluxing in a mixture of toluene and water, indicating higher activation energies<br />

of both incorporation and escape [412]. Functional groups around the opening<br />

also affected the inclusion property of a water molecule.<br />

Carbon monoxide was incorporated into 55 by heating a solution of a mixture<br />

of H 2O@55 and 55 in the presence of CO [413]. The fraction of CO@55 reached<br />

84% under 9.0 MPa of CO and at 100 °C. CO@55 gradually released the trapped<br />

CO and reverted to a mixture of 55 and H 2O@55 under ambient conditions. This<br />

is in contrast to the spontaneous H 2O encapsulation into 55, indicating that H 2O<br />

binds more tightly than does CO.<br />

5.5.3<br />

Properties<br />

5.5.3.1 Host Fullerenes<br />

Basically, complete endohedral fullerenes are stable under ambient conditions. No<br />

decomposition was observed on H 2 @C 60 at 500 °C [407, 408]. Thermal escape of<br />

helium and neon from He@C 60 or Ne@C 60 only took place above 600 °C [381–383,<br />

414]. Since the energy required to escape from the pristine cage is too high to<br />

produce it thermally, it has been proposed that one or two fullerene bonds are<br />

reversibly cleaved to form a temporary window to allow the guest release. On the<br />

other hand, escape of the nitrogen atom in N@C 60 took place at a relatively low<br />

temperature (260 °C) [392, 394]. A different pathway through the insertion of the<br />

trapped nitrogen atom into a fullerene bond has been proposed to account for<br />

the difference. Endohedral open-cage complexes release the guests by heating,<br />

as described earlier.<br />

There are only weak van der Waals’ interactions between the guests and the<br />

fullerene cages. The 13 C NMR chemical shifts of the X@C 60 cages (X = guest)<br />

are representatives as well as poor separation factors in HPLC (Table 5.5). In all<br />

cases, the signals appeared downfield relative to that of the empty C 60 cage. The<br />

shift increased with the size of the guest but was less than 1 ppm even for the<br />

maximum case of Xe@C 60 .<br />

Slight changes were observed in the IR and UV/Vis spectra of X@C 60 as<br />

shown below, but further study is necessary for systematic considerations. In<br />

the IR spectrum of Kr@C 60 , three of the four well-known C 60 bands (528, 1183,<br />

and 1429 cm –1 ) shifted by 8–16 cm –1 to higher frequencies [386]. The remaining<br />

absorption band at 577 cm –1 became significantly weak, but the reason for this<br />

observation is not clear. The IR spectrum of Ar@C 60 also showed small (2–4 cm –1 )<br />

shifts below 600 cm –1 in the same direction, but opposite shifts were observed on<br />

other higher frequency bands [385]. For H 2 @C 60 , only a small shift was observed<br />

on the absorption band at 577 cm –1 [408]. The UV/Vis spectrum of Kr@C 60<br />

287


288 5 Fullerenes<br />

showed red shifts by 1–3 nm in the range of 580–640 nm [386], but no detectable<br />

change was observed on either Ar@C 60 or H 2@C 60 [385, 408]. N@C 60 displays<br />

a yellowish-brown color in solution, unlike the magenta color of C 60 [395]. Accordingly,<br />

broad absorption bands of C 60 in the range of 440–640 nm are notably<br />

suppressed in the case of N@C 60.<br />

The chemical reactivity of the fullerene cage is not affected by such a guest, as<br />

expected. Organic reactions developed for fullerenes are reproducible under the<br />

same conditions [383, 415, 416]. The availability of the products will be shown in<br />

the next paragraph.<br />

5.5.3.2 Guest Substrates<br />

In contrast to the insensitivity of host fullerenes, guests are quite sensitive to the<br />

cage size, electronic state, and organic addends of the hosts. This property has<br />

enabled their valuable applications as chemical probes and markers. The smallest<br />

helium atom has played a major role in NMR studies [382, 338]. The abundant 4 He<br />

is not an NMR active nucleus, but isotopic 3 He has a spin of I = 1/2 and exhibits<br />

a good NMR sensitivity. Generally, signals of the trapped atoms and molecules<br />

shift upfield due to magnetic shielding by the fullerene cage. Helium-3 in 3 He@<br />

C60 and 3 He@C70 appeared at � = –6.4 and –28.8 ppm, respectively, relative to the<br />

dissolved 3 He gas [383, 417]. Chemical shifts of the 3 He atoms trapped in higher<br />

fullerenes vary in the range between those of 3 He@C60 and 3 He@C70 . Higher<br />

fullerenes often have several isomeric structures with different symmetries, and<br />

each shows multiple signals in the 13 C NMR spectra. Also, multiple addition<br />

reactions on C60 generally produce an inseparable mixture of up to 8 regioisomers,<br />

and each adduct shows multiple signals in 13 C NMR due to a lowering of<br />

molecular symmetry. In contrast, a trapped 3 He atom in any structure shows only<br />

one single line in the 3 He NMR spectrum. Thus, 3 He complexes can be used as<br />

markers in the analyses of mixtures of higher fullerenes and multiple adducts<br />

of fullerenes [383, 417, 418]. Their use for studying structures of hydrogenated<br />

fullerenes will be discussed in Section 5.6.<br />

Organic addends on 3 He@C60 generally produce additional upfield shifts [383].<br />

A [5,6]-opened homofullerene structure retains the original 60 � elelectronic structure,<br />

and the internal 3 He atom shows only a small shift (�� = –0.2 ppm) (Table 5.7).<br />

On the other hand, a [6,6]-closed structure accompanies a loss of �-electrons, and<br />

the trapped He atoms show much larger shifts (�� = –1.8 – –3.2 ppm) than those in<br />

the former [5,6]-opened structure. Adding electrons to 3 H@C60 produces a drastic<br />

6–<br />

upfield shift [417]. The hexa-anion, He@C60 showed the signal at � = –48.7 ppm<br />

corresponding to �� = –42.4 ppm relative to that of He@C60 . It should be noted<br />

that He@C70 shows opposite downfield shifts by both organic addends and electronic<br />

reductions. For example, 3 6–<br />

He@C70 showed the signal at � = +8.3 ppm,<br />

which corresponded to �� = +37.1 ppm relative to He@C70 .<br />

Recent success in the macroscopic synthesis of H2@C60 has allowed measurement<br />

of the most familiar 1 H NMR spectra. The trends observed were the same<br />

as in the case of 3 He@C60 . The trapped H2 molecule in H2@C60 appeared at<br />

� = –1.44 ppm, which corresponded to �� = –5.97 ppm relative to the dissolved H2


5.5 Endohedral Fullerenes with Neutral Atoms and Molecules<br />

Table 5.7 3 He and 1 H NMR chemical shifts (in ppm) of the functionalized endohedral<br />

fullerenes with 3 He atom or H 2 molecule inside [383, 408].<br />

3 He NMR –6.40 –6.63 –8.11 –8.11 –9.45<br />

1 H NMR a)<br />

a) Relative to dissolved H2 gas (not TMS).<br />

–5.97 – – –7.80 –9.17<br />

gas (Table 5.7) [407, 408]. Organic addends induced additional upfield shifts [407,<br />

408, 416]. The relaxation time (T 1 ) of H 2 in H 2 @C 60 was in the range of 0.04–0.12 s,<br />

10–20 times shorter than that of the dissolved H 2 (0.84–1.44 s) [419].<br />

With respect to the guests in open-cage C 60 derivatives, creating an opening<br />

on the C 60 surface induces an upfield shift similar to those observed in organic<br />

addends. Shifts become larger with increase in the size of the openings in both 3 He<br />

and 1 H NMR. 3 He@52 and 3 He@53 showed upfield shifts of �� = –3.70 ppm and<br />

�� = –5.46 ppm, respectively, relative to 3 He@C 60 [404, 405] (Table 5.6). Similarly,<br />

H 2 @52–54 showed upfield shifts of �� = –3.99 ppm (H 2 @52), �� = –5.81 ppm<br />

(H 2 @53), and �� = –5.90 ppm (H 2 @54) relative to H 2 @C 60 [404, 406, 408, 410].<br />

The proton signals of water molecule in H 2 O@55 appeared at � = –11.4 ppm in<br />

the 1 H NMR spectrum [411]. Although H 2 O@C 60 has not still been synthesized, it<br />

corresponds to the upfield shift by �� = –13 ppm relative to the residual water in<br />

the solvent (� = 1.6 ppm). The signal was a sharp singlet indicating free rotation of<br />

the trapped H 2 O in the C 60 cage. The carbon signal of the trapped CO in CO@55<br />

appeared at � = 174.6 ppm shifted by �� = –10 ppm relative to the dissolved free<br />

CO gas (� = 184.6 ppm) [413]. The signal appeared as a singlet line even at –80 °C,<br />

indicating that CO rotates rapidly on the NMR time scale. In the IR spectrum,<br />

in contrast, two CO absorption bands were observed at � = 2125 and 2112 cm –1 .<br />

These are shifted by –18 and –31 cm –1 to lower frequencies, respectively, from<br />

the CO gas frequency (� = 2143 cm –1 ). The observed difference between NMR<br />

and IR can be interpreted in terms of the difference in the time scale of these<br />

measurements.<br />

As for N@C 60 , atomic nitrogen has unpaired electrons and is paramagnetic. The<br />

excellent sensitivity of EPR (electron paramagnetic resonance) spectroscopy has<br />

allowed the detection and analysis of N@C 60 even at an extremely low fraction.<br />

289


290 5 Fullerenes<br />

The ESR spectrum of N@C60 shows three lines originating from the hyperfine<br />

inter action of the electron spin with the nuclear spin of 14 N (I = 1) [392–395],<br />

indicating that the trapped nitrogen atom keeps its quartet spin ground state<br />

configuration without an electron transfer to the cage. Regarding cage size,<br />

electronic state, and organic addends, trends observed were the same as in the<br />

cases of 3 He and 1 H NMR. Compared with N@C60, the central line of the triplet<br />

of N@C70 shifted upfield by 0.0065 mT in the magnetic field corresponding to<br />

6–<br />

–19 ppm [392]. The hexa-anion N@C60 showed an upfield shift by –35 ppm<br />

6–<br />

relative to N@C60, but N@C70 did a downfield shift by +41 ppm relative to N@<br />

C70. Organic addends also caused an upfield shift for N@C60 but a downfield<br />

shift for N@C70 [392].<br />

5.5.4<br />

Binding Energies, Theoretical Investigations<br />

To maximize yield of the endohedral fullerenes, and to construct an efficient host<br />

system using open-cage fullerene derivatives, it is important to understand the<br />

interaction between the guest and the fullerene cage. Generally, trapped guests<br />

have been believed to have attractive interactions with the fullerene cage, because<br />

molecular incorporation is an entropically disfavored process. Various theoretical<br />

studies have been carried out along with synthetic experiments [420–427].<br />

The binding energies of noble gas complexes of C 60 were estimated by ab<br />

initio and DFT (density functional theory) calculations. At the MP2 (secondorder<br />

Møller–Plesset perturbation) theory, a series of noble gas complexes, from<br />

He@C 60 to Xe@C 60 , were computed to have stabilization energies of –0.3 to<br />

–7.5 kcal mol –1 , indicating attractive interactions in all cases [421, 422]. In contrast,<br />

the frequently used B3LYP (Becke three-parameter exchange functional coupled<br />

with the Lee-Yang-Parr correction) functional in DFT studies generally underestimates<br />

weak molecular interaction and predicts repulsive destabilization [421,<br />

423]. For example, H 2 @C 60 was estimated to be unstable by +1.6 kcal mol –1 at<br />

the B3LYP/6-31G(d,p) level, whereas it was stabilized by –4.0 kcal mol –1 at the<br />

MP2/6-31G(d,p) [423, 424]. Recently, the MPWB1K functional (modified Perdew-<br />

Wang and Becke functionals) of DFT was reported to give results consistent with<br />

the MP2 calculations (–7.9 kcal mol –1 for H 2 @C 60 ) [424]. In a real experiment,<br />

however, the DSC (differential scanning calorimetry) measurement for the H 2<br />

escape from the open-cage H 2 @54 suggested that the trapped H 2 destabilizes<br />

the cage [410]. Also, simple molecular mechanics and semiempirical quantum<br />

calculations have been discussed over a variety of endohedral complexes including<br />

multiple guest systems [425–427]. The possibility of housing more than one H 2<br />

molecule in the C 60 and higher fullerene cages has been examined by Dodziuk<br />

[426]. Further experiments with a variety of guests are needed to reach reliable<br />

and systematic conclusions.


5.5.5<br />

Summary<br />

5.6 Hydrogenated Fullerenes<br />

Endohedral nonmetal complexes of fullerenes allow us to study the inside of the<br />

fullerene sphere. The properties of host fullerenes scarcely change in response<br />

to the neutral guest substrates. In contrast, guests are quite sensitive to the host<br />

size, structure, electronic state, and organic addends, including chemical openings.<br />

Along with first generations produced by forced gas penetrations, new endohedral<br />

complexes are beginning to be synthesized by the molecular surgery method,<br />

which will continue to expand the library of endohedral fullerenes in macroscopic<br />

quantities and at high incorporation levels, leading to practical applications.<br />

5.6<br />

Hydrogenated Fullerenes<br />

Mark S. Meier<br />

Fullerenes are among the most highly strained aromatic compounds known, approaching<br />

500 kcal mol –1 in strain energy. Pyramidalization of sp 2 carbon atoms,<br />

by ~8° – 12° in fullerenes, is required to form the closed shells, but also results<br />

in a degree of rehybridization [428] that diminishes overlap in the �-system. The<br />

pyramidalization induces strain, and in turn that strain introduces reactivity<br />

[429] not normally seen in other aromatic systems. The interplay between strain,<br />

aromaticity [430], thermodynamics, and kinetics leads to a very rich chemistry,<br />

producing some beautiful structures in situations where one might expect intractable<br />

mixtures of products. Fullerene reactions are also discussed in more<br />

general terms in Section 5.2.<br />

The hydrogenation of fullerenes [431] provides a clear context for the study of<br />

fullerene reactivity. The continuous series of C 60H 2n, with n ranging (at least in<br />

principle) from 1 to 30, covers a molecular formula range from C 60 to C 60H 60.<br />

Hydrogen (followed closely by fluorine) is the least bulky covalent addend that<br />

may be bonded to a fullerene, and hence provides an excellent probe for revealing<br />

the reactivity of the fullerene core without interference from steric effects.<br />

In practice, only a handful of the 30 different C 60H 2n molecular formulae are<br />

formed in the course of most hydrogenations of C 60. This result is striking in its<br />

own right, but the selectivity for a limited number of isomers, when vast numbers<br />

are possible in most cases, is truly amazing.<br />

5.6.1<br />

Synthesis and Structure<br />

The simplest of the hydrogenated fullerenes is C 60 H 2 . This compound has been<br />

prepared by a long list of methods, including methods as diverse as hydroboration<br />

[432], electrochemical reduction/protonation [433–435], dissolving metal<br />

291


292 5 Fullerenes<br />

Figure 5.51 C 60 H 2 (57), shown in a 3-dimensional view and in a planar projection<br />

(Schlegel diagram).<br />

reduction/protonation [436], borohydride reduction [437], hydrozirconation [438],<br />

thermal [437] and sonochemical [439] transfer hydrogenation, direct hydrogenation<br />

with metal catalysts [440–444], and with diimide [440]. This compound has<br />

been made indirectly through decomposition of a Si-bridged C 60 dimer [445], as<br />

well as by hydrolysis of acylfullerenes [446], and it has been detected in fullereneproducing<br />

flames [447].<br />

While a considerable number of isomers of C 60 H 2 are possible [448, 449],<br />

only one has been prepared, purified, and fully characterized. The 1,2-isomer<br />

(57, Figure 5.51) appears to be the most thermodynamically stable isomer, and<br />

given the high acidity of fullerene C–H bonds [450, 451], it seems likely that any<br />

unstable isomers that may be formed would be transient and easily undergo<br />

isomerization.<br />

C 60 H 2 forms a brown solution in toluene, strikingly different from the magenta<br />

color of C 60 solutions. The 1 H NMR spectrum is quite simple – a singlet at<br />

5.12 ppm (CS 2 solution). The significant downfield shift of this resonance is likely<br />

caused by large ring currents in the remaining �-electron system [452]. The 13 C<br />

NMR spectrum is consistent with the C s -symmetry, with a single sp 3 resonance<br />

at 54.2 ppm [453]. Notably, the 1 H-coupled 13 C spectrum reveals the 2-bond H-C<br />

coupling constant (6.7 Hz), which is a useful indicator of a 1,2 arrangement of<br />

hydrogen atoms on the fullerene surface.<br />

Essentially any reagent that is capable of hydrogenating C 60 is likely to be capable<br />

of hydrogenating C 60 H 2 to C 60 H 4 or farther. An elegant study by Cahill produced<br />

and characterized several isomers of C 60 H 4 [454]. Hydroboration–protonolysis of<br />

C 60 produces a mixture of six isomers (Figure 5.52) of which two could be isolated<br />

and definitively identified, with 58a being the major product. A mixture of C 60 H 4<br />

compounds could be isomerized on platinum to the 1,2,3,4-isomer (cis-1) 58a,<br />

indicating that this not only the kinetic isomer but is likely the thermodynamically<br />

most stable isomer as well. This crucial experiment validates the computational<br />

prediction that the cis-1 isomer (58a) is one of the lowest energy structures studied.<br />

Dissolving metal reduction also produces C 60 H 4 , with 58b and 58c being identified<br />

in the Zn(Cu) reduction [455].<br />

Dissolving metal reduction of C 60 using Zn(Cu) results in the formation of a<br />

major isomer 59 (trans-3, trans-3) and 2 minor isomers [455]. The remarkably<br />

simple 1 H (singlet) and 13 C NMR spectra (10 resonances) indicates a structure


Figure 5.52 Isomers of C 60 H 4 .<br />

Figure 5.53 C 60 H 6 (59).<br />

5.6 Hydrogenated Fullerenes<br />

with D 3 -symmetry, and the measured 6.8 Hz 2-bond H–C–C coupling constant<br />

(indicating a 1,2-arrangement of hydrogens in 3 equivalent sets) permits assignment<br />

of the structure (Figure 5.53).<br />

Chromatography suggests that a second, minor isomer is formed along with<br />

59, and investigation of the 3 He NMR spectra of 3 HeC 60 H 4 proves that there are at<br />

least two minor isomers present in the Zn(Cu) reduction [456], although the structures<br />

have not been determined. Clearly the major isomer 59 does not result from<br />

hydrogenation of the most stable isomer of C 60 H 4 (58a) but from one of the less<br />

stable ones (58b). Computational work [457] has suggested that most stable C 60 H 6<br />

structures include one related to the thermodynamic isomer of C 60 H 4 (58a) and<br />

one that is analogous to the C 60 Cl 6 structure [458]. This result highlights the fact<br />

that fullerene chemistry is a kinetic world not always ruled by thermodynamics.<br />

Beyond C 60 H 6 lies a homologous series of compounds (C 60 H 8 , C 60 H 10 , C 60 H 12 ,<br />

etc.) in which no one species emerges as significantly more stable than others<br />

in the immediate neighborhood. Accordingly, to date none of these species has<br />

been isolated in good yield in pure form. The remarkable exceptions to this rule<br />

are C 60 H 18 and C 60 H 36 . The fact that specific isomers of these two molecular<br />

formulae appear as islands in a vast sea of possible formulae and isomers is one<br />

of the most startling results in fullerene chemistry.<br />

293


294 5 Fullerenes<br />

Figure 5.54 C 3v -C 60 H 18 60. Left and middle: bottom and side views; right: planar projection.<br />

C 60H 18, like C 60H 36 (below), is more stable than the nearby homologs (C 60H 16,<br />

C 60H 20, etc), appearing as a significant product in the Birch reduction [459],<br />

transfer hydrogenation [460], and in the Zn/HCl reduction [461] of C 60. These<br />

reactions could produce every adjacent oxidation state, but C 60H 18 is formed to<br />

the near exclusion of the adjacent formulae. This compound is formed as a C 3v<br />

‘turtle shell’ isomer 60 [462]. This structure contains a nearly planar benzene ring<br />

at the base (Figure 5.54).<br />

A wide array of reaction conditions lead to this same C 3v isomer of C 60 H 18 .<br />

These methods include reduction of C 60 with amines [463], hydrogenation [464,<br />

465], and dehydrogenation of C 60 H 36 [466].<br />

Reduction of C 60 with lithium in ammonia containing t-BuOH (Birch reduction)<br />

results in the formation of C 60 H 36 [459]. This result is striking, as there are<br />

numerous adjacent oxidation states (… C 60 H 34 , C 60 H 38 , etc) that might be formed,<br />

but C 60 H 36 is formed as a major product to the near exclusion of other products.<br />

Other reagents also lead to C 60 H 36 , including transfer hydrogenation [467, 468],<br />

reduction by Zn metal in acidic solution [469], and catalytic hydrogenation<br />

[441].<br />

The number of possible isomers for C 60 H 36 is truly staggering, exceeding 10 14<br />

by one report [448]. A number of methods for preparation of this compound have<br />

been reported [441, 459, 460, 467, 469, 470], but most result in complex mixtures<br />

of species from which C 60 H 36 must be isolated. Transfer hydrogenation [460, 467,<br />

468, 471] using 9,10-dihydroanthracene produces a mixture of three isomers,<br />

with a C 1 isomer 61a being strongly dominant, although a smaller amount of a<br />

C 3 isomer 61b and a trace of T isomer 61c are observed [466]. Detailed NMR work<br />

on this mixture has led to the structural assignments shown (Figure 5.55).<br />

It is believed that the formation of this limited array of structures, out of the<br />

huge number possible, is the result of rapid isomerization under the conditions<br />

of transfer hydrogenation. Further heating of this mixture results in conversion<br />

of the T and C 1 isomers to the C 3 isomer, supporting both the notion of thermal<br />

isomerization as well as supporting calculations that suggest that the C 3 isomer<br />

is the most stable.<br />

Reduction of C 60 H 36 with lithium in diethylamine leads to a mixture of highly hydrogenated<br />

products, composed of C 60 H 38 though C 60 H 44 [472]. These compounds<br />

were detected by MS, but despite extensive HPLC work no one compound could<br />

be isolated. 3 He NMR spectra of the Li/diethylamine reduction of He@C 60 H 36


Figure 5.55 C 60 H 36 structures.<br />

5.6 Hydrogenated Fullerenes<br />

confirms that numerous different compounds are present. A small amount of<br />

C 60 H 48 was observed in the Zn/HCl reduction of C 60 , but again no structural assignment<br />

has been made [469]. Isolation and definitive structural assignment of<br />

any member of this class of compounds is a major challenge, as they are relatively<br />

insoluble (even among fullerenes) as well as fragile, decomposing in solution in<br />

air and light.<br />

The endpoint of hydrogenation of C 60 , C 60 H 60 , has not been isolated or characterized.<br />

This is not surprising, as the large number of eclipsing interactions<br />

would make an all-exo structure impossibly high in energy. Structures with some<br />

of the hydrogens on the interior would be significantly more stable [473, 474], but<br />

synthesis is a challenge. These structures are discussed in Section 2.4.7.1.<br />

5.6.2<br />

C 70 Chemistry<br />

The reactivity of C 70 is similar to that of C 60 and as expected, the less pyramidalized<br />

[429] carbons along the equator are typically less reactive than the more pyramidalized<br />

carbon at the poles. As is the case with the reduction of C 60 to C 60 H 2 ,<br />

reduction of C 70 to C 70 H 2 can be accomplished by a variety of methods and has<br />

been demonstrated with many of the same reagents used with C 60 [440].<br />

The first step in the hydrogenation of C 70 results in the formation of 1,2-C 70 H 2<br />

(62a, Figure 5.56), resulting from hydrogenation of the most highly strained part<br />

of the molecule [475, 476]. This regiochemistry is the common result for addition<br />

to one C-C bond in C 70 , being observed as the major product in Diels–Alder,<br />

Hirsch-Bingel, Prato, and other additions. The 2-bond H-C-C coupling constants<br />

(4.7–5.2 Hz) in 62a are somewhat smaller than the analogous coupling in C 60 H 2 .<br />

A smaller amount of the 5,6-isomer 62b is formed along with 62a, resulting from<br />

reduction of the second most strained site in C 70 .<br />

295


296 5 Fullerenes<br />

Figure 5.56 1,2-C 70 H 2 (6a), 5,6-C 70 H 2 (62b), and C 70 H 4 (63a).<br />

The second hydrogenation step is an interesting one, because if strain determines<br />

the regiochemistry of addition then at least five different isomers of C 70 H 4<br />

could result from the reduction of C 70 H 2 (62a), resulting from addition to any one<br />

of the remaining 6–6 ring fusions at the poles. In practice, the Zn(Cu) reduction<br />

of C 70 results in the formation of C 70 H 4 (63a) from C 70 H 2 (62a) [476].<br />

The Zn(Cu) reduction of C 70 produces C 70 H 8 (64) as a major product (Figure<br />

5.57) [476]. This compound does not form by reduction of C 70 H 4 (63a), but<br />

instead from an ‘unseen’ pathway starting at C 70 and proceeding through a series<br />

of intermediates (presumably C 70 H 2-6 species) that are reactive and rapidly reduce<br />

again, until this relatively stable isomer is formed. At that point, 64 is relatively<br />

slow to reduce again, and hence it builds in concentration.<br />

Structure 64 is analogous to the structure of C 70 Cl 10 [477], but is unusual for the<br />

1,4-pattern of hydrogenation. A 1,2-pattern is found is virtually all other examples.<br />

The crowded 1 H NMR resonances are more upfield than seen in 62a and 63a,<br />

consistent with hydrogenation away from the poles. The 1,4-pattern is suggested<br />

from the absence of the familiar 2-bond C–H coupling constant, replaced instead<br />

by smaller coupling constants.<br />

Figure 5.57 C 70 H 8 (64).


Figure 5.58 Polar views of C 70 H 8 (64) and C 70 H 10 (65).<br />

5.6 Hydrogenated Fullerenes<br />

Further reduction of 64 leads primarily to C 70H 10 (65), although two minor<br />

and unidentified isomers are also formed (Figure 5.58). The 13 C spectrum of 65<br />

indicates a plane of symmetry, and a typical 2-bond C–H coupling constant (5.6 Hz)<br />

appears again. This structure results from reduction of 64 in such a manner as to<br />

cause minimal disruption of the remaining aromatic system.<br />

As with C 60, powerful hydrogenating conditions result in highly hydrogenated<br />

compounds. With extended reaction times, the Zn(Cu) reduction produces<br />

material beyond C 70 H 10 , but numerous compounds result. The Zn/HCl reduction<br />

can also produce highly hydrogenated materials,<br />

Strongly reducing conditions (catalytic hydrogenation [441], Zn/HCl reduction,<br />

transfer hydrogenation [460, 468]) leads to the formation of C 70 H 36 , in a close<br />

parallel to the fate of C 60 under the same conditions.<br />

5.6.3<br />

Higher Fullerenes<br />

Reduction of C 76 (as a mixture of isomers) with Zn in HCl leads to a mixture<br />

of products, ranging from C 76 H 46 to C 76 H 50 . Likewise, treatment of C 78 and C 84<br />

(again, both as mixtures of isomers) under these conditions produces complex<br />

mixtures of products. The reduction of C 78 produces a spectrum of products, with<br />

C 78 H 36 being a major product accompanied by C 78 H 48 .<br />

Reduction of the higher fullerenes has been reported to be accompanied by a<br />

measure of breakdown of the fullerene cages, as evidenced by the observation<br />

C 60 H 36 and/or C 70 H 36 species [478].<br />

5.6.4<br />

Reactivity of Hydrogenated Fullerenes<br />

Hydrogenation of fullerenes is an easily reversible process, with dehydrogenation<br />

back to the parent fullerene being brought about by reagents such as DDQ<br />

[459] and metal hydrogenation catalysts [479]. The most remarkable aspect of the<br />

chemistry of the higher fullerenes is the pronounced acidity of the C-H bonds.<br />

It is possible to deprotonate t-BuC 60 H with weak bases such as acetate, and the<br />

pK a of the fullerene C-H bond was estimated to be 5.7 [450]. In DMSO, the pK a<br />

of C 60 H 2 was determined to be 4.6, quite acidic for a C n H n compound [451].<br />

297


298 5 Fullerenes<br />

The acidity of other hydrogenated fullerenes is pronounced [480, 481], and a<br />

prediction of 8.0 has been made for C 60H 6 [480].<br />

In practice, deprotonation of hydrogenated fullerenes is a facile method for<br />

formation of fullerene anions. These anions react readily with electron deficient<br />

alkenes such as acrylonitriles to form aminocyclopentenes [482]. Treatment of<br />

C 60H 2 with weak bases and alkyl halides produces 1,2-monoalkyl- or 1,4-dialkyl<br />

fullerenes [483] although yields are mediocre, presumably because the anions<br />

undergo rapid oxidation by adventitious oxidants. Alkylation with alkyl tosylates<br />

is not effective, because fullerene anions typically react by electron transfer [484]<br />

rather than by a standard S n2 mechanism.<br />

Dialkylation of C 60H 2 with CH 3I results in a mixture of 1,2- and 1,4-dialkylation,<br />

which is consistent with 1,2-addition being nominally the most stable arrangement,<br />

but the crowding of addends destabilizes this arrangement, leading to the<br />

second alkylation step proceeding at an allylic position.<br />

Treatment of C 70H 2 (63a) with one equivalent of base and ethyl bromoacetate<br />

results in a 37 : 1 ratio of C-1 alkylation 66a to C-2 alkylation 66b (Figure 5.59)<br />

[485]. The selectivity here probably results from more rapid deprotonation at C-1,<br />

a reaction that probably produces a more stable, delocalized anion than would be<br />

produced by deprotonation at C-2.<br />

Figure 5.59 Monobenzyl C 70 from alkylation at C-1 (66a) and from alkylation at C-2 (66b).<br />

Treatment of C 70 H 2 (62a) with two equivalents of base and alkyl halide produces<br />

a mixture of products (Figure 5.60). While a trace amount of the 1,2-bibenzyl<br />

product 67a is obtained the dominant product, resulting from alkylation near the<br />

equator 67b, is obtained in only 10% yield [483].<br />

2–<br />

Figure 5.60 Products of dibenzylation of C70.


5.7<br />

Applications of Fullerenes<br />

Rossimiriam Pereira de Freitas and Jean-François Nierengarten<br />

5.7.1<br />

Introduction<br />

5.7 Applications of Fullerenes<br />

When fullerenes were first discovered more than two decades ago [486], there was<br />

much excitement about possible applications for this new molecular material. In<br />

spite of initial speculation, the fullerenes could not be extensively studied until<br />

1990 when Krätschmer et al. [29] made C 60 available in macroscopic quantities.<br />

Since then, the unusual properties of this fascinating allotropic carbon form have<br />

been intensively investigated. Among the most spectacular findings, C 60 was<br />

found to behave like an electronegative molecule able to reversibly accept up to six<br />

electrons [487], to become a supraconductor in M 3 C 60 species (M = alkali metals)<br />

[488, 489] or to be an interesting material with nonlinear optical properties [490].<br />

Although possessing exceptional properties, C 60 is difficult to handle because<br />

it forms aggregates and is insoluble or only sparingly soluble in most solvents<br />

[491]. This serious obstacle for practical applications can be overcome, at least in<br />

part, with the help of organic modification. Effectively, the recent developments<br />

in the functionalization of fullerenes allow the preparation of highly soluble C 60<br />

derivatives which are easier to handle, and whose electronic properties, such as<br />

facile multiple reducibility, optical nonlinearity or efficient photosensitization<br />

that are characteristic of the parent fullerene, are maintained for most of the C 60<br />

derivatives [492–497].<br />

After several years of research, fullerene chemistry is now a well-established field<br />

and the knowledge acquired has revealed both potentials and limitations of this<br />

class of compounds. The present chapter illustrates some of the most promising<br />

applications for chemically modified fullerenes. The objective is not to present<br />

an exhaustive review but to describe some of the most illustrative examples in<br />

materials science and biology.<br />

5.7.2<br />

Applications in Materials Science<br />

5.7.2.1 C 60 Derivatives for Optical Limiting Applications<br />

Optical Limiting (OL) is a nonlinear phenomenon in which the absorption of a<br />

material increases when the incident radiation intensity increases. Materials or<br />

devices with transmission that decreases with light level are called optical limiters<br />

and can potentially be used to protect optical sensors, including the human eye,<br />

from dangerous laser beams. Recently much effort has been invested in the<br />

research of organic materials that can behave as nonlinear absorbers because<br />

they are, in general, easily integrated into optical devices.<br />

C 60 itself has been widely investigated for potential application in the field of<br />

optical limiting [490, 498, 499]. Effectively, the transmission of fullerene solutions<br />

299


300 5 Fullerenes<br />

decreases by increasing the light intensity. For short pulses (ps), the limiting action<br />

is ascribed to pure reverse saturable absorption (RSA), whereas for longer pulses<br />

(ns-µs) thermal effects are also invoked [500–505]. Even if fullerene solutions are<br />

efficient optical limiters, the use of solid devices is largely preferred for practical<br />

applications because they are easier to handle. Therefore, crystalline films of C60 have been studied, but found to be inefficient against pulses longer than tens of<br />

ps. This result is ascribed to a fast de-excitation of the laser-created excited state<br />

due to the interactions of neighboring C60 molecules in the solid phase [501]. In<br />

contrast, it has been shown that C60 keeps its limiting properties after inclusion in<br />

solid matrices such as sol–gel glasses [506–508], polymethylmethacrylate (PMMA)<br />

matrices [509] and glass–polymer composite samples [510]. As far as sol–gel glasses<br />

are concerned, special procedures have to be employed since good solvents for<br />

fullerenes are incompatible with the sol–gel process [500, 511–513]. Actually, incorporation<br />

of fullerenes is typically achieved by soaking mesoporous silica glasses<br />

with a solution of C60 [506–508]. Another efficient approach is to incorporate<br />

water-soluble C – 60 derivatives, compatible with the sol–gel process, directly into<br />

Figure 5.61 Transmission versus incident fluence at 532 nm of a sol-gel sample containing<br />

compound 68.


5.7 Applications of Fullerenes<br />

the sol [514]. For example, the water soluble methanofullerene 68 (Figure 5.61)<br />

has been synthesized [514] and successfully included in a sol–gel during the<br />

gelation process. The transmission of the sample at 532 nm as a function of the<br />

incoming laser fluence is shown in Figure 5.61. With increasing pulse energy, the<br />

transmission of the sample clearly decreases. The threshold for the onset of the<br />

limiting action is located at about 3 mJ cm –2 , a value comparable or even slightly<br />

lower than that obtained with inclusions of C 60 in sol–gel matrices [501]. The<br />

damage threshold of samples is about 200 mJ cm –2 . Up to this fluence, the effect<br />

is fully reversible. For higher values, cumulative damage of the C 60 molecules was<br />

observed in their glass environment. With the same pulse lengths, the threshold<br />

value is considerably smaller than that of materials showing simultaneous twophoton-absorption.<br />

Even among materials showing RSA, this value is quite good<br />

[501] and similar to that found in C 60 solutions, confirming the fact that C 60 keeps<br />

its favorable limiting properties even after chemical modification.<br />

However, for sol–gel glasses doped with plain C 60 or methanofullerene 68, faster<br />

de-excitation dynamics and reduced triplet yields were observed when compared<br />

with the solutions. The latter observations have been mainly explained by two<br />

factors: (1) perturbation of the molecular energy levels by the interactions with the<br />

sol–gel matrix and (2) interactions between neighboring fullerene spheres caused<br />

by aggregation [500, 514]. To prevent such undesirable effects, fullerodendrimers<br />

69–70 (Figure 5.62) in which the C 60 core is buried in the middle of a dendritic<br />

structure [514–519] were prepared and incorporated in sol–gel glasses by soaking<br />

mesoporous silica glasses with a solution of 69 and 70 [516].<br />

Figure 5.62 Fullerodendrimers 69 and 70.<br />

301


302 5 Fullerenes<br />

Figure 5.63 Transmission versus incident fluence at 532 nm of a sol-gel sample containing<br />

compound 70.<br />

The resulting samples contain only well-dispersed fullerodendrimer molecules.<br />

Measurements on the resulting doped samples have revealed efficient optical<br />

limiting properties [516]. The transmission as a function of the fluence of the laser<br />

pulses is shown in Figure 5.63. It remains nearly constant for fluences lower than<br />

5 mJ cm –2 . When the intensity increases above this threshold, the effect of induced<br />

absorption appears, and the transmission diminishes rapidly, thus showing the<br />

potential of these materials for optical limiting applications.<br />

Stable sol–gel glasses have been also prepared from several fullerene derivatives<br />

containing both solubilizing chains and siloxane groups (Figure 5.64) [520, 521].<br />

The optical-limiting properties of these derivatives have been investigated both<br />

in solution and in sol–gel glasses. By irradiation at 652 nm compounds 71–75<br />

gave better results than nonfunctionalized fullerene, while at 532 nm C 60 gave<br />

the best result.<br />

The photoinduced optical second-harmonic generation (PISHG) and the transmission<br />

versus the intensity of the laser beam have been studied in solution and<br />

in the solid state for a series of C 60 -tetrathiafulvalene (TTF) dyads (Figure 5.65)<br />

[522–524]. Compounds 76a–b and 77 revealed improved nonlinear optical (NLO)<br />

properties when compared with C 60 . Using molecular dynamics simulations and<br />

quantum chemical calculations, the authors have shown that an asymmetry in the<br />

excited states influences the NLO properties of these systems as a result of a photoinduced<br />

charge redistribution both on the intra- and intermolecular levels.<br />

The synthesis and electronic properties of a series of multiple C 60 terminated<br />

oligo(p-phenylene ethynylene) (OPE) hybrid compounds obtained by an in situ<br />

ethynylation method [525] was reported by Tour and coworkers (Figure 5.66). There<br />

was evidence of a synergistic interaction between the conjugated OPE backbone<br />

and fullerene. C 60 -OPE hybrid 77 presented an enhanced NLO performance<br />

relative to its OPE precursor; this behavior is presumably due to the occurrence<br />

of periconjugation and/or charge transfer effects in the excited state.


Figure 5.64 Fullerene derivatives containing siloxane groups.<br />

Figure 5.65 Fullerene-TTF dyads studied as optical limiters.<br />

Figure 5.66 Fullerene-OPE hybrid 78.<br />

5.7 Applications of Fullerenes<br />

303


304 5 Fullerenes<br />

Figure 5.67 Combination of phthalocyanine and fullerene moieties for optical limiting.<br />

In the optical limiting field, combination of phtalocyanine and fullerene moieties<br />

has been made by different researchers [526, 527] in order to provide improved<br />

performances through a synergistic effect. The optical limiting behavior of copper<br />

phthalocyanine-fullerene dyad 78 [526] (Figure 5.67) has been investigated in<br />

solution and as a nanoparticle dispersion using nanosecond laser pulses at<br />

532 nm. An enhanced optical limiting performance of the nanoparticle sample<br />

compared with that of the solution sample has been observed. The formation of<br />

ordered aggregates with a well-defined ‘face-to-face’ packing fashion is proposed<br />

to be responsible for the enhancement of the optical limiting performance of the<br />

nanoparticle sample.<br />

5.7.2.2 C 60 Derivatives for Photovoltaic Applications<br />

The interaction of C 60 with light has attracted considerable interest in the exploration<br />

of applications related to photophysical, photochemical and photoinduced<br />

charge transfer properties of [60]fullerene derivatives. Following the observation<br />

of ultrafast photoinduced electron transfer from �-conjugated polymers to C 60<br />

core [528] giving rise to long-lived charge-separated states [529], intensive research<br />

programs are focused on the use of fullerene derivatives acting as electron<br />

acceptors in organic solar cells. The development of these devices was stimulated<br />

by the inherent advantages of organic materials such as their low weight and cost,<br />

and by the possibility of fabricating large active surfaces thanks to their processability.<br />

After only ten years of studies, it is now clearly established that [60]fullerenebased<br />

materials are among the most important candidates for the expansion of<br />

plastic solar cells and for renewable sources of electrical energy [530].<br />

Similarly to the photosynthesis process in plants, organic solar cells are based<br />

on absorption of light followed by cascades of energy and electron transfer events.<br />

Typically, organic solar cells are constituted of at least four distinct layers [531],<br />

not counting the substrate, which may be glass or some flexible, transparent<br />

polymer (Figure 5.68). On top of the substrate the cathode is laid. Indium tin oxide<br />

(ITO), is a popular cathodic material because it is transparent, and glass substrate<br />

coated with ITO is commercially available. The cathode is very often aluminum<br />

(calcium, magnesium, gold are also used). Inserted (or sandwiched) between the


5.7 Applications of Fullerenes<br />

Figure 5.68 Typical structure of an organic solar cell (a) and production of photocurrent (b).<br />

two electrodes, the photoactive layer is responsible for light absorption, exciton<br />

generation/dissociation and charge carrier diffusion. These heterojunctions are<br />

typically fabricated using p-type donor (D) and n-type acceptor (A) semiconductors.<br />

A layer of the conductive polymer mixture poly(3,4-ethylenedioxythiophene)/<br />

poly(styrenesulfonate) (PEDOT-PSS) may be applied between the cathode and the<br />

active layer. The PEDOT-PSS layer serves several functions. Not only does it serve<br />

as a hole transporter and exciton blocker, but it also smooths out the ITO surface,<br />

seals the active layer from oxygen, and keeps cathode material from diffusing into<br />

the active layer, which can lead to unwanted trap sites. Under illumination, electron<br />

transfer from the donor to the acceptor and generation of excitons followed by<br />

charge separation and transport of carriers to the electrodes induces a photocurrent<br />

(see Figure 5.69). At present, thanks to the invention of N.S. Sariciftci and<br />

A.J. Heeger [532], one of the most used acceptors in heterojunction photovoltaic<br />

cells is plain C 60 or fullerene derivatives.<br />

From the theoretical point of view, the principal feature of the p–n heterojunction<br />

is the built-in potential at the interface between both materials presenting a<br />

difference of electronegativities [532]. In fact, the absorption of light induces the<br />

promotion of an electron from the highest occupied molecular orbital (HOMO)<br />

[or the valence band (VB)] of the donor to the lowest unoccupied molecular orbital<br />

(LUMO) [or the conducting band (CB)] of the acceptor, generating an exciton at<br />

the interface of the junction. Then the built-in potential and the associated difference<br />

in electronegativities of materials allow the exciton dissociation [533]. The<br />

charge separation occurs at D/A interfaces and free charge carriers are transported<br />

through semiconducting materials with the electron reaching the cathode (Al)<br />

and the hole reaching the anode (ITO).<br />

The first heterojunction with a conjugated polymer and C 60 was reported in<br />

1993 [534]. In this work, the device ITO/MEH-PPV C 60 /Au (or Al) was fabricated<br />

by sublimation of fullerene onto a MEH-PPV (poly[2-methoxy-5-(2’-ethylhexyloxy)-<br />

1,4-phenylenevinylene]) layer spin-coated on ITO-covered glass. This solar cell<br />

showed a relative high FF of 0.48 and a power conversion efficiency (PCE) of<br />

0.04% under monochromatic illumination.<br />

305


306 5 Fullerenes<br />

In such devices, the interaction between the electron donor and the electron<br />

acceptor materials is only effective at the interface, and is limited by the diffusion<br />

length of the exciton (near 20 nm maximum). As a consequence, low short-circuit<br />

photocurrent (Isc) values and conversion efficiency were obtained. In order to<br />

overcome these deficiencies, interpenetrating networks were developed as ideal<br />

photovoltaic materials for a high-efficiency photovoltaic conversion. A crucial and<br />

major breakthrough towards efficient organic devices was realized by Heeger, Wudl<br />

and coworkers with the development of the ‘bulk-heterojunction’ concept [535];<br />

that is an interpenetrating network of a (p-type) donor conjugated polymer and<br />

C 60 or another fullerene derivative as (n-type) acceptor material. Consequently, the<br />

photoactive layer of these solar cells consists of blending the conjugated polymer<br />

and the fullerene derivative.<br />

The effective interaction between the donor and the acceptor compounds within<br />

these so-called ‘bulk-heterojunction’ solar cells can take place in the volume of the<br />

entire device. Subsequently, the separated charge carriers are transported to the<br />

electrodes via an interpenetrating network. A major shortcoming of this kind of<br />

device is the tendency, especially for pristine C 60 , to phase separate and then to<br />

crystallize. This aggregation phenomenon imposes important consequences on<br />

the solubility of C 60 within a conjugated polymer matrix. For that reason, intensive<br />

efforts on organic solar cells rapidly focused on interpenetrated networks of conjugated<br />

polymers with the 1-(3-methoxycarbonyl)propyl-1-1-phenyl-[6,6]methanofullerene<br />

80 ([60]PCBM) (Figure 5.69). This compound, initially synthesized<br />

by Wudl and coworkers [536], is more soluble in organic solvents than pristine<br />

C 60 . The first example of blend between MEH-PPV and [60]PCBM 80 exhibited a<br />

PCE of 2.9% [530], but under monochromatic low intensity light [535].<br />

Figure 5.69 Structure of [60]PCBM 80 and some conjugated polymers commonly used for the<br />

preparation of organic solar cells.


5.7 Applications of Fullerenes<br />

Remarkably, the efficiency of the bulk heterojunction devices consisting of<br />

poly[2-methoxy-5-(3,7-dimethyloctyloxy)]-p-phenylene vinylene (MDMO-PPV) and<br />

[60]PCBM was increased to 3% by the group of Sariciftci [537]. In the recent years,<br />

regioregular polyalkylthiophenes (PAT) which combine the potential advantages of<br />

a better photostability and a smaller bandgap than PPV derivatives, were the most<br />

widely used �-donor polymers associated with [60]PCBM. Bulk-heterojunction<br />

solar cells consisting of poly(3-hexylthiophene), P3HT and [60]PCBM have reached<br />

a PCE around 5% [538, 539].<br />

However, for commercial use the efficiency and the stability of the organic<br />

photodiodes have to be improved dramatically. For these purposes, �-conjugated<br />

polymers with a strong absorption in all visible range and a good stability towards<br />

light are needed. Additionally, the initially formed phases between the donor<br />

and acceptor have to be fixed by either crosslinking the two compounds or using<br />

polymer/polymer mixtures with high T g since the two phases tend to separate as<br />

a result of the the operational heat through illumination, thus reducing progressively<br />

the performances of the device. Hence, in parallel to the development of<br />

the polymer/fullerene derivative bulk-heterojunctions, C 60 functionalized macromolecules<br />

[540–544] have been investigated for the preparation of all-polymer<br />

solar cells (Figure 5.70) [540, 541]. The controlled incorporation of fullerenes into<br />

well-defined linear polymers was achieved by polycondensation of a bifunctional<br />

fullerene with diols affording the C 60 -based polymer 81 with an average polymerization<br />

degree of 25 [540]. Photovoltaic cells were prepared by blending this<br />

soluble C 60 -polymer 81 with MDMO-PPV. The device ITO/PEDOT-PSS/MDMO-<br />

PPV:C 60 -polymer(1:5)/LiF/Al showed clear photovoltaic behavior (V oc = 0.36 V,<br />

Figure 5.70 C 60 -functionalized macromolecules for the preparation of all-polymer solar cells.<br />

307


308 5 Fullerenes<br />

J sc = 0.7 µA cm –2 , FF = 0.36) but these weak values of short- and open-circuit<br />

currents might be ascribed to the low conductivity of the fullerene polymer. The<br />

presence of large solubilizing groups may inhibit interactions between fullerenes,<br />

thus decreasing the charge transport [545].<br />

The performances of bulk heterojunction devices obtained from �-conjugated<br />

polymers and C 60 derivatives are very sensitive to the morphology of the blend<br />

[546]. Ideally (to ensure efficient exciton dissociation), an acceptor species should<br />

be within the exciton diffusion range from any donor species and vice versa.<br />

Moreover, both the donor and the acceptor phases should form a bi-continuous<br />

microphase separated network to allow bipolar charge transport. However, the<br />

donor and acceptor molecules are usually incompatible and tend to undergo uncontrolled<br />

macrophase separation. In particular, phase separation and clustering<br />

of the fullerene can occur caused by the operational heat through illumination,<br />

thus reducing the effective donor/acceptor interfacial area and the efficiency of<br />

the devices [546]. In order to prevent such undesirable effects, it was proposed<br />

that the bicontinuous network could be simply obtained by chemically linking a<br />

hole-conducting moiety to the electron-conducting fullerene subunit [547]. Based<br />

on these considerations, compound 82 in which an oligophenylenevinylene (OPV)<br />

moiety is covalently linked to the fullerene sphere (Figure 5.71) was prepared [547].<br />

The use of a fulleropyrrolidine derivative attached to an oligophenylenevinylene<br />

can be considered as the first example of a molecular heterojunction specifically<br />

designed for photovoltaic conversion. The fulleropyrrolidine OPV-C 60 82 was<br />

incorporated in a photovoltaic device by spin-casting between aluminum and<br />

ITO electrodes. The device ITO/OPV-C 60 82/Al delivered a low PCE of 0.01%<br />

(V oc = 0.46 V, J sc = 10 µA cm –2 , FF = 0.3) under monochromatic irradiation<br />

(400 nm, 12 mW cm –2 ). This study showed that plastic solar cells can be obtained<br />

by chemically linking the hole-conducting and the electron-conducting units. The<br />

length of the OPV was increased from three to four units and the performances of<br />

the ITO//OPV-C 60 83/Al were significantly improved with a monochromatic PCE<br />

of 0.03%. The limited efficiency of these devices was attributed to the competition<br />

between energy transfer and electron transfer [547].<br />

The OPV-C 60 hybrid 84 was tested as an active material in photovoltaic cells<br />

and the device ITO/PEDOT-PSS/84/Al presented enhanced I/V characteristics<br />

(V oc = 0.65 V, J sc = 235 µA cm –2 under white light illumination at 65 mW cm –2 )<br />

Figure 5.71 OPV-C 60 hybrids as molecular heterojunctions for photovoltaic conversion.


5.7 Applications of Fullerenes<br />

(Figure 5.72). Even if these solar cells are not prepared under similar conditions<br />

used for OPV-C 60 derivatives 82 and 83, the increased donating ability of the<br />

OPV moiety is an important argument for the improvement as that of the device<br />

performance [548]. Photovoltaic devices were also prepared from oligophenyleneethynylene<br />

– C 60 (OPE-C 60) oligomers 85 and 86. Under light, both devices show<br />

clear photovoltaic behavior. Interestingly, the performances of the devices prepared<br />

from the N,N-dialkylaniline terminated derivative 86 are significantly improved<br />

when compared with those obtained with 85. The latter observation can be related<br />

to the differences in their first oxidation potentials. Effectively, due to the increased<br />

donating ability of the OPE moiety in 86 when compared with 85, the energy level<br />

of the charge separated states resulting from a photoinduced electron transfer is<br />

significantly lower in energy. Therefore, the thermodynamic driving force is more<br />

favorable, thus electron transfer which is one of the key step for the photocurrent<br />

production must be more efficient for 86. As a result, the power conversion efficiency<br />

of the devices is increased by one order of magnitude [549]. This clearly<br />

demonstrated the interest of the molecular approach, which allows to establish<br />

structure/activity relationships.<br />

Figure 5.72 OPV- and OPE-C 60 hybrids tested as active material in photovoltaic cells.<br />

309


310 5 Fullerenes<br />

5.7.3<br />

Biological Applications<br />

[60]fullerene derivatives have attracted attention regarding their pharmacological<br />

properties since their discovery, but the low solubility of this material in aqueous<br />

media was initially an obvious problem for biological studies. Over the past few<br />

years, several strategies have been developed to overcome the natural repulsion<br />

of fullerenes for water and render them biocompatible. Generally, these strategies<br />

including chemical covalent modification of their surface with polar groups<br />

as terminal amines, alcohols, carboxylic acids, amino acids [550, 551] and sugars<br />

[552], or preparation of water soluble supramolecular complexes with macrocyclic<br />

host systems such as cyclodextrin, cyclotriveratrylene or calixerene derivatives<br />

[553–556]. Once in solution, fullerenes and derivatives exhibit a wide range of<br />

biological activity [557–564].<br />

A potential application of fullerene derivatives is related to the easy photoexcitation<br />

of C 60 by visible light. The resulting singlet excited-state 1 C 60 is readily<br />

converted to the long-lived triplet 3 C 60 via intersystem crossing. In the presence<br />

of molecular oxygen, the fullerene can decay from its triplet to the ground state,<br />

transferring its energy to O 2, generating singlet oxygen 1 O 2, known to be a highly<br />

cytotoxic species. Therefore fullerenes constitute an excellent photosensitizer to<br />

be used in photodynamic therapy (PDT) [565]. There are two different pathways<br />

of DNA photocleavage acting mainly at guanine sites. The generation of singlet<br />

oxygen (type II photosensitization) as well as the energy transfer from the triplet<br />

excited state of fullerene to bases (type I photosensitization) can be responsible of<br />

the oxidation of guanosines and these modifications increase the instability of the<br />

phosphodiesteric bond that becomes easily susceptible of alkaline hydrolysis.<br />

The first report on the DNA photocleaving activity of fullerenes was made in<br />

1993 by Nakamura and coworkers [566]. The brief light exposure of a mixture of<br />

compound 87 (Figure 5.73) and a plasmid DNA resulted in single-strand nicking,<br />

and longer exposure led to double-strand cleavage, largely at guanine sites. No<br />

cleavage was observed in the dark.<br />

Since then, numerous approaches [557–564] have been developed to obtain<br />

fullerenes derivatives capable of cleaving the DNA under photoirradiation. Among<br />

recent reports, there is the preparation of derivatives 88–91 a–e containing mono-<br />

and disaccharides (Figure 5.74) [552]. The introduction of sugar moieties improves<br />

the biological properties of fullerene because they increase its solubility and they<br />

play an important role in cell–cell interaction. The production of singlet oxygen<br />

was analyzed by measuring near IR emission at 1270 nm. The bisadducts were<br />

Figure 5.73 Example of a fullerene derivative used for DNA photocleavage.


5.7 Applications of Fullerenes<br />

Figure 5.74 Water soluble derivatives containing mono- and disaccharides substituents.<br />

less effective in generating singlet oxygen. The treatment of HeLa cells with these<br />

derivatives was almost not effective in dark condition, while phototoxicity was<br />

more efficient upon incubation with monoadducts.<br />

Fullerene derivative 92 presenting a water-soluble �-cyclodextrin as appendage<br />

was able to act as DNA-cleavage agent after photoirradiation (Figure 5.75) [567].<br />

The authors studied the mechanism of action and demonstrated the necessity of<br />

oxygen in the system. Moreover, EPR studies evidenced the presence of 1 O 2 as<br />

active species in the cleavage process.<br />

Figure 5.75 Fullerene derivative substituted with a �-cyclodextrin subunit.<br />

311


312 5 Fullerenes<br />

Figure 5.76 Fullerene-phorphyrin hybrids.<br />

Fullerene-phorphyrin hybrids possess attractive photoabsorption properties and<br />

application of this material in PDT has been realized by some researchers. Comparison<br />

between porphyrin, porphyrin-fullerene dyad (P-C 60 ) 93 and metalated<br />

porphyrin-fullerene dyad (ZnP-C 60 ) 94 (Figure 5.76) showed that the phototoxic<br />

activity decreased from P-C 60 to ZnP-C 60 and to porphyrin alone. This behavior<br />

was found also in anaerobic conditions, demonstrating that in this case both Type<br />

I and Type II mechanisms were involved [568].<br />

Fullerene is an excellent acceptor in the ground state and can accept, reversibly,<br />

up to six electrons in solution. This property renders C 60 an excellent radical<br />

scavenger and provides a possible therapeutic approach for some neurodegenerative<br />

disorders such as Parkinson’s, Alzheimer’s and Lou Gehrig’s diseases. There<br />

is evidence to suggest that these diseases are due to hyper-production of reactive<br />

oxygen species (ROS).<br />

An important class of fullerene derivatives studied mainly as neuroprotective<br />

agents and radical scavengers comprises carboxyfullerenes such as the tris-malonic<br />

acid derivatives [569] with C 3 or D 3 symmetry 95 and 96 (Figure 5.77). Many researchers<br />

have investigated the antioxidant mechanism of these two regioisomers<br />

Figure 5.77 Regioisomers of tris-malonic acid derivatives with C3 and D3 symmetry.


5.7 Applications of Fullerenes<br />

[570, 571]. For the compound C 3 the neuroprotective effect was not only related to<br />

its ability of scavenging free radicals but also to the nitric oxide synthase inhibition<br />

and the possible suppression of toxic cytokines.<br />

Polyhydroxylated fullerenes named fullerenols [C 60(OH) n] have been shown<br />

to be excellent antioxidants. Djordjevic and coworkers analyzed the mechanism<br />

of action of C 60(OH) 2–26 by ESR in presence of 2,2-diphenyl-1-picryhydrazyl<br />

(DPPH) free radical and OH radicals generated by Fenton reaction [572]. The<br />

addition of fullerenol to the DPPH solution decreased the radical concentration<br />

as demonstrated by reduction of EPR signal of DPPH. The same behavior, with<br />

better results, was found in the case of OH radicals produced by Fenton reaction.<br />

Electron or hydrogen atom donation from fullerenol to free radicals was the<br />

proposed mechanism. Fullerenols C 60(OH) 22, C 60(OH) 7±2 and C 60(OH) 24 have<br />

been also tested as scavenger of ROS and nitric oxide [573–575].<br />

The ability of �-alanine fullerene derivatives 97–99 in scavenging OH radicals<br />

has been studied by chemiluminescence (Figure 5.78) [576]. The radical sponge<br />

action is dose-dependent and the three adducts presented a variable scavenger<br />

activity. Compound 98 was the best, followed by 99, while 97 was the less effective<br />

in the series.<br />

Figure 5.78 �-alanine fullerene derivatives with scavenger activity.<br />

The antibacterial activity of fullerene derivatives was first reported in 1996 for<br />

fulleropyrrolidinium salts 100a–c (Figure 5.79) [577]. The hypothesized mechanism<br />

of action was attributed to the cell membrane disruption by the bulky carbon<br />

cage, which seemed not really adaptable to planar cellular surface. Since this first<br />

report, numerous papers have related the potential antimicrobial effects of C 60 and<br />

derivatives. More recently, Mashino and coworkers [578] studied the antibacterial<br />

activity of C 60 -bis(N,N-dimethylpyrrolidinium salts) regioisomers 101–103 against<br />

Escherichia coli and Gram-positive bacteria such as Enterococcus faecalis. The three<br />

compounds demonstrated excellent and not significantly different antibacterial<br />

activity, which was comparable with that of vancomycin. The mechanism of action<br />

seemed to be inhibition of the respiratory chain.<br />

In 1993 Wuld and coworkers [579] reported for the first time the inhibition of<br />

HIV-protease probably by interaction of fullerene with the hydrophobic active site<br />

of the enzyme. Since then, many studies on different fullerene derivatives have<br />

been performed [580–584]. Recently Marchesan and coworkers obtained good<br />

results against HIV on CEM cells infected by HIV-1(III B ) or HIV-2 (ROD) using<br />

a series of N,N-dimethyl bis fulleropyrrolidinium salts. However, the mechanism<br />

involved in the antiviral activity was not elucidated. Analyses in this sense have<br />

been performed by Mashino and coworkers. They identified efficient inhibitors of<br />

313


314 5 Fullerenes<br />

Figure 5.79 Fulleropyrrolidinium salts with antibacterial activity.<br />

the HIV reverse transcriptase, more active than nevirapine. The same compounds<br />

were found to be active also against hepatitis C virus RNA polymerases [585].<br />

Recent studies have shown that chemically modified fullerenes can still be<br />

considered potentially interesting systems for drug delivery [586] and gene<br />

therapy [587]. Finally, radiolabeled fullerenes [588, 589] can became nano-vehicles<br />

for imaging, diagnosis, therapy and microsurgery because they can be used as<br />

contrast agents or radiotracers.<br />

5.7.4<br />

Conclusions<br />

Recent progress in the chemistry of C 60 allowed the synthesis of a large variety<br />

of fullerene derivatives for various applications in materials science and biology.<br />

In the first part of this chapter we have shown that fullerenes are efficient optical<br />

limiters and interesting acceptors for photovoltaic applications. In the second part,<br />

the biological properties of fullerene derivatives have been summarized. Despite<br />

some remarkable recent achievements, it is clear that the examples discussed<br />

herein represent only the first steps towards the design of commercial drugs or<br />

fullerene-based materials which can display functionality at the macroscopic level.<br />

More research in these areas is clearly needed to fully explore the possibilities<br />

offered by these compounds.


References<br />

1 Kroto, H. Angew. Chem. Int. Ed. 1997, 36,<br />

1579–1593.<br />

2 Herzberg, G., 1945, p. 12.<br />

3 Botchvar, D. E.; Galpern, E. G. Dokl. AN<br />

SSSR 1973, 209, 610.<br />

4 Osawa, E. Kagaku (Kyoto) 1970, 25, 854.<br />

5 Rohlfing, E. A.; Cox, D. M.; Kaldor, A.<br />

J. Chem. Phys., 1984, 81, 3322.<br />

6 Krätschmer, W.; Fostiropoulos, K.;<br />

Huffmann, D. Chem. Phys. Lett. 1990,<br />

170, 167.<br />

7 Yannoni, C. S.; Bernier, P. P.; Meier, G.;<br />

Salem, J. R. J. Am. Chem. Soc. 1991, 113,<br />

3205.<br />

8 Thilgen, C.; Diederich, F. Chem. Rev.<br />

2006, 106, 5049.<br />

9 Becker, L.; R. J. Poreda; Bunch, T. E.<br />

Proc. Natl. Acad. Sci. USA 2000, 97, 2979.<br />

10 Heymann, D.; Jenneskens, L. W.;<br />

Jehlicka, J.; Koper, C.; Vlietstra, E.<br />

Full. Nanotubes Carb. Nanostr. 2003, 11,<br />

333.<br />

11 Dresselhaus, M. S.; Dresselhaus, G.;<br />

Eklund, P. C. Science of Fullerenes and<br />

Carbon Nanotubes; Academic Press: San<br />

Diego, 1996.<br />

12 Fowler, P. W.; Manolopoulos, D. E.<br />

An Atlas of Fullerenes; Clarendon Press:<br />

Oxford, 1995.<br />

13 Prinzbach, H.; Weiler, A.;<br />

Landenberger, P.; Wahl, F.;<br />

Worth, J.; Scott, L. T.; Gelmont, M.;<br />

von Olevano, D.; Issendorff, B.<br />

Nature 2000, 407, 60.<br />

14 Wahl, F.; Weiler, A.; Landenberger, P.;<br />

Sackers, E.; Voss, T.; Haas, A.; Lieb, M.;<br />

Hunkler, D.; Worth, J.; Knothe, L.;<br />

Prinzbach, H. Chem. Eur. J. 2006, 12,<br />

6255.<br />

15 Sackers, E.; Obwald, T.; Weber, K.;<br />

Keller, M.; Hunkler, D.; Worth, J.;<br />

Knothe, L.; Prinzbach, H. Chem. Eur. J.<br />

2006, 12, 6242.<br />

16 Lehn, J.-M. Supramolecular Chemistry,<br />

p. 140, VCH: Weinheim, 1995.<br />

17 Seiders, T. J.; Baldridge, K. K.;<br />

Siegel, J. S. Tetrahedron 2001, 57, 3737.<br />

18 Sygula, A.; Fronczek, F. R.; Rabideau,<br />

P. W. Tetrahedron Lett. 1997, 38, 5095.<br />

19 Amaya, T.; Mori, K.; Wu, H. L.;<br />

Ishida, S.; Nakamura, J. I.; Murata, K.;<br />

Hirao, T. Chem. Commun. 2007, 1902.<br />

References<br />

20 Stoddart, J. F. Angew. Chem. Int. Ed.<br />

1991, 30, 70.<br />

21 Taylor, R.; Avent, A. G.; Dennis, T. J.;<br />

Hare, J. P.; Kroto, H. W.;<br />

Walton, D. R. M.; Holloway, J. H.;<br />

Hope, E. G.; Langley, G. J. Nature 1992,<br />

355, 27.<br />

22 Dodziuk, H.; Nowinski, K. Chem. Phys.<br />

Lett. 1996, 249, 406.<br />

23 Dodziuk, H.; Dolgonos, G.; Lukin, O.<br />

Carbon 2001, 39, 1907.<br />

24 Shustova, N. B.; Popov, A. A.; Newell,<br />

B. S.; Miller, S. M.; Anderson, O. P.;<br />

Seppelt, K.; Bolskar, R. D.;<br />

Boltalina, O. V.; Strauss, S. H. Angew.<br />

Chem. Int. Ed. 2007, 46, 4111–4114.<br />

25 Arndt, M.; Nairz, O.; Voss-Andreae, J.;<br />

Keller, C.; van der Zouw, G.; Zeilinger, A.<br />

Nature 1999, 401, 608–682.<br />

26 Nairz, O.; Brezger, B.; Arndt, M.;<br />

Zeilinger, A. Phys. Rev. Lett. 2001, 87,<br />

160401.<br />

27 Bandaru, P. R.; Daraio, C.; Jin, S.;<br />

Rao, A. M. Nature Mat. 2005, 4, 663–666.<br />

28 a) Kroto, H. W.; Heath, J. R.;<br />

O’Brien, S. C.; Curl, R. F.; Smalley, R. E.<br />

Nature, 1985, 318, 162. b) Kroto, H. W.<br />

Angew. Chem., Int. Ed. Engl., 1992, 31,<br />

111.<br />

29 Krätschmer, W.; Lamb, L. D.;<br />

Fostiropoulos, K.; Huffman, D. R.<br />

Nature, 1990, 347, 354.<br />

30 a) Buckminsterfullerenes, Ed. Billups,<br />

W. E.; Ciufolini, M. A., VCH, New York,<br />

(1993). b) Chemistry of the Fullerenes,<br />

Hirsch, A., Thieme, Stuttgart (1994).<br />

c) Science of Fullerenes and Carbon<br />

Nanotubes, Ed. Dresselhaus, M. S.;<br />

Dresselhaus, G.; Eklund, P. C.,<br />

Academic, San Diego (1996).<br />

d) Fullerenes and Related Structures,<br />

Ed. Hirsch, A. Top. Curr. Chem., 199,<br />

Springer, Berlin (1999). e) Fullerenes:<br />

Chemistry, Physics, and Technology, Ed.<br />

Kadish, K. M.; Ruoff, R. S., Wiley, New<br />

York (2000).<br />

31 a) Irie, K.; Nakamura, Y.; Ohigashi, H.;<br />

Tokuyama, H.; Yamago, S.;<br />

Nakamura, E. Biosci. Biotech. & Biochem.,<br />

1996, 60, 1359; b) Zhao, X.; Striolo, A.;<br />

Cummings, P. T. Biophys. J., 2005, 89,<br />

3856.<br />

315


316 5 Fullerenes<br />

32 Diener, M. D.; Alford, M. J.; Nabity, J.,<br />

Pat. WO 2003050040, 2003 (see also<br />

C.&EN, 2003, 81, Aug. 11, 13).<br />

33 Powell, W. H.; Cozzi, F.; Moss, G. P.;<br />

Thilgen, C.; Hwu, R. J.-R.; Yerin, A.<br />

Pure Appl. Chem., 2002, 74, 629.<br />

34 Nakamura, Y.; Okawa, K.; Nishimura, T.;<br />

Yashima, E.; Nishimura, J. J. Org. Chem.,<br />

2003, 68, 3251.<br />

35 A fullerene having a cyclobutane ring<br />

was made: see Qian, W.; Chuang, S.-C.;<br />

Amador, R. B.; Jarrosson, T.; Sander, M.;<br />

Pieniazek, S.; Khan, S. I.; Rubin, Y.<br />

J. Am. Chem. Soc., 2003, 125, 2066<br />

36 Hirsch, A. Chem. Record, 2005, 5, 196.<br />

37 a) Fagan, P. J.; Krusic, P. J.;<br />

Evans, D. H.; Lerke, S. A.; Johnston, E.<br />

J. Am. Chem. Soc., 1992, 114, 9697;<br />

b) Hirsch, A.; Soi, A.; Karfunkel, H. R.<br />

Angew. Chem. Int. Ed. Engl., 1992,<br />

31, 766; c) Hirsch, A.; Gröser, T.;<br />

Skiebe, A.; Soi, A. Chem. Ber., 1993,<br />

126, 1061; d) Vail, S. A.; Krawczuk, P. J.;<br />

Guldi, D. M.; Palkar, A.; Echegoyen, L.;<br />

Tome, J. P. C.; Fazio, M. A.;<br />

Schuster, D. I. Chem. Eur. J., 2005, 11,<br />

3375 and references cited therein.<br />

38 a) Nagashima, H.; Terasaki, H.;<br />

Kimura, E.; Nakajima, K.; Itoh, K.<br />

J. Org. Chem., 1994, 59, 1246;<br />

b) Nagashima, H.; Terasaki, H.; Saito, Y.;<br />

Jinno, K.; Itoh, K. J. Org. Chem., 1995,<br />

60, 4966; c) Nagashima, H.; Saito, M.;<br />

Kato, Y.; Goto, H.; Osawa, E.; Haga, H.;<br />

Itoh, K. Tetrahedron, 1996, 52, 5053;<br />

d) Nakamura, E.; Sawamura, M. Pure<br />

and Applied Chem., 2001, 73, 355;<br />

e) Nakamura, E.; Isobe, H. Acc. Chem.<br />

Res., 2003, 36, 807; f) Nakamura, E.<br />

Pure Appl. Chem., 2003, 75, 427.<br />

39 Bingel, C. Chem. Ber., 1993, 126, 1957.<br />

40 Keshavarz-K., M.; Knight, B.;<br />

Srdanov, G.; Wudl, F. J. Am. Chem. Soc.,<br />

1995, 117, 11371.<br />

41 Subramanian, R.; Kadish, K. M.;<br />

Vijayashree, M. N.; Gao, X.; Jones, M. T.;<br />

Miller, M. D.; Krause, K. L.; Suenobu, T.;<br />

Fukuzumi, S. J. Phys. Chem., 1996, 100,<br />

16327.<br />

42 McEwen, C. N.; McKay, R. G.; Larsen,<br />

B. S. J. Am. Chem. Soc., 1992, 114, 4412.<br />

43 a) Chiang, L. Y.; Swirczewski, J. W.;<br />

Hsu, C. S.; Chowdhury, S. K.;<br />

Cameron, S.; Creegan, K. J. Chem.<br />

Soc., Chem. Commun., 1992, 1791;<br />

b) Chiang, L. Y.; Upasani, R. B.;<br />

Swirczewski, J. W.; Soled, S.<br />

J. Am. Chem. Soc., 1993, 115, 5453;<br />

c) Anantharaj, V.; Bhonsle, J.;<br />

Canteevala, T.; Chiang, L. Y. J. Chem.<br />

Soc., Perkin Trans. 1, 1999, 31.<br />

44 Olah, G. A.; Bucsi, I.; Lambert, C.;<br />

Aniszfeld, R.; Triverdi, N. J.;<br />

Sensharma, D. K.; Prakash, G. K. S.<br />

J. Am. Chem. Soc., 1991, 113, 9387.<br />

45 a) Hirsch, A.; Lamparth, I.;<br />

Karfunkel, H. R. Angew. Chem., Int.<br />

Ed. Engl., 1994, 33, 437; b) Hirsch, A.<br />

Top. Curr. Chem., 1999, 199, 1;<br />

c) Diederich, F.; Kessinger, R. Acc. Chem.<br />

Res., 1999, 32, 537.<br />

46 a) Ros, T. D.; Prato, M.; Lucchini, V.<br />

J. Org. Chem., 2000, 65, 4289;<br />

b) Nishimura, T.; Tsuchiya, K.;<br />

Ohsawa, S.; Maeda, K.; Yashima, E.;<br />

Nakamura, Y.; Nishimura, J. J. Am.<br />

Chem. Soc., 2004, 126, 11711.<br />

47 Nakamura, Y.; Okawa, K.;<br />

Matsumoto, M.; Nishimura, J.<br />

Tetrahedron, 2000, 56, 5429.<br />

48 Nakamura, Y.; Takano, N.; Nishimura, T.;<br />

Yashima, E.; Sato, M.; Kudo, T.;<br />

Nishimura, J. Org. Lett., 2001, 3, 1193.<br />

49 Djojo, F.; Herzog, A.; Lamparth, I.;<br />

Hampel, F.; Hirsch, A. Chem. Eur. J.,<br />

1996, 2, 1537.<br />

50 a) Fowler, P. W. J. Chem. Soc., Faraday<br />

Trans., 1995, 91, 2241; b) Kordatos, K.;<br />

Bosi, S.; Ros, T. D.; Zambon, A.;<br />

Lucchini, V.; Prato, M. J. Org. Chem.,<br />

2001, 66, 2802.<br />

51 Hirsch, A.; Vostrowsky, O. Eur. J. Org.<br />

Chem., 2001, 829.<br />

52 Nakamura, E.; Isobe, H.; Tokuyama, H.;<br />

Sawamura, M. J. Chem. Soc., Chem.<br />

Commun., 1996, 1747.<br />

53 Zhou, Z.; Schuster, D. I.; Wilson, S. R.<br />

J. Org. Chem., 2006, 71, 1545.<br />

54 Taki, M.; Sugita, S.; Nakamura, Y.;<br />

Kasashima, E.; Yashima, E.; Okamoto, Y.;<br />

Nishimura, J. J. Am. Chem. Soc., 1997,<br />

119, 926.<br />

55 Lamparth, I.; Maichle-Moessmer, C.;<br />

Hirsch, A. Angew. Chem., Int. Ed. Engl.,<br />

1995, 34, 1607.<br />

56 Li, H.; Haque, S. A.; Kitaygorodskiy, A.;<br />

Meziani, M. J.; Torres-Castillo, M.;<br />

Sun, Y.-P. Org. Lett., 2006, 8, 5641.


57 a) Skiebe, A.; Hirsch, A.; Klos, H.;<br />

Gotschy, B. Chem. Phys. Lett., 1994,<br />

220, 138; b) Hirsch, A.; Li, Q.; Wudl, F.<br />

Angew. Chem., Int. Ed. Engl., 1991, 30,<br />

1309.<br />

58 Mikami, K.; Matsumoto, S.;<br />

Ishida, A.; Takamuku, S.; Suenobu, T.;<br />

Fukuzumi, S. J. Am. Chem. Soc., 1995,<br />

117, 11134 and references cited therein.<br />

59 Kessinger, R.; Crassous, J.;<br />

Herrmann, A.; Ruuttimann, M.;<br />

Echegoyen, L.; Diederich, F. Angew.<br />

Chem., Int. Ed., 1998, 37, 1919.<br />

60 Martín, N.; Altable, M.; Filippone, S.;<br />

Martín-Domenech, A.; Echegoyen, L.;<br />

Cardona, C. M. Angew. Chem. Int. Ed.,<br />

2006, 45, 110.<br />

61 Meidine, M. F.; Avent, A. G.;<br />

Darwish, A. D.; Kroto, H. W. ;<br />

Ohashi, O.; Taylor, R.; Walton, D. R. M.;<br />

J. Chem. Soc., Perkin Trans. 2, 1994, 1189.<br />

62 Nossal, J.; Saini, R. K.; Alemany, L. B.;<br />

Meier, M.; Billups, W. E. Eur. J. Org.<br />

Chem., 2001, 4167 and references cited<br />

therein.<br />

63 a) Saunders, M. Science, 1991, 253, 330;<br />

b) Dodziuk, H.; Nowi�ski, K. Horror<br />

vacui or topological in-out isomerism<br />

in perhydrogenated fullerenes.<br />

Part 1. C 60 H 60 and monoalkylated<br />

perhydrogenated fullerenes, Chem. Phys.<br />

Lett., 1996, 249, 406.<br />

64 a) Kalsbeck, W. A.; Thorp, H. H.<br />

J. Electroanal. Chem., 1991, 314, 363;<br />

b) Creegan, K. M.; Robbins, J. L.;<br />

Robbins, W. K.; Millar, J. M.;<br />

Sherwood, R. D.; Tindall, P. J.;<br />

Cox, D. M.; Smith, III, A. B.;<br />

McCauley, J. P., Jr.; Jones, D. R.;<br />

Gallagher, R. T. J. Am. Soc. Chem.,<br />

1992, 114, 1103; c) Werner, H.;<br />

Schedel-Niedrig, Th.; Wohlers, M.;<br />

Herein, D.; Herzog, B.; Chlogl, R.;<br />

Keil, M.; Bradshaw, A. M.; Kirshner, J.<br />

J. Chem. Soc., Faraday Trans., 1994,<br />

90, 403; d) Camp, A. G.; Ford, W. T.;<br />

Lary, A.; Sensharma, D. K.; Chang, Y. H.;<br />

Hercules, D. M.; Williams, J. B.<br />

Fullerene Sci. Tech., 1997, 5, 527;.<br />

e) Malhotra, R.; Kumar, S.; Satyam, A.<br />

J. Chem. Soc., Chem. Commun., 1994,<br />

1339; f) Elemes, Y.; Silverman, S. K.;<br />

Sheu, C.; Kao, M.; Foote, C. S.;<br />

Alvarez, M. M.; Whetten, R. L. Angew.<br />

References<br />

Chem., Int. Ed. Engl., 1992, 31, 351;<br />

g) Balch, A. L.; Costa, D. A.; Noll, B. C.;<br />

Olmstead, M. M. J. Am. Chem. Soc.,<br />

1995, 117, 8926; h) Hamano, P.;<br />

Mashino, T.; Hirobe, M. J. Chem. Soc.,<br />

Chem. Commun., 1995, 1537.<br />

65 Shigemitsu, Y.; Kaneko, M.; Tajima, Y.;<br />

Takeuchi, K. Chem. Lett., 2004, 33, 1604.<br />

66 Roy, S.; Sarkar, S. J. Chem. Soc., Chem.<br />

Commun., 1994, 275.<br />

67 Langer, J. J.; Gibinski, T. Adv. Mater. Opt.<br />

Electr., 2000, 9, 273.<br />

68 Chiang, L. Y.; Upasani, R. B.;<br />

Swirczewski, J. W. J. Am. Chem. Soc.,<br />

1992, 114, 10154.<br />

69 Troshin, P. A.; Astakhova, A. S.;<br />

Lyubovskaya, R. N. Fullerenes, Nanotubes,<br />

and Carbon Nanostructures, 2005, 13, 331.<br />

70 a) Komatsu, K.; Murata, Y.; Takimoto, N.;<br />

Mori, S.; Sugita, N.; Wan, T. S. M.<br />

J. Org. Chem., 1994, 59, 6101;<br />

b) Timmerman, P.; Witschel, L. E.;<br />

Diederich, F.; Boudon, C.;<br />

Giesselbrecht, J.-P.; Gross, M.<br />

Helv. Chim. Acta, 1996, 79, 6 and<br />

references cited therein; c) Murata, Y.;<br />

Motoyama, K.; Komatsu, K.;<br />

Wan, T. S. M. Tetrahedron, 1996, 52,<br />

5077; d) Murata, Y.; Komatsu, K.;<br />

Wan, T. S. M. Tetrahedron Lett., 1996,<br />

37, 7061; e) Fujiwara, K.; Murata, Y.;<br />

Wan, T. S. M.; Komatsu, K. Tetrahedron,<br />

1998, 54, 2049; f) Cardin, D. J.;<br />

Gibson, T.; Pell, J. A. Fulleren Sci, Tech.,<br />

1997, 5, 681.<br />

71 Sawamura, M.; Iikura, H.; Nakamura, E.<br />

J. Am. Chem. Soc., 1996, 118, 12850.<br />

72 Nakamura, E.; Tahara, K.; Matsuo, Y.;<br />

Sawamura, M. J. Am. Chem. Soc., 2003,<br />

125, 2834<br />

73 Wudl, F. J. Mater. Chem., 2002, 12, 1959.<br />

74 Thilgen, C.; Sergeyev, S.; Diederich, F.<br />

Top. Curr. Chem., 2005, 248, 1.<br />

75 Nakamura, Y.; Minami, S.; Iizuka, K.;<br />

Nishimura, J. Angew. Chem., Int. Ed.,<br />

2003, 42, 3158.<br />

76 Qian, W.; Rubin, Y. Angew. Chem., Int.<br />

Ed. Engl., 1999, 38, 2356.<br />

77 Kessinger, R.; Gomez-Lopez, M.;<br />

Boudon, C.; Gisselbrecht, J.-P.;<br />

Gross, M.; Echegoyen, L.; Diederich, F.<br />

J. Am. Chem. Soc., 1998, 120, 8545.<br />

78 Nierengarten, J.-F.; Nicoud, J.-F.<br />

Tetrahedron Lett., 1997, 38, 7737.<br />

317


318 5 Fullerenes<br />

79 Benito, A. M.; Darwish, A. D.;<br />

Kroto, H. W.; Meidine, M. F.; Taylor, R.;<br />

Walton, D. R. M. Tetrahedron Lett., 1996,<br />

37, 1085.<br />

80 Shu, L.-H.; Wang, G.-W.; Wu, S.-H.;<br />

Wu, H.-M. J. Chem. Soc., Chem.<br />

Commun., 1995, 367.<br />

81 Bestmann, H. J.; Hadawi, D.; Röder, T.;<br />

Moll, C. Tetrahedron Lett., 1994, 35, 9017.<br />

82 a) Wang, Y.; Cao, J.; Schuster, D. I.;<br />

Wilson, S. R. Tetrahedron Lett., 1995, 36,<br />

6843; b) Li, J.; Yoshizawa, T.; Ikuta, M.;<br />

Ozawa, M.; Nakahara, K.; Hasegawa, T.;<br />

Kitazawa, K.; Hayashi, M.; Kinbara, K.;<br />

Nohara, M.; Saigo, K. Chem. Lett., 1997,<br />

1037.<br />

83 a) Ito, H.; Tada, T.; Sudo, M.; Ishida, Y.;<br />

Hino, T.; Saigo, K. Org. Lett., 2003, 5,<br />

2643; b) Tada, T.; Ishida, Y.; Saigo, K.<br />

J. Org. Chem., 2006, 71, 1633.<br />

84 Nakamura, Y.; Inamura, K.;<br />

Oomuro, R.; Laurenco, R.; Tidwell, T. T.;<br />

Nishimura, J. Org. Biomol. Chem., 2005,<br />

3, 3032 and references cited therein.<br />

85 Suzuki, T.; Li, Q.; Khemani, K. C.;<br />

Wudl, F. J. Am. Chem. Soc., 1992, 114,<br />

7301.<br />

86 Wakahara, T.; Niino, Y.; Kato, T.;<br />

Maeda, Y.; Akasaka, T.; Liu, M. T. H.;<br />

Kobayashi, K.; Nagase, S. J. Am. Chem.<br />

Soc., 2002, 124, 9465.<br />

87 Ishida, T.; Tanaka, K.; Nogami, T. Chem.<br />

Lett., 1994, 561.<br />

88 Prato, M.; Li, Q. C.; Wudl, F. J. Am.<br />

Chem. Soc., 1993, 115, 1148.<br />

89 Reuther, U.; Hirsch, A. Carbon, 2000, 38,<br />

1539 and references cited therein.<br />

90 Akasaka, T.; Ando, A.; Kobayashi, K.;<br />

Nagase, S. J. Am. Chem. Soc., 1993, 115,<br />

1605.<br />

91 Wang, G. W.; Komatsu, K.; Murata, Y.;<br />

Shiro, M. Nature, 1997, 387, 583.<br />

92 McClenaghan, N. D.; Absalon, C.;<br />

Bassani, D. M. J. Am. Chem. Soc., 2003,<br />

125, 13004.<br />

93 a) Vassilikogiannakis, G.; Chronakis, N.<br />

Orfanopoulos, M. J. Am. Chem. Soc.,<br />

1998, 120, 9911; b) Vassilikogiannakis,<br />

G.; Hatzimarinaki, M.; Orfanopoulos, M.<br />

J. Org. Chem., 2000, 65, 8180;<br />

c) Schuster, D. I.; Cao, J.; Kaprinidis, N.;<br />

Wu, Y.; Jensen, A. W.; Lu, Q.; Wang, H.;<br />

Wilson, S. R. J. Am. Chem. Soc., 1996,<br />

118, 5639 and references cited therein.<br />

94 a) Zhang, X.; Foote, C. S. J. Am. Chem.<br />

Soc., 1995, 117, 4271; b) Yamago, S.;<br />

Takeichi, A.; Nakamura, E. J. Am. Chem.<br />

Soc., 1994, 116, 1123; c) Matsui, S.;<br />

Kinbara, K.; Saigo, K. Tetrahedron<br />

Lett., 1999, 40, 899; d) Nair, V.;<br />

Sethumadhavan, D.; Nair, S. M.;<br />

Shanmugam, P.; Treesa, P. M.;<br />

Eigendorf, G. K. Synthesis, 2002, 1655.<br />

95 a) Hoke, II, S. H.; Molstad, J.;<br />

Dilettato, D.; Jay, M. J.; Carlson, D.;<br />

Kahr, B.; Cooks, R. G. J. Org. Chem.,<br />

1992, 57, 5069; b) Tsuda, M.; Ishida, T.;<br />

Nogami, T.; Kurono, S.; Ohashi, M.<br />

Chem. Lett., 1992, 2333; c) Meier, M. S.;<br />

Wang, G.-W.; Haddon, R. C.;<br />

Brock, C. P.; Lloyd, M. A.; Selegue, J. P.<br />

J. Am. Chem. Soc., 1998, 120, 2337.<br />

96 Maggini, M.; Scorrano, G.; Prato, M.<br />

J. Am. Chem. Soc., 1993, 115, 9798.<br />

97 Georgakilas, V.; Kordatos, K.; Prato, M.;<br />

Guldi, D. M.; Holzinger, M.; Hirsch, A.<br />

J. Am. Chem. Soc., 2002, 124, 760.<br />

98 Yamashiro, T.; Aso, Y.; Otsubo, T.;<br />

Tang, H.; Harima, Y.; Yamashita, K.<br />

Chem. Lett., 1999, 28, 443.<br />

99 Imahori, H.; Sekiguchi, Y.;<br />

Kashiwagi, Y.; Sato, T.; Araki, Y.; Ito, O.;<br />

Yamada, H.; Fukuzumi, S. Chem. Eur. J.,<br />

2004, 10, 3184.<br />

100 Hutchison, K.; Gao, J.; Schick, G.;<br />

Rubin, Y.; Wudl, F. J. Am. Chem. Soc.,<br />

1999, 121, 5611.<br />

101 a) Averdung, J.; Mattay, J. Tetrahedron,<br />

1996, 52, 5407; b) Matsubara, Y.;<br />

Tada, H.; Nagase, S.; Yoshida, Z.<br />

J. Org. Chem., 1995, 60, 5372;<br />

c) Irngartinger, H.; Weber, A.; Escher, T.<br />

Liebigs Ann., 1996, 1845.<br />

102 Ohno, M.; Yashiro, A.; Eguchi, S. Synlett,<br />

1996, 815.<br />

103 Jagerovic, N.; Elguero, J.; Aubagnac, J.-L.<br />

J. Chem. Soc., Perkin Trans. 1, 1996, 499.<br />

104 Ohno, M.; Yashiro, A.; Eguchi, S. J. Chem.<br />

Soc., Chem. Commun., 1996, 291.<br />

105 a) Lawson, G. E.; Kitaygorodskiy, A.;<br />

Ma, B.; Bunker, C. E.; Sun, Y.-P. J. Chem.<br />

Soc., Chem. Commun., 1995, 2225.<br />

b) Nakamura, Y.; Suzuki, M.; O-kawa, K.;<br />

Konno, T.; Nishimura, J. J. Org. Chem.,<br />

2005, 70, 8472.<br />

106 Akasaka, T.; Ando, W.; Kobayashi, K.;<br />

Nagase, S. J. Am. Chem. Soc., 1993, 115,<br />

10366.


107 Wang, G.-W.; Chen, X.-P.; Cheng, X.<br />

Chem. Eur. J., 2006, 12, 7246.<br />

108 a) Rubin, Y.; Khan, S.; Freedberg, D. I.;<br />

Yeretzian, C. J. Am. Chem. Soc., 1993,<br />

115, 344; b) Belik, P.; Guegel, A.;<br />

Spickermann, J.; Müllen, K. Angew.<br />

Chem., Int. Ed. Engl., 1993, 32, 78;<br />

c) Tago, T.; Minowa, T.; Okada, Y.;<br />

Nishimura, J. Tetrahedron Lett., 1993,<br />

34, 8461; d) Guegel, A.; Kraus, A.;<br />

Spickermann, J.; Belik, P.; Müllen, K.<br />

Angew. Chem., Int. Ed. Engl.,<br />

1994, 33, 559; e) Ganapathi, P. S.;<br />

Friedman, S. H.; Kenyon, G. L.;<br />

Rubin, Y. J. Org. Chem., 1995, 60, 2954.<br />

109 Takeshita, H.; Liu, J. F.; Kato, N.;<br />

Mori, A.; Isobe, R. Chem. Lett., 1995, 377.<br />

110 Wang, G.-W.; Saunders, M.; Khong, A.;<br />

Cross, R. J. J. Am. Chem. Soc., 2000, 122,<br />

3216.<br />

111 a) Zhang, X.; Foote, C. S. J. Org. Chem.,<br />

1994, 59, 5235; b) Nakamura, Y.;<br />

O-kawa, K.; Minami, S.; Ogawa, T.;<br />

Tobita, S.; Nishimura, J. J. Org. Chem.,<br />

2002, 67, 1247.<br />

112 a) Kampe, K. D.; Egger, N.; Vogel, M.<br />

Angew. Chem., Int. Ed. Engl., 1993,<br />

32, 1174; b) Kampe, K. D.; Egger, N.<br />

Liebigs Ann., 1995, 115; c) Balch, A. L.;<br />

Ginwalla, A. S.; Olmstead, M. M.;<br />

Herbst-Irmer, R. Tetrahedron, 1996, 52,<br />

5021.<br />

113 Llacay, J.; Veciana, J.; Gancedo, J. V.;<br />

Bourdelande, J. L.; Moreno, R. G.;<br />

Rovira, C. J. Org. Chem., 1998, 63, 5201.<br />

114 Nakamura, Y.; Taki, M.; Tobita, S.;<br />

Shizuka, H.; Yokoi, H.; Ishiguro, K.;<br />

Sawaki, Y.; Nishimura, J. J. Chem. Soc.,<br />

Perkin Trans. 2, 1999, 127.<br />

115 Fowler, P. W.; Manolopoulous, D. E.<br />

An Atlas of Fullerenes; Clarendon: Oxford,<br />

2006 (revised).<br />

116 Kroto, H. W. Nature 1987, 329, 529–531.<br />

117 Powell, W. H.; Cozzi, F.; Moss, G. P.;<br />

Thilgen, C.; Hwu, R. J. R.; Yerin, A.<br />

Pure Appl. Chem. 2002, 74, 629–695.<br />

118 Hirsch, A.; Brettreich, M. Fullerenes:<br />

Chemistry and Reactions; Wiley-VCH:<br />

Weinheim, 2005.<br />

119 Shustova, N. B.; Kuvychko, I. V.;<br />

Bolskar, R. D.; Seppelt, K.;<br />

Strauss, S. H.; Popov, A. A.;<br />

Boltalina, O. V. J. Am. Chem. Soc. 2006,<br />

128, 15793–15798.<br />

References<br />

120 Hitchcock, P. B.; Avent, A. G.;<br />

Martsinovich, N.; Troshin, P. A.;<br />

Taylor, R. Chem. Commun. 2005, 75–77.<br />

121 Goryunkov, A. A.; Markov, V. Y.;<br />

Ioffe, I. N.; Sidorov, L. N.; Bolskar, R. D.;<br />

Diener, M. D.; Kuvychko, I. V.;<br />

Strauss, S. H.; Boltalina, O. V. Angew.<br />

Chem. Int. Ed. 2004, 43, 997–1000.<br />

122 Lukoyanova, O.; Cardona, C. M.;<br />

Rivera, J.; Lugo-Morales, L. Z.;<br />

Chancellor, C. J.; Olmstead, M. M.;<br />

Balch, A. L.; Echegoyen, L. J. Am. Chem.<br />

Soc. 2007, 129, 10423–10430.<br />

123 Neretin, I. S.; Slovokhotov, Y. L. Russ.<br />

Chem. Rev. 2004, 73, 455–486.<br />

124 Balch, A. L. In Fullerenes: Chemistry,<br />

Physics, and Technology, K. M. Kadish and<br />

R. S. Ruoff, (Eds.) John Wiley & Sons<br />

Inc. New York, 2000, pp. 177–223.<br />

125 Haddon, R. C. Science 1993, 261,<br />

1545–1550.<br />

126 Popov, A. A.; Kareev, I. E.;<br />

Shustova, N. B.; Stukalin, E. B.;<br />

Lebedkin, S. F.; Seppelt, K.;<br />

Strauss, S. H.; Boltalina, O. V.;<br />

Dunsch, L. J. Am. Chem. Soc. 2007, 129,<br />

in press (doi 10.1021/ja073181e).<br />

127 Popov, A. A.; Kareev, I. E.;<br />

Shustova, N. B.; Lebedkin, S. F.;<br />

Strauss, S. H.; Boltalina, O. V.; Dunsch, L.<br />

Chem. Eur. J. 2007, 13, in press.<br />

128 Saunders, M.; Jarrosson, T.;<br />

Chuang, S.-C.; Khan, S. I.; Rubin, Y.<br />

J. Org. Chem. 2006, 72, 2724–2731.<br />

129 Wang, F.; Xiao, Z.; Gan, L.; Jia, Z.;<br />

Jiang, Z.; Zhang, S.; Zheng, B.; Gu, Y.<br />

Org. Lett. 2007, 9, 1741–1743.<br />

130 Yannoni, C. S.; Bernier, P. P.;<br />

Bethune, D. S.; Meijer, G.; Salem, J. R.<br />

J. Am. Chem. Soc 1991, 113, 3190–3192.<br />

131 Burgi, H.-B.; Blanc, E.; Schwarzenbach,<br />

D.; Liu, S.; Lu, Y.-J.; Kappes, M. M.;<br />

Ibers, J. A. Angew. Chem. Int. Ed. Engl.<br />

1992, 31, 640–643.<br />

132 David, W. I. F.; Ibberson, R. M.;<br />

Dennis, T. J.; Hare, J. P.; Prassides, K.<br />

Europhys. Lett. 1992, 18, 219–225.<br />

133 Schiebel, P.; Wulf, K.; Prandl, W.;<br />

Heger, G.; Papoular, R.; Paulus, W.<br />

Acta Cryst. 1996, A52, 176–188.<br />

134 Zuo, T.; Beavers, C. M.; Duchamp, J. C.;<br />

Campbell, A.; Dorn, H. C.;<br />

Olmstead, M. M.; Balch, A. L. J. Am.<br />

Chem. Soc. 2007, 129, 2035–2043.<br />

319


320 5 Fullerenes<br />

135 Olmstead, M. M.; de Bettencourt-Dias,<br />

A.; Lee, H. M.; Pham, D.; Balch, A. L.<br />

Dalton Trans. 2003, 3227–3232.<br />

136 Cai, T.; Xu, L.; Gibson, H. W.; Dorn, H. C.;<br />

Chancellor, C. J.; Olmstead, M. M.;<br />

Balch, A. L. J. Am. Chem. Soc. 2007, 129<br />

(35), 10795–10800.<br />

137 Beavers, C. M.; Zuo, T.; Duchamp, J. C.;<br />

Harich, K.; Dorn, H. C.; Olmstead,<br />

M. M.; Balch, A. L. J. Am. Chem. Soc.<br />

2006, 128, 11352–11353.<br />

138 Popov, A. A.; Dunsch, L. J. Am. Chem.<br />

Soc 2007, 129 (38), 11835–11849.<br />

139 Raghavachari, K. Chem. Phys. Lett. 1992,<br />

190, 397–400.<br />

140 Kareev, I. E.; Kuvychko, I. V.; Popov,<br />

A. A.; Lebedkin, S. F.; Miller, S. M.;<br />

Anderson, O. P.; Strauss, S. H.;<br />

Boltalina, O. V. Angew. Chem. Int. Ed.<br />

2005, 44, 7984–7987.<br />

141 Shustova, N. B.; Popov, A. A.; Newell,<br />

B. S.; Miller, S. M.; Anderson, O. P.;<br />

Seppelt, K.; Bolskar, R. D.;<br />

Boltalina, O. V.; Strauss, S. H. Angew.<br />

Chem. Int. Ed. 2007, 46, 4111–4114.<br />

142 Furche, F.; Ahlrichs, R. J. Chem. Phys.<br />

2001, 114, 10362–10367.<br />

143 Kareev, I. E.; Lebedkin, S. F.;<br />

Miller, S. M.; Anderson, O. P.;<br />

Strauss, S. H.; Boltalina, O. V. Acta<br />

Crystallogr. 2006, E62, o1498-o1500.<br />

144 Kareev, I. E.; Kuvychko, I. V.;<br />

Lebedkin, S. F.; Miller, S. M.;<br />

Anderson, O. P.; Seppelt, K.;<br />

Strauss, S. H.; Boltalina, O. V.<br />

J. Am. Chem. Soc. 2005, 127,<br />

8362–8375.<br />

145 Kareev, I. E.; Lebedkin, S. F.; Popov,<br />

A. A.; Miller, S. M.; Anderson, O. P.;<br />

Strauss, S. H.; Boltalina, O. V. Acta<br />

Crystallogr. 2006, E62, o1501-o1503.<br />

146 Shustova, N. B.; Peryshkov, D. V.;<br />

Popov, A. A.; Boltalina, O. V.;<br />

Strauss, S. H. Acta Cryst. 2007, E63,<br />

o3129.<br />

147 Troyanov, S. I.; Dimitrov, A.;<br />

Kemnitz, E. Angew. Chem. Int. Ed. 2006,<br />

45, 1971–1974.<br />

148 Kareev, I. E.; Shustova, N. B.;<br />

Peryshkov, D. V.; Lebedkin, S. F.;<br />

Miller, S. M.; Anderson, O. P.;<br />

Popov, A. A.; Boltalina, O. V.;<br />

Strauss, S. H. Chem. Commun. 2007,<br />

1650–1652.<br />

149 Troyanov, S. I.; Boltalina, O. V.;<br />

Kuvychko, I. V.; Troshin, P. A.;<br />

Kemnitz, E.; Hitchcock, P. B.; Taylor, R.<br />

Fullerenes Nanotubes Carbon Nanostruct.<br />

2002, 10, 243–260.<br />

150 Troyanov, S., I.; Goryunkov, A. A.;<br />

Dorozhkin, E. I.; Ignat’eva, D. V.;<br />

Tamm, N. B.; Avdoshenko, S. M.;<br />

Ioffe, I. N.; Markov, V. Y.; Sidorov, L. N.;<br />

Scheural, K.; Kemnitz, E. J. Fluor. Chem.<br />

2007, 128, 545–551.<br />

151 Hitchcock, P. B.; Taylor, R. Chem.<br />

Commun. 2002, 2078–2079.<br />

152 Sun, C. H.; Lu, G. Q.; Cheng, H. M.<br />

J. Phys. Chem. B 2006, 110, 218–221.<br />

153 Wakahara, T.; Nikawa, H.; Kikuchi, K.;<br />

Nakahodo, T.; Aminur Rahman, G. M.;<br />

Tsuchiya, T.; Maeda, Y.; Akasaka, T.;<br />

Yoza, K.; Horn, E.; Yamamoto, K.;<br />

Mizorogi, N.; Slanina, Z.; Nagase, S.<br />

J. Am. Chem. Soc. 2006, 128,<br />

14228–14229.<br />

154 Olmstead, M. M.; Lee, H. M.;<br />

Duchamp, J. C.; Stevenson, S.;<br />

Marciu, D.; Dorn, H. C.; Balch, A. L.<br />

Angew. Chem. Int. Ed. 2003, 42, 900–903.<br />

155 Matsuzawa, N.; Dixon, D. A.;<br />

Fukunaga, T. J. Phys. Chem. 1992, 96,<br />

7594–7604.<br />

156 Matsuzawa, N.; Fukunaga, T.;<br />

Dixon, D. A. J. Phys. Chem. 1992, 96,<br />

10747–10756.<br />

157 Roth, G.; Adelmann, P. J. Phys. I Fr.<br />

1992, 2, 1541–1548.<br />

158 Burgi, H.-B.; Venugopalan, P.;<br />

Schwarzenbach, D.; Diederich, F.;<br />

Thilgen, C. Helv. Chim. Acta 1993, 76,<br />

2155–2159.<br />

159 Michel, R. H.; Kappes, M. M.;<br />

Adelmann, P.; Roth, G. Angew. Chem.<br />

Int. Ed. 1994, 33, 1651–1654.<br />

160 Olmstead, M. M.; Nurco, D. J. Cryst.<br />

Growth Design 2006, 6, 109–113.<br />

161 Reich, A.; Panthofer, M.; Modrow, H.;<br />

Wedig, U.; Jansen, M. J. Am. Chem. Soc.<br />

2004, 126, 14428–14434.<br />

162 Olmstead, M. M.; de Bettencourt-<br />

Dias, A.; Duchamp, J. C.; Stevenson, S.;<br />

Marciu, D.; Dorn, H. C.; Balch, A. L.<br />

Angew. Chem. Int. Ed. 2001, 40,<br />

1223–1225.<br />

163 Pham, D.; Ceron-Bertran, J.;<br />

Olmstead, M. M.; Mascal, M.; Balch, A. L.<br />

Cryst. Growth Design 2007, 7, 75–82.


164 Fowler, P. W.; Sandall, J. P. B.;<br />

Austin, S. J.; Manolopoulous, D. E.;<br />

Lawrenson, P. D. M.; Smallwood, J. M.<br />

Synth. Metals 1996, 77.<br />

165 Fowler, P. W.; Rogers, K. M.;<br />

Somers, K. R.; Troisi, A. J. Chem. Soc.,<br />

Perkin Trans. 2 1999, 2023–2027.<br />

166 Balaban, A. T.; Klein, D. J.; Ivanciuc, O.<br />

Fullerenes Nanotubes and Carbon Nanostr.<br />

2005, 13, 109–129.<br />

167 Bakowies, D.; Thiel, W. J. Am. Chem. Soc<br />

1991, 113, 3704–3714.<br />

168 Fowler, P. W. In From Chemical<br />

Topology to Three-Dimensional Geometry,<br />

Balaban, A. T., Plenum Press: New York,<br />

1997, pp. 237–262.<br />

169 Kratschmer, W.; Fostiropoulos, K.;<br />

Huffman, D. R. Chem. Phys. Lett. 1990,<br />

170, (2–3), 167–170.<br />

170 Kratschmer, W.; Lamb, L. D.;<br />

Fostiropoulos, K.; Huffman, D. R.<br />

Nature 1990, 347, (6291), 354–358.<br />

171 Menendez, J.; Page, J. B., Vibrational<br />

spectroscopy of C-60. in Light Scattering<br />

in Solids VIII, Springer-Verlag Berlin:<br />

Berlin, 2000; Vol. 76, pp. 27–95.<br />

172 Schettino, V.; Pagliai, M.; Ciabini, L.;<br />

Cardini, G. J. Phys. Chem. A 2001, 105,<br />

(50), 11192–11196.<br />

173 Martin, M. C.; Du, X. Q.; Kwon, J.;<br />

Mihaly, L. Phys. Rev. B 1994, 50, (1),<br />

173–183.<br />

174 Vanloosdrecht, P. H. M.; Vanbentum,<br />

P. J. M.; Verheijen, M. A.; Meijer, G.<br />

Chem. Phys. Lett. 1992, 198, (6), 587–595.<br />

175 Sassara, A.; Zerza, G.; Chergui, M.;<br />

Negri, F.; Orlandi, G. J. Chem. Phys.<br />

1997, 107, (21), 8731–8741.<br />

176 Orlandi, G.; Negri, F. Photochem.<br />

Photobiol. Sci. 2002, 1, (5), 289–308.<br />

177 Coulombeau, C.; Jobic, H.; Carlile, C. J.;<br />

Bennington, S. M.; Fabre, C.; Rassat, A.<br />

Fullerene Sci. Technol. 1994, 2, (3),<br />

247–254.<br />

178 Heid, R.; Pintschovius, L.; Godard, J. M.<br />

Phys. Rev. B 1997, 56, (10), 5925–5936.<br />

179 Nissen, M. K.; Wilson, S. M.;<br />

Thewalt, M. L. W. Phys. Rev. Lett. 1992,<br />

69, (16), 2423–2426.<br />

180 Gensterblum, G. J. Electron Spectrosc.<br />

Relat. Phenom. 1996, 81, (2), 89–223.<br />

181 Schmidt, S.; Hunt, M. R. C.; Miao, P.;<br />

Palmer, R. E. Phys. Rev. B 1997, 56, (15),<br />

9918–9924.<br />

References<br />

182 Popov, A. A. Ph. D. Thesis. Moscow<br />

State University, 2003.<br />

183 Christides, C.; Nikolaev, A. V.;<br />

Dennis, T. J. S.; Prassides, K.; Negri, F.;<br />

Orlandi, G.; Zerbetto, F. J. Phys. Chem.<br />

1993, 97, (15), 3641–3643.<br />

184 Horoyski, P. J.; Thewalt, M. L. W.<br />

Chem. Phys. Lett. 1994, 217, (4), 409–412.<br />

185 Michel, R. H.; Schreiber, H.; Gierden, R.;<br />

Hennrich, F.; Rockenberger, J.;<br />

Beck, R. D.; Kappes, M. M.; Lehner, C.;<br />

Adelmann, P.; Armbruster, J. F. Ber.<br />

Bunsen-Ges. Phys. Chem. Chem. Phys.<br />

1994, 98, (7), 975–978.<br />

186 Benz, M.; Fanti, M.; Fowler, P. W.;<br />

Fuchs, D.; Kappes, M. M.; Lehner, C.;<br />

Michel, R. H.; Orlandi, G.; Zerbetto, F.<br />

J. Phys. Chem. 1996, 100, (32),<br />

13399–13407.<br />

187 Dennis, T. J. S.; Hulman, M.;<br />

Kuzmany, H.; Shinohara, H. J. Phys.<br />

Chem. B 2000, 104, (23), 5411–5413.<br />

188 Krause, M.; Hulman, M.; Kuzmany, H.;<br />

Dennis, T. J. S.; Inakuma, M.;<br />

Shinohara, H. J. Chem. Phys. 1999, 111,<br />

(17), 7976–7984.<br />

189 Eisler, H. J.; Gilb, S.; Hennrich, F. H.;<br />

Kappes, M. M. J. Phys. Chem. A 2000,<br />

104, (8), 1762–1768.<br />

190 Eisler, H. J.; Hennrich, F. H.; Gilb, S.;<br />

Kappes, M. M. J. Phys. Chem. A 2000,<br />

104, (8), 1769–1773.<br />

191 Saito, S.; Oshiyama, A. Phys. Rev. Lett.<br />

1991, 66, (20), 2637–2640.<br />

192 Golden, M. S.; Knupfer, M.; Fink, J.;<br />

Armbruster, J. F.; Cummins, T. R.;<br />

Romberg, H. A.; Roth, M.; Sing, M.;<br />

Schmidt, M.; Sohmen, E. J. Phys.-Condes.<br />

Matter 1995, 7, (43), 8219–8247.<br />

193 Rudolf, P.; Golden, M. S.;<br />

Bruhwiler, P. A. J. Electron Spectrosc.<br />

Relat. Phenom. 1999, 100, 409–433.<br />

194 Lichtenberger, D. L.; Jatcko, M. E.;<br />

Nebesney, K. W.; Ray, C. D.;<br />

Huffman, D. R.; Lamb, L. D.<br />

In Clusters and Cluster-Assembled<br />

Interfaces, MRS Symposia Proceedings,<br />

1991; Averback, R. S.; Bernholc, J.;<br />

Nelson, D. L., Eds. Material Research<br />

Society, Pittsburgh, 1991, pp. 673–678.<br />

195 Jost, M. B.; Troullier, N.; Poirier, D. M.;<br />

Martins, J. L.; Weaver, J. H.;<br />

Chibante, L. P. F.; Smalley, R. E.<br />

Phys. Rev. B 1991, 44, (4), 1966–1969.<br />

321


322 5 Fullerenes<br />

196 Terminello, L. J.; Shuh, D. K.;<br />

Himpsel, F. J.; Lapianosmith, D. A.;<br />

Stohr, J.; Bethune, D. S.; Meijer, G.<br />

Chem. Phys. Lett. 1991, 182, (5), 491–496.<br />

197 Krummacher, S.; Biermann, M.;<br />

Neeb, M.; Liebsch, A.; Eberhardt, W.<br />

Phys. Rev. B 1993, 48, (11), 8424–8429.<br />

198 Wang, X. B.; Woo, H. K.; Wang, L. S.<br />

J. Chem. Phys. 2005, 123, (5), 4.<br />

199 Shustova, N. B.; Kuvychko, I. V.;<br />

Bolskar, R. D.; Seppelt, K.; Strauss, S. H.;<br />

Popov, A. A.; Boltalina, O. V. J. Am.<br />

Chem. Soc. 2006, 128, (49), 15793–15798.<br />

200 Yang, S. H.; Pettiette, C. L.;<br />

Conceicao, J.; Cheshnovsky, O.;<br />

Smalley, R. E. Chem. Phys. Lett. 1987,<br />

139, (3–4), 233–238.<br />

201 Kietzmann, H.; Rochow, R.;<br />

Gantefor, G.; Eberhardt, W.; Vietze, K.;<br />

Seifert, G.; Fowler, P. W. Phys. Rev. Lett.<br />

1998, 81, (24), 5378–5381.<br />

202 Prinzbach, H.; Weller, A.;<br />

Landenberger, P.; Wahl, F.; Worth, J.;<br />

Scott, L. T.; Gelmont, M.; Olevano, D.;<br />

von Issendorff, B. Nature 2000, 407,<br />

(6800), 60–63.<br />

203 Wang, X. B.; Woo, H. K.; Huang, X.;<br />

Kappes, M. M.; Wang, L. S. Phys. Rev.<br />

Lett. 2006, 96, (14), 4.<br />

204 Leach, S.; Vervloet, M.; Despres, A.;<br />

Breheret, E.; Hare, J. P.; Dennis, T. J.;<br />

Kroto, H. W.; Taylor, R.; Walton, D. R. M.<br />

Chem. Phys. 1992, 160, (3), 451–466.<br />

205 Bauernschmitt, R.; Ahlrichs, R.;<br />

Hennrich, F. H.; Kappes, M. M. J. Am.<br />

Chem. Soc. 1998, 120, (20), 5052–5059.<br />

206 Haufler, R. E.; Chai, Y.;<br />

Chibante, L. P. F.; Fraelich, M. R.;<br />

Weisman, R. B.; Curl, R. F.;<br />

Smalley, R. E. J. Chem. Phys. 1991, 95,<br />

(3), 2197–2199.<br />

207 Close, J. D.; Federmann, F.;<br />

Hoffmann, K.; Quaas, N. Chem. Phys.<br />

Lett. 1997, 276, (5–6), 393–398.<br />

208 Vandenheuvel, D. J.; Vandenberg,<br />

G. J. B.; Groenen, E. J. J.; Schmidt, J.;<br />

Holleman, I.; Meijer, G. J. Phys. Chem.<br />

1995, 99, (30), 11644–11649.<br />

209 Negri, F.; Orlandi, G.; Zerbetto, F. Chem.<br />

Phys. Lett. 1988, 144, (1), 31–37.<br />

210 Haufler, R. E.; Wang, L. S.; Chibante,<br />

L. P. F.; Jin, C. M.; Conceicao, J.;<br />

Chai, Y.; Smalley, R. E. Chem. Phys. Lett.<br />

1991, 179, (5–6), 449–454.<br />

211 Diederich, F.; Whetten, R. L. Accounts<br />

Chem. Res. 1992, 25, (3), 119–126.<br />

212 Hennrich, F. H.; Michel, R. H.;<br />

Fischer, A.; Richard-Schneider, S.;<br />

Gilb, S.; Kappes, M. M.; Fuchs, D.;<br />

Burk, M.; Kobayashi, K.; Nagase, S.<br />

Angew. Chem.-Int. Edit. Engl. 1996, 35,<br />

(15), 1732–1734.<br />

213 Wang, C. R.; Sugai, T.; Kai, T.;<br />

Tomiyama, T.; Shinohara, H.<br />

Chem. Commun. 2000, (7), 557–558.<br />

214 Dennis, T. J. S.; Kai, T.; Asato, K.;<br />

Tomiyama, T.; Shinohara, H. J. Phys.<br />

Chem. A 1999, 103, (44), 8747–8752.<br />

215 Levitt, H. Spin Dynamics: Basics of<br />

Nuclear Magnetic Resonance. John Wiley<br />

& Sons, Chichester, 2001.<br />

216 Slichter, C. P. Principles of Magnetic<br />

Resonance. Springer Verlag, New York,<br />

3rd edn., 1990.<br />

217 Knight, W. D. Electron Paramagnetism<br />

and Nuclear Magnetic Resonance in<br />

Metals, Vol. 2 of Solid State Physics-<br />

Advances in Research and Applications,<br />

pp. 93–136. Academic Press Inc., San<br />

Diego, CA, 1956.<br />

218 Ashbrook, S. E.; Duer, M. J. Concepts<br />

Magn. Reson. Part A, 2006, 28A(3),<br />

183–248.<br />

219 Brewer, J. H. Muon Spin Rotation/<br />

Relaxation/Resonance, Vol. 11 of<br />

Encyclopedia of Applied Physics,<br />

pp. 23–53. Wiley -VCH, Weinheim, 1994.<br />

220 Blundell, S. J. Contemp. Phys., 1999,<br />

40(3), 175–192.<br />

221 Johnson, R. D.; Bethune, D. S.;<br />

Yannoni, C. S. Acc. Chem. Res., 1992, 25,<br />

169–175.<br />

222 Mizoguchi, K. J. Phys. Chem. Sol., 1993,<br />

54 (12), 1693–1698.<br />

223 Tycko, R.. Solid State Nucl. Magn. Reson.,<br />

1994, 3, 303–314.<br />

224 Pennington, C. H.; Stenger, V. A. Rev.<br />

Mod. Phys., 1996, 68, 855–910.<br />

225 Taylor, R.; Hare, J. P.; Abdul Sada, A. K.;<br />

Kroto, H. W. J. Chem. Soc. Chem.<br />

Commun., 1990, 20, 1423–1425.<br />

226 Johnson, R. D.; Meijer, G.; Salem, J. R.;<br />

Bethune, D. S. J. Am. Chem. Soc., 1991,<br />

113, 3619–3621.<br />

227 Zimmer, G.; Helmle, M.; Mehring, M.;<br />

Rachdi, F.; Reichenbach, J.; Firlej, L.;<br />

Bernier, P. Europhys. Lett., 1993, 24(1),<br />

59–64.


228 Tycko, R.; Dabbagh, G.;<br />

Rosseinsky, M. J.; Murphy, D. W.;<br />

Ramirez, A. P.; Fleming, R. M.<br />

Phys. Rev. Lett., 1992, 68 (12), 1912–1915.<br />

229 Mehring, M.; Rachdi, F.; Zimmer, G.<br />

Phil. Mag. B, 1994, 70(3), 787–794.<br />

230 Maniwa, Y.; Sugiura, D.; Kume, K.;<br />

Kikuchi, K.; Suzuki, S.; Achiba, Y.;<br />

Hirosawa, I.; Tanigaki, K.; Shimoda, H.;<br />

Iwasa, Y. Phys. Rev. B, 1996, 54(10),<br />

R6861–R6864.<br />

231 Goze Bac, C.; Latil, S.; Lauginie, P.;<br />

Jourdain, V.; Conard, J.; Duclaux, L.;<br />

Rubio, A.; Bernier, P. Carbon, 2002,<br />

40(10), 1825–1842.<br />

232 Haddon, R. C. J. Am. Chem. Soc., 1990,<br />

112(9), 3385–3389.<br />

233 Haddon, R. C. Acc. Chem. Res., 1992,<br />

25(3), 127–133.<br />

234 Hawkins, J. M.; Loren, S.; Meyer, A.;<br />

Nunlist, R. J. Am. Chem. Soc., 1991,<br />

113(20), 7770–7771.<br />

235 Yannoni, C. S.; Bernier, P. P.; Bethune,<br />

D. S.; Meijer, G.; Salem, J. R. J. Am.<br />

Chem. Soc., 1991, 113(8), 3190–3192.<br />

236 Saunders, M.; Jiménez-Vázquez, H. A.;<br />

Bangerter, B. W.; Cross, R. J.;<br />

Mroczkowski, S.; Freedberg, D. I.;<br />

Anet, F. A. L. J. Am. Chem. Soc., 1994,<br />

116(8), 3621–3622.<br />

237 Yannoni, C. S.; Johnson, R. D.;<br />

Meijer, G.; Bethune, D. S.; Salem, J. R.<br />

J. Phys. Chem., 1991, 95(1), 9–10.<br />

238 Tycko, R.; Haddon, R. C.; Dabbagh, G.;<br />

Glarum, S. H.; Douglass, D. C.;<br />

Mujsce, A. M. J. Phys. Chem., 1991,<br />

95(2), 518–520.<br />

239 Coffey, T.; Krim, J. Phys. Rev. Lett.,<br />

96(18), 186104, 2006.<br />

240 Heiney, P. A.; Fischer, J. E.;<br />

McGhie, A. R.; Romanow, W. J.;<br />

Denenstein, A. M.; McCauley, J. P., Jr.;<br />

Smith, A. B.; Cox, D. E. Phys. Rev. Lett.,<br />

1991, 66(22), 2911–2914.<br />

241 Tycko, T.; Dabbagh, G.; Fleming, R. M.;<br />

Haddon, R. C.; Makhija, A. V.;<br />

Zahurak, S. M. Phys. Rev. Lett., 1991,<br />

67(14), 1886–1889.<br />

242 Johnson, R. D.; Yannoni, C. S.;<br />

Dorn, H. C.; Salem, J. R.; Bethune, D. S.<br />

Science, 1992, 255, 1235–1238.<br />

243 Alloul, H.; Brouet, V.; Lafontaine, E.;<br />

Malier, L.; Forro, L. Phys. Rev. Lett., 1996,<br />

76(16), 2922–2925.<br />

References<br />

244 Thier, K. F.; Mehring, M.; Rachdi, F.<br />

Phys. Rev. B, 1997, 55(1), 124–126.<br />

245 Riccò, M.; Shiroka, T.; Belli, M.;<br />

Pontiroli, D.; Pagliari, M.; Ruani, G.;<br />

Palles, D.; Margadonna, S.;<br />

Tomaselli, M. Phys. Rev. B, 2005, 72(15),<br />

155437.<br />

246 Herzfeld, J.; Berger, A. E. J. Chem. Phys.,<br />

1980, 73(12), 6021–6030.<br />

247 Schaefer, J.; Stejskal, E. O. J. Am. Chem.<br />

Soc., 1976, 98(4), 1031–1032.<br />

248 Maniwa, Y.; Saito, T.; Ohi, A.;<br />

Mizoguchi, K.; Kume, K.; Kikuchi, K.;<br />

Ikemoto, I.; Suzuki, S.; Achiba, Y.;<br />

Kosaka, M.; Tanigaki, K.; Ebbesen, T. W.<br />

J. Phys. Soc. Jap., 1994, 63(3), 1139–1148.<br />

249 Forró, L.; Mihály, L. Rep. Prog. Phys.,<br />

64(5), 649–699, 2001.<br />

250 Korringa, J. Physica, 1950, 16(7–8),<br />

601–610.<br />

251 Antropov, V. P.; Mazin, I. I.;<br />

Andersen, O. K.; Liechtenstein, A. I.;<br />

Jepsen, O. Phys. Rev. B, 1993, 47(18),<br />

12373–12376.<br />

252 Pennington, C. H.; Stenger, V. A.;<br />

Recchia, C. H.; Hahm, C.; Gorny, K.;<br />

Nandor, V.; Buffinger, D. R.; Lee, S. M.;<br />

Ziebarth, R. P. Phys. Rev. B, 1996, 53(6),<br />

R2967–R2970.<br />

253 Hebel, L. C.; Slichter, C. P. Phys. Rev.,<br />

1959, 113(6), 1504–1519.<br />

254 Kiefl, R. K.; MacFarlane, W. A.;<br />

Chow, K. H.; Dunsiger, S.; Duty, T. L.;<br />

Johnston, T. M. S.; Schneider, J. W.;<br />

Sonier, J.; Brard, L.; Strongin, R. M.;<br />

Fischer, J. E.; Smith, A. B. Phys. Rev.<br />

Lett., 1993, 70(25), 3987–3990.<br />

255 MacFarlane, W. A.; Kiefl, R. F.;<br />

Dunsiger, S.; Sonier, J. E.; Chakhalian, J.;<br />

Fischer, J. E.; Yildirim, T.; Chow, K. H.<br />

Phys.Rev. B, 1998, 58(2), 1004–1024.<br />

256 Stenger, V. A.; Pennington, C. H.;<br />

Buffinger, D. R.; Ziebarth, R. P. Phys.<br />

Rev. Lett., 1995, 74(9), 1649–1652.<br />

257 Chlistunoff, J.; Cliffel, D.; Bard, A. J. in<br />

Handbook of Conductive Molecules and<br />

Polymers, Vol. 1 Ch. 7, Charge Transfer<br />

Salts, Fullerenes and Photoconductors,<br />

Nalwa, H. S. (Ed.) Wiley, Chichester<br />

1997, 333.<br />

258 Echegoyen, L.; Echegoyen, L. E. Acc.<br />

Chem. Res. 1998, 31, 593<br />

259 Martin, N.; Sanchez, L.; Illescas, B.;<br />

Perez, I. Chem. Rev., 1998, 98, 2527.<br />

323


324 5 Fullerenes<br />

260 Reed, C. A.; Bolskar, R. D. Chem. Rev.,<br />

2000, 100, 1075.<br />

261 Echegoyen, L.; Herranz, M. A. in<br />

Fullerene Electrochemistry in Fullerenes:<br />

from Synthesis to Optoelectronic Properties<br />

Guldi, D. M.; Martin, N. (Eds.) Kluwer,<br />

Dordrecht 2002, 276.<br />

262 Sherigara, B. S.; Kutner, W.; D’Souza, F.;<br />

Electroanalysis 2003, 15, 753.<br />

263 Echegoyen, L.; Diederich, F.;<br />

Echegoyen, L. E. in Fullerenes: Chemistry,<br />

Physics, and Technology, Kadish, K. M.;<br />

Ruoff, R. S. (Eds.) John Wiley & Sons<br />

Inc. New York, 2000, p. 1.<br />

264 Allemand, P.-M.; Koch, A.; Wudl, F.;<br />

Rubin, Y.; Diederich, F.; Alvarez, M. M.;<br />

Anz, S. J.; Whetten, R. L. J. Am.Chem.<br />

Soc., 1991, 113, 1050.<br />

265 Arias, F.; Xie, Q.; Lu, Q.; Wilson, S. R.;<br />

Echegoyen, L. J. Am. Chem. Soc., 1994,<br />

116, 6388.<br />

266 Xie, Q.; Perez-Cordero, E.; Echegoyen, L.<br />

J. Am. Chem. Soc. 1992, 114, 3978.<br />

267 Cox, D. M.; Bethal, S.; Disco, M.;<br />

Gorun, S. M. D.; Greaney, M.;<br />

Hsu, C. S.; Kollin, E. B.; Millar, J.;<br />

Robins, J.; Robins, W.; Sherwood, R. D.;<br />

Tindall, P. J. Am. Chem. Soc., 1991, 113,<br />

2940.<br />

268 Dubois, D.; Moninot, G.; Kutner, W.;<br />

Jones, M. T.; Kadish, K. M. J. Phys.<br />

Chem. 1992, 96, 7137.<br />

269 Krishman, V.; Moninot, G.; Dubois, D.;<br />

Kutner, W.; Kadish, K. M. J. Electroanal.<br />

Chem., 1993, 356, 93.<br />

270 Stinchcombe, J.; Penicaud, A.;<br />

Bhyrappa, P.; Boyd, P. D. W.; Reed, C. A.<br />

J. Am. Chem. Soc., 1993, 115, 5217.<br />

271 Fawcett, W. R.; Opallo, M.; Fedurco, M.;<br />

Lee, J. W. J. Am. Chem. Soc. 1993, 115,<br />

196.<br />

272 Soucaze-Guillous, B.; Kutner, W.;<br />

Jones, M. T.; Kadish, K. M.<br />

J. Electrochem. Soc. 1996, 143, 550.<br />

273 Noviandri, I.; Bolskar, R. D.; Lay, P. A.;<br />

Reed, C. A. J. Phys. Chem. B 1997, 101,<br />

6350.<br />

274 Reed, C. A.; Bolskar, R. D. Chem. Rev.<br />

2000, 100, 1075.<br />

275 Bolskar, R. D.; Gallaghar, S. H.;<br />

Armstrong, R. S.; Lay, P. A.; Leed, C. A.<br />

Chem. Phys. Lett., 1995, 247, 57.<br />

276 Koga, N.; Morokuma, K. Chem. Phys.<br />

Lett., 1992, 196.<br />

277 Compton, R. G.; Spackman, R. A.;<br />

Riley, D. J.; Wellington, R.; Eklund, J. C.;<br />

Fisher, A. C.; Green, M. L. H.;<br />

Douthwaite, R. E.; Stephens, A. H. H.;<br />

Turner, J. J. Electroanal. Chem., 1993,<br />

344, 235.<br />

278 Chlistunoff, J.; Cliffel, D.; Bard, A.<br />

Thin Solid Films 1995, 257, 166.<br />

279 Mirkin, C. A.; Caldwell, W. B.<br />

Tetrahedron 1996, 52, 5113.<br />

280 Compton, R. G.; Spackman, R. A.;<br />

Wellington, R. G.; Green, M. L. H.;<br />

Turner, J. J. Electroanal.Chem., 1992, 327,<br />

337.<br />

281 Soucaze-Guillous, B.; Kutner, W.;<br />

Jones, M. T.; Kadish, K. M.<br />

J. Electrochem. Soc., 1996, 143, 550.<br />

282 Bruno, C.; Doubitski, I.; Marcaccio, M.;<br />

Paolucci, F.; Paolucci, D.; Zaopo, A.<br />

J. Am. Chem. Soc., 2003, 125, 15738.<br />

283 Lem, G.; Schuster, D. I.; Courtney, S. H.;<br />

Lu, Q.; Wilson, S. R. J. Am. Chem. Soc.,<br />

1995, 117, 554.<br />

284 Yang, Y. F.; Arias, F.; Echegoyen, L.;<br />

Chibante, L. P. F.; Flanagan, S.;<br />

Robertson, A.; Wilson, L. J. Am. Chem.<br />

Soc., 1995, 117, 7801.<br />

285 Webster, R. D.; Heath, G. A. Phys. Chem.<br />

Chem. Phys., 2001, 3, 2588.<br />

286 Boudon, C.; Gisselbrecht, J.-P.;<br />

Gross, M.; Herrmann, A.; Rutimann, M.;<br />

Crassous, J.; Cardullo, F.; Echegoyen, L.;<br />

Diederich, F. J. Am. Chem. Soc., 1998,<br />

120, 7860.<br />

287 Pénicaud, A.; Azamar-Barrios, J. A.<br />

Fullerene Sci. Technol., 1998, 6, 743.<br />

288 Illescas, B. M.; Martin, N. C. R. Chimie,<br />

2006, 9, 1038.<br />

289 D’Souza, F.; Chitta, R.; Gadde, S.;<br />

Zandler, M. E.; McCaty, A. L.;<br />

Sandanayaka, A. S. D.; Araki, Y.; Ito, O.<br />

J. Phys. Chem. A, 2006, 110, 4338.<br />

290 Margadonna, S.; Prassides, K. J. Solid<br />

State Commun., 2002, 168, 639.<br />

291 Diekers, M.; Hirsch, A.; Pyo, S.; Rivers, J.;<br />

Echegoyen, L. Eur.Org.Chem., 1998, 1111.<br />

292 Suzuki, T.; Maruyama, Y.; Akasaka, T.;<br />

Ando, W.; Kobayashi, K.; Nagase, S.<br />

J. Am. Chem. Soc., 1994, 116, 1359.<br />

293 Echegoyen, L.; Echegoyen, L. E.<br />

Acc. Chem. Res., 1998, 31, 593.<br />

294 Keshavarz-K., M.; Knight, B.;<br />

Srdanov, G.; Wudl, F. J. Am. Chem. Soc.,<br />

1995, 117, 11371.


295 Keshawarz-K., M.; Knight, B.;<br />

Haddon, R. C.; Wudl, F. Tetrahedron,<br />

1996, 52, 5149.<br />

296 Taylor, R. Chem. Eur. J., 2001, 7, 4075.<br />

297 Ohkubo, K.; Taylor, R.; Boltalina, O. V.;<br />

Ogo, S.; Fukuzumi, S. Chem. Commun.<br />

2002, 1952.<br />

298 Guldi, D. M. J. Phys. Chem. B, 2000, 104,<br />

1483.<br />

299 Trzcinska, K.; Bilkova, P.; Szydlowska, J.;<br />

Mieczkowski, J.; Bilewicz, R.;<br />

Pociecha, D.; Gorecka, E. Pol. J. Chem.,<br />

2006, 80, 1899.<br />

300 Knight, B.; Martin, N.; Ohno, T.; Orti, E.;<br />

Rovira, C.; Veciana, J.; Vidal-Gancedo, J.;<br />

Viruela, P.; Wudl, F. J. Am. Chem. Soc.,<br />

1997, 119, 9871.<br />

301 Oçafrain, M.; Herranz, M. Á.; Marx, L.;<br />

Thilgen, C.; Diederich, F.; Echegoyen, L.<br />

Chem. Eur. J. 2003, 9, 4811.<br />

302 Bourgeois, J.-P.; Echegoyen, L.;<br />

Fibbioli, M.; Pretsch, E.; Diederich, F.<br />

Angew. Chem. Int. Ed. 1998, 37, 2118.<br />

303 Smith, P. M.; McCarty, A. L.;<br />

Nguyen, N. Y.; Zandler, M.; D’Souza, F.<br />

Chem. Commun. 2003, 8783.<br />

304 Popov, A. A.; Kareev, I. E.;<br />

Shustova, N. B.; Stukalin, E. B.;<br />

Lebedkin, S. F.; Seppelt, K.; Strauss,<br />

S. H.; Boltalina, O. V.; Dunsch, L.<br />

J. Am. Chem. Soc. 2007; 129 11551.<br />

305 Echegoyen, L. E.; Djojo, F. D.;<br />

Hirrsch, A.; Echegoyen, L. J. Org. Chem.<br />

2000, 65, 4994.<br />

306 Avent, A. G.; Birkett, P. R.; Carano, M.;<br />

Darwish, A. D.; Kroto, H. W.;<br />

Lopez, J. O.; Paolucci, D.; Roffia, S.;<br />

Taylor, R.; Wachter, N.; Walton, D. R. M.;<br />

Zerbetto, F. J. Chem. Soc., Perkin Trans. 2<br />

2001, 140.<br />

307 Suresh, C.; Koga, N. J. Am. Chem. Soc.<br />

2002, 124, 1790.<br />

308 Campanelli, A. R.; Domenicano, A.;<br />

Ramondo, F.; Hargittai, I. J. Phys.<br />

Chem. A 2004, 108, 4940.<br />

309 Popov, A. A.; Kareev, I. E.;<br />

Shustova, N. B.; Lebedkin, S. F.;<br />

Seppelt, K.; Strauss, S. H.;<br />

Boltalina, O. V.; Dunsch, L. Chem. Eur. J.<br />

2007. doi:10.1002/chem.200700970.<br />

310 Bendikov, M.; Wudl, F.; Perepichka, D. F.<br />

Chem. Rev. 2004, 104, 4891.<br />

311 Guldi, D. M.; Maggini, M.; Martin, N.;<br />

Prato, M. Carbon 2000, 38, 1615.<br />

References<br />

312 Mataga, N.; Miyasakain, H. Electron<br />

Transfer, Jortner, J.; Bixon, M. (Eds.),<br />

John Wiley & Sons, New York, 1999,<br />

p. 431.<br />

313 Guldi, D. M.; Kamat, P. V. in Fullerenes:<br />

Chemistry, Physics, and Technology,<br />

Kadish, K. M.; Ruoff, R. S. (Eds.),<br />

John Wiley & Sons, New York, 2000,<br />

p. 225.<br />

314 Llacay, J.; Veciana, J.; Vidal-Gancedo, J.;<br />

Bourdelande, J. L.; González-Moreno, R.;<br />

Rovira, C. J. Org. Chem, 1998, 63, 5201.<br />

315 Segura, M.; Sanchez, L.; de Mendoza, J.;<br />

Martin, N.; Guldi, D. M. J. Am. Chem.<br />

Soc., 2003, 125, 15093.<br />

316 Kreher, D.; Cariou, M.; Liu, S.-G.;<br />

Levillain, E.; Veciana, J.; Rovira, C.;<br />

Gorgues, A.; Hudhomme, P. J. Mater.<br />

Chem., 2002, 12, 2137.<br />

317 Guldi, D. M.; Rahman, G. M. A.;<br />

Marczak, R.; Matsuo, Y.; Yamananka, M.;<br />

Nakamura, E. J. Am. Chem. Soc. 2006,<br />

128, 9420.<br />

318 Narutaki, M.; Takimiya, K.; Otsubo, T.;<br />

Harima, Y.; Zhang, H.; Araki, Y.; Ito, O.<br />

J. Org. Chem., 2006, 71, 1761.<br />

319 Zhou, Z.; Sarova, G. H.; Zhang, S.;<br />

Ou, Z.; Tat, F. T.; Kadish, K. M.;<br />

Echegoyen, L.; Guldi,. D. M.;<br />

Schuster, D. I.; Wilson, S. R. Chem. Eur.<br />

J. 2006, 12, 4241.<br />

320 Li, W.-S.; Kim, K. S.; Jiang, D.-L.;<br />

Tanaka, H.; Kawai, T.; Kwon, J. H.;<br />

Kim, D.; Aida, T. J. Am. Chem. Soc.,<br />

2006, 128, 10527.<br />

321 Huang, Y.; Wayner, D. M. J. Am. Chem.<br />

Soc., 1993, 115, 367.<br />

322 D’Souza, F.; Choi, J. P.; Hsieh, Y.-Y.;<br />

Shriver, K.; Kutner, W. J. Phys. Chem. B,<br />

1998, 102, 212.<br />

323 D’Souza, F.; Choi, J. P.; Kutner, W.<br />

J. Phys. Chem. B, 1998, 102, 4247.<br />

324 Allard, E.; Riviere, L.; Delaunay, J.;<br />

Dubois, D.; Cousseau, J. Tetrahedron<br />

Lett., 1999, 40, 7223.<br />

325 Bard, A. J.; Faulkner, L. R. Electrochemical<br />

Methods Fundamentals and<br />

Applications, John Wiley, New York,<br />

1980. Chap. 11.<br />

326 Saveant, J.-M. Elements of Molecular and<br />

Biomolecular Electrochemistry, Wiley &<br />

Sons Inc, Hoboken, NJ, 2006, Ch. 3.<br />

327 D’Souza, F.; Choi, J. P.; Kutner, W.<br />

J. Phys. Chem. B 1999, 103, 2892.<br />

325


326 5 Fullerenes<br />

328 Niyazymbetov, M. E.; Evans, D. H.<br />

J. Electrochem. Soc., J. Electrochem. Soc.,<br />

1995, 142, 2655.<br />

329 Willner, I.; Katz, E. Bioelectronics,<br />

Wiley-VCH Weinheim 2005.<br />

330 Patolsky, F.; Tao, G.; Katz, E.; Willner, I.<br />

J. Electroanal. Chem., 1998, 454, 9.<br />

331 D’Souza, F.; Rogers, L. M.;<br />

O’Dell, E. S.; Kochman, A.; Kutner, W.<br />

Bioelectrochemistry, 2005, 66, 35.<br />

332 Brabec, C. J.; Sariciftci, N. S.;<br />

Hummelen, J. C. Adv. Funct. Mater.<br />

2001, 11, 15–26.<br />

333 Wang, P.; Metzger, R. M.; Bandow, S.;<br />

Maruyama, Y. J. Phys. Chem. 1993, 97,<br />

2926–2927.<br />

334 Margadonna, S.; Prassides, K. J. Solid<br />

State Chem. 2002, 168, 639–652.<br />

335 Jehoulet, C.; Obeng, Y. S.; Kim. Y.-T.;<br />

Zhou, F.; Bard, A. J. J. Am Chem. Soc.<br />

1992, 114, 4237–4247.<br />

336 Luo, C.; Huang, C.; Gan, L.; Zhou, L.;<br />

Xia, W.; Zhuang, Q.; Zhao, Y.; Huang, Y.<br />

J. Phys. Chem. 1996, 100, 16685–16689.<br />

337 Zhang, W.; Shi, Y.; Gan, L.; Wu, N.;<br />

Huang, C.; Wu, D. Langmuir 1999, 15,<br />

6921–6924.<br />

338 Imahori, H.; Fukuzumi, S. Adv. Funct.<br />

Mater. 2004, 14, 525–536.<br />

339 Guldi, D. M.; Prato, M. Acc. Chem. Res.<br />

2000, 33, 695–703.<br />

340 Guldi, D. M.; Martin, N. J. Mater. Chem.<br />

2002, 12, 1978–1992.<br />

341 Kvarnström, C.; Kulovaara, H.; Damlin,<br />

P.; Vuorinen, T.; Lemme tyinen, H.;<br />

Ivaska, A. Synth. Met. 2005, 149, 39–45.<br />

342 Peterson, I. R. J. Phys. D: Appl. Phys.<br />

1990, 23, 379–395.<br />

343 Langmuir–Blodgett Films; Roberts, G.,<br />

Ed.; Plenum Press: New York, 1990.<br />

344 Hebard, A. F.; Haddon, R. C.;<br />

Fleming, R. M.; Kortan, A. R. Appl.<br />

Phys. Lett. 1991, 59, 2109–2111.<br />

345 Kataura, H.; Endo, Y.; Achiba, Y.;<br />

Kikuchi, K.; Hanyu, T.; Yamaguchi, S.<br />

J. Phys. Chem. Solids 1997, 58, 1913–1917.<br />

346 Al-Mohammad, A.; Allaf, A. W.<br />

Synth. Met. 1999, 104, 39–44.<br />

347 Zhang, H.; Ding, Y.; Zhong, S.; Wu, C.;<br />

Ning, X.; He, Y.; Liang, Y.; Hong, J.<br />

Thin Solid Films 2005, 492, 41–44.<br />

348 Kamat, P. V.; Barazzouk, S.; Thomas, G.;<br />

Hotchandani, S. J. Phys. Chem. B 2000,<br />

104, 4014–4017.<br />

349 Barazzouk, S.; Hotchandani, S.; Kamat,<br />

P. V. Adv. Mater. 2001, 13, 1614–1617.<br />

350 Sun, Y.-P.; Ma, B.; Bunker, C. E.;<br />

Liu, B. J. Am. Chem. Soc. 1995, 117,<br />

12705–12711.<br />

351 Bulhões, L. O. S.; Obeng, Y. S.; Bard, A.<br />

J. Chem. Mater. 1993, 5, 110–114.<br />

352 Berzina, T. S.; Troitsky, V. I.;<br />

Neilands, O. Ya.; Sudmale, I. V.;<br />

Nicolini, C. Thin Solid Films 1995, 256,<br />

186–191.<br />

353 Castillo, R.; Ramos, S.; Ruiz-Carzia, J.<br />

J. Phys. Chem. B 1996, 100,<br />

15235–15241.<br />

354 Evans, A. K. J. Phys. Chem. B 1998, 102,<br />

7016–7022.<br />

355 Effing, J.; Jonas, U.; Jullien, L.;<br />

Plesnivy, T.; Ringsdorf, H.; Diederich, F.;<br />

Thilgen, C.; Weinstein, D. Angew. Chem.<br />

Int. Ed Engl. 1992, 31, 1599–1602.<br />

356 Dei, L.; LoNostro, P.; Capuzzi, G.;<br />

Baglioni, P. Langmuir 1998, 14,<br />

4143–4147.<br />

357 Vuorimaa, E.; Vuorinen, T.;<br />

Tkachenko, N.; Cramariuc, O.;<br />

Hukka, T.; Nummelin, S.; Shivanyuk, A.;<br />

Rissanen, K.; Lemmetyinen, H.<br />

Langmuir 2001, 17, 7327–7331.<br />

358 Berzina, T. S.; Troisky, V. I.;<br />

Neilands, O. Ya.; Sudmale, I. V.;<br />

Nicolini, C. Thin Solid Films 1995, 256,<br />

186–191.<br />

359 Hawker, C. J.; Saville, P. M.; White, J. W.<br />

J. Org. Chem. 1994, 59, 3503–3505.<br />

360 Maggini, M.; Pasimeni, L.; Prato, M.;<br />

Scorrano, G.; Valli, L. Langmuir 1994, 10,<br />

4164–4166.<br />

361 Guldi, D. M.; Maggini, M.; Mondini, S.;<br />

Guérin, F.; Fendler, J. H. Langmuir 2000,<br />

16, 1311–1318.<br />

362 Nierengarten, J.-F. New J. Chem. 2004,<br />

28, 1177–1191.<br />

363 Nierengarten, J.-F.; Schall, C.;<br />

Nicoud, J.-F.; Heinrich, B.; Guillon, D.<br />

Tetrahedron Lett. 1998, 39, 5747–5750.<br />

364 Cardullo, F.; Diederich, F.;<br />

Echegoyen, L.; Habisher, T.;<br />

Jayaraman, N.; Leblanc, R. M.;<br />

Stoddart, J. F.; Wang, S. Langmuir 1998,<br />

14, 1955–1959.<br />

365 Nierengarten, J.-F.; Eckert, J.-F.; Rio, Y.;<br />

del Pilar Carreon, M.; Gallani, J.-L.;<br />

Guillon, D. J. Am. Chem. Soc. 2001, 123,<br />

9743–9748.


366 Liddell, P. A.; Sunida, J. P.;<br />

McPherson, A. N.; Noss, L.; Seely, G. R.;<br />

Clark, K. N.; Moore, A. L.; Moore, T. A.;<br />

Gust, D. Photochem. Photobiol. 1994, 60,<br />

537–541.<br />

367 Guldi, D. M.; Zilbermann, I.;<br />

Gouloumis, A.; Vázquez, P.; Torres, T.<br />

J. Phys. Chem. B 2004, 108, 18485–18494.<br />

368 Heljala, J.; Tauber, A. Y.; Abel, Y.;<br />

Tkachenko, N. V.; Lemmetyinen, H.;<br />

Kilpeläinen, I.; Hynninen, P. H. J. Chem.<br />

Soc., Pekin Trans. 1 1999, 2403–2408.<br />

369 Tkachenko, N. V.; Vuorimaa, E.;<br />

Kesti, T.; Alekseev, A. S.; Tauber, A. Y.;<br />

Hynninen, P. H.; Lemmetyinen, H.<br />

J. Phys. Chem. B 2000, 104, 6371–6379.<br />

370 Efimov, A.; Vainiotalo, P.;<br />

Tkachenko, N. V.; Lemmetyinen, H.<br />

J. Porphyrins Phthalocyanines 2003, 7,<br />

610–616.<br />

371 Vuorinen, T.; Kaunisto, K.;<br />

Tkachenko, N. V.; Efimov, A.;<br />

Lemmetyinen, H.; Alekseev, A. S.;<br />

Hosomizu, K. Imahori, H. Langmuir<br />

2005, 21, 5383–5390.<br />

372 Guldi, D. M.; Zilbermann, I.;<br />

Anderson, G. A.; Kordatos, K.; Prato, M.;<br />

Tafuro, R.; Valli, L. J. Mater. Chem. 2004,<br />

14, 303–309.<br />

373 Xie, Q.; PPerez-Cordero, E.;<br />

Echegoyen, L. J. Am. Chem. Soc. 1992,<br />

114, 3978–3980.<br />

374 Echegoyen, L.; Echegoyen, L. E.<br />

Acc. Chem. Res. 1998, 31, 593–601.<br />

375 Jehoulet, C.; Bard, A. J.; Wudl, F. J. Am.<br />

Chem. Soc. 1991, 113, 5456–5457.<br />

376 Ji, W.; Tang, S. H.; Xu, G. Q.;<br />

Chan, H. S. O.; Ng, S. C.; Ng, W. W.<br />

J. Appl. Phys. 1993, 74, 3669–3672.<br />

377 Liu, H.; Jia, W.; Lin F.; Mao, S. J.<br />

Luminesc. 1996, 66–67, 128–132.<br />

378 Chollet, P. A.; Kajzar, F. J. Sol-Gel Sci.<br />

Technol. 2001, 22, 255–265,<br />

379 Brusatin, G.; Signorini, R. J. Mater.<br />

Chem. 2002, 12, 1964–1977.<br />

380 Bensasson, R. V.; Hill, T.; Lambert, C.;<br />

Land, E. J.; Leach, S.; Truscott, T. G.<br />

Chem. Phys. Lett. 1993, 201, 326–335.<br />

381 Saunders, M.; Jiménez-Vázquez, H. A.;<br />

Cross, R. J.; Poreda, R. J. Science 1993,<br />

259, 1428–1430.<br />

382 Saunders, M.; Cross, R. J. in Endofullerenes:<br />

A New Family of Carbon<br />

Clusters, Akasaka, T.; Nagase, S. (Eds.),<br />

References<br />

Kluwer Academic Publishers, Dordrecht,<br />

2002, pp. 1–11.<br />

383 Saunders, M.; Cross, R. J.;<br />

Jiménez-Vázquez, H. A.; Shimshi, R.;<br />

Khong, A. Science 1996, 271, 1693–1697.<br />

384 DiCamillo, B. A.; Hettich, R. L.;<br />

Guiochon, G.; Compton, R. N.;<br />

Saunders, M.; Jiménez-Vázquez, H. A.;<br />

Khong, A.; Cross, R. J. J. Phys. Chem.<br />

1996, 100, 9197–9201.<br />

385 Yakigaya, K.; Takeda, A.; Yokoyama, Y.;<br />

Ito, S.; Miyazaki, T.; Suetsuna, T.;<br />

Shimotani, H.; Kakiuchi, T.; Sawa, H.;<br />

Takagi, H.; Kitazawa, K.; Dragoe, N.<br />

New J. Chem. 2007, 31, 973–979.<br />

386 Yamamoto, K.; Saunders, M.; Khong, A.;<br />

Cross, R. J.; Grayson, M.; Gross, M. L.;<br />

Benedetto, A. F.; Weisman, R. B.<br />

J. Am. Chem. Soc. 1999, 121, 1591–1596.<br />

387 Lee, H. M.; Olmstead, M. M.;<br />

Suetsuna, T.; Shimotani, H.; Dragoe, N.;<br />

Cross, R. J.; Kitazawa, K.; Balch, A. L.<br />

Chem. Commun. 2002, 1352–1353.<br />

388 Syamala, M. S.; Cross, R. J.; Saunders, M.<br />

J. Am. Chem. Soc. 2002, 124, 6216–6219.<br />

389 Sternfeld, T.; Hoffman, R. E.;<br />

Saunders, M.; Cross, R. J.;<br />

Syamala, M. S.; Rabinovitz, M. J. Am.<br />

Chem. Soc. 2002, 124, 8786–8787.<br />

390 Krapp, A.; Frenking, G. Chem. Eur. J.<br />

2007, 13, 8256–8270.<br />

391 Peres, T.; Cao, B.; Cui, W.; Khong, A.;<br />

Cross, R. J.; Saunders, M.; Lifshitz, C.<br />

Int. J. Mass Spectrom. 2001, 210, 241–247.<br />

392 Pietzak, B.; Weidinger, A.; Dinse, K.-P.;<br />

Hirsch, A. in Endofullerenes: A New<br />

Family of Carbon Clusters, Akasaka, T.;<br />

Nagase, S. (Eds.), Kluwer Academic<br />

Publishers, Dordrecht, 2002, pp. 13–65.<br />

393 Weidinger, A.; Waiblinger, M.;<br />

Pietzak, B.; Murphy, T. A. Appl. Phys. A<br />

1998, 66, 287–292.<br />

394 Waiblinger, M.; Lips, K.; Harneit, W.;<br />

Weidinger, A. Phys. Rev. B. 2001, 64,<br />

159901.<br />

395 Kanai, M.; Porfyrakis, K.; Andrew, G.;<br />

Briggs, D.; Dennis, T. J. S. Chem.<br />

Commun. 2004, 210–211.<br />

396 Suetsuna, T.; Dragoe, N.; Harneit, W.;<br />

Weidinger, A.; Shimotani, H.; Ito, S.;<br />

Takagi, H.; Kitazawa, K. Chem-A Eur J.<br />

2002, 8, 5079–5083.<br />

397 Khong, A.; Cross, R. J.; Saunders, M.<br />

J. Phys. Chem. A 2000, 104, 3940–3943.<br />

327


328 5 Fullerenes<br />

398 Braun, T.; Rausch, H. Chem. Phys. Lett.<br />

1995, 237, 443–447.<br />

399 Gadd, G. E.; Schmidt, P.; Bowles, C.;<br />

McOrist, G.; Evans, P. J.; Wood, J.;<br />

Smith, L.; Dixon, A.; Easey, J. J. Am.<br />

Chem. Soc. 1998, 120, 10322–10325.<br />

400 Watanabe, S.; Ishioka, N. S.;<br />

Sekine, T.; Kudo, H.; Shimomura, H.;<br />

Muramatsu, H.; Kume, T. J. Radioanal.<br />

Nucl. Chem. 2005, 266, 499–502.<br />

401 Hummelen, J. C.; Prato, M.; Wudl, F.<br />

J. Am. Chem. Soc. 1995, 117, 7003–7004.<br />

402 Rubin, Y. Chem.-A Eur. J. 1997, 3,<br />

1009–1016.<br />

403 Kitagawa, T.; Murata, Y.; Komatsu, K.<br />

in Carbon-Rich Compounds: From<br />

Molecules to Materials, Haley, M. M.;<br />

Tykwinski, R. R. (Eds.), Wiley-VCH,<br />

Weinheim, 2006, pp. 383–420.<br />

404 Rubin, Y.; Jarrosson, T.; Wang, G.-W.;<br />

Bartberger, M. D.; Houk, K. N.;<br />

Schick, G.; Saunders, M.; Cross, R. J.<br />

Angew. Chem. Int. Ed. 2001, 40,<br />

1543–1546.<br />

405 Stanisky, C. M.; Cross, R. J.;<br />

Saunders, M.; Murata, M.; Murata, Y.;<br />

Komatsu, K. J. Am. Chem. Soc. 2005,<br />

127, 299–302.<br />

406 Murata, M.; Murata, Y.; Komatsu, K.<br />

J. Am. Chem. Soc. 2003, 125, 7152–7153.<br />

407 Komatsu, K.; Murata, M.; Murata, Y.<br />

Science 2005, 307, 238–240.<br />

408 Murata, M.; Murata, Y.; Komatsu, K.<br />

J. Am. Chem. Soc. 2006, 128, 8024–8033.<br />

409 Chuang, S.-C.; Murata, Y.; Murata, M.;<br />

Komatsu, K. J. Org. Chem. 2007, 72,<br />

6447–6453.<br />

410 Iwamatsu, S.-I.; Murata, S.; Andoh, Y.;<br />

Minoura, M.; Kobayashi, K.;<br />

Mizorogi, N.; Nagase, S. J. Org. Chem.<br />

2005, 70, 4820–4825.<br />

411 Iwamatsu, S.-I.; Uozaki, T.; Kobayashi, K.;<br />

Suyong, R.; Nagase, S.; Murata, S. J. Am.<br />

Chem. Soc. 2004, 126, 2668–2669.<br />

412 Iwamatsu, S.-I.; Murata, S. Synlett. 2005,<br />

2117–2129.<br />

413 Iwamatsu, S.-I.; Stanisky, C. M.;<br />

Cross, R. J.; Saunders, M.; Mizorogi, N.;<br />

Nagase, S.; Murata, S. Angew. Chem. Int.<br />

Ed. 2006, 45, 5337–5340.<br />

414 Shimshi, R.; Khong, A.;<br />

Jiménez-Vázquez, H.-A.; Cross, R. J.;<br />

Saunders, M. Tetrahedron 1996, 52,<br />

5143–5148.<br />

415 Dietel, E.; Hirsch, A.; Pietzak, B.;<br />

Waiblinger, M.; Lips, L.; Weidinger, A.;<br />

Gruss, A.; Dinse, K.-P. J. Am. Chem. Soc.<br />

1999, 121, 2432–2437.<br />

416 Matsuo, Y.; Isobe, H.; Tanaka, T.;<br />

Murata, Y.; Murata, M.; Komatsu, K.;<br />

Nakamura, E. J. Am. Chem. Soc. 2005,<br />

127, 17148–17149.<br />

417 Sternfeld, T.; Saunders, M.; Cross, R. J.;<br />

Rabinovitz, M. Angew. Chem. Int. Ed.<br />

2003, 42, 3136–3139.<br />

418 Ruttimann, M.; Haldimann, R. F.;<br />

Isaacs, L.; Diederich, F.; Khong, A.;<br />

Jiménez-Vázquez, H. A.; Cross, R. J.;<br />

Saunders, M. Chem. Eur. J. 1997, 3,<br />

1071–1076.<br />

419 Sartori, E.; Ruzzi, M.; Turro, N. J.;<br />

Decatur, J. D.; Doetschman, D. C.;<br />

Lawler, R. G.; Buchachenko, A. L.;<br />

Murata, Y.; Komatsu, K. J. Am. Chem.<br />

Soc. 2006, 128, 14752–14753.<br />

420 Cioslowski, J. J. Am. Chem. Soc. 1991,<br />

113, 4139–4141.<br />

421 Patchkovskii, S.; Thiel, W. J. Chem. Phys.<br />

1997, 105, 1796–1799.<br />

422 Bühl, M.; Patchkovskii, S.; Thiel, W.;<br />

Chem. Phys. Lett. 1997, 275, 14–18.<br />

423 Hu, Y. H.; Ruckenstein, E. J. Chem. Phys.<br />

2005, 123, 144303.<br />

424 Slanina, Z.; Pulay, P.; Nagase, S. J. Chem.<br />

Theory Comput. 2006, 2, 782–785.<br />

425 Dodziuk, H.; Dolgonos, G.; Lukin, O.<br />

Carbon 2001, 39, 1907–1911.<br />

426 Dodziuk, H. Chem. Phys. Lett. 2005, 410,<br />

39–41.<br />

427 Dodziuk, H. J. Nanosci. Nanotechnol.<br />

2007, 7, 1–10.<br />

428 Haddon, R. C.; Brus, L. E.; Raghavachari,<br />

K. Chem. Phys. Lett. 1986, 131,<br />

165.<br />

429 Haddon, R. C. Science 1993, 261, 1545.<br />

430 Bühl, M.; Hirsch, A. Chem. Rev. 2001,<br />

101, 1153.<br />

431 Nossal, J.; Saini, R. K.; Alemany, L. B.;<br />

Meier, M. S.; Billups, W. E. Eur. J. Chem.<br />

2001, 4167.<br />

432 Henderson, C. C.; Cahill, P. A. Science<br />

1993, 259, 1885.<br />

433 Cliffel, D. E.; Bard, A. J. J. Phys. Chem.<br />

1994, 98, 8140.<br />

434 Deronzier, A.; Moutet, J.-C.; Seta, P.<br />

J. Am. Chem. Soc. 1994, 116, 5019.<br />

435 Tkachenko, L. I.; Lobach, A. S.; Strelets,<br />

V. V. Russ.Chem. Bulletin 1998, 47, 1105.


436 Meier, M. S.; Vance, V. K.; Corbin, P. K.;<br />

Clayton, M.; Mollman, M.; Poplawska, M.<br />

Tetrahedron Lett. 1994, 35, 5789.<br />

437 Wang, G.-W.; Li, Y.-J.; Li, F.-B.; Liu, Y.-C.<br />

Letters in Organic Chemistry 2005, 2, 595.<br />

438 Ballenweg, S.; Gleiter, R.;<br />

Krätschmer, W. Tetrahedron Lett. 1993,<br />

34, 3737.<br />

439 Mandrus, D.; Kele, M.; Hettich, R. L.;<br />

Guiochon, G.; Sales, B. C.; Boatner, L. A.<br />

J. of Phys. Chem. B 1997, 101, 123.<br />

440 Avent, A. G.; Darwish, A. D.;<br />

Heimbach, D. K.; Kroto, H. W.;<br />

Meidine, M. F.; Parsons, J. P.;<br />

Remars, C.; Roers, R.; Ohashi, O.;<br />

Taylor, R.; Walton, D. R. M. J. Chem. Soc.<br />

Perkin Trans. 2 1994, 15.<br />

441 Shigematsu, K.; Abe, K. Chem. Express<br />

1992, 7, 905.<br />

442 Jin, C.; Hettich, R.; Compton, R.;<br />

Joyce, D.; Blencoe, J.; Burch, T.<br />

J. Phys. Chem. 1994, 98, 4215.<br />

443 Becker, L.; Evans, T. P.; Bada, J. L.<br />

J. Org. Chem. 1993, 58, 7630.<br />

444 Morosin, B.; Henderson, C.;<br />

Schirber, J. E. Appl. Phys. A 1994, 59, 178.<br />

445 Fujiwara, K.; Komatsu, K. Org. Lett.<br />

2002, 4, 1039.<br />

446 Siedschlag, C.; Luftmann, H.; Wolff, C.;<br />

Mattay, J. Tetrahedron 1999, 55, 7805.<br />

447 Bachmann, M.; Greisheimer, J.;<br />

Homann, K.-H. Chem. Phys. Lett. 1994,<br />

223, 506.<br />

448 Balasubramanian, K. Chem. Phys. Lett.<br />

1992, 182, 257.<br />

449 Henderson, C. C.; Cahill, P. A. Chem.<br />

Phys. Lett. 1992, 198, 570.<br />

450 Fagan, P. J.; Krusic, P. J.; Evans, D. H.;<br />

Lerke, S. A.; Johnston, E. J. Am. Chem.<br />

Soc. 1992, 114, 9697.<br />

451 Niyazymbetov, M. E.; Evans, D. H.;<br />

Lerke, S. A.; Cahill, P. A.;<br />

Henderson, C. C. J. Phys. Chem. 1994,<br />

98, 13093.<br />

452 Meier, M. S.; Spielmann, H. P.;<br />

Haddon, R. C.; Bergosh, R. G.<br />

J. Am. Chem. Soc. 2002, 124, 8090.<br />

453 Bergosh, R. G.; Laske Cooke, J. A.;<br />

Meier, M. S.; Spielmann, H. P.;<br />

Weedon, B. R. J. Org. Chem. 1997, 62,<br />

7667.<br />

454 Henderson, C. C.; Rohlfing, C. M.;<br />

Assink, R. A.; Cahill, P. A. Angew. Chem.<br />

Int. Ed. Engl. 1994, 33, 786.<br />

References<br />

455 Meier, M. S.; Weedon, B. R.;<br />

Spielmann, H. P. J. Am. Chem. Soc.<br />

1996, 118, 11682.<br />

456 Wang, G.-W.; Weedon, B. R.;<br />

Meier, M. S.; Saunders, M.; Cross, R. J.<br />

Org. Lett. 2000, 2, 2241.<br />

457 Cahill, P. A. Chem. Phys. Lett. 1996, 254,<br />

257.<br />

458 Birkett, P. R.; Avent, A. G.;<br />

Darwish, A. D.; Kroto, H. W.; Taylor, R.;<br />

Walton, D. R. M. J. Chem. Soc., Chem.<br />

Commun. 1993, 1230.<br />

459 Haufler, R. E.; Conceicao, J.;<br />

Chibante, L. P. F.; Chai, Y.; Byrne, N. E.;<br />

Flanagan, S.; Haley, M. M.;<br />

O’Brien, S. C.; Pan, C.; Xiao, Z.;<br />

Billups, W. E.; Ciufolini, M. A.;<br />

Hauge, R. H.; Margrave, J. L.;<br />

Wilson, L. J.; Curl, R. F.; Smalley, R. E.<br />

J. Phys. Chem. 1990, 94, 8634.<br />

460 Lobach, A. S.; Shul’ga, Y. M.;<br />

Roshchupikina, O. S.; Rebrov, A. I.;<br />

Perov, A. A.; Morozov, Y. G.;<br />

Spector, V. N.; Ovchinnikov, A. A.<br />

Fullerene Sci. Tech. 1998, 6, 375.<br />

461 Darwish, A. D.; Abdul-Sada, A. K.;<br />

Langley, G. J.; Kroto, H. W.; Taylor, R.;<br />

Walton, D. R. M. Synth. Met. 1996, 77,<br />

303.<br />

462 Darwish, A. D.; Avent, A. G.; Taylor, R.;<br />

Walton, D. R. M. J. Chem. Soc. Perkin<br />

Trans. 2 1996, 2051.<br />

463 Briggs, J. B.; Montgomery, M.; Silva, L. L.;<br />

Miller, G., P. Org. Lett. 2005, 7, 5553.<br />

464 Wågberg, T.; Johnels, D.; Peera, A.;<br />

Hedenström, M.; Schulga, Y. M.;<br />

Tsybin, Y. O.; Purcell, J. M.;<br />

Marshall, A. G.; Noreus, D.; Sato, T.;<br />

Talyzin, A. V. Org. Lett. 2005, 7, 5557.<br />

465 Vieira, S. M. C.; Ahmed, W.;<br />

Birkett, P. R.; Rego, C. A. Chem. Phys.<br />

Lett. 2001, 347, 355.<br />

466 Gakh, A. A.; Romanovich, A. Yu; Bax, A.<br />

J. Am. Chem. Soc. 2003, 125, 7902.<br />

467 Rüchardt, C.; Gerst, M.; Ebenhoch, J.;<br />

Beckhaus, H.-D.; Campbell, E. E. B.;<br />

Tellgmann, R.; Schwarz, H.; Weiske, T.;<br />

Pitter, S. Angew. Chem. Int. Ed. Engl.<br />

1993, 32, 584.<br />

468 Nossal, J. R.; Saini, R. K.; Sadana, A. K.;<br />

Bettinger, H. F.; Alemany, L. B.;<br />

Scuseria, G. E.; Billups, E. W.;<br />

Saunders, M.; Khong, A.; Weisemann, R.<br />

J. Am. Chem. Soc. 2001, 123, 8482.<br />

329


330 5 Fullerenes<br />

469 Darwish, A. D.; Abdul-Sada, A. K.;<br />

Langley, J.; Kroto, H. W.; Taylor, R.;<br />

Walton, D. R. M. J. Chem. Soc., Perkin<br />

Trans. 2 1995, 2359.<br />

470 Attalla, M. I.; Vassallo, A. M.;<br />

Tattam, B. N.; Hanna, J. V. J. Phys. Chem.<br />

1993, 97, 6329.<br />

471 Billups, W. E.; Gonzalez, A.;<br />

Gesenberg, C.; Luo, W.; Marriot, T.;<br />

Alemany, L. B.; Saunders, M.;<br />

Jiménez-Vásquez, H. A.; Khong, A.<br />

Tetrahedron Lett. 1997, 38, 175.<br />

472 Peera, A.; Saini, R. K.; Alemany, L. B.;<br />

Billups, W. E.; Saunders, M.; Khong, A.;<br />

Syamala, M. S.; Cross, R. J. Eur. J. Org.<br />

Chem. 2003, 4140.<br />

473 Dodziuk, H.; Nowinski, K. Chem. Phys.<br />

Lett. 1996, 249, 406.<br />

474 Guo, T.; Scuseria, G. E. Chem. Phys. Lett.<br />

1992, 191, 527.<br />

475 Henderson, C. C.; Rohlfing, C. M.;<br />

Gillen, K. T.; Cahill, P. A. Science 1994,<br />

264, 397.<br />

476 Spielmann, H. P.; Wang, G.-W.;<br />

Meier, M. S.; Weedon, B. R. J. Org.<br />

Chem. 1998, 9865.<br />

477 Birkett, P. R.; Avent, A. G.;<br />

Darwish, A. D.; Kroto, H. W.; Taylor, R.;<br />

Walton, D. R. M. J. Chem. Soc., Chem.<br />

Commun. 1995, 683.<br />

478 Darwish, A. D.; Kroto, H. W.;<br />

Taylor, R. T.; Walton, D. R. M.<br />

J. Chem. Soc. Perkin Trans. 2 1996, 1415.<br />

479 Shigematsu, K.; Abe, K.; Mitani, M.;<br />

Nakao, M.; Tanaka, K. Chem. Express<br />

1993, 8, 669.<br />

480 Choho, K.; Van Lier, G.;<br />

Van de Woude, G.; Geerlings, P.<br />

J. Chem. Soc. Perkin Trans. 2 1996, 1723.<br />

481 Van Lier, G.; Geerlings, P. Chemical<br />

Physics Letters 1998, 289, 591.<br />

482 Li, Y.-J.; Wang, G.-W.; Li, J.-X.; Liu, Y.-C.<br />

New J. Chem. 2004, 28, 1043.<br />

483 Meier, M. S.; Bergosh, R. G.;<br />

Gallagher, M. E.; Spielmann, H. P.;<br />

Wang, Z. J. Org. Chem. 2002, 67,<br />

5946.<br />

484 D’Souza, F.; Choi, J.-P.; Kutner, W.<br />

J. Phys. Chem. B. 1998, 102, 4247.<br />

485 Wang, Z.; Meier, M. S. J. Org. Chem.<br />

2003, 69, 2178.<br />

486 Kroto, H. W.; Heath, J. R.; O’Brien, S. C.;<br />

Curl, R. F.; Smalley, R. E. Nature 1985,<br />

318, 162.<br />

487 Echegoyen, L.; Echegoyen, L. E.<br />

Acc. Chem. Res. 1998, 31, 593.<br />

488 Haddon, R. C.; Hebard, A. F.;<br />

Rosseinski, M. J.; Murphy, D. W.;<br />

Duclos, S. J.; Lyons, K. B.; Miller, B.;<br />

Zahurak, J. M.; Tycko, R.; Dabbagh, G.;<br />

Thiel, F. A. Nature 1991, 350, 320.<br />

489 Holczer, K.; Klein, O.; Huang, S.-M.;<br />

Kaner, R. B.; Fu, K.-J.; Whetten, R. L.;<br />

Diederich, F. Science 1991, 252, 1154.<br />

490 Tutt, L. W.; Krost, A. Nature 1992, 356,<br />

225.<br />

491 Ruoff, R. S.; Tse, D. S.; Malhotra, R.;<br />

Lorents, D. C. J. Phys. Chem. 1993, 97,<br />

3379.<br />

492 Hirsch, A. The Chemistry Of The<br />

Fullerenes, Thieme, Stuttgart, 1994.<br />

493 Diederich, F.; Isaacs, L.; Philp, D. Chem.<br />

Soc. Rev. 1994, 23, 243.<br />

494 Diederich, F.; Thilgen, C. Science 1996,<br />

271, 317.<br />

495 Prato, M.; Maggini, M. Acc. Chem. Res.<br />

1998, 31, 519.<br />

496 Diederich, F.; Kessinger, R. Acc. Chem.<br />

Res. 1999, 32, 537.<br />

497 Nierengarten, J.-F. Chem. Eur. J. 2000, 6,<br />

3667.<br />

498 Vincent, D.; Cruickshank, J. Appl.Opt.,<br />

1997, 36, 7794.<br />

499 Sun, Y.-P.; Lawson, G. E.; Riggs, J. E.;<br />

Ma, B.; Wang, N.; Moton, D. K. J. Phys.<br />

Chem.A 1998, 102, 5520.<br />

500 Schell, J.; Felder, D.; Nierengarten, J.-F.;<br />

Rehspringer, J.-L.; Lévy, R.;<br />

Hönerlage, B. J. Sol-Gel Science And<br />

Technology, 2001, 22, 225.<br />

501 Henari, F.; Callaghan, J.; Stiel, H.;<br />

Blau, W.; Cardin, D. J. Chem. Phys. Lett.<br />

1992, 199, 14.<br />

502 Mclean, D. G.; Sutherland, R. L.;<br />

Brant, M. C.; Brandelik, D. M. Opt. Lett.<br />

1993, 18, 858.<br />

503 Li, C.; Zhang, L.; Wang, R.; Song, Y. L.;<br />

Wang, Y. J. Opt. Soc. Am. B 1994, 11,<br />

1356.<br />

504 Couris, S.; Koudoumas, E.; Ruth, A. A.;<br />

Leach, S. J. Phys. B 1995, 28, 4537.<br />

505 Li, C.; Si, J. H.; Yang, M.; Wang, R. B.;<br />

Zhang, L. Phys. Rev. A 1995, 51, 569.<br />

506 Schell, J.; Brinkmann, D.; Ohlmann, D.;<br />

Hönerlage, B.; Levy, R.; Joucla, M.;<br />

Rehspringer, J.-L.; Serughetti, J.;<br />

Bovier, C. J. Chem. Phys. 1998, 108,<br />

8599.


507 Schell, J.; Ohlmann, D.; Brinkmann, D.;<br />

Lévy, R.; Joucla, M.; Rehspringer, J.-L.;<br />

Hönerlage, B. J. Chem. Phys. 1999, 111,<br />

5929.<br />

508 Bentivegna, F.; Canva, M.; Georges, P.;<br />

Brun, A.; Chaput, F.; Malier, L.;<br />

Boilot, J.-P. Applied Physics Letters, 1993,<br />

62, 1721.<br />

509 Krost, A.; Tutt, L.; Klein, M. B.;<br />

Dougherty, T. K.; Elias, W. E. Opt. Lett.<br />

1993, 18, 334.<br />

510 Gvishi, R.; Bhawalkar, J. D.;<br />

Kumar, N. D.; Ruland, G.; Narang, U.;<br />

Prasad, P. N.; Reinhardt, B. A. Chem.<br />

Mater. 1995, 7, 2199.<br />

511 Signorini, R.; Zerbetto, M.;<br />

Meneghetti, M.; Bozio, R.; Maggini, M.;<br />

De Faveri, C.; Prato, M.; Scorrano, G.<br />

Chem. Commun. 1996, 1891.<br />

512 Maggini, M.; De Faveri, C.; Scorrano, G.;<br />

Prato, M.; Brusatin, G.; Guglielmi, M.;<br />

Meneghetti, M.; Signorini, R.; Bozio, R.<br />

Chem. Eur. J. 1999, 5, 2501.<br />

513 Sun, Y.-P.; Lawson, G. E.; Riggs, J. E.;<br />

Ma, B.; Wang, N.; Moton, D. K. J. Phys.<br />

Chem. A 1998, 102, 5520.<br />

514 Felder, D.; Guillon, D.; Lévy, R.;<br />

Mathis, A.; Nicoud, J.-F.;<br />

Nierengarten, J.-F.; Rehspringer, J.-L.;<br />

Schell, J. J. Mater. Chem. 2000, 10, 887.<br />

515 Rio, Y.; Nicoud, J.-F.; Rehspringer, J.-L.;<br />

Nierengarten, J.-F. Tetrahedron Lett.<br />

2000, 41, 10207.<br />

516 Rio, Y.; Accorsi, G.; Nierengarten, H.;<br />

Rehspringer, J.-L.; Hönerlage, B.;<br />

Kopitkovas, G.; Chugreev, A.;<br />

Van Dorsselaer, A.; Armaroli, N.; Nierengarten,<br />

J.-F. New J. Chem. 2002, 26, 1146.<br />

517 Nierengarten, J.-F.; Armaroli, N.;<br />

Accorsi, G.; Rio, Y.; Eckert, J.-F. Chem.<br />

Eur. J. 2003, 9, 36.<br />

518 Nierengarten, J.-F.; Topics In Current<br />

Chem. 2003, 228, 87.<br />

519 Rio, Y.; Accorsi, G.; Nierengarten, H.;<br />

Bourgogne, C.; Strub, J.-M.;<br />

Van Dorsselaer, A.; Armaroli, N.;<br />

Nierengarten, J.-F. Tetrahedron 2003, 59,<br />

3833.<br />

520 Maggini, M.; De Faveri, C.; Scorrano, G.;<br />

Prato, M.; Brusatin, G.; Guglielmi, M.;<br />

Meneghetti, M.; Signorini, R.; Bozio, R.<br />

Chem.Eur.J. 1999, 5, 2501.<br />

521 Brusatin, G.; Signorini, R. J. Mater.<br />

Chem. 2002, 12, 1964.<br />

References<br />

522 Sahraoui, B.; Kityk, I. V.;<br />

Hudhomme, P.; Gorgues, A. J.Phys.<br />

Chem.B 2001, 105, 6295.<br />

523 Kreher, D.; Cariou, M.; Liu, S.-G.;<br />

Levillain, E.; Veciana, J.; Rovira, C.;<br />

Gorgues, A.; Hudhomme, P. J. Mater.<br />

Chem. 2002, 12, 2137.<br />

524 Allard, E.; Delaunay, J.; Cheng, F.;<br />

Cousseau, J.; Orduna, J.; Garín, J.<br />

Org. Lett. 2001, 3, 3503.<br />

525 Zhao, Y.; Shirai, Y.; Slepkov, A. D.;<br />

Cheng, L.; Alemany, L. B.; Sasaki, T.;<br />

Hegmann, F. A.; Tour, J. M. Chem. Eur.<br />

J. 2005, 11, 3643.<br />

526 Tian, Z.; He, C.; Liu, C.; Yang, W.;<br />

Yao, J.; Nie, Y.; Gong, Q.; Liu, Y.<br />

Mater. Chem.Phys. 2005, 94, 444.<br />

527 Doyle, J.; Ballesteros, B.; De La Torre, G.;<br />

Mcgovern, D.; Kelly, J.; Torres, T.;<br />

Blau, W. Chem. Phys. Lett. 2006, 428,<br />

307.<br />

528 Sariciftci, N. S.; Smilowitz, L.;<br />

Heeger, A. J.; Wudl, F. Science 1992, 258,<br />

1474.<br />

529 Sariciftci, N. S.; Heeger, A. J. in<br />

Handbook Of Organic Conductive<br />

Molecules And Polymers, Vol. 1,<br />

Nalwa, H. S. (Ed.), Wiley, New York<br />

1997, 8, 413.<br />

530 Brabec, C. J.; Sariciftci, N. S.;<br />

Hummelen, J. C. Adv. Funct. Mater.<br />

2001, 11, 15.<br />

531 Benati, T. L.; Venkataraman, D.<br />

Photosynth. Res. 2006, 87, 73.<br />

532 Sariciftci, N. S.; Heeger, A. J.<br />

U.S. Patent 005331183a, 1994.<br />

533 Spanggaard, H.; Krebs, F. C. Solar<br />

Energy Mater. Solar Cells 2004, 83, 125.<br />

534 Sariciftci, N. S.; Braun, D.; Zhang, C.;<br />

Srdanov, V. I.; Heeger, A. J.; Stucky, G.;<br />

Wudl, F. Appl. Phys. Lett. 1993, 62, 585.<br />

535 Yu, G.; Gao, J.; Hummelen, J. C.;<br />

Wudl, F.; Heeger, A. J. Science 1995, 270,<br />

1789.<br />

536 Hummelen, J. C.; Knight, B. W.;<br />

Lepeq, F.; Wudl, F. J. Org. Chem. 1995,<br />

60, 532.<br />

537 Shaheen, E.; Brabec, C. J.;<br />

Sariciftci, N. S.; Pradinger, F.;<br />

Fromherz, T.; Hummelen, J. C.<br />

Appl. Phys. Lett. 2001, 78, 841.<br />

538 Ma, W.; Yang, C.; Gong, X.; Lee, K.;<br />

Heeger, A. J. Adv. Funct. Mater. 2005, 15,<br />

1617.<br />

331


332 5 Fullerenes<br />

539 Huang, C.-H.; Mcclenaghan, N. D.;<br />

Kuhn, A.; Hofstraat, J. W.;<br />

Bassani, D. M. Org. Lett. 2005, 7, 3409.<br />

540 Gutiérrez-Nava, M. Setayesh, S.;<br />

Rameau, A.; Masson, P.; Nierengarten,<br />

J.-F. New J. Chem. 2002, 26, 1584.<br />

541 Gutiérrez-Nava, M.; Masson, P.;<br />

Nierengarten, J.-F. Tetrahedron Lett.<br />

2003, 44, 4487.<br />

542 Eklund, P. C.; Rao, A. M. Fullerene<br />

Polymers And Fullerene Polymer Composites,<br />

Springer Series In Materials<br />

Science, Vol. 38, Springer Verlag, Berlin,<br />

2000.<br />

543 Chiang, L. Y.; Wang, L. Y.; Tseng, S.-G.;<br />

Wu, J.-S.; Hsieh, K.-H. Chem. Commun.<br />

1994, 2675.<br />

544 Ederlé, Y.; Mathis, C. Macromolecules<br />

1999, 32, 554.<br />

545 Nierengarten, J.-F. New. J. Chem. 2004,<br />

28, 1177.<br />

546 Brabec, C. J.; Sariciftci, N. S.;<br />

Hummelen, J. C. Adv. Funct. Mater.<br />

2001, 11, 15 And References Therein.<br />

547 Nierengarten, J.-F.; Eckert, J.-F.;<br />

Nicoud, J.-F.; Ouali, L.; Krasnikov, V.;<br />

Hadziioannou, G. Chem. Commun. 1999,<br />

617.<br />

548 Yamashita, Y.; Takashima, W.; Kaneto, K.<br />

Jpn J. Appl. Phys. 1993, 32, 1017.<br />

549 Peumans, P.; Forrest, S. R. Appl. Phys.<br />

Lett. 2001, 79, 126.<br />

550 Bogdanov, G. N.; Kotel’nikova, R. A.;<br />

Frog, E. S.; Shtol’ko, V. N.;<br />

Romanova, V. S.; Bubnov, Y. N.<br />

Dokl. Biochem. Biophys. 2004, 396, 165.<br />

551 Enes, R. F.; Tome, A. C.; Cavaleiro, J. A. S.<br />

Tetrahedron 2005, 61, 1423.<br />

552 Mikata, Y.; Takagi, S.; Tanahashi, M.;<br />

Ishii, S.; Obata, M.; Miyamoto, Y.;<br />

Wakita, K.; Nishisaka, T.; Hirano, T.;<br />

Ito, T.; Hoshino, M.; Ohtsuki, C.;<br />

Tanihara, M.; Yano, S. Bioorg. Med.<br />

Chem. Lett. 2003, 13, 3289.<br />

553 Murthy, C. N.; Geckeler, K. E. Chem.<br />

Comm. 2001, 1194.<br />

554 Takekuma, S.-I.; Takekuma, H.;<br />

Matsumoto, T.; Yoshida, Z.-I. Tetrahedron<br />

Lett. 2000, 41, 4909.<br />

555 Ikeda, A.; Hatano, T.; Kawaguchi, M.;<br />

Suenaga, H.; Shinkai, S. Chem. Comm.<br />

1999, 1403.<br />

556 Rio, Y.; Nierengarten, J.-F. Tetrahedron<br />

Lett. 2002, 43, 4321.<br />

557 Pantarotto, D.; Tagmatarchis, N.;<br />

Bianco, A.; Prato, M. Mini Rev. Med.<br />

Chem. 2004, 4, 805.<br />

558 Nakamura, E.; Isobe, H. Acc. Chem. Res.<br />

2003, 36, 807.<br />

559 Bosi, S.; Da Ros, T.; Spalluto, G.; Prato, M.<br />

Eur. J. Med. Chem. 2003, 38, 913.<br />

560 Tagmatarchis, N.; Shinohara, H.<br />

Mini Rev. Med. Chem. 2001, 1, 339.<br />

561 Bianco, A.; Da Ros, T.; Prato, M.;<br />

Toniolo, C. J. Pept. Sci. 2001, 7, 208.<br />

562 Da Ros, T.; Prato, M. Chem. Commun.<br />

1999, 663.<br />

563 Jensen, A. W.; Wilson, S. R.;<br />

Schuster, D. I. Bioorg. Med. Chem. 1996,<br />

4, 767.<br />

564 Mitsutoshi, S.; Takayanagi, I.<br />

J. Pharmacol. Sci. 2006, 100, 513.<br />

565 Wang, S.; Ruomei, G.; Zhou, F.;<br />

Selke, M. J. Mater. Chem. 2004, 14, 487.<br />

566 Tokuyama, H.; Yamago, S.;<br />

Nakamura, E.; Shiraki, T.; Sugiura, Y.<br />

J. Am. Chem. Soc. 1993, 115, 6506.<br />

567 Liu, Y.; Zhao, Y. L.; Chen, Y.; Liang, P.;<br />

Li, L. Tetrahedron Lett. 2005, 46, 2507.<br />

568 Milanesio, M.; Alvarez, M.; Rivarola, V.;<br />

Silber, J.; Durantini, E. Photochem.<br />

Photobiol. 2005, 81, 891.<br />

569 Hirsch, A.; Lamparth, I.;<br />

Karfunkel, H. R. Angew. Chem. Int. Ed.<br />

Engl. 1994, 33, 1994.<br />

570 Dugan, L. L.; Lovett, E.; Cuddihy, S.;<br />

Ma, B.-W.; Lin, T.-S.; Choi, D. W. In<br />

‘Carboxyfullerenes As Neuroprotective<br />

Antioxidants’, Kadish, K. M.; Ruoff, R. S.<br />

(Eds.), 2000.<br />

571 Ali, S. S.; Hardt, J. I.; Quick, K. L.;<br />

Kim-Han, J. S.; Erlanger, B. F.;<br />

Huang, T. T.; Epstein, C. J.; Dugan, L. L.,<br />

Free Radical Bio. Med. 2004, 37, 1191.<br />

572 Djordjevic, A. N.; Canadanovic-<br />

Brunet, J. M.; Voijnovic-Miloradov, M.;<br />

Bogdanovic, G. M. Oxid. Commun. 2004,<br />

27, 806.<br />

573 Chen, Y. W.; Hwang, K. C.; Yen, C. C.;<br />

Lai, Y. L. Am. J. Physiol.-Reg. I. 2004, 287,<br />

R21.<br />

574 Bogdanovic, G. M.; Kojic, V.;<br />

Dordevic, A.; Canadanovic-Brunet, J.-M.;<br />

Vojinovic-Miloradov, M. B.; Baltic, V. V.<br />

Toxicol. In Vitro 2004, 18, 629.<br />

575 Mirkov, S. M.; Djordjevic, A. N.;<br />

Andric, N. L.; Andric, S. A.; Kostic, T. S.;<br />

Bogdanovic, G. M.; Vojinovic-Miloradov,


M. B.; Kovacevic, R. Z. Nitric Oxide-Biol.<br />

Ch. 2004, 11, 201.<br />

576 Sun, T.; Jia, Z.; Xu, Z. Bioorg. Med.<br />

Chem. Lett. 2004, 14, 1779.<br />

577 Bosi, S.; Da Ros, T.; Castellano, S.;<br />

Banfi, E.; Prato, M. Bioorg. Med. Chem.<br />

Lett. 2000, 10, 1043.<br />

578 Mashino, T.; Nishikawa, D.;<br />

Takahashi, K.; Usui, N.; Yamori, T.;<br />

Seki, M.; Endo, T.; Mochizuchi, M.<br />

Bioorg. Med. Chem. Lett., 2003, 13,<br />

4395.<br />

579 Schinazi, R. F.; Sijbesma, R. P.;<br />

Srdanov, G.; Hill, C. L.; Wudl, F.<br />

Antimicrob. Agents Ch. 1993, 37, 1707.<br />

580 Nakamura, E.; Tokuyama, H.;<br />

Yamago, S.; Shiraki, T.; Sugiura, Y.<br />

Bull. Chem. Soc. Jpn. 1996, 69, 2143.<br />

581 Friedman, S. H.; Ganapathi, P. S.;<br />

Rubin, Y.; Kenyon, G. L. J. Med. Chem.<br />

1998, 41, 2424.<br />

582 Marcorin, G.; Da Ros, T.; Castellano, S.;<br />

Stefancich, G.; Borin, I.; Miertus, S.;<br />

Prato, M. Org. Lett. 2000, 2, 3955.<br />

583 Schuster, D. I.; Wilson, L. J.;<br />

Kirschner, A. N.; Schinazi, R. F.;<br />

Schlueter-Wirtz, S.; Tharnish, P.;<br />

Barnett, T.; Ermolieff, J.; Tang, J.;<br />

Brettreich, M.; Hirsch, A. in Fullerene<br />

References<br />

2000, Vol. 9: Functionalized Fullerenes,<br />

Maggini, M.; Martin, N.; Guldi, D. M.<br />

(Eds.) The Electrochemical Society,<br />

Pennington, NJ, USA, 2000, 267.<br />

584 Bosi, S.; Da Ros, T.; Spalluto, G.;<br />

Balzarini, J.; Prato, M. Bioorg. Med.<br />

Chem. Lett. 2003, 13, 4437.<br />

585 Mashino, T.; Shimotohno, K.;<br />

Ikegami, N.; Nishikawa, D.; Okuda, K.;<br />

Takahashi, K.; Nakamura, S.;<br />

Mochizuki, M. Bioorg. Med. Chem. Lett.<br />

2005, 15, 1107.<br />

586 Georgakilas, V.; Pellarini, F.; Prato, M.;<br />

Guldi, D.; Melle-Franco, M.; Zerbetto, F.<br />

P. Natl. Acad. Sci. Usa, 2002, 99, 5075.<br />

587 Nakamura, E.; Isobe, H.; Tomita, N.;<br />

Sawamura, M.; Jinno, S.; Okayama, H.<br />

Angew. Chem., Int. Ed. 2000, 39, 4254.<br />

588 Chan, H. B. S.; Ellis, B. L.;<br />

Sharma, H. L.; Frost, W.; Caps, V.;<br />

Shields, R. A.; Tsang, S. C. Adv. Mater.<br />

2004, 16, 144.<br />

589 Wharton, T.; Wilson, L. Bioorg. Med.<br />

Chem. 2002, 10, 3545.<br />

590 Zalibera, M.; Popov, A. A.; Kalbac, M.;<br />

Rapta, P.; Dunsch, L. Chem. Eur. J. 2008,<br />

14, 9960.<br />

591 Tao, Y.-Z.; Liao, Z.-J., et al., Nature<br />

Materials 2008, 7, 790–794.<br />

333


6<br />

Carbon Nanotubes<br />

6.1<br />

The Structure and Properties of Carbon Nanotubes<br />

Anke Krueger<br />

6.1.1<br />

Introduction<br />

Another class of strained (hydro)carbon structures consists of only carbon in<br />

its pristine state. Tubular objects with a graphitic structure can be considered<br />

an extremely unsaturated form of an aromatic hydrocarbon. These so-called<br />

carbon nanotubes (CNTs) had been discovered long before they became a major<br />

research object for chemists and physicists alike. M. Endo described tubular<br />

carbon structures with several walls and coined the name ‘carbon nanotubes’<br />

for these interesting objects (Figure 6.1a) and S. Iijima recognized the concentrically<br />

rolled shape of these objects [1]. Originally, they were thought to be just<br />

another type of carbon fiber before their unique properties were discovered. It<br />

was only in the 1990s that another type of CNT, the single-walled parent systems<br />

of the more complex MWNT (multi-walled nanotubes), were described by Iijima<br />

(Figure 6.1b) [2]. Their structure was unambiguously determined by electron<br />

microscopy (HRTEM, STEM; STM, AFM) and their spectroscopic properties<br />

investigated [3]. To understand these materials the knowledge of their structure<br />

is of utmost importance. This chapter will discuss the structure of single-walled<br />

nanotubes (SWNT) as well as their multi-walled counterparts. Additionally, the<br />

aspect of aromaticity will be reviewed.<br />

6.1.2<br />

The Structure of Single-walled Carbon Nanotubes<br />

The basic structural element of SWNT is the graphene plane. This is the basic<br />

unit of graphite, which can be formally rolled up to yield a tubular structure in<br />

the case of CNTs (Figure 6.1). Of course, the real production of CNTs is carried<br />

out in different ways. Various methods including CVD techniques, laser ablation,<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

335


336 6 Carbon Nanotubes<br />

Figure 6.1 HRTEM images of (a) MWNT (from Nature 1991, 354, 56),<br />

(b) SWNT (courtesy of F. Banhart) and (c) the formal roll-up of graphene to form tubes.<br />

arc discharge, the HiPCo process etc. have been developed for different demands<br />

of quality, size and purity of CNT samples [3].<br />

The seamless tubular objects obtained are nevertheless directly related to<br />

their planar counterpart (graphene or graphite) concerning their symmetry,<br />

spectroscopic and chemical properties, but their unique structure causes various<br />

peculiarities as well. In order to understand these features the structure of the<br />

different types of CNTs has to be known in detail.<br />

Dresselhaus defined a logical system for all types of CNT in order to approach the<br />

structural description systematically [4]. The basis of this method is the graphene<br />

layer. Depending on the way of seamless roll-up, one obtains different structure<br />

types of CNT. There are two privileged directions on the graphene sheet. They are<br />

represented by the unit vectors of its unit cell and form an angle of 60° (Figure 6.2).<br />

When a nanotube is formed from a graphene sheet, the roll-up is defined by a<br />

vector C that corresponds to its circumference after the tube formation (Figure 6.2).<br />

The direction of this vector defines the roll-up direction of the graphene sheet<br />

with the tube axis being perpendicular to C.<br />

The vector C can be described as the vector sum of multiples of the unit vectors<br />

a 1 and a 2 of the graphene lattice with<br />

C = n ⋅ a + m ⋅ a with n ≥ m and m ≥ 0<br />

(6.1)<br />

1 2<br />

A CNT is then unambiguously described by the set of descriptors n and m. Only<br />

its length needs to be defined, as all tubes with the same diameter and orientation<br />

but different lengths possess the same pair (n, m). It turns out that the length<br />

does not have a major importance for most of the real CNTs as they are very<br />

long compared to their diameter (up to several microns compared to a few nano-


6.1 The Structure and Properties of Carbon Nanotubes<br />

Figure 6.2 Construction of CNTs from graphene (top: the vector for a (6,3) tube is inscribed),<br />

and chiral angle � and length of the translational unit cell of a (6,3) nanotube (bottom).<br />

meters) [5]. Only in the case of very short tubes do the effects of the ends, such as<br />

wall bending, have to be taken into account [6]. The diameter and the length of<br />

CNTs are quantified according to the quantification of the underlying graphene<br />

lattice. The diameter needs to yield a seamless tube; the length is determined by<br />

the numbers of hexagons along the tube axis.<br />

Some rules exist for the unambiguous description of CNT by this system.<br />

Firstly, m and n are real numbers; secondly, m must be greater than or equal to n<br />

in order to avoid redundancy in the description, e.g. a (6,0) nanotube is identical<br />

to a (0,6) nanotube. In this case only the orientation of the rolled-up graphene<br />

sheet is changed by 60°.<br />

Three different types of SWNT can be distinguished (Figure 6.3): When n = 0,<br />

the orientation of the vector C is parallel to a 1 . This type is called zigzag nanotube<br />

because it has a zigzag structure at either end. The other archetype nanotube with<br />

m = n is called armchair nanotube. Its ends show a pattern that resembles the<br />

armchair conformation of cyclic hydrocarbons.<br />

337


338 6 Carbon Nanotubes<br />

Figure 6.3 Zigzag, armchair and chiral carbon nanotubes.<br />

In between these two extremes, all sets of m and n shape chiral nanotubes. The<br />

chirality originates from a helical roll-up of the graphene sheet and the chiral<br />

angle � is defined as<br />

⎛ 3 m ⎞<br />

� = sin ⎜ ⎟<br />

−1 2 2<br />

⎝n + n m + m ⎠ (6.2)<br />

These chiral tubes exist in two enantiomeric forms (left-handed and right-handed<br />

roll-up), which have been separated after noncovalent functionalization [7].<br />

The diameter of the nanotube is also defined by the structure parameter set:<br />

1<br />

2 2<br />

dCNT = ⋅ a ⋅ n + n m + m<br />

(6.3)<br />

�<br />

According to Equation (6.3), a (5,5) nanotube, a (7,3) nanotube and a (9,0) tube<br />

have a very similar diameter of ~0.7 nm, close to the diameter of the archetypical<br />

fullerene C60 . Although indistinguishable by their size (e.g. in a HRTEM image),<br />

these tubes exhibit completely different properties. Whereas the (5,5) armchair<br />

tube is electrically conducting, the zigzag (9,0) tube is a semiconductor at absolute<br />

zero but conducting at room temperature, whereas the (7,3) nanotube is a semiconductor.<br />

The different symmetry of chiral, armchair and zigzag tubes is also<br />

responsible for their different Raman spectra [8].<br />

In general, all properties depend on the respective nanotube symmetry. The<br />

density of states (see Figure 6.4 for examples) and consequently the electrical<br />

conductivity depend on the following conditions. The conductivity is governed<br />

by the existence of a gap at the Fermi level. All armchair nanotubes are metallic.<br />

In the case of semiconducting tubes a gap is opening between the valence and<br />

the conduction band [9]. Those nanotubes with n – m = 3 q (q � 0) possess a small<br />

gap at the absolute zero; at room temperature they behave like metallic tubes<br />

(Table 6.1) [9].


6.1 The Structure and Properties of Carbon Nanotubes<br />

Figure 6.4 Density of states (DOS) of a (9,0) and a (10,0) carbon nanotube [12].<br />

In the case of the (9,0) tube the density of states is not zero at the Fermi level.<br />

Table 6.1 Electrical conductivity as a function of structure parameters m and n [9].<br />

Nanotube Conditions<br />

Electrically conducting metallic nanotubes n = m<br />

Semiconducting nanotubes with a small band gap<br />

(conducting at room temperature)<br />

n – m = 3 q (q � 0)<br />

Semiconducting nanotubes n – m � 3 q<br />

In real samples of CNTs these types coexist, and so far it has not been possible<br />

to produce a certain type or even tubes with a defined set (m,n) selectively [10].<br />

Only the enrichment was achieved by different methods including electrophoresis<br />

[11] and selective functionalization (see next section).<br />

Another important aspect of the CNT structure is the construction of the tips.<br />

The formal roll-up of a graphene sheet would lead to unsaturated bonds at the<br />

339


340 6 Carbon Nanotubes<br />

Figure 6.5 Short carbon nanotubes with functionalized tips; (a) hydrogenated, (b) oxidized tip.<br />

In the case of the short hydrogenated tube the widening of the ends is highlighted.<br />

terminal carbon atoms at both rims and therefore to an elevated and unfavorable<br />

energetic state. In reality, the ends of CNTs show many different ways to overcome<br />

the dangling bonds and to produce a valence-saturated structure. The simplest<br />

would be the saturation of all dangling bonds with hydrogen (Figure 6.5a). Calculations<br />

by Schleyer et al. showed that short fragments of CNTs with hydrogenated<br />

ends are widened at their tips and no longer exhibit a perfect tubular shape [6].<br />

Other functional groups can be also present at the tips of CNTs (Figure 6.5b). They<br />

are usually introduced by oxidative reagents, e.g. mineral acids, and are often used<br />

to introduce more complex moieties at the CNT surface (see Section 6.2).<br />

After production, the tips of SWNTs are usually closed by caps of different shape<br />

(see Figure 6.6). In principle, it is possible to postulate partial fullerene structures<br />

that fit exactly to the rim to be covered with a cap. Fujita and Dresselhaus have developed<br />

a formalism for the construction of CNT tips by arranging five-membered<br />

rings on the graphene sheet at the ends of the tube shell (Figure 6.6b) [13]. The<br />

five-membered rings are responsible for the curvature of the cap and the formation<br />

of a partial fullerene structure. The opening angle of the cap depends linearly on<br />

the number of five-membered rings in the fullerenic cap (Table 6.2) [14].<br />

In the case of the archetypical zigzag (9,0) and armchair (5,5) carbon nanotubes<br />

it is theoretically possible to cover their ends with cut C 60 molecules. The fullerene<br />

hemisphere has to be selected according to the rim structure of the nanotube<br />

(Figure 6.7). Although highly symmetrical, those structures are rarely formed.<br />

Most often somewhat irregular nanotube caps are observed, often having sharp<br />

ends [15].<br />

Real CNTs are usually not defect free (Figure 6.8). In addition to the imperfect<br />

structures at their ends they have also defects in the sidewalls. Several types of<br />

these defects are known. Missing carbon atoms create holes in the graphene<br />

network of the wall, disturbing the electronic properties as well as the �-conjugation<br />

Table 6.2 Opening angle of the nanotube cap as a function of five-membered rings in the cap.<br />

Number of five-membered ring 1 2 3 4 5<br />

Opening angle (deg.) 112.9 83.6 60.0 38.9 19.2


6.1 The Structure and Properties of Carbon Nanotubes<br />

Figure 6.6 (a) HRTEM image of open (arrows) and closed SWNTs (from Chem. Phys. Lett.<br />

2000, 316, 349). (b) An example for the construction of a theoretically possible cap according<br />

to the method of Fujita and Dresselhaus. Highlighted hexagons represent the positions of<br />

pentagons in the cap (following the scheme in the lower right corner a segment from these<br />

hexagons is removed). Superimposing the fields with equal numbers results in the formation of<br />

the curved cap.<br />

Figure 6.7 Capping of a (5,5) and a (9,0) carbon nanotube by appropriate half-spheres of C 60 .<br />

The rims of the tubes are marked grey.<br />

341


342 6 Carbon Nanotubes<br />

Figure 6.8 Defects in the side wall of CNTs. (a) Holes; (b) Stone–Wales defect (side and<br />

top view) with the highest reactivity at the borders of the defect (highlighted in grey);<br />

(c) the migration of Stone–Wales defects (top) and Stone–Wales rearrangement (bottom).<br />

(see below) [16]. Another defect consists in the substitution of six-membered rings<br />

by pairs of five-membered and seven-membered rings. This so-called Stone–Wales<br />

defect is usually built by two sets of these rings (Figure 6.8) [17]. It can be produced<br />

either by a rearrangement where a bond between two six-membered rings is turned<br />

by 90° or by the addition of a C 2 unit to the graphene network [18].<br />

In summary, the effect of a Stone–Wales defect on the curvature is zero as the<br />

convex contribution of the pentagons is compensated by the concave contribution<br />

of the heptagons. These rings do not necessarily occupy neighboring positions.<br />

In a rearrangement reaction including a four-membered Hückel transition state<br />

they migrate away from each other leading to bent or even coiled nanotubes,<br />

if a certain number of these defects is positioned in a regular manner [19]. For<br />

example, a pentagon on one (convex) side of the nanotube and a heptagon on the<br />

other (concave) side leads to a kink in the respective nanotube.<br />

In the case of a complete Stone–Wales defect the curvature is not evenly distributed<br />

over the whole defect structure. Thus the reactivity of the different bonds<br />

depends on the prehybridization and strain exerted by the additional curvature<br />

(see Figure 6.8).<br />

6.1.3<br />

The Structure of Multi-walled Carbon Nanotubes<br />

The structure of MWNTs is closely related to that of SWNTs. Each wall of the<br />

MWNT represents a SWNT, only their diameters differ. The whole object is a


6.1 The Structure and Properties of Carbon Nanotubes<br />

combination of these walls and the main question is whether there are interactions<br />

between the single layers of a MWNT. The inter-wall distance in MWNT is mostly<br />

slightly larger than the distance between the graphene sheets in graphite [20].<br />

This indicates that the interactions between the walls should be somewhat weaker<br />

than in graphite. Inspection of the possible interactions between the walls in<br />

different types of MWNT reveals a fundamental difficulty: The interlayer distance<br />

of 0.34 nm defines a certain increase of � u = 2.14 nm in circumference (because<br />

of the increase of � d = 0.68 nm in diameter). Depending on the type of CNT this<br />

can be expressed in multiples of the basic unit of the circumferential vector C.<br />

In the case of armchair nanotubes this corresponds well to � u = 5 · 0.426 nm. On<br />

the other hand, it is difficult to produce a good fit for zigzag tubes as the nearest<br />

multiple is � u = 9 · 0.246 nm = 2.214 nm. With this increase in diameter the interwall<br />

distance increases to 0.352 nm.<br />

A real MWNT will rarely be formed of just one type of CNT, making the situation<br />

even more complex. Chiral tubes are very unlikely to fit with their neighboring<br />

tubes and a mixture of zigzag, armchair and chiral tubes is usually present in real<br />

MWNTs [21]. As the interaction between the walls is rather weak, the mobility of<br />

the inner walls is only slightly hindered and rotation is possible [22]. This is also<br />

responsible for the ‘sword in a sheath’ phenomenon which is observed when a<br />

MWNT filled composite is mechanically strained in the direction of the MWNT<br />

[23]. Multi-walled carbon nanotubes show mixed electronic behavior [24]. Some<br />

authors state that the outermost wall determines whether a tube is metallic or<br />

semiconducting [25].<br />

In similarity to single-walled CNTs, MWNTs exhibit even more complex<br />

structures at the ends of the cylindrical body. Especially, the presence of sevenmembered<br />

rings in addition to the usual five-membered rings, opens many<br />

possibilities for tip structures. Mostly, slightly unsymmetrical tips are observed,<br />

beak-like structures as well as so-called ‘acute’ tips with sharp ends can also be<br />

found (Figure 6.9) [26].<br />

Another feature of real MWNTs is the occurrence of internal structures.<br />

These include inner caps as well as bamboo-like structures (Figure 6.10). These<br />

features can be formed by one or more internal walls of the MWNT [27]. In<br />

some cases one observes toroidal structures at the tips of opened MWNT.<br />

The neighboring walls are joined by C–C bonds, diminishing the number of<br />

dangling bonds [28].<br />

Figure 6.9 Cap structures of real MWNTs.<br />

343


344 6 Carbon Nanotubes<br />

Figure 6.10 Internal structures of multi-walled carbon nanotubes. Internal caps at a bend (left) and<br />

HRTEM image of bamboo-like carbon structures (right). (From Chem. Phys. Lett. 2000, 323, 560).<br />

Figure 6.11 Electron-microscopic images of coiled multi-walled carbon nanotubes<br />

(from J. Mater. Res. 2000, 15, 808 (top) and Appl. Phys. Lett. 2002, 81, 3567 (bottom)).<br />

Coiled and bent MWNTs have been also found in nanotube samples. These<br />

defects are formed by the existence of five- and seven-membered rings in the<br />

graphene network (see above) [19]. With appropriate conditions, strongly coiled<br />

MWNTs can be produced in macroscopic amounts (Figure 6.11) [29].<br />

In comparison to SWNTs the defectiveness of the outer walls of a MWNT does<br />

not affect the mechanical and chemical properties in the same dramatic way.<br />

Even if the outermost walls have holes, the tube shape remains still closed as<br />

the inner walls are still intact. On the other hand, the electronic properties are


6.1 The Structure and Properties of Carbon Nanotubes<br />

affected more seriously because of the major contribution of the outer shell to<br />

the electrical transport [30].<br />

6.1.4<br />

The Aromaticity of Carbon Nanotubes<br />

When we regard carbon nanotubes as a type of highly unsaturated form of hydrocarbon<br />

the issue of aromaticity needs to be discussed.<br />

The concept of aromaticity is difficult to apply to carbon nanotubes as the conjugated<br />

system shows a significant curvature along its circumference. On the other<br />

hand there is no curvature in the direction of the tube axis. Is a nanotube aromatic<br />

or not? The question has to be answered whether a sufficient delocalization of the<br />

�-electrons is realized or not. According to the bond lengths, conjugation reaches a<br />

high level as there is no significant alternation of single and double bonds [31].<br />

For the discussion it is useful to neglect the effects of the tips in the first instance<br />

as they represent defects in the perfect structure of a straight, defect-free CNT. The<br />

endless graphene layer, from which a hypothetical perfect nanotube is formally<br />

derived, possesses a planar, fully delocalized �-electron cloud and therefore can<br />

be considered to be aromatic [32]. In the case of the rolled-up nanotube the participating<br />

electrons and their orbitals are still the same. Only the orientation and<br />

to some extent the shape of the �-orbitals is different. A 3D structure is formed<br />

and the conjugation is weakened by the radial distribution of the �-orbitals<br />

(Figure 6.12). Along the tube axis the extent of conjugation remains the same but<br />

in the circumferential direction it changes the more, the smaller the diameter of<br />

the nanotube [31, 33].<br />

Figure 6.12 Orientation and conjugation of �-orbitals in graphene and CNTs.<br />

It has been shown that the form of possible resonance structures depends on<br />

the set of structure parameters (m, n). Metallic tubes with m – n mod 3 = R and<br />

R = 0 (n mod x yields the remainder of the division of n by x) can be described<br />

with a Clar resonance structure consisting only of complete benzene rings. These<br />

are fully conjugated [34]. All other tubes with R = 1 or R = 2 (m – n not divisible<br />

by 3) have at least one seam of isolated double bonds in their resonance structure<br />

(Figure 6.13) [34]. These tubes are semiconducting. A special case are the tubes<br />

with (m, 0) and R = 1. Here it would be possible to construct a chiral Clar resonance<br />

structure with the double bond seam as the chiral element. To avoid this, a structure<br />

is formulated that shows an achiral quinoidal seam (Figure 6.13, right).<br />

345


346 6 Carbon Nanotubes<br />

Figure 6.13 Resonance structures for a (12,9), a (12,8), a (12,7) and a (19,0) nanotube<br />

(from J. Org. Chem. 2004, 69, 4287).<br />

The calculation of NICS (nucleus independent chemical shift) values by Ormsby<br />

and King has shown the validity of the model and the dependence of the electronic<br />

properties on the tube diameter and the graphene sheet orientation [35].<br />

In reality, the nanotubes are not endless and therefore some restrictions apply<br />

to the number of their �-electrons and the options for fully delocalized systems.<br />

To assess the �-electron delocalization, the resonance structures should be<br />

inspected for finite pieces of CNTs. Nakamura showed that the length of the tube<br />

fragment plays an important role in the full delocalization of the �-electrons in the<br />

case of armchair CNTs [36]. Only in the case of a length that accommodates full<br />

aromatic sextets can a complete conjugation be achieved. Otherwise, complete<br />

and incomplete Clar structures best represent the situation in finite CNTs. Interestingly,<br />

the HOMO-LUMO gap and the frontier orbital energy oscillate as well<br />

(Figure 6.14).<br />

Figure 6.14 Resonance structures in finite armchair CNTs.<br />

Functionalization has an important influence on the conjugation of the<br />

�-electrons [37]. Depending on the addition mode of reactants, the conjugation can<br />

be diminished or even completely destroyed. In all cases this has an influence on<br />

the electronic properties of the tube as well. Metallic CNTs become semiconducting<br />

when they are functionalized on the side walls. The conjugation is interrupted<br />

at the position of an addition reaction.


6.2 The Functionalization of Carbon Nanotubes<br />

Similar effects are observed for defects in the side wall. The conjugation is<br />

broken at places where carbon atoms are missing or where the graphene network<br />

is interrupted by the existence of ring defects or functional groups [38].<br />

6.1.5<br />

Conclusions<br />

In summary, the tubular structure of CNTs is unique in the field of highly unsaturated<br />

hydrocarbons. Depending on the structure parameters quite different<br />

properties are observed. Fully conjugated �-systems show metallic conductivity<br />

whereas semiconductivity occurs with isolated double bonds in the resonance<br />

structure of the respective tube. The curvature induces an increased reactivity of<br />

the carbon atoms.<br />

In the next section the reactivity of these highly inspiring molecules will be<br />

discussed. The three-dimensional �-system allows for a variety of useful reactions<br />

and the application of CNTs in various field, e.g electronics, composites, and<br />

biomedical applications.<br />

In order to achieve these high expectations, it will be necessary to produce<br />

homogeneous samples of CNTs with defined electronic properties, i.e. defined<br />

structure parameters. So far only mixtures are available, but significant progress<br />

has been made towards the separation of different types of nanotubes.<br />

6.2<br />

The Functionalization of Carbon Nanotubes<br />

Anke Krueger<br />

6.2.1<br />

Introduction<br />

Functionalization of materials is one of the most important research areas. It<br />

is only after appropriate modification of the pristine material’s properties that<br />

many applications open up for the new compounds or composites. Especially for<br />

biological applications such as drug delivery or sensing applications, a suitable<br />

and stable functionalization is of importance. It is only natural that for carbon<br />

nanotubes (CNT) the research on their surface modification began just shortly<br />

after their production in sufficient amounts. The work has continued ever since<br />

then, as a vast range of reactions has been tested and various new materials have<br />

been developed.<br />

In the following sections we will discuss the different types of surface modification<br />

for CNTs. In this, MWNTs will only be separately mentioned if a significant<br />

difference is observed compared with SWNTs. Normally, the behavior of these two<br />

classes of CNTs is rather similar. Only large MWNTs with decreased curvature show<br />

lower reactivity because there is a smaller prehybridization of the carbon atoms,<br />

i.e. a smaller deviation from sp 2 towards sp 3 hybridization caused by curvature.<br />

347


348 6 Carbon Nanotubes<br />

Figure 6.15 Functionalization types for carbon nanotubes.<br />

The topology of CNTs permits different kinds of modification: reaction at the<br />

tips, at the outer or inner wall or the inclusion of objects in the internal space<br />

(Figure 6.15). Functionalization can be realized by covalent bonds or by noncovalent<br />

inter actions, e.g. electrostatic forces [39].<br />

Compared with the fullerenes, there are some structural differences. In fullerenes<br />

the six-membered rings are partially connected to five-membered rings<br />

inducing three-dimensional curvature and strain, whereas CNTs are not curved<br />

along their tube axis. Additionally, the diameter of the tubes is often (especially in<br />

the case of MWNTs) larger than for the archetypical fullerene C 60 . Therefore, the<br />

reactivity of CNTs should be reduced compared with C 60 , which is actually the case<br />

[40]. Many of the reactions described for fullerenes can be carried out with CNTs<br />

as well but need harsher conditions and/or longer reaction times [41]. Only the<br />

tips of carbon nanotubes exhibit a comparable reactivity because of the existence<br />

of pentagons [42]. A similar effect is observed for the so-called Stone–Wales defect.<br />

They consist of two pentagons and two heptagons (see former Section) and some<br />

of their bonds are much more strained than the usual bonds in CNT [43].<br />

6.2.2<br />

Functionalization of the Nanotube Tips<br />

The first procedures for the functionalization of CNTs were closely related to their<br />

purification. It was observed that reaction with oxidizing acids or piranha water<br />

(mixture of sulfuric acid and hydrogen peroxide) not only opened the nanotube tips<br />

but introduced carboxylic groups at the rims of the cylindrical structures [44].<br />

It is obvious why this reaction preferably takes place at the ends of the tubes.<br />

The higher strain and prehybridization induced by the five-membered rings in<br />

the caps cause a higher reactivity. The carbon is oxidized and the caps removed.<br />

The resulting dangling bonds are saturated with carboxylic groups, which can<br />

further react with, for example, amines or alcohols to form amides or esters [45].<br />

Using long-chain alkyl alcohols or amines for this purpose results in solubility of<br />

the modified tubes in organic solvents (Figure 6.16) [46].<br />

Solubility of the CNTs in reaction or physiological media (important for future<br />

bioapplications) is one of the most challenging and important issues of CNT<br />

research [47]. Certain reactions only take place in solution; homogeneous distribution<br />

is possible and applications in blood or serum depend on stable solutions<br />

of the nanotubes. For nanotubes there are several approaches to this problem:<br />

Addition of surface active compounds [48] or coating of the tube with alkyl chains<br />

[49] or hydrophilic groups [50] result in improved dispersibility.


Figure 6.16 Functionalization of the tips of a CNT.<br />

6.2 The Functionalization of Carbon Nanotubes<br />

Under the reaction conditions that are necessary to remove the caps of CNTs,<br />

a large number of defects is created in the side walls of the tubes [51]. Especially<br />

at the position of already existing small holes, larger defects are produced by<br />

oxidative removal of carbon atoms. The rims of these holes are also covered with<br />

carboxylic groups (Figure 6.17).<br />

Figure 6.17 Defects in the side wall carry carboxylic groups at the rims after oxidative treatment.<br />

Further grafting of e.g. alkyl chains via amidation or esterification leads to improved<br />

dispersibility.<br />

6.2.3<br />

Non-covalent Functionalization of Carbon Nanotubes<br />

As fully unsaturated hydrocarbons, carbon nanotubes are strongly hydrophobic.<br />

They are therefore not dispersible in any polar solvent or water. But they can<br />

undergo strong interactions with other hydrophobic compounds, especially when<br />

these are able to form �–� interactions [52]. Benzene and other small aromatic<br />

compounds do not establish very strong interactions with the nanotube surface.<br />

349


350 6 Carbon Nanotubes<br />

Figure 6.18 Arrangement of pyrene molecules along the nanotube axis (left);<br />

a larger planar aromatic system does not improve the interaction with the nanotube (right).<br />

But condensed aromatic compounds such as pyrene are highly suitable for this<br />

kind of surface modification [53]. Pyrene with its elongated shape is able to<br />

establish �–� interactions throughout its �-system. It can arrange itself with its long<br />

axis parallel to the tube axis in order to maximize the binding force (Figure 6.18).<br />

Bigger planar aromatics are not necessarily more strongly bound because of the<br />

increasing distance between the �-system of the nanotube and the ends of the<br />

aromatic compound (Figure 6.18) [54]. A perfect fit would be achieved with a<br />

curved structure that imitates the curvature of the respective nanotube. In this<br />

case it would be possible to observe selective interactions with the nanotubes of<br />

the best-fitting diameter.<br />

Depending on the functionalization of the pyrenes that are interacting with<br />

the nanotubes it is possible to tune the hydrophilicity of the resulting composite<br />

resulting ultimately in water solubility (Figure 6.19) [55], debundling [56] or solubility<br />

in organic solvents if the pyrenes carry long side chains [57]. At the same<br />

time they influence the band structure of the functionalized nanotube as can be<br />

seen from the resulting absorption spectra (Figure 6.20) [58]. Polymers bearing<br />

pyrenes in their side chains can be mixed with nanotubes to yield a non-covalently<br />

bound composite material that unites the properties of the nanotubes with those<br />

of the polymer [59].<br />

Figure 6.19 Functionalization of a CNT with pyrene derivatives. The pyrene moieties are<br />

arranged along the tube axis. Depending on the side chain the conjugates are dispersible in<br />

different media.


6.2 The Functionalization of Carbon Nanotubes<br />

Figure 6.20 Absorption spectra of CNTs before and after functionalization with DomP<br />

(from J. Am. Chem. Soc. 2004, 126, 10234).<br />

Similar to pyrenes, phthalocyanins and porphyrins interact with the CNT surface<br />

via van der Waals interactions and �-stacking [60]. They can be used to study the<br />

energy transfer from the porphyrin skeleton to the nanotube. In 2004 Sun et al.<br />

reported on the selective interaction of semiconducting nanotubes while the<br />

metallic ones remained unchanged and insoluble [61].<br />

Not only organic molecules can be deposited on the nanotube surface but also<br />

metallic clusters. Usually, this requires the addition of a reducing agent (except<br />

for electrodeposition) and a thorough surface cleaning of the nanotube surface<br />

[62]. Deposited clusters include gold, platinum and palladium nanoparticles,<br />

zinc and magnesium oxide [63]. These composites are interesting for catalytic<br />

applications.<br />

Another topologically completely different approach is taken with long, chainlike<br />

molecules, in general polymeric materials. These can be wrapped around<br />

single nanotubes or small bundles [64]. Usually the interaction is based on electrostatic<br />

forces or �-stacking. Examples (Figure 6.21) for this functionalization method<br />

include the wrapping of CNTs with amylose, peptides, polyaniline and polymers<br />

Figure 6.21 Non-covalent wrapping of a CNT with hydrophilic amylose (lower left) or<br />

hydrophobic PmPV (lower right).<br />

351


352 6 Carbon Nanotubes<br />

such as PVP (polyvinylpyrrolidone) or PmPV (poly-m-phenylenevinylidene) [65].<br />

The latter produces interesting composites with CNTs where the electrical conductivity<br />

of the composite is significantly increased whereas the luminescence<br />

properties of PmPV remain basically unchanged [66]. The electrical conductivity<br />

is also increased in polyaniline (PANI) composites [67]. The circular and sizeselective<br />

complexation of CNTs is achieved with cyclodextrins [68, 69].<br />

6.2.4<br />

Covalent Side-wall Functionalization of Carbon Nanotubes<br />

Instead of using electrostatic interactions and �-stacking, CNTs can also be functionalized<br />

covalently at their side-walls. The conceptually easiest way of modifying<br />

the surface structure of a CNT consists in reducing the number of unsaturated<br />

bonds, and therefore its hydrogenation (Figure 6.22). So far, a completely hydrogenated<br />

CNT has not been reported. Only partial reduction was observed<br />

after reaction in a hydrogen plasma or with lithium in liquid ammonia (Birch<br />

type reaction), after high temperature treatment in a hydrogen atmosphere or<br />

electrochemical hydrogenation [70, 71]. According to calculations a completely<br />

hydrogenated CNT should be stable up to a diameter of 1.25 nm [72].<br />

Figure 6.22 Hydrogenation of CNTs.<br />

In close resemblance to the hydrogenation, nanotubes can also be halogenated.<br />

Especially the reaction with fluorine yields samples with a high halogen content<br />

of up to 100% [73]. This does not mean that each carbon is carrying exactly one<br />

fluorine atom, but there are positions like the tips and defects where the fluorination<br />

grade is elevated. The reaction is reversible; treatment with hydrazine yields<br />

the clean nanotubes after annealing [74]. The fluorination reaction was used to<br />

study the addition mechanism of the addition to CNTs. Usually, the fluorine atoms<br />

are positioned close to each other and further fluorination continues along the<br />

circumference, not the axis of the tube [75]. This results in a nonregular distribution<br />

of highly fluorinated areas and some with almost no fluorine. One explanation<br />

for this behavior is the preferred 1,4-addition (over 1,2-addition), which is also<br />

corroborated by computational results (Figure 6.23) [76]. The electronic properties<br />

of fluorinated CNT as very good insulators strengthen the argumentation,<br />

too [77]. The 1,4-addition destroys the conjugation of the �-bonds over the whole<br />

diameter. Fluorinated CNTs can be used in further reactions such as nucleophilic<br />

substitutions with amines or carbanions [78]. Other halogens are much less<br />

reactive towards CNTs. Only an electrochemical method for the halogenation<br />

has been described [79].


Figure 6.23 Fluorination of CNTs (1,4- vs. 1,2 reaction).<br />

6.2 The Functionalization of Carbon Nanotubes<br />

Figure 6.24 Bingel reaction on CNTs and further modification of the ester groups.<br />

Another important reaction of CNTs is the nucleophilic attack by bromomalonates<br />

in the presence of a base (Figure 6.24). This so-called Bingel reaction is<br />

one of the most versatile reactions in fullerene chemistry and can be applied<br />

to CNTs, although only under harsher conditions [80]. The ester groups of the<br />

malonate can be modified in various ways, enabling a vast range of further CNT<br />

derivatives and improved solubility [81, 82].<br />

In general, the reaction with nucleophiles as described for fullerenes can be<br />

transferred to CNTs (e.g. addition of carbenes and nitrenes [83]), but the lower<br />

reactivity in some cases remains an obstacle to efficient functionalization.<br />

Radical reactions are another option for the surface grafting of organic moieties.<br />

Perfluoroalkyl radicals as well as acyloyls radical have been used for the modification<br />

of the nanotubes surface [84]. The reaction with diazonium salts is a very<br />

important reaction of CNTs. The intermediate phenyl radicals directly bind to the<br />

tube by C–C bonds [85]. Strano et al. reported on the selective functionalization<br />

of metallic tubes with this reaction [86].<br />

CNTs are good candidates for cycloaddition reactions because they have a system<br />

of conjugated double bonds (Figure 6.25). They always react as the -ene component,<br />

even in [4+2] cycloadditions [87]. It is rather unlikely that a reaction as the diene<br />

component would be observed, because the nanotube has unfavorable geometry<br />

and also is electron deficient [88]. The most common cycloaddition on CNTs is<br />

353


354 6 Carbon Nanotubes<br />

Figure 6.25 Cycloaddition reactions of CNTs.<br />

the Prato reaction, that has also been extensively studied in the case of fullerenes<br />

[89]. This 1,3-dipolar cyclodaddition of azomethineylides yields functionalized<br />

pyrrolidine rings on the nanotube surface. This reaction accepts a broad range<br />

of functional groups in the sidechains. That is why this reaction is often used to<br />

immobilize biological structures like DNA, antibodies or viruses on CNTs for applications<br />

such as drug or antibody delivery [90]. Solubility in physiological media<br />

is achieved with bridging oligoethylenglycol moieties [91].<br />

Additionally, the reactions with nitrile imines and ozone are [3+2] cycloadditions.<br />

The latter can be used to establish carboxylic, keto or hydroxylic groups at<br />

the nanotube surface depending on the selected work-up [92].<br />

Compared to [3+2] cycloadditions, the [4+2] addition has rarely been reported.<br />

Only a few examples using o-chinodimethanes have been described [93]. Fluorinated<br />

CNTs are better dienophiles because of the strain induced by the sp 3<br />

carbon atoms that exist in the neighborhood of the remaining double bonds and<br />

the electron-withdrawing properties of the fluorine substituents [94]. [2+2] cycloadditions<br />

to CNTs have not been reported so far.


6.2 The Functionalization of Carbon Nanotubes<br />

Figure 6.26 Transition metal complexes with CNTs have so far been reported only via functional<br />

groups (right). Dihapto and hexahapto complexes are difficult to achieve due to the weak donor<br />

character of the tubes and the unfavorable geometry.<br />

Altogether the cycloaddition chemistry of CNTs is less pronounced than for<br />

fullerenes. The lower curvature and the lower electrophilicity play the key roles<br />

for the lower reactivity.<br />

An interesting question is whether CNTs are able to form coordination<br />

complexes with transition metals. Their double bonds could act as �-donors in<br />

these compounds. So far, only complexes via functional groups of the nanotube<br />

have been described, such as the CNT–Wilkinson complex conjugate (Figure 6.26)<br />

[95]. These derivatives are interesting objects for catalyst research. The reasons for<br />

the inability to complex the metal centers directly at the nanotube are the better<br />

conjugation of the double bonds, the absence of five-membered rings and the<br />

larger HOMO-LUMO gap compared with C 60 [96].<br />

6.2.5<br />

Endohedral Functionalization of Carbon Nanotubes<br />

After the removal of the caps from the nanotube tips the inner space becomes<br />

accessible. Many reports on the endohedral functionalization of carbon nanotubes<br />

have been published [97]. Most of these results have been obtained for SWNTs.<br />

Small molecules as well as metals [98] or metal oxides [99] can be included in<br />

the tube.<br />

Besides, organic molecules can be filled into CNTs. These include large<br />

molecules such as polypyrrole or carotene [100] and small structures like metallocenes<br />

[101] and aromatic compounds [102, 103]. A very special example for the<br />

endohedral functionalization of SWNTs is the inclusion of fullerenes inside CNTs<br />

[104]. The resulting material is a ‘carbon-only’ compound with a tube-like shell<br />

and dot-shaped inclusions.<br />

For a long time it was believed that covalent functionalization of the nanotube<br />

wall is prevented by the small orbital lobes inside the tube. However, Gao et al.<br />

reported on the grafting of amino groups inside CNTs [105].<br />

355


356 6 Carbon Nanotubes<br />

6.2.6<br />

Conclusions<br />

In summary, the chemistry of carbon nanotubes is a vast research area. There are<br />

many options for the modification of both SWNTs and MWNTs. These include the<br />

removal of the caps with subsequent formation of carboxylic groups, the endohedral<br />

inclusion of small clusters and molecules, and the side-wall functionalization<br />

at defects or regular double bonds. This can be done with surface active agents<br />

leading to non-covalent composites of nanotubes with organic molecules. Metal<br />

clusters can be deposited on the CNT surface and wrapping with macromolecular<br />

compounds results in better dispersibility of the tubes. In the case of SWNTs this<br />

can even break up the bundles of tubes. Covalent surface modification can make<br />

use of existing functional groups or directly use the conjugated double bonds of the<br />

graphene lattice. Cycloadditions, radical additions of diazonium salt etc. and the<br />

Bingel reaction are examples for the efficient modification of the carbon nanotube<br />

surface. All these reactions lead to new materials with interesting properties, e.g.<br />

the selective functionalization of metallic or semiconducting tubes. For new applications<br />

in composites or electronics these features are very useful.<br />

6.3<br />

Applications of Carbon Nanotubes<br />

Marc Monthioux<br />

6.3.1<br />

Introduction<br />

Carbon nanofilaments – or nanotubes – were actually very probably the core of<br />

the much larger vapor-grown carbon filaments used by Edison to operate the<br />

early version of his light-bulb. However, although carbon nanotubes (CNTs) were<br />

unambiguously revealed as early as 1952 [1], and shown to exhibit diameters as<br />

low as 3–4 nm as early as 1976 [1], they have only been thought to be useful for a<br />

wide range of applications from the 1990s, i.e. from the time of the well-known<br />

paper by Iijima [1]. In this regard, CNTs are a good example of availability being<br />

not the only condition for a material to be considered of some interest from the<br />

technological point of view. Both technology and the minds (of scientists and<br />

engineers) have to be prepared, specifically when downsizing is involved [106].<br />

Fifty years ago, vapor-grown carbon (nano)filaments were mostly undesirable<br />

by-products of industrial processing that were studied to be avoided. Nanotubes<br />

are nowadays the most popular nanomaterials, as measured from the number of<br />

scientific articles published yearly, and from the amount of investment dedicated<br />

to related R&D. Consequently, applications involving CNTs are now coming to the<br />

market although they are probably still too few with respect to the expectations.<br />

This section will briefly address these aspects, along with providing examples of<br />

applications.


6.3.2<br />

Properties of CNTs<br />

6.3 Applications of Carbon Nanotubes<br />

6.3.2.1 Which CNT for Which Application?<br />

CNTs were considered very promising because they exhibit extreme properties<br />

in almost every aspect. Table 6.3 provides the most important ones, as gathered<br />

from literature.<br />

As with other kinds of graphene-based carbon materials (otherwise called polyaromatic<br />

carbons, including graphite), the properties of CNTs are determined<br />

by the level of anisotropy, i.e. the extent to which the graphenes are aligned with<br />

respect to the direction at which the property is measured, because graphene and<br />

graphene stacks exhibit high intrinsic anisotropy (Table 6.4).<br />

Table 6.3 Some extreme properties of CNTs.<br />

Property Values Comments<br />

Aspect ratio ~1000 Diameter > 0.4 nm, length up to cm<br />

Specific surface area ~2700 m 2 g –1<br />

The highest!<br />

Tensile strength > 45 GPa The highest ever!<br />

Tensile modulus 1–1.3 TPa The highest!<br />

Tensile strain > 40% Provides the highest toughness ever!<br />

Thermal stability > 3000 °C In oxygen-free atmosphere<br />

Electrical conductivity 10 4 –10 7 S cm –1<br />

Better than that of copper<br />

Transport regime Ballistic Induced superconductivity with T c < 1 K<br />

Thermal conductivity ~6000 W mK –1<br />

Electron emission 10 6 –10 12 A cm –2<br />

Higher than those of diamond or graphite<br />

Highest current density!<br />

Table 6.4 Properties for a stack of perfect graphenes (except the tensile strength-to-failure,<br />

calculated for a single graphene) as measured parallel or perpendicular to the graphene plane<br />

direction.<br />

Direction in graphene (stack) Parallel Perpendicular<br />

Bond energy kJ mole –1<br />

Thermal expansion °C –1<br />

Thermal conductivity W (m · K) –1<br />

524 7<br />

1.5 · 10 –6<br />

Electrical resistivity � cm 3.8 · 10 –5<br />

3000 8<br />

Young modulus GPa > 1000 ~50<br />

Strength-to-failure GPa ~100 ?<br />

27 · 10 –6<br />

10 –2<br />

357


358 6 Carbon Nanotubes<br />

Hence, properties are maximized along the axis of single-wall CNTs (SWNTs)<br />

because anisotropy is at maximum (or ‘assumes maximum value’) on the one<br />

hand, and SWNT structure is close to perfection on the other hand, with only<br />

pentagons or heptagons (see the preceding section) as possible structural defects<br />

(vacancies are minor in pristine SWNTs). The fact that SWNTs can be as perfect<br />

as a (macro)molecule makes a major difference with regular materials.<br />

The situation is quite different for multi-wall CNTs (MWNTs), in which all<br />

kinds of orientation of graphene stacks can be found with respect to the nanofilament<br />

axis, from fully parallel (as in MWNTs with concentric texture) to fully<br />

perpendicular (as in nanofibers with platelet texture), including combinations of<br />

both (as in MWNTs with bamboo texture). In addition, they may be affected by<br />

a much larger variety of possible structural defects, such as sp 3 carbons, heteroatoms,<br />

disclinations, in addition to other than six-membered rings as in SWNTs,<br />

making graphenes in graphene stacks and the mutual orientation of the latter<br />

more or less perfect. Hence, a much wider range of property values can be found<br />

for MWNTs.<br />

An interesting property of CNTs is their intrinsic biocompatibility, which is<br />

of importance not only for medical applications but also for safety issues. Such<br />

biocompatibility has long been known for pyrolytic carbon or graphite, because<br />

they contain a limited amount of accessible graphene edges, which actually are<br />

the reactive sites, as opposed to the graphene plane. In this regard, although the<br />

reactivity of SWNTs or concentric MWNTs is low (for being free of graphene<br />

edges), using high purity materials is compulsory, otherwise carbon or noncarbon<br />

impurities will affect the reactivity, and hence the biocompatibility, of the<br />

whole.<br />

These brief comments explain why research efforts regarding the synthesis<br />

of nanotubes ultimately aim at high-yield production of SWNTs, preferably to<br />

MWNTs, although production capacity is still much larger for the latter so far. It<br />

also illustrates that, considering carbon nanotube properties (physical as well as<br />

chemical, including reactivity), it is of utmost importance to define which kind<br />

of nanotubes is involved.<br />

6.3.2.2 Why is ‘Nano’ Beautiful?<br />

As mentioned before, extreme properties of CNTs come from their high anisotropy.<br />

But other highly anisotropic polyaromatic carbon materials, carbon fibers, known<br />

for long time, also exhibit extreme properties (sometimes close to those of CNTs,<br />

e.g. Young’s modulus, thermal conductivity, etc.). They have already found many<br />

applications, from daily-life ones, such as sports goods, to high-tech uses in space<br />

shuttles. One might then wonder whether it is worth investing millions of dollars<br />

worldwide in the R&D of CNTs. Reasons for this are multiple:<br />

� CNTs exhibit a much higher specific surface area, thus a much higher surface<br />

of interaction, than carbon fibers. This is especially useful not only for composites<br />

but also for applications where the specific reactivity of CNTs is exploited<br />

(e.g. sensors).


6.3 Applications of Carbon Nanotubes<br />

� CNTs exhibit a much higher aspect ratio than carbon fibers. This feature is very<br />

favorable for applications where they are used as a network, e.g. as fillers in electrically<br />

or thermally conductive composites built with an insulating matrix (e.g.<br />

most of polymers). It is actually well-known that the percolation threshold (i.e.<br />

the proportion of filler from which the conductivity suddenly jumps) decreases<br />

as the aspect ratio increases.<br />

� CNTs have a much better propensity to structural perfection than carbon<br />

fibers, especially when considering SWNTs or possibly concentric MWNTs.<br />

This provides some of the as-prepared CNTs with ultimate properties, as<br />

opposed to carbon fibers which contain numerous and various kinds of defects<br />

(Figure 6.27).<br />

� Rather simple, mostly single step, fabrication processes can be used to prepare<br />

CNTs whereas carbon fibers require complex, multi-step processes. This makes<br />

the synthesis of CNTs a much more affordable technology, regarding both the<br />

needed equipment and the level of manpower qualification.<br />

� Contrary to the reagents used in production of carbon fibers, feedstocks for<br />

carbon nanotube synthesis are basic compounds with perfectly defined chemical<br />

composition, easy availability, and constant quality (CH4 , C2H2 , CO, graphite,<br />

pure metal catalysts, …).<br />

Whereas the first three reasons relate to the intrinsic superiority of CNTs over<br />

carbon fibers, the last two relate to the superiority of their synthesis processes,<br />

which are likely to ultimately result in much cheaper materials.<br />

Figure 6.27 Tensile mechanical properties of the fibers available on the market, among the<br />

most performing ones. Circles and triangles are carbon fiber, polyacrilonitrile-based and<br />

pitch-based, respectively. Only carbon nanotubes (SWNT) are able to exhibit both high modulus<br />

(1–1.2 TPa) and high strength (45 GPa or more). (Modified from [107]).<br />

359


360 6 Carbon Nanotubes<br />

6.3.2.3 Potential Problems Related to the Use of CNTs<br />

However, there are obviously still unsolved problems that prevent CNTs from<br />

overwhelming the market:<br />

� The most important one is quality. Many marketed CNTs exhibit low purity,<br />

since they come along with other carbon nanophases (amorphous carbon,<br />

fullerenoids, polyaromatic carbon shells, nano-horns …) and catalyst and/or<br />

solvent remnants (Figure 6.28). Enriching in CNT content is a technological<br />

challenge because purification processes based on chemicals are obviously more<br />

or less harmful to the CNT structure, because the CNTs are chemically similar<br />

to some of the various phases to be discarded.<br />

Likewise, all marketed nanotube grades exhibit poor selectivity, e.g. regarding<br />

the helicity of SWNTs (which controls the metallic vs semi-conductor behavior),<br />

CNT diameters, number of walls in MWNTs … (Figure 6.29).<br />

� Processing is another major issue because the aspect ratio of nanotubes makes<br />

them entangle (except for the carpet-type growth mode, where CNTs are aligned<br />

and perpendicular to the substrate). Meanwhile, CNT nanosize makes them<br />

sensitive to weak surface forces and clump upon any attempt to mix and disperse<br />

them into liquids (e.g. solvents, matrix precursors) using regular mixing procedures<br />

(e.g. see Figure 6.30, left). This is highly detrimental to the properties of<br />

the resulting composite materials. Various solutions are however being studied<br />

(e.g. see Figure 6.30, right).<br />

� The low surface reactivity of CNTs causes a problem in some cases, e.g. when<br />

strong nanotube/matrix interactions are needed as in composites for structural<br />

applications (in which stress transfer from the matrix to the nanotubes would<br />

then be minimized), or when there will be a need to interconnect them, (e.g.<br />

when used as electronic components). CNT functionalization discussed in the<br />

preceding section is a way to overcome this problem.<br />

� Cost is still a major issue, as for any recently developed material. Marketed prices<br />

are in the range 80–2000 $/g for SWNTs, and higher than 0.05 $/g for MWNTs,<br />

depending on the synthesis process, number of post-treatments, and resulting<br />

purity, structural quality, etc. Obviously, this is by far too high for incorporating<br />

most of nanotube grades in any mass application. But prices tend to get lower<br />

every year, and were already divided by 2 to 4 within the past decade.<br />

� Health and safety issues are nowadays a major concern for any nanosized<br />

material. The toxicity of CNTs is supposed to increase as the CNT aspect ratio,<br />

specific surface area, and surface reactivity (when any, e.g. for functionalized<br />

nanotubes) increase. But results in literature are still contradictory [110], and<br />

there is no certainty yet regarding the actual cyto- and eco-toxicity of CNTs.<br />

A significant reason is the very large variety of CNTs on the one hand, and<br />

the relatively poor characterization of the nanotubes investigated on the other<br />

hand, which make results barely comparable and understandable. Another<br />

significant reason is the absence of standard procedures for the evaluation of<br />

toxicity. Investigations are in progress worldwide in this regard.


6.3 Applications of Carbon Nanotubes<br />

Figure 6.28 Example of the typical<br />

aspect of an as-prepared nanotube<br />

material from the arc discharge<br />

process (from [108]). Round morphologies<br />

with dark contrast are metal<br />

catalyst remnants. Every other phase<br />

is carbon-based.<br />

Figure 6.29 Example from purified<br />

CNTs, formerly prepared via catalystenhanced<br />

chemical vapour deposition<br />

(CCVD). Even rather clean, nanotubes<br />

exhibit discrepancies in diameter,<br />

and number of walls (one or two).<br />

Figure 6.30 In spite of the addition of a surfactant such as sodium dodecyl sulfate (SDS), the<br />

dispersion of as-prepared SWNTs (from arc discharge) is poor and forms agglomerates visible<br />

with an optical microscope (left). If the material is pre-treated with freeze-drying, agglomeration<br />

occurs only at a much lower scale, invisible to the optical microscope (right) (from [109]).<br />

361


362 6 Carbon Nanotubes<br />

6.3.3<br />

Applications of CNTs<br />

Applications of nanotubes are currently at various levels of development. Considering<br />

the variety and excellence of the physical properties of nanotubes, areas<br />

in which they can be applied are too numerous to be listed here (Figure 6.31).<br />

Hence, only examples will be provided, including (1) prospective applications;<br />

(2) applications under development; and (3) applications already on the market.<br />

Figure 6.31 The ‘application-tree’ for CNTs. Not exhaustive (by courtesy of M. Endo, modified<br />

from original).<br />

6.3.3.1 Prospective Applications<br />

Applications from this category still have to face one or several drawbacks from<br />

those listed in Section 6.3.2.3 (beside cost) to such an extent which severely – yet not<br />

definitively – prevents their development. They are therefore still in the research<br />

laboratories, although examples of prototypes can be found, mostly fabricated by<br />

companies for high-tech advertising purposes rather than for real advances in<br />

performance with respect to the existing technology (e.g. sport goods).


6.3 Applications of Carbon Nanotubes<br />

Structural composites. While the benefit of CNTs for tribological applications is<br />

still being investigated (for which both the mechanical and thermal performances<br />

are likely to be useful), the benefit expected from incorporating nanotubes into<br />

various matrices to make composites for structural applications appears obvious<br />

from considering Figure 6.27. Difficulty regarding processing (dispersion) is the<br />

main problem still to be solved, as well as the low reactivity that prevents good<br />

bonding to the matrix. Consequently, even if CNTs (e.g. SWNT, and concentric-<br />

MWNT types) are actually known to reinforce matrices efficiently [111, 112], the<br />

resulting performances still hardly beat those of composites reinforced with regular<br />

carbon fibers available on market. Cost and safety are also issues that contribute to<br />

delaying the development. Meanwhile, various ways to make nanotube-based fibers<br />

are being investigated: (1) as a filler to fiber precursors at soften state (polymer<br />

[e.g. 113] or pitch [114]), (2) as a dry-spun [115] or wet-spun [116] nanotube fiber<br />

from as-prepared nanotubes. The interest of making micrometer-sized fibers out<br />

of nanotubes is to allow weaving fabrics, that can be put in composites or to be<br />

used as such.<br />

Once the pending problems are solved, nanotube-containing composites will<br />

lead to mass applications, starting with sport goods (e.g. golf club shafts, tennis<br />

racket frame, fishing rods) then structural parts for transportation vehicles of all<br />

kinds, from bicycle to space shuttle, and many other areas. Fabrics prepared with<br />

nanotube-based fibers (i.e. with no matrix involved) will be also used for functional<br />

clothes (e.g. bullet-proof vest), cables, stain-resistant textiles, etc.<br />

Electronic parts. Thanks to the amazing versatility of electronic behavior of SWNTs,<br />

they could be components of miniaturized computers, or, in a more distant future,<br />

in ultra-fast quantum computers, e.g. as [117]: (1) connectors, using metallic-type<br />

SWNTs, with a possibility of building 3D circuitry for better compactness; (2) transistors,<br />

using semi-conducting SWNTs (ten times faster than regular metal-oxidesemiconductor<br />

transistor); (3) diodes, using combined metallic/semi-conductor<br />

SWNT junctions via pentagons and heptagons; (4) superconducting transistors<br />

[118]. A common advantage of such carbon-based nano-components for electronics<br />

is low energy consumption and their ability to withstand high temperatures, thus<br />

not needing cooling systems. Most current obstacles to development are the lack<br />

of sources for SWNTs with specific transport behavior, the quality and feasibility<br />

of contacts, and more generally the technical difficulty to build electronic circuits<br />

with nanosized components.<br />

Actuators. Based on the finding that the lattice constants of graphene expand differently<br />

when doped with anions and when doped with cations, the idea of using<br />

doped nanotube-based fibers as current-powered actuators was proposed in 1999<br />

[119]. Making artificial muscles is expected, thanks to both the biocompatibility<br />

and the unprecedented mechanical properties of CNTs. Such artificial muscles<br />

could be used for repairing surgery (and robotics), if internally implanted, but<br />

they also could be used for military purpose and hard structures, if supported by<br />

an external frame providing extra strength to anyone wearing it as some kind of<br />

363


364 6 Carbon Nanotubes<br />

suit. Problems are many and studies are in progress, in particular at the University<br />

of Texas (R. Baughman’s group) and in France (P. Poulin’s group). Voltagepowered<br />

nano- or micro-scaled tools, although not necessarily requiring previous<br />

ion-doping, such as micro-tweezers [120], have also been proposed. Related applications<br />

are mechanical memory elements (as developed by NANTERO), and<br />

nanoscale electric motors.<br />

Membranes and low-friction surfaces. Filtering membranes can be made which<br />

are able to let water molecules go through the inner channels while large ions<br />

are blocked. Thanks to the perfect smoothness of the graphene surface, high<br />

speed transfer of fluids throughout the inner cavity of nanotubes, five orders of<br />

magnitude higher than predicted by theory, was recently demonstrated [121].<br />

Likewise, again thanks to the low friction, but also to the low reactivity and thermal<br />

resistance, slick surfaces (slicker than Teflon) can be envisaged.<br />

6.3.3.2 Applications Under Development<br />

Applications from this category are being developed by profit-based companies,<br />

which have invested enough to prepare prototypes that exhibit one or several<br />

benefits with respect to existing materials and devices. Reaching the market<br />

is however still prevented by one or several drawbacks from those listed in<br />

Section 6.3.2.3, generally including cost issues.<br />

Conductive composites (part 1). They mostly correspond to any application in<br />

which static electricity is a problem [112]. Thanks to the huge aspect ratio of<br />

CNTs (in the range of 1000), loading an insulating polymer matrix with them<br />

allows reaching the percolation threshold with a CNT content as low as 1 wt%<br />

or less (depending on the type of CNTs used), as opposed to ~15 wt% for regular<br />

nanosized carbon fillers such as carbon blacks. One of the benefits is that the<br />

color of the resulting composite is not affected, as opposed to becoming black, and<br />

even photo-transparent polymers remain so. The development of nanotube-loaded<br />

conductive composites is less problematic than that of structural composites<br />

because there is no need for a specific bonding with the matrix. Consequently,<br />

transparent films are prepared as an alternative to regular indium tin oxide films<br />

to make flexible polymer-based transistors, touch-screens, and displays (Eiko,<br />

Unidym). Good thermal conductivity brought by incorporating nanotubes into<br />

materials is also of use. Nanotubes loaded polymer composites are also planned<br />

to be applied in tanks dedicated to the storage of spark-sensitive, explosive or inflammable<br />

compounds (Arkema). Other related applications are electro-magnetic<br />

shielding, radar-absorbing materials, electric motor brushes, heat exchangers and<br />

dissipaters, etc.<br />

Bio-related materials. Other nanotube-loaded polymer-based manufactured materials<br />

are also under development, yet not with the goal to make them electrically<br />

conductive but biocompatible. Polymer-based surgery wires (catheters) were successfully<br />

tested as an alternative to currently used animal-derived wires (‘catguts’).


6.3 Applications of Carbon Nanotubes<br />

Prosthetic coatings are also considered because cells were shown not to adhere<br />

to (some of) nanotubes, on the one hand, while graphene surface has a very low<br />

friction coefficient on the other hand. The same properties made them applicable<br />

for anti-fouling coatings for boat hulls. Likewise, water and air filtration devices<br />

made from nanotubes are supposed to be more resistant to fungus and bacteria<br />

colonization.<br />

Batteries and supercapacitors. Appropriate features for energy storage device electrodes<br />

are both a high amount of mesoporosity (allowing for an easy circulation<br />

of the electrolyte) and a high amount of microporosity (presenting a high surface<br />

area of charge exchange to the electrolyte which is required for fast current delivery<br />

and high speed charging), as well as good electrical conductivity. Appropriate pore<br />

size distribution was actually previously achieved using other kinds of carbon<br />

materials. However, the preparation of the latter involved chemical treatments<br />

(activation) detrimental to the material conductivity. Based on the early work by<br />

Niu et al. [122], CNTs were found to be able to intrinsically exhibit all the required<br />

features while requiring limited pre-treatments. Development is in progress [123],<br />

but nanotubes-based supercapacitors (with capacitance > 300 F/g, exhibiting faster<br />

speed charging with respect to other carbon-based supercapacitors) are now built,<br />

e.g. by Montena) as well as fuel cell with nanotube electrodes (Nec, Motorola,<br />

Intematix). Recently, a foldable, postage-stamp-sized supercapacitor with a voltage<br />

of ~2.5 V was developed from a nanotube reinforced cellulose paper [124].<br />

Support for catalysts. Based on a pioneering demonstration by Planeix and<br />

coworkers [125], nanotubes were also found to be good catalyst supports for catalyst<br />

particles for heavy chemical industrial process, with promises for needs of mass<br />

production (Arkema). High surface area, low chemical reactivity, large mesopore<br />

network, and tailorable surface energetics (for retaining catalyst nanoparticles<br />

and preventing them from coalescence) are among the suitable properties of<br />

nanotubes for this application. With respect to regular ceramic-based catalyst<br />

supports (alumina, typically), CNTs also exhibit much higher mechanical strength,<br />

which goes with higher durability.<br />

Chemical sensors. Based on pioneering works published in 2000 [126, 127] CNTs<br />

can be used as chemical sensors, either under the form of a single SWNT (semiconductor<br />

type), or a paper-like SWNT network. Because electron transport in<br />

CNTs is a surface phenomenon, the conductance response is affected by the occurrence<br />

of molecule(s) adsorbed onto the nanotube surface. More interestingly, the<br />

response can be specific since it differs depending on the nature of the molecule<br />

adsorbed. Highly sensitive (sometimes in the range of ppm) chemical sensors are<br />

therefore currently developed (Nanomix, Motorola) for use in atmosphere control<br />

for instance. Among other companies involved are Applied Nanotech, Carbon<br />

Nanoprobes, Honeywell International, Nanosensors, etc.<br />

365


366 6 Carbon Nanotubes<br />

Electron emitters (part 1). Thanks to their ability to withstand enormous current<br />

density (see Table 6.3), it was proposed as early as 1995 [128] that CNTs could be<br />

used as powerful field emitters [128, 129]. Alternatively, they are able to carry the<br />

same current density as regular emitters (e.g. in Mo or Si) but with a much lower<br />

extraction voltage threshold, typically of several orders of magnitude. Depending<br />

on the target onto which the emitted electrons are projected, various applications<br />

are proposed. If projected onto a phosphor layer, nanotube-based emitter arrays can<br />

serve for low consumption displays for TV sets or computers, or high autonomy<br />

portable devices (e.g. cell-phones). Samsung was first to develop a prototype of TV,<br />

then Motorola, Futaba Corp., Copytele Inc., etc. However, it seems that the fact<br />

that no nanotube-based flat screen TV has yet been marketed, while prototypes<br />

were shown several years ago, could be because its life-time is still insufficient.<br />

On the other hand, instead of considering a periodic array of nanotube emitters,<br />

a single nanotube-based tip can be used as a field emission electron source<br />

with high coherence and brightness for electron microscopes [129], as currently<br />

developed by FEI).<br />

6.3.3.3 Applications on the Market<br />

Applications from this category are those available on market or used by companies<br />

to process manufactured products. There are still few, and considering arguments<br />

given in Section 6.3.2.3, no wonder if they correspond to applications with low<br />

requirements regarding purity and selectivity, and/or with low requirement<br />

regarding quantity.<br />

Electron emitters (part 2). If the target is a metal cathode (Cu, Mo), the very<br />

high current density carried by a single nanotube tip allows building portable<br />

X-ray generators. As opposed to electron sources for electron microscopes (see<br />

Section 6.3.3.2), ultimate, expensive technology (e.g. ultra-high vacuum) is not<br />

required in that case, which has allowed e.g. Oxford Instruments to put such an<br />

application on the market.<br />

Near-field microscopy probes. In 1996 Dai et al. [130] were first to propose to use<br />

a single carbon nanotube as a tip for atomic force microscopy, with several advantages<br />

with respect to regular ceramic tips (Si, Si 3 N 4 ), typically a much higher<br />

mechanical resistance, providing a longer life-time, a much higher aspect ratio,<br />

and a better lateral resolution, and minimizing tip artefacts. Benefits are high<br />

enough to compensate still prohibitive prices, in the range of several hundred<br />

dollars per tip (Nanoscience Instruments).<br />

Conductive composites (part 2). An application of nanotube-loaded conductive polymers<br />

that can be considered on the market is electro-painting (in which applying an<br />

electric potential difference between the painting nozzle and the piece to be painted<br />

prevents electrostatic repulsion effects and allows for a good adhesion of the paint),<br />

since General Motors has already applied it to some car parts. Electro-painting of<br />

polymer-based parts is also proposed by Hyperion for a good surface finish.


6.3.4<br />

Conclusions<br />

6.3 Applications of Carbon Nanotubes<br />

CNTs dominate R&D in the field of nanotechnology. This is fully justified by their<br />

large panel of amazing properties. On the other hand, only a few applications are<br />

on the market so far, even though about 15 years have elapsed from the moment<br />

when the scientists realized the full potential of nanotubes. This has been enough<br />

to inspire recent pessimistic comments regarding the ultimate usefulness of<br />

nanotubes that would finally not fulfill the expectations regarding their ubiquitous<br />

applicability. Such doubts are also inspired by the previous disappointment related<br />

to fullerenes, whose discovery in 1985 was acknowledged with a Nobel Prize in<br />

1996, which have generated millions of dollars investments in research programs<br />

worldwide, but have finally resulted in only a few examples of applications.<br />

Such an analysis is misleading because it ignores the major difference between<br />

fullerenes and nanotubes, which is the high morphological, textural, nanotextural,<br />

and structural versatility of the latter with respect to the former. It also ignores<br />

the time usually needed to bring the idea to the market in the field of advanced<br />

materials. In this regard, the example of carbon fibers teaches us that CNTs are<br />

just following the regular path (Figure 6.32).<br />

Carbon fibers were actually invented in the 1960s, were first used in sports goods<br />

about 12 years later, then started being used in aircraft and space industries after<br />

10 more years. Finally, it took about 40 years in total for the market to become<br />

mature and demanding for large quantities, synonymous of large profits. CNTs<br />

do not behave differently, and they still have the potential to ultimately dramatically<br />

influence our daily life.<br />

Figure 6.32 Evolution of the world market for carbon fibers (by courtesy of M. Endo).<br />

367


368 6 Carbon Nanotubes<br />

References<br />

1 A. Oberlin, M. Endo, T. Koyama,<br />

J. Cryst. Growth 1976, 32, 335; S. Iijima,<br />

Nature 1991, 354, 56; earlier reports<br />

on tubular filamentous carbon include<br />

L. V. Radushkevich, V. M. Lukyanovich,<br />

Zhurn. Fis. Khim. 1952, 26, 88.<br />

2 S. Iijima, T. Ichihashi, Nature 1993, 363,<br />

603.<br />

3 See for instance: M. Meyyappan<br />

(Ed.), Carbon Nanotubes: Science<br />

and Applications, CRC Press, Boca<br />

Raton 2005; M. J. O’Connell (Ed.),<br />

Carbon Nanotubes, CRC Press,<br />

Boca Raton 2006; M. S. Dresselhaus,<br />

G. Dresselhaus, Ph. Avouris (Eds.),<br />

Carbon Nanotubes: synthesis, structure,<br />

properties and applications, Top. Appl.<br />

Phys., Vol. 80, Springer, Berlin 2001,<br />

P. J. F. Harris, Carbon Nanotubes and<br />

Related Structures, Cambridge University<br />

Press, Cambridge 1999.<br />

4 M. Fujita, R. Saito, G. Dresselhaus,<br />

M. S. Dresselhaus, Phys. Rev. B. 1992,<br />

45, 13834; R. Saito, M. S. Dresselhaus,<br />

G. Dresselhaus, Physical properties of<br />

carbon nanotubes, Imperial College<br />

Press, London 1998.<br />

5 C. Branca, V. Magazu, A. Mangione,<br />

Diamond Relat.Mater. 2005, 14, 846, and<br />

references therein.<br />

6 X. Lu, Z. Chen, P. v. R. Schleyer, J. Am.<br />

Chem. Soc. 2005, 127, 20; Z. Chen,<br />

S. Nagase, A. Hirsch, R. C. Haddon,<br />

W. Thiel, P. v. R. Schleyer, Angew. Chem.<br />

2004, 116, 1578; Angew. Chem. Int. Ed.<br />

2004, 43, 1552.<br />

7 X. Peng, N. Komatsu, S. Bhattacharya,<br />

T. Shimawaki, S. Aonuma, T. Kimura, A.<br />

Osuka Nature Nanotechnol. 2007, 2, 361.<br />

8 S. Reich, C. Thomsen, J. Maultzsch,<br />

Carbon Nanotubes; Wiley-VCH,<br />

Weinheim 2004; M. Damnjanovi�,<br />

I. Miloševi�, T. Vukovi�, R. Sredanovi�,<br />

Phys. Rev. B 1999, 60, 2728.<br />

9 Ph. Avouris, M. Radosavljevi�,<br />

S. J. Wind, Carbon Nanotube Electronics<br />

and Optoelectronics, In: S. V. Rotkin,<br />

S. Subramoney (Eds.), Applied Physics of<br />

Carbon Nanotubes; Springer, Berlin 2005,<br />

and references therein.<br />

10 E. Nakamura, K. Tahara, Y. Matsuo,<br />

M. Sawamura, J. Am. Chem. Soc. 2003,<br />

125, 2834; J. L. O’Loughlin, C.-H. Kiang,<br />

C. H. Wallace, T. K. Reynolds,<br />

L. Rao, R. B. Kaner, J. Phys. Chem. B<br />

2001, 105, 1921; S. Kammermaier,<br />

R. Herges, Angew. Chem. 1996, 108,<br />

470, Angew. Chem. Int. Ed. Engl.<br />

1996, 35, 417; R. E. Smalley, Y. Li,<br />

V. C. Moore, B. K. Price, R. Colorado,<br />

Jr., H. K. Schmidt, R. H. Hauge,<br />

A. R. Barron, J. M. Tour, J. Am. Chem.<br />

Soc. 2006, 128, 15824.<br />

11 R. Krupke, F. Hennrich, H. v. Löhneysen,<br />

M. M. Kappes, Science 2003, 301, 344;<br />

M. S. Arnold, A. A. Green, J. F. Hulvat,<br />

S. I. Stupp, M. C. Hersam, Nature<br />

Nanotechnol. 2006, 1, 60.<br />

12 R. Saito, M. Fujita, G. Dresselhaus,<br />

M. S. Dresselhaus, Appl. Phys. Lett. 1992,<br />

60, 2204.<br />

13 M. Fujita, R. Saito, G. Dresselhaus,<br />

M. S. Dresselhaus, Phys. Rev. B 1992, 45,<br />

13834.<br />

14 M. Endo, K. Takeuchi, K. Kobori,<br />

K. Takahashi, H. W. Kroto, A. Sarkar,<br />

Carbon 1995, 33, 873.<br />

15 S. Reich, L. Li, J. Robertson, Phys.<br />

Rev. B 2005, 72, 165423; K. A. Dean,<br />

B. R. Chalamala, J. Vac. Sci. Technol. B<br />

2003, 21, 868; S. C. Tsang, P. de Oliveira,<br />

J. J. Davis, M. L. H. Green, H. A. O. Hill,<br />

Chem. Phys. Lett. 1996, 249, 413<br />

16 J.-C. Charlier, Acc. Chem. Res. 2002,<br />

35, 1063; Y. Fan, B. R. Goldsmith,<br />

P. G. Collins, Nature Mater. 2005, 4,<br />

906; M. Grujicic, G. Cao, R. Singh,<br />

Appl. Surf. Sci. 2003, 211, 166;<br />

A. Bachtold, C. Strunk, J.-P. Salvetat,<br />

J.-M. Bonard, L. Forró, T. Nussbaumer,<br />

C. Schönenberger, Nature 1999, 397,<br />

673; J.-C. Charlier, T. W. Ebbesen,<br />

P. Lambin, Phys. Rev. B 1996, 53, 11108.<br />

17 A. J. Stone, D. J. Wales, Chem. Phys. Lett.<br />

1986, 128, 501.<br />

18 M. Buongiorno-Nardelli, B. I. Yakobson,<br />

J. Bernholc, Phys. Rev. Lett. 1998, 81,<br />

4656; B. I. Yakobson, Appl. Phys. Lett.<br />

1998, 72, 918.<br />

19 S. Yang, X. Chen, S. Motojima, Appl.<br />

Phys. Lett. 2002, 81, 3567–3569;<br />

P. Lambin, A. Fonseca, J. P. Vigneron,<br />

J. B. Nagy, A. A. Lucas, Chem. Phys. Lett.<br />

1995, 245, 85.


20 X. F. Zhang, X. B. Zhang,<br />

G. Van Tendeloo, S. Amelinckx,<br />

M. Op de Beeck, J. Van Landuyt, J. Cryst.<br />

Growth 1993, 130, 368; J. G. Lavin,<br />

S. Subramoney, R. S. Ruoff, S. Berber,<br />

D. Tomanek, Carbon 2002, 40, 11230.<br />

21 D. Reznik, C. H. Olk, D. A. Neumann,<br />

J. R. D. Copley, Phys. Rev. B 1995, 52, 116.<br />

22 J.-C. Charlier, J.-P. Michenaud, Phys. Rev.<br />

Lett. 1993, 70, 1858.<br />

23 M. Yu, O. Lourie, M. Dyer, K. Moloni,<br />

T. Kelly, R. S. Ruoff, Science 2000, 287,<br />

637.<br />

24 L. Forró, C. Schönenberger, Physical<br />

Properties of Multi-wall Nanotubes, in<br />

M. S. Dresselhaus, G. Dresselhaus,<br />

Ph. Avouris (Eds.), Carbon Nanotubes:<br />

synthesis, structure, properties and applications,<br />

Top. Appl. Phys., Vol. 80, Springer,<br />

Berlin 2001, pp. 329 and references<br />

therein.<br />

25 S. Frank, P. Poncharal, Z. L. Wang,<br />

W. A. de Heer, Science 1998, 280, 1744;<br />

A. Bachtold, C. Strunk, J. P. Salvetat,<br />

J. M. Bonard, L. Forró, T. Nussbaumer,<br />

C. Schönenberger, Nature 1999, 397,<br />

673–675.<br />

26 S. Iijima, T. Ichihashi, Y. Ando, Nature<br />

1992, 356, 776; H. P. Boehm, Carbon<br />

1997, 35, 581.<br />

27 Y. F. Li, J. S. Qiu, Z. B. Zhao,<br />

T. H. Wang, Y. P. Wang, W. Li, Chem.<br />

Phys. Lett. 2002, 366, 544.<br />

28 S. Iijima, Mater. Sci. Eng. B 1993, 19, 172.<br />

29 H. Hou, Z. Jun, F. Weller, A. Greiner,<br />

Chem. Mater. 2003, 15, 3170; S. Ihara,<br />

S. Itoh, J. Kitakami, Phys. Rev. B 1993,<br />

48, 5643; X. Chen, W. In-Hwang,<br />

S. Shimada, M. Fujii, H. Iwanaga,<br />

S. Motojima, J. Mater. Res. 2000, 15, 808.<br />

30 Y.-G. Yoon, P. Delaney, S. G. Louie, Phys.<br />

Rev. B 2002, 66, 073407; P. J. de Pablo,<br />

S. Howell, S. Crittenden, B. Walsh,<br />

E. Graugnard, R. Reifenberger, Appl.<br />

Phys. Lett. 1999, 75, 3941<br />

31 X. Lu, Z. Chen, Chem. Rev. 2005, 105,<br />

3643 and ref. therein; J. W. Mintmire,<br />

B. I. Dunlap, C. T. White, Phys. Rev. Lett.<br />

1992, 68, 631;<br />

32 J. Aihara, T. Yamabe, H. Hosoya,<br />

Synth. Met. 1994, 64, 309; Z. Zhou,<br />

M. Steigerwald, M. Hybertsen, L. Brus,<br />

R. A. Friesner, J. Am. Chem. Soc. 2004,<br />

126, 3597.<br />

References<br />

33 X. Blase, L. X. Benedict, E. L. Shirley,<br />

S. G. Louie, Phys. Rev. Lett. 1994, 72,<br />

1878.<br />

34 J. L. Ormsby, B. T. King, J. Org. Chem.<br />

2004, 69, 4287.<br />

35 J. L.Ormsby, B. T. King, J. Org. Chem.<br />

2007, 72, 4035.<br />

36 Y. Matsuo, K. Tahara, E. Nakamura,<br />

Org. Lett. 2003, 5, 3181.<br />

37 V. N. Khabashesku, W. E. Billups,<br />

J. L. Margrave, Acc. Chem. Res. 2002, 35,<br />

1087.<br />

38 S. Banerjee, T. Hemraj-Benny,<br />

S. S. Wong, Adv. Mater. 2005, 17, 17;<br />

N. Chandra, S. Namilae, C. Shet,<br />

Phys. Rev. B 2004, 69, 094101.<br />

39 D. Tasis, N. Tagmatarchis, A. Bianco,<br />

M. Prato, Chem. Rev. 2006, 106, 1106;<br />

A. Hirsch, O. Vostrowsky, Top. Curr.<br />

Chem. 2006, 245, 193, and references<br />

therein.<br />

40 S. Niyogi, M. A. Hamon, H. Hu,<br />

B. Zhao, P. Bhowmik, R. Sen,<br />

M. E. Itkis, R. C. Haddon, Acc. Chem.<br />

Res. 2002, 35, 1105; S. B. Sinnot,<br />

J. Nanosci. Nanotech. 2002, 2, 113;<br />

F. Béguin, E. Flahaut, A. Linares-Solano,<br />

J. Pinson, Surface Properties, Porosity,<br />

Chemical and Electrochemical Applications,<br />

in: A. Loiseau, P. Launois, P. Petit,<br />

S. Roche, J.-P. Salvetat, Understanding<br />

Carbon Nanotubes, Lect. Notes Phys.<br />

677, Springer Berlin, 2006, p. 509;<br />

K. Mylvaganam, L. C. Zhang, J. Phys.<br />

Chem. B 2004, 108, 5217.<br />

41 Z. Shen, W. Thiel, A. Hirsch, Chem. Phys.<br />

Chem. 2003, 4, 93; J. Bahr, J. M. Tour,<br />

J. Mater. Chem. 2002, 12, 1952.<br />

42 M. Burghard, Small 2005, 1, 1148;<br />

43 K. Balasubramanian, M. Burghard,<br />

Small 2005, 1, 180; S. Banerjee,<br />

T. Benny-Hemraj, S. S. Wong,<br />

Adv. Mater. 2005, 17, 17.<br />

44 J. Chen, M. A. Hamon, H. Hu,<br />

Y. S. Chen, A. M. Rao, P. C. Eklund,<br />

R. C. Haddon, Science 1998, 282, 95.<br />

45 J. Liu, A. G. Rinzler, H. J. Dai,<br />

J. H. Hafner, R. K. Bradley, P. J. Boul,<br />

A. Lu, T. Iverson, K. Shelimov,<br />

C. B. Huffman, F. Rodriguez-Macias,<br />

Y. S. Shon, T. R. Lee, D. T. Colbert,<br />

R. E. Smalley, Science 1998, 280, 1253;<br />

S. Heinze, J. Tersoff, P. Avouris, Appl.<br />

Phys. Lett. 2003, 83, 5038; M. A. Hamon,<br />

369


370 6 Carbon Nanotubes<br />

H. Hui, P. Bhowmik, M. E. Itkis,<br />

R. C. Haddon, Appl Phys A 2002, 74, 333.<br />

46 M. A. Hamon, J. Chen, H. Hu,<br />

Y. S. Chen, M. E. Itkis, A. M. Rao,<br />

P. C. Eklund, R. C. Haddon, Adv. Mater.<br />

1999, 11, 834, 13, 559<br />

47 D. Tasis, N. Tagmatarchis, V. Georgakilas,<br />

M. Prato, Chem. Eur. J. 2003, 9, 4000.<br />

48 S. Bandow, A. M. Rao, K. A. Williams,<br />

A. Thess, R. E. Smalley, P. C. Eklund,<br />

J. Phys. Chem. B 1997, 101, 8839;<br />

G. S. Duesberg, M. Burghard, J. Muster,<br />

G. Philipp, S. Roth, Chem. Commun.<br />

1998, 435; V. Krstic, G. S. Duesberg,<br />

J. Muster, M. Burghard, S. Roth, Chem.<br />

Mater. 1998, 10, 2338; M. J. O’Connell,<br />

S. M. Bachilo, C. B. Huffman,<br />

V. C. Moore, M. S. Strano, E. H. Haroz,<br />

K. L. Rialon, P. J. Boul, W. H. Noon,<br />

C. Kittrell, J. Ma, R. H. Hauge,<br />

R. B. Weisman, R. E. Smalley, Science<br />

2002, 297, 593; M. F. Islam, E. Rojas,<br />

D. M. Bergey, A. T. Johnson, A. G. Yodh,<br />

Nano Lett. 2003, 3, 269; M. S. Strano,<br />

V. C. Moore, M. K. Miller, M. J. Allen,<br />

E. H. Haroz, C. Kittrell, R. H. Hauge,<br />

R. E. Smalley, J. Nanosci. Nanotechnol.<br />

2003, 3, 81.<br />

49 J. Kong, H. Dai, J. Phys. Chem. B 2001,<br />

105, 2890.<br />

50 J. Chen, A. M. Rao, S. Lyuksyutov,<br />

M. E. Itkis, M. A. Hamon, H. Hu,<br />

R. W. Cohn, P. C. Eklund, D. T. Colbert,<br />

R. E. Smalley, R. C. Haddon, J. Phys.<br />

Chem. B 2001, 105, 2525; F. Pompeo,<br />

D. E. Resasco, Nano Lett. 2002, 2, 369,<br />

K. A. S. Fernando, Y. Lin, Y. P. Sun,<br />

Langmuir 2004, 20, 4777.<br />

51 M. Monthioux, B. W. Smith,<br />

B. Burteaux, A. Claye, J. E: Fischer,<br />

D. E. Luzzi, Carbon 2001, 39, 1251,<br />

J. Chen, M. A. Hamon, H. Hu, Y. Chen,<br />

A. M. Rao, P. C. Eklund, R. C. Haddon,<br />

Science 1998, 282, 95.<br />

52 J. Zhao, J. P. Lu, J. Han,C. K. Yang,<br />

Appl. Phys. Lett. 2003, 82, 3746;<br />

S. Gotovac, C.-M. Yang, Y. Hattori,<br />

K. Takahashi, H. Kanoh, K. Kaneko,<br />

J. Colloid Interface Sci. 2007, 314, 18.<br />

53 R. J. Chen, Y. Zhang, D. Wang, H. Dai,<br />

J. Am. Chem. Soc. 2001, 123, 3838.<br />

54 S. Gotovac, H. Honda, Y. Hattori,<br />

K. Takahashi, H. Kanoh, K. Kaneko,<br />

NanoLett. 2007, 7, 583.<br />

55 N. Nakashima, Y. Tomonari,<br />

H. Murakami, Chem. Lett. 2002, 638;<br />

A. B. Artyukhin, O. Bakajin,<br />

P. Stroeve, A. Noy, Langmuir 2004, 20,<br />

1442.<br />

56 G. J. Bahun, C. Wang, A. Adronov,<br />

J. Polymer Sci. A 2006, 44, 1941.<br />

57 R. B. Martin, L. Qu, Y. Lin, B. A. Harruff,<br />

C. E. Bunker, J. R. Gord, L. F. Allard,<br />

Y.-P- Sun, J. Phys. Chem. B 2004, 108,<br />

11447.<br />

58 K. A. S. Fernando, Y. Lin, W. Wang,<br />

S. Kumar, B. Zhou, S.-Y. Xie,<br />

L. T. Cureton, Y.-P. Sun, J. Am. Chem.<br />

Soc. 2004, 126, 10234.<br />

59 P. Petrov, F. Stassin, C. Pagnoulle,<br />

R. Jerome, Chem. Commun. 2003,<br />

2904; X. Lou, R. Daussin, S. Cuenot,<br />

A.-S. Duwez, C. Pagnoulle,<br />

C. Detrembleur, C. Bailly, R. Jerome,<br />

Chem. Mater. 2004, 16, 4005.<br />

60 D. M. Guldi, G. M. A. Rahman,<br />

F. Zerbetto, M. Prato, Acc. Chem. Res.<br />

2005, 38, 871.<br />

61 H. Li, B. Zhou, Y. Lin, L. Gu, W. Wang,<br />

K. A. S. Fernando, S. Kumar, L. F. Allard,<br />

Y.-P. Sun, J. Am. Chem. Soc. 2004, 126,<br />

1014.<br />

62 K. Y. Lee, M. Kim, Y. W. Lee, J.-J. Lee,<br />

S. W. Han, Chem. Phys. Lett.2007,<br />

440, 249; B. M. Quinn, C. Dekker,<br />

S. G. Lemay, J. Am Chem. Soc. 2005, 127,<br />

6146; G. Ren, Y. Xing, Nanotechnology<br />

2006, 17, 5596.<br />

63 C. Kim, Y. J. Kim, Y. A. Kim,<br />

T. Yanagisawa, K. C. Park, M. Endo,<br />

M. S. Dresselhaus, J. Appl. Phys. 2004,<br />

96, 5903; L. Liu, T. Wang, J. Li,Z. X. Guo,<br />

L. Dai, D. Zhang, D. Zhu, Chem. Phys.<br />

Lett. 2003, 367, 747; J. Sun, L. Gao,<br />

M. Iwasa, Chem. Commun. 2004, 832;<br />

Q. Fu, C. Lu, J. Liu, Nano Lett. 2002, 2,<br />

329; W. Han, A. Zettl, Nano Lett. 2003,<br />

3, 681; S. Lee, W. M. Sigmund, Chem.<br />

Commun. 2003, 780; J. Sun, L. Gao,<br />

W. Li, Chem. Mater. 2002, 14, 5169;<br />

E. A. Whitsitt, A. R. Barron, Nano Lett.<br />

2003, 3, 775; T. Seeger, Th. Kohler, Th.<br />

Frauenheim, N. Grobert, M. Ruhle,<br />

M. Terrones, G. Seifert, Chem. Commun.<br />

2002, 34; K. Hernadi, E. Ljubovic,<br />

J. W. Seo, L. Forro, Acta Mater. 2003, 51,<br />

1447; J. Sun, L. Gao, Carbon 2003, 41,<br />

1063.


64 S. A. Carran, P. M. Ajayan, W. J. Blau,<br />

D. L. Carroll, J. N. Coleman,<br />

A. B. Dalton, A. P. Davey, A. Drury,<br />

B. McCarthy, S. Maier, A. Strevens,<br />

Adv. Mater. 1998, 10, 1091; B. Z. Tang,<br />

H. Xu, Macromolecules 1999, 32, 2569;<br />

M. J. O’Connell, P. Boul, L. M. Ericson,<br />

C. B. Huffman, Y. H. Wang, E. Haroz,<br />

C. Kuper, J. M. Tour, K. D. Ausman,<br />

R. E. Smalley, Chem. Phys. Lett. 2001,<br />

342, 265; R. Bandyopadhyaya, F. Nativ-<br />

Roth, O. Regev, R. Yerushalmi-Rozen,<br />

Nano Lett. 2002, 2, 25.<br />

65 A. Star, D. W. Steuerman, J. R. Heath,<br />

J. F. Stoddart, Angew. Chem. 2002,<br />

114, 2618; Angew. Chem. Int. Ed. 2002,<br />

41, 2508; M. S. Arnold, M. O. Guler,<br />

M. C. Hersam, S. I. Stupp, Langmuir<br />

2005, 21, 4705; Z. Wei, M. Wan, T. Lin,<br />

L. Dai, Adv. Mater. 2003, 15, 136;<br />

M. J. O’Connell, P. Boul, L. M. Ericson,<br />

C. Huffman, Y. H. Wang, E. Haroz,<br />

C. Kuper, J. Tour, K. D. Ausman,<br />

R. E. Smalley, Chem. Phys. Lett., 2001,<br />

342, 265.<br />

66 A. Star, J. F. Stoddart, D. Steuerman,<br />

M. Diehl, A. Boukai, E. W. Wong,<br />

X. Yang, S.-W. Chung, H. Choi,<br />

J. R. Heath, Angew. Chem. 2001, 113,<br />

1771; Angew. Chem. 2001, 40, 1721.<br />

67 M. Cochet, W. K. Maser, A. M. Benito,<br />

M. A. Callejas, M. T. Martinez,<br />

J. M. Benoit, Chem. Commun. 2001, 1450.<br />

68 A. Ikeda, K. Hayashi, T. Konishi,<br />

J. Kikuchi, Chem. Commun. 2004, 1334.<br />

69 H. Dodziuk, A. Ejchart, W. Anczewski,<br />

H. Ueda, E. Krinichnaya, G. Dolgonosa,<br />

W. Kutner, Chem. Commun. 2003, 986.<br />

70 S. Pekker, J.-P. Salvetat, E. Jakab,<br />

J.-M. Bonard, L. J. Forro, Phys.<br />

Chem. B 2001, 105, 7938; B. N. Khare,<br />

M. Meyyappan, A. M. Cassell,<br />

C. V. Nguyen, J. Han, Nano Lett. 2002, 2,<br />

73; B. N. Khare, M. Meyyappan, J. Kralj,<br />

P. Wilhite, M. Sisay, H. Imanaka,<br />

J. Koehne, C. W. Bauschlicher, Appl.<br />

Phys. Lett. 2002, 81, 5237; K. S. Kim,<br />

D. J. Bae, J. R. Kim, K. A. Park,<br />

S. C. Lim, J.-J. Kim, W. B. Choi,<br />

C. Y. Park, Y. H. Lee, Adv. Mater. 2002,<br />

14, 1818.<br />

71 F. J. Owens, Z. Iqbal, 23rd Army Science<br />

Conf. 2002, (http://www.asc2002.com/<br />

summaries/l/LP-11.pdf).<br />

References<br />

72 T. Yildirim, O. Gülseren, S. Ciraci,<br />

Phys. Rev. B 2001, 64, 075404.<br />

73 V. N. Khabashesku, J. L. Margrave,<br />

E. V. Barrera, Diamond Relat. Mater.<br />

2005, 14, 859; E. T. Mickelson,<br />

C. B. Huffman, A. G. Rinzler,<br />

R. E. Smalley, R. H. Hauge,<br />

J. L. Margrave, Chem. Phys. Lett. 1998,<br />

296, 188.<br />

74 P. R. Marcoux, J. Schreiber, P. Batail,<br />

S. Lefrant, J. Renouard, G. Jacob,<br />

D. Albertini, J.-Y. Mevellec, Phys. Chem.<br />

Chem. Phys. 2002, 4, 2278.<br />

75 V. N. Khabashesku, W. E. Billups,<br />

J. L. Margrave, Acc. Chem. Res. 2002, 35,<br />

1087; H. F. Bettinger, Chem. Phys. Chem.<br />

2003, 4, 1283.<br />

76 K. F. Kelly, I. W. Chiang, E. T. Mickelson,<br />

R. H. Hauge, J. L. Margrave,<br />

X. Wang, G. E. Scusseria, C. Radloff,<br />

N. J. Halas, Chem. Phys. Lett.1999, 313,<br />

445; K. N. Kudun, H. F. Bettinger,<br />

G. E. Scuseria, Phys. Rev. B 2001, 63,<br />

045413.<br />

77 E. T. Mickelson, I. W. Chiang,<br />

J. L. Zimmerman, P. J. Boul, J. Lozano,<br />

J. Liu, R. E. Smalley, R. H. Hauge,<br />

J. L. Margrave, J. Phys. Chem. B 1999,<br />

103, 4318.<br />

78 R. K. Saini, I. W. Chiang, H. Peng,<br />

R. E. Smalley, W. E. Billups,<br />

R. H. Hauge, J. L. Margrave, J. Am.<br />

Chem. Soc. 2003, 125, 3617; P. J. Boul,<br />

J. Liu, E. T. Mickelson, C. B. Huffman,<br />

L. M. Ericson, I. W. Chiang,<br />

K. A. Smith, D. T. Colbert, R. H. Hauge,<br />

J. L. Margrave, R. E. Smalley, Chem.<br />

Phys. Lett. 1999, 310, 367; J. L. Stevens,<br />

A. Y. Huang, H. Peng, I. W. Chiang,<br />

V. N. Khabashesku, J. L. Margrave,<br />

Nano Lett. 2003, 3, 331.<br />

79 E. Unger, A. Graham, F. Kreupl,<br />

M. Liebau, M. Hoenlein, Curr. Appl.<br />

Phys. 2002, 2, 107.<br />

80 K. A. Worsley, K. R. Moonoosawmy,<br />

P. Kruse, Nano Lett. 2004, 4, 1541;<br />

T. Umeyama, N. Tezuka, M. Fujita,<br />

Y. Matano, N. Takeda, K. Murakoshi,<br />

K. Yoshida, S. Isoda, H. Imahori,<br />

J. Phys. Chem. C 2007, 111, 9734.<br />

81 J. M. Ashcroft, K. B. Hartman,<br />

Y. Mackeyev, C. Hofmann, S. Pheasant,<br />

L. B. Alemany, L. J. Wilson, Nanotechnol.<br />

2006, 17, 5033.<br />

371


372 6 Carbon Nanotubes<br />

82 K. S. Coleman, S. R. Bailey, S. Fogden,<br />

M. L. H. Green, J. Am. Chem. Soc. 2003,<br />

125, 8722.<br />

83 K. Kamaras, M. E. Itkis, H. Hu, B. Zhao,<br />

R. C. Haddon, Science 2003, 301,<br />

1501; H. Hu, B. Zhao, M. A. Hamon,<br />

K. Kamaras, M. E. Itkis, R. C. Haddon,<br />

J. Am. Chem. Soc. 2003, 125, 14893;<br />

M. Holzinger, J. Abraham, P. Whelan,<br />

R. Graupner, L. Ley, F. Hennrich,<br />

M. Kappes, A. Hirsch, J. Am. Chem.<br />

Soc. 2003, 125, 8566; M. Holzinger,<br />

J. Steinmetz, D. Samaille, M. Glerup,<br />

M. Paillet, P. Bernier, L. Ley,<br />

R. Graupner, Carbon 2004, 42, 941.<br />

84 M. Holzinger, O. Vostrowsky, A. Hirsch,<br />

F. Hennrich, M. Kappes, R.Weiss,<br />

F. Jellen, Angew. Chem. 2001, 113,<br />

4132; Angew. Chem. Int. Ed. 2001, 40,<br />

4002; Y. Ying, R. K. Saini, F. Liang,<br />

A. K. Sadana, W. E. Billups, Org. Lett.<br />

2003, 5, 1471.<br />

85 J. M. Tour, C. A. Dyke, Chem. Eur. J.<br />

2004, 10, 812.<br />

86 M. S. Strano, C. A. Dyke, M. L. Usrey,<br />

P. W. Barone, M. J. Allen, Ho. Shan,<br />

C. Kittrell, R H. Hauge, J. M. Tour,<br />

R. E. Smalley, Science 2003, 301, 1519.<br />

87 C. Ménard-Moyon, F. Dumas, E. Doris,<br />

C. Mioskowski, J. Am. Chem. Soc. 2006,<br />

128, 14764.<br />

88 X. Lu, F. Tian, N. Wang, Q. Zhang,<br />

Org. Lett. 2002, 4, 4313.<br />

89 V. Georgakilas, K. Kordatos, M. Prato,<br />

D. M. Guldi, M. Holzinger, A. Hirsch,<br />

J. Am. Chem. Soc. 2002, 124, 760;<br />

C. Ménard-Moyon, N. Izard, E. Doris,<br />

C. Mioskowski, J. Am. Chem. Soc. 2006,<br />

128, 6552.<br />

90 A. Bianco, K. Kostarelos, C. D. Partidos,<br />

M. Prato, Chem. Commun. 2005, 571.<br />

91 V. Georgakilas, N. Tagmatarchis,<br />

D. Pantarotto, A. Bianco, J.-P. Briand,<br />

M. Prato, Chem. Commun. 2002, 3050.<br />

92 M. Alvaro, P. Atienzar, P. de la Cruz,<br />

J. L. Delgado, H. Garcia, F. J. Langa,<br />

Phys. Chem. B 2004, 108, 12691;<br />

Y. Wang, Z. Iqbal, S. Mitra, Carbon<br />

2005, 43, 1015; S. Banerjee, S. S. Wong,<br />

J. Phys. Chem. B 2002, 106, 12144.<br />

93 J. L. Delgado, P. de la Cruz, F. Langa,<br />

A. Urbina, J. Casadoc, J. T. López<br />

Navarrete, Chem. Commun. 2004,<br />

1734.<br />

94 L. Zhang, J. Yang, C. L. Edwards,<br />

L. B. Alemany, V. N. Khabashesku,<br />

A. R. Barron, Chem. Commun. 2005,<br />

3265.<br />

95 S. Banerjee, M. G. C. Kahn, S. S. Wong,<br />

Chem. Eur. J. 2003, 9, 1898.<br />

96 F. Nunzi, F. Mercuri, A. Sgamellotti,<br />

N. Re, J. Phys. Chem. B 2002, 106, 10622;<br />

F. Nunzi, F. Mercuri, A. Sgamellotti,<br />

Mol. Phys. 2003, 101, 2047.<br />

97 A. N. Khlobystov, D. A. Britz,<br />

G. A. D. Briggs, Acc. Chem. Res. 2005,<br />

38, 901.<br />

98 D. Ugarte, T. Stöckli, J. M. Bonard,<br />

A. Châtelain, W. A. de Heer, Appl. Phys.<br />

A 1998, 67, 101.<br />

99 J. Sloan, A. I. Kirkland, J. L. Hutchison,<br />

M. L. H. Green, Acc. Chem. Res. 2002,<br />

35, 1054.<br />

100 J. Steinmetz, S. Kwon, H.-J. Lee,<br />

E. Abou-Hamad, R. Almairac,<br />

C. Goze-Bac, H. Kim, Y.-W. Park, Chem.<br />

Phys. Lett. 2006, 431, 139; K. Yanagi,<br />

Y. Miyata, H. Kataura, Adv. Mater. 2006,<br />

18, 437.<br />

101 L.-J. Li, A. N. Khlobystov, J. G. Wiltshire,<br />

G. A. D. Briggs, R. J. Nicholas, Nature<br />

Mater. 2005, 4, 481.<br />

102 F. Stercel, N. M. Nemes, J. E. Fischer,<br />

D. E. Luzzi, Mater. Res. Soc. Symp. Proc.<br />

2002, 706, 245; H. Kataura, Y. Maniwa,<br />

M. Abe, A. Fujiwara, T. Kodama,<br />

K. Kikuchi, H. Imahori, Y. Misaki,<br />

S. Suzuki, Y. Achiba, Appl. Phys. A 2002,<br />

74, 349.<br />

103 Y. Fujita, S. Bandow, S. Iijima, Chem.<br />

Phys. Lett. 2005, 413, 410.<br />

104 B. W. Smith, M. Monthioux, D. E. Luzzi,<br />

Nature 1998, 396, 323.<br />

105 L. Jiang, L. Gao, Carbon 2003, 41, 2923.<br />

106 M. Monthioux, V. L. Kuznetsov, Carbon<br />

2006, 44, 1621.<br />

107 M. Monthioux, P. Serp, E. Flahaut,<br />

M. Razafinimanana, C. Laurent,<br />

A. Peigney, W. Bacsa, J.-M. Broto,<br />

in: Nanotechnology Handbook<br />

(ed. B. Bhushan), 2nd Edition (revised),<br />

Springer-Verlag, Heidelberg, Germany,<br />

2006, pp. 43.<br />

108 A. Mansour, M. Razafinimanana,<br />

M. Monthioux, M. Pacheco, A. Gleizes,<br />

Carbon 2007, 45, 1651.<br />

109 M. Maugey, W. Neri, C. Zakri, A. Derré,<br />

A. Pénicaud, L. Noé, M. Chorro,


P. Launois, M. Monthioux, P. Poulin,<br />

J. Nanosci. Nanotechnol. 2007, 7,<br />

2633.<br />

110 R. H. Hurt, M. Monthioux., A. Kane,<br />

Eds: Special issue on the toxicity of<br />

carbon nanomaterials, Carbon 2006, 44,<br />

Issue 6.<br />

111 J. N. Coleman, U. Khan, W. J. Blau,<br />

Y. K. Gun’ko, Carbon 2006, 44, 1624.<br />

112 G. G. Tibbetts, M. L. Lake, K. L. Strong,<br />

B. P. Rice, Compos. Sci. Technol. 2007,<br />

67, 1709.<br />

113 R. Haggenmueller, H. H. Gommans,<br />

A. G. Rinzler, J. E. Fischer, K. I. Winey,<br />

Chem. Phys. Lett. 2000, 330, 219.<br />

114 R. Andrews, D. Jacques, A. M. Rao,<br />

T. Rantell, F. Derbyshire, Y. Chen,<br />

J. Chen, R. C. Haddon, Appl. Phys. Lett.<br />

1999, 75, 1329.<br />

115 M. Zhang, S. Fang, A. A. Zakhidov,<br />

S. B. Lee, A. E. Aliev, C. D. Williams,<br />

K. R. Atkinson, R. H. Baughman, Science<br />

2004, 306, 1356.<br />

116 V. Pichot, S. Badaire, P. A. Albouy,<br />

C. Zakri, P. Poulin, P. Launois, Phys.<br />

Rev. B 2006, 74, 245416.<br />

117 P. Avouris, Z. Chen, V. Perebeinos,<br />

Nature Nanotechnol. 2007, 2, 605.<br />

118 J.-P. Cleuziou,.W. Wernsdorfer,<br />

V. Bouchiat, T. Ondarçuhu,<br />

M. Monthioux, Nature Nanotechnol.<br />

2006, 1, 53.<br />

119 R. H. Baughman, C. Cui,<br />

A. A. Zakhidov, Z. Iqbal, J. N. Barisci,<br />

G. M. Spinks, G. G. Wallace,<br />

References<br />

A. Mazzoldi, D. DeRossi, A. G. Rinzler,<br />

O. Jaschinski, S. Roth, M. Kertesz,<br />

Science 1999, 284, 1340.<br />

120 P. Kim, C. M. Lieber, Science 1999, 286,<br />

2148.<br />

121 M. Whitby, N. Quirke, Nature<br />

Nanotechnol. 2007, 2, 87.<br />

122 C. Niu, E. K. Sichel, R. Hoch, D. Moy,<br />

H. Tennent, Appl. Phys. Lett. 1997, 70,<br />

1480.<br />

123 M. Endo,Y. J. Kim, T. Chino, O. Shinya,<br />

Y. Matsuzawa, H. suezaki, K. Tantrakarn,<br />

M. S. Dresselhaus, Appl. Phys. A 2006,<br />

82, 559.<br />

124 V. Pushparaj, M. M. Shaijumon,<br />

A. Kumar*, S. Murugesan†, L. Ci,<br />

R. Vajtai, R. J. Linhardt, O. Nalamasu,<br />

P. M. Ajayan, Proc. Nat. Acad. Sci. USA<br />

2007, 104, 13574.<br />

125 J. M. Planeix, N. Coustel, B. Coq,<br />

V. Brotons, P. S. Kumbhar, R. Dutartre,<br />

P. Geneste, P. Bernier, P. M. Ajayan,<br />

J. Am. Chem. Soc. 1994, 116, 7935.<br />

126 J. Kong, N. Franklin, C. Zhou,<br />

M. Chapline, S. Peng, K. Cho, H. Dai,<br />

Science 2000, 287, 622.<br />

127 P. G. Collins, K. Bradley, M. Ishigami,<br />

A. Zettl, Science 2000, 287, 1801.<br />

128 W. A. de Heer, A. Chatelain, D. Ugarte,<br />

Science 1995, 270, 1179.<br />

129 N. De Jonge, J.-M. Bonard, Phil. Trans.<br />

R. Soc. Lond. A 2004, 362, 2239.<br />

130 H. Dai, J. H. Hafner, A. G. Rinzler,<br />

D. T. Colbert, R. E. Smalley, Nature 1996,<br />

384, 147.<br />

373


7<br />

Angle-strained Cycloalkynes<br />

Henning Hopf and Jörg Grunenberg<br />

7.1<br />

Introduction<br />

According to the sp-hybridization model, the carbon–carbon triple bond possesses<br />

a linear structure, and indeed, many alkynes fulfill this prediction. High level ab<br />

initio calculations [1] of isolated alkyne molecules, as well as gas phase electron<br />

diffraction experiments [2] reveal their linear minimum structure. On the other<br />

hand, given that the C�C–R bending force constant becomes intrinsically smaller<br />

for substituted alkynes and polyynes, crystal packing effects may lead to deviations<br />

from strictly linear geometries [3]. From a dynamic point of view even in the case<br />

of acetylene one has to include the – albeit reduced – flexibility of the carbon–<br />

carbon triple bond in order to describe, for example, the vinylidene acetylene<br />

rearrangement (see below) [4]. Further, the linear sp-hybrid picture only holds true<br />

for the electronic ground state. Cis- and trans-bent configurations are known in<br />

the case of low-lying electronic states [5]. Nevertheless, many alkynes correspond<br />

to the prediction of linearity due to the sp-hybrid model, and we have collected<br />

a selection of typical examples, reaching from the parent compound acetylene 1<br />

via a few alkyl 2–4 and aryl acetylenes 5–7 to a functionalized acetylene, dimethyl<br />

acetylenedicarboxylate 8, in Scheme 7.1, together with the appropriate references<br />

[1, 2, 6–11], all dating from the recent literature.<br />

The easiest way to distort a triple bond is to incorporate it into a sufficiently small<br />

ring structure. The question, of course, is what this ring size is, and from what<br />

ring size on the respective cycloalkynes can we isolate compounds, that can be<br />

worked with under normal laboratory conditions [12]. Rather than using the nonspecific<br />

term ‘distorted cycloalkyne’ Krebs [12] prefers to speak of ‘angle-strained<br />

cycloalkynes’. Since it is easier to deform a C�C–C arrangement than its more<br />

hydrogenated olefinic and saturated analogs, relatively large angle deformations<br />

are possible in cycloalkynes without significant changes in energy. Krebs arbitrarily<br />

considers all cycloalkynes in which the C�C–C angle is deformed by more than<br />

10° as angle-strained [12]. Accepting this definition, the stable cyclononyne (see<br />

below) is an angle-strained cycloalkyne, whereas cyclodecyne is not.<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

375


376 7 Angle-strained Cycloalkynes<br />

Scheme 7.1<br />

7.2<br />

Cyclopropyne and Cyclobutyne: Speculations and Calculations<br />

on Non-isolable Cycloalkynes<br />

7.2.1<br />

Cyclopropyne and Related Systems<br />

The smallest conceivable cyclic acetylene is cyclopropynylidene (9, Scheme 7.2),<br />

the lowest member of the cyclocarbons (see below).<br />

Scheme 7.2<br />

This structure has been discussed in the chemical literature [13], but only in the<br />

context of theoretical investigations. On the other hand, cyclopropyne (general<br />

structure 10) has been invoked both in spectroscopic [14] and theoretical studies<br />

[15]. According to computational evidence, it appears likely that the tetragonal<br />

form of this hydrocarbon, 14, is a transition state in the automerization of propadienylidene<br />

13; whereas planar cyclopropyne 15 could be involved as a transition<br />

state in the rearrangement of 13 into cylopropenylidene (11, Scheme 7.3) [15], a<br />

C 3H 2 isomer that has already been prepared in matrix in 1984 [16] and has been<br />

discussed in connection with interstellar chemistry [17].<br />

The olefinic system, corresponding to 10, cyclopropene 12, known for a long<br />

time, has often been used as an addition partner in preparative chemistry [18].


7.2 Cyclopropyne and Cyclobutyne: Speculations and Calculations on Non-isolable Cycloalkynes<br />

Scheme 7.3<br />

Replacement of the methylene group in 10 by a SiH 2 moiety results in silacyclopropyne<br />

19, a cycloalkyne that has been generated as summarized in<br />

Scheme 7.4 [19].<br />

Pulsed flash pyrolysis of the precursor 16 results in the generation of the<br />

silylene 17 that undergoes rearrangement to the silacyclopropene 18. Photolysis<br />

of either 17 or 18 furnished 19, the first formal cyclopropyne ever to be prepared<br />

and characterized experimentally by matrix IR spectroscopy and theoretically by<br />

ab initio calculations.<br />

The even simpler SiC 2 20 has been prepared by laser vaporization of a silicon<br />

carbide rod within a pulsed supersonic nozzle [20]. By spectroscopic analysis it was<br />

shown that the molecule is triangular in both the ground and excited electronic<br />

states. The carbon–carbon length in the ground state is 125 pm and the C–Si–C<br />

angle amounts to 40°; i.e. the cycloalkynes possesses C 2v symmetry.<br />

Didehydrooxirene (21, oxacyclopropyne) has been discussed in connection with<br />

mass spectrometric studies of the hydroxypropynal molecular ion [21].<br />

Scheme 7.4<br />

377


378 7 Angle-strained Cycloalkynes<br />

7.2.2<br />

Cyclobutyne<br />

Experiments to generate and trap cyclobutyne have been scarce until the present<br />

day although there were several early attempts to generate this next higher cycloalkyne<br />

homolog [22].<br />

According to theoretical studies, cyclobutyne may be the limiting ring size for<br />

simple cycloalkynes: in ab initio calculations both singlet and triplet cyclobutyne<br />

correspond to energy minima [23]; the structural parameters of a calculated<br />

structure, 22, are shown in Scheme 7.5.<br />

Scheme 7.5<br />

Two isomerization processes have been considered likely for the highly reactive<br />

22: thermal isomerization to butatriene 23, and ring contraction to cyclopropylidene<br />

methylene 24 [24]. Whereas the (calculated) activation energy for the<br />

unimolecular rearrangement of 22 to 23 should be substantial [25], cyclobutyne<br />

22 should isomerize to 24 with little or no barrier making the direct observation<br />

of this cycloalkyne difficult or impossible [26].<br />

These calculations are supported by the results summarized in Scheme 7.6<br />

[27].<br />

Scheme 7.6


7.3 Cyclopentyne, Cyclohexyne, Cycloheptyne: from Reactive Intermediates to Isolable Compounds<br />

When the halides 25 are treated with lithium diethylamide (LDA) in THF in<br />

the presence of lithium thiophenolate, the two thio ethers 27 and 29 are obtained<br />

after work-up. To rationalize these findings it has been proposed that initially a<br />

cyclo butyne derivative, bicyclo[3.2.0]-hept-6-yne 26 is generated, that subsequently<br />

either adds the trapping agent to form 29 or undergoes the ring-contraction to 28<br />

mentioned, that then reacts with the thiophenolate to give 27.<br />

Annelation with an aromatic system has often been employed to stabilize<br />

reactive organic molecules; however, in the case of synthesizing benzannulated<br />

cyclo butynes this approach failed [28]. Another method of influencing the stability<br />

and reactivity of hydrocarbons consists in exchange of all hydrogen substituents<br />

for fluorine. However, in the case of cyclobutyne, ab initio calculations suggest that<br />

fluorine substitution leads to destabilization with respect to cyclobutene [29].<br />

The ability of metal atoms to complex and stabilize highly reactive molecules is<br />

well known [30], and metal complexes containing the ‘cyclobutyne’ ligand such as<br />

Os 3(CO) 9(µ-� 2 -cyclo-butyne)(µ-SPh)(µ-H) have been described [31].<br />

7.3<br />

Cyclopentyne, Cyclohexyne, Cycloheptyne: from Reactive Intermediates<br />

to Isolable Compounds<br />

7.3.1<br />

Cyclopentyne and its Derivatives<br />

Reliable experimental ground is reached with cyclopentyne 34, which has been<br />

generated by various routes [32]; three conceptually different ones are summarized<br />

in Scheme 7.7.<br />

For example, on phenyl lithium treatment, 30 is debrominated to carbene 33,<br />

that stabilizes itself by the already discussed vinylidenecarbene � acetylene rearrangement<br />

to 34 [26, 33].<br />

That 34 has actually been produced was shown by different trapping experiments:<br />

for example, in the presence of trans-2-butene the cyclobutene 36 was<br />

obtained [34]. In another route cyclopropenone 31 was prepared by photodecomposition<br />

of 2,6-diazocyclohexanone in an argon matrix and was decarbonylated on<br />

further irradiation at 8 K to 34 [35]. Originally, only the allene 37 as a secondary<br />

product of cyclopentyne was observed; in later experiments the IR spectrum of<br />

34 could be recorded in matrix [36]. Finally, cyclobutanone 32 was converted to<br />

34 via 35 and (presumably) 33. In this case trapping with trans-1-methoxypropene<br />

yielded the trans-cycloadduct 38 [37].<br />

Cyclopentyne 34 can be stabilized in various ways (Scheme 7.8).<br />

4-Thia-3,3,5,5-tetramethylcyclopentyne 39 has been obtained by oxidation of the<br />

corresponding bis-hydrazone and has been trapped by various addition reagents<br />

[38]; the intermediate is presumably less strained because of the longer C–S bonds<br />

and consequently larger C�C–C bond angles, and the neighboring gem-dimethyl<br />

groups provide further protection (see below).<br />

379


380 7 Angle-strained Cycloalkynes<br />

Scheme 7.7<br />

Scheme 7.8�


7.3 Cyclopentyne, Cyclohexyne, Cycloheptyne: from Reactive Intermediates to Isolable Compounds<br />

Metal complexation provides another means of stabilization. Thus the platinum<br />

complex 40 is reduced by sodium amalgam to give a colorless, very reactive solid<br />

for which structure 41 has been proposed on the basis of NMR and IR data [39].<br />

On methanol treatment 41 gives instantaneously the dinuclear complex 42, the<br />

structure of which was established by X-ray crystallography [39].<br />

Stabilization by condensed aromatic rings is realized in the case of the aromatic<br />

hydrocarbon acenaphthyne 45, which has been prepared by several routes as<br />

shown in Scheme 7.9.<br />

Thus photoextrusion of nitrogen from 43 [35] or nitrogen and carbon monoxide<br />

from the precursor 44 in a matrix led to 45 – allowing its UV and IR spectra to be<br />

recorded at 15 K – as did the Ramberg–Bäcklund rearrangement of the dibromo<br />

sulfone 48 [40]. Chemical proof that 45 has been generated in these transformations<br />

was established by its interception with oxygen leading to the quinone 46,<br />

and self-trapping furnished the trimer decacyclene (47, see below) [35].<br />

The bicyclic acetylene norbornyne 51 was first prepared [41] by metalating 49<br />

with n-butyllithium; the resulting organolithio intermediate 50 on mild warming<br />

lost lithium chloride to provide 51 that trimerizes in poor yield (ca. 10%) to the<br />

polycyclic aromatic stereoisomers 52 and 53 (Scheme 7.10) [42].<br />

Scheme 7.9 (a) h�, Ar, 15 K; (b) h�, Ar, 15 K; (c) t-BuOK, t-BuOH, room temp.; (d) room temp.<br />

(5% from 44); (e) O 2 .<br />

381


382 7 Angle-strained Cycloalkynes<br />

Scheme 7.10<br />

The still higher unsaturated cyclopentyne derivative bicyclo[2.2.1]hept-2-en-<br />

5-yne [43] has also been generated by an elimination reaction [43, 44]. With a<br />

calculated strain energy of ca. 97 kcal mol –1 it is approximately 30 kcal mol –1 more<br />

strained than cyclopentyne itself [44]. Lower bicyclo[2.1.1]hex-2-yne [45] and higher<br />

homologs of 51 have also been produced (see below) [46]; like norbornyne 51 they<br />

trimerize to aromatic compounds of type 52/53.<br />

7.3.2<br />

Cyclohexyne and its Derivatives<br />

Among all angle-strained cycloalkynes cyclohexyne 56 has received the greatest<br />

attention. Since this classical work has been reviewed several times [47], only the<br />

most important approaches to 56 will be discussed here (Scheme 7.11).<br />

In the debromination of 1,2-dibromocyclohexene 54 [48] and the dehydrochlorination<br />

of 1-chlorocyclohexene 55 [49] the relationship between precursor<br />

and the desired hydrocarbon is obvious. The oxidation of the bis-hydrazone 57<br />

to 56 can be rationalized by postulating a bis-carbene intermediate [50]. And the<br />

last two approaches [51] could also proceed via a vinylidene carbene, 59. Since<br />

cyclohexyne 56 is still too reactive to be isolated as such, the chemical proofs of its<br />

intermediate existence again rest on trapping and matrix isolation experiments. For<br />

example, it has been reacted with tris(triphenylphosphine)platinum(0) to provide a<br />

stable platinum complex [52], and trimerization leads to dodeca hydrotriphenylene<br />

(see below) [40, 48].<br />

Just as in the case of the lower homolog, matrix isolation has allowed recording<br />

of the IR spectrum of 56 (Scheme 7.12) [53, 54].<br />

Flash vacuum pyrolysis of 4-cyclopentylidene-3-methyl-isoxazol-5(4H)-one 61<br />

at 700–800 °C and 10 –4 torr causes fragmentation to the carbene 59 that isomerizes<br />

to 56. The IR spectrum recorded at 77 K shows absorption bands at 2090 and<br />

2105 cm –1 that were assigned to the stretching vibrations of the deformed triple<br />

bond. These signals disappeared on warm-up between –110 and –100 °C, and<br />

when room temperature had been reached, 62 had been formed.


7.3 Cyclopentyne, Cyclohexyne, Cycloheptyne: from Reactive Intermediates to Isolable Compounds<br />

Scheme 7.11 (a) Li/Hg, THF; (b) PhLi, ether; (c) HgO, benzene; (d) 480–640 °C; (e) tBuOK.<br />

Scheme 7.12<br />

The tetramethyl flanked cyclohexyne 64 is produced on heating 63 at 370 °C and<br />

then quenching the pyrolysate at 12 K in an Ar matrix [55]. Again, the vibrational<br />

spectrum could be recorded at low temperature, but, disappointingly, the C�C<br />

stretching vibration of 64 could not be identified; very likely because it is hidden<br />

under one of the (intense) CO-bands at 2139 and 2149 cm –1 . On warming the<br />

reaction mixture to 45 K dimerization of 64 to the enyne 65 took place.<br />

383


384 7 Angle-strained Cycloalkynes<br />

Scheme 7.13<br />

As in the case of the bicyclic cyclopentyne 51 a relatively large number of<br />

cyclo hexyne derivatives is known [12, 56] in which this strained moiety has been<br />

incorporated into a polycyclic framework.<br />

The surprisingly stable tetrasila cyclohexyne 68 is obtained in good yield when<br />

the tetrasilane 66 is reacted with acetylene bis-Grignard (67, Scheme 7.13).<br />

This cyclohexyne is stable at room temperature and has a half life of 8 h at<br />

174 °C in decane; clearly, it is stabilized both by steric protection through the<br />

methyl substituents and the long carbon–silicon and silicon–silicon bonds and<br />

large C�C–C bond angles, respectively [57]. Still, the acetylene–vinylidenecarbene<br />

equilibrium can evidently take place here as well, since in the presence of diphenyldiazomethane,<br />

68 is converted into the allene 70 in quantitative yield, presumably<br />

by forming the carbene intermediate 69 first.<br />

7.3.3<br />

Cycloheptyne and its Derivatives<br />

The methods used to prepare cycloheptyne do not differ from those employed to<br />

prepare the lower homologs (see above). However, with this hydrocarbon we are<br />

beginning to reach the shores of stability. In dilute dichloromethane solution at<br />

–25 °C this cycloalkyne (72, Scheme 7.14) has a half-life of less than a minute, but<br />

at –78 °C this has already increased to one hour [58].<br />

Irradiation of the cyclopropenone 71 in an Ar matrix at 17 K provided 72 as<br />

expected and allowed the measurement of the vibrational spectrum: the triple<br />

bond stretching frequency is registered at 2121 cm –1 [59].<br />

The blockage of the immediate vicinity of the triple bond by methyl groups leads<br />

to a drastic increase in stability. 3,3,7,7-Tetramethylcycloheptyne 73 is a stable hydrocarbon<br />

at room temperature with a half-life for dimer formation of one day at<br />

25 °C [60]. Compared to the parent hydrocarbon 72 the methyl protected derivative


Scheme 7.14<br />

7.4 The Isolable Angle-strained Cycloalkynes: Cyclooctyne, Cyclononyne, and Beyond<br />

dimerizes 10 7 to 10 8 times more slowly. Krebs and co-workers have synthesized<br />

numerous cycloheptyne derivatives of the general structure 74 [12]; most of them<br />

are more stable than 72. For example the thio ether (74, X = S) shows no tendency<br />

to di- or oligomerization even at 140 °C [61]. Because of the stability of the sulfur<br />

compound it was possible to determine its molecular structure by electron diffraction;<br />

with 145.8° for the C�C–C angle the deformation is about 12.7° higher<br />

than that observed for cyclooctyne (see below) [62].<br />

As described for cyclohexyne (see above), the platinum complex 75 was<br />

obtained from intermediately generated 72 [52]. Its X-ray structural analysis<br />

shows that the triple bond deviates from linearity by ca. 40°. The dimeric metal<br />

complex 76 is produced if 73 is reacted with CuCl·Me 2 S. X-ray crystallography<br />

reveals that the ‘deformation angle’ at the triple bond amounts to 34.4° in this<br />

case [63].<br />

7.4<br />

The Isolable Angle-strained Cycloalkynes: Cyclooctyne, Cyclononyne, and Beyond<br />

7.4.1<br />

Cyclooctyne and its Derivatives<br />

With cyclooctyne (79, Scheme 7.15) we have definitely reached the region of stable<br />

acetylenes, and, in fact, the hydrocarbon is so readily available [64], that it has<br />

become a versatile building block in organic synthesis [65]. The presently most<br />

convenient methods to prepare 79 are summarized in Scheme 7.15.<br />

385


386 7 Angle-strained Cycloalkynes<br />

Scheme 7.15 (a) Br 2 ,CH 2 Cl 2 , –40°C; (b) t-BuOK, ether, THF, 0–15 °C; (c) LDA, hexane,<br />

THF, –25 – +15 °C (80–84%); (d) n-BuLi, –70 °C (X = S; 70%. X = Se; 85%).<br />

Starting from cyclooctene 77 bromination first provides the expected dibromide<br />

which is dehydrobrominated as shown in the scheme to the vinylbromide 78. After<br />

a base switch to LDA a second elimination step furnishes 79 that can be obtained by<br />

this route in 30-g lots [66]. Alternatively, the also readily available 1,2,3-thiadiazole<br />

(80, X = S) or 1,2,3-selenodiazole (80, X = Se) on treatment with n-butyl lithium<br />

at low temperature furnish the cycloalkyne in good yield as well [67].<br />

Yet cyclooctyne still possesses a quite strongly deformed C–C�C–C-unit.<br />

Structure determination by electron diffraction shows a reduced C�C–C angle of<br />

154.5° [68]. Whereas many medium-ring cycloalkynes have an overall flat structure<br />

(see below), cyclooctyne is twisted.<br />

Because of the strained nature of their triple bonds, cyclooctynes form metal<br />

complexes readily [69].<br />

A sizeable number of derivatives of cyclooctyne is known, and many of these<br />

are summarized in compilations by Krebs and Meier and their coworkers [12,<br />

70–72].<br />

7.4.2<br />

Cyclononyne and Cyclodecyne<br />

With cyclononyne and cyclodecyne we are ending our journey through the<br />

homologous series of cycloalkynes. Both hydrocarbons had been obtained by<br />

various elimination reactions from the appropriate precursors but were often<br />

contaminated with their allenic and dienic isomers. They were finally obtained in<br />

pure form by employing the Curtius route (oxidation of 1,2-bis-hydrazones, see<br />

above) to cyclonona- and cyclodeca-1,2-dione [73]. Still, cyclononyne is not free<br />

from angle-strain. With a C�C–C angle of 160.2° – determined by electron diffraction<br />

[74] – this is only slightly larger than in cycloctyne (see above), while the<br />

triple bond has a normal length in cyclononyne. No experimental structural data<br />

of cyclodecyne seem to be known. According to molecular mechanics calculations<br />

the above angle is 171.6°, i.e. still not yet 180° [75]. It has been pointed out, though,<br />

that because of non-bonded interactions, additional deformation and hence additional<br />

strain is produced, making it unlikely that cyclodecyne will possess a linear<br />

arrangement of the triple bond and the two carbon atoms flanking it [76].


7.5<br />

Cyclic Polyacetylenes<br />

7.5 Cyclic Polyacetylenes<br />

As cyclic polyacetylenes we want to define all cycloalkynes possessing at least two<br />

triple bonds, whether these are conjugated or not. Many of these systems possess<br />

nonlinear acetylene or diacetylene moieties, and the number of know representatives<br />

of this class is growing rapidly. The following compilation does not intend to<br />

be comprehensive; it only aims at demonstrating the large variation in structures<br />

one is encountering in this sub-group.<br />

One the most famous hydrocarbons in this class is 1,5-cyclooctadiyne 81,<br />

originally prepared in small yield by the dimerization of butatriene 23 [77], and<br />

later with far better success from the dibromide 82 under the conditions shown<br />

in Scheme 7.16 [78]. The hydrocarbon is essentially planar, the deviation from the<br />

180° degree geometry of the acyclic model compounds summarized in Scheme 7.1<br />

amounts to 20.7°, and the distance between the parallel double bonds is 259.7 pm.<br />

Compared to cyclooctyne 79 (see above) the angle deformation at the triple bond<br />

is less pronounced. The two benzannelated hydrocarbons 83 and 84 have also<br />

been investigated by X-ray crystallography. In 83 the C�C–C angle is 155.8° and<br />

the distance between the triple bonds amounts to 261 pm. The molecule is nearly<br />

planar, but the two benzene rings are slightly folded out of the plane of the eightmembered<br />

ring by 2°, a distortion that has been related to packing effects [79].<br />

For 84 the corresponding structural data are: acetylene angle distortion 154°,<br />

(average) distance between the triple bonds 285 pm [80].<br />

Scheme 7.16<br />

387


388 7 Angle-strained Cycloalkynes<br />

The interesting tetrasila analog of 81, the octamethyl derivative 86, is obtained,<br />

when 85 is subjected to flash vacuum pyrolysis at 680 °C or to irradiation [81].<br />

The compound is practically planar, its deformation from linearity at the triple<br />

bonds is smaller than that for 81 (average angle 166°), and the intra-annular<br />

distance between the acetylene moieties has increased to 324 pm, a consequence<br />

of the longer bonds to silicon. When 86 is pyrolyzed a second time, another<br />

die methylsilylene unit is split off and the seven-membered diacetylene 87 is<br />

produced.<br />

Gleiter and co-workers have prepared the symmetrical bis homolog of 81, the<br />

decadiyne 88 [82], and determined its structural parameters. As indicated in<br />

Scheme 7.17 the molecule possesses a chair conformation and the molecular arrangement<br />

at the triple bonds (intraannular distance 299 pm) is clearly nonlinear<br />

(C�C–C angle 171.7°). In the higher homolog, the dodecadiyne 89, the triple bonds<br />

are crossed (crossing angle 24°) and still not quite linear (C�C–C angle 173.8°)<br />

[82]. The introduction of heteroatoms into the carbon framework of 89 has no<br />

significant influence on the structure. In all systems 90 chair-like conformations<br />

prevail, the triple bonds are always slightly bent, the deviation from 180° varying<br />

between 6 and 10°, and the distance between the triple bonds lies between 290<br />

and 310 pm [83].<br />

As expected, ring enlargement causes stronger linearization of the C–C�C–C<br />

fragment. For example, in the 12-membered hydrocarbon 91 the deviation from<br />

180° amounts to 5.9° only [84], and in the 14-membered bis-ether 94, obtained<br />

from the dibromide 92 via the bis-acetylene 93 (Scheme 7.17), the four carbon<br />

atoms lie practically on one single line (C�C–C angle 178.7°) [85].<br />

Numerous cyclic and polycyclic systems have been obtained in the meantime<br />

that contain more than two triple bonds [86, 87]. A classical example is shown in<br />

Scheme 7.18 in which the cyclic polyacetylene 95 is isomerized by base treatment<br />

to a fully conjugated dehydroannulene 96. Although many of these intermediates<br />

and products are highly instable and could not be studied by X-ray crystallography,<br />

in some cases single crystals for X-ray investigations could be obtained, and they<br />

showed that these compounds did indeed contain distorted acetylene or diacetylene<br />

moieties as subunits. For example, in the tetrayne 97, a hydrocarbon possessing a<br />

Scheme 7.17


Scheme 7.18<br />

7.5 Cyclic Polyacetylenes<br />

chair conformation, the angle �C–C�C amounts to 176.7°, whereas the C�C–CH 2<br />

angle is 172.9°; the C1–C11 distance is 309.8 pm, whereas the C2–C10 distance is<br />

slightly extended to 339.0 pm [88]. In the even more strained tetrayne 98 the two<br />

mentioned distortion angles are 166.6° and 166.3°, respectively, and the C1–C10<br />

distance is 269.6 pm, whereas the C2–C9 separation amounts to 323.6 pm [89].<br />

Overall these molecules adopt an ellipsoidal shape.<br />

Under the influence of the discovery of new allotropes of carbon (fullerenes)<br />

there has been a veritable renaissance of annulene chemistry and a huge number<br />

of new carbon-rich compounds has been prepared. Since there are very recent<br />

reviews available in this area, these studies will not be discussed here [90, 91].<br />

A small selection of stable and intermediately produced distorted alkynes are<br />

summarized in Scheme 7.19 to illustrate where the development is going.<br />

The octaacetylene 99 (Scheme 7.19) has, as expected, a planar structure [92]. That<br />

there must be considerable strain in the molecule is indicated by the butadiyne<br />

moieties which have bond angles as low as 164.5°.<br />

More and more examples are reported in the chemical literature in which<br />

distorted alkyne units are incorporated in three-dimensional structures. For<br />

example, the vinyl bromide 100 on treatment with potassium tert-butoxide<br />

evidently yields the (simplest) cyclophyne, 101. This has not only been inferred<br />

from the isolation of its trimer, trifoliaphane 102 [93], but also by a metal complexation<br />

experiment similar to the one discussed above for cyclopentyne (cyclophanes<br />

are presented in Section 4.2). Thus treatment of dibromide 103 with<br />

sodium amalgam in the presence of platinum tris(triphenylphosphine) afforded<br />

the complex 104 in good yield. As shown by X-ray structural analysis, the C–Pt–C<br />

angle amounts to 38.9°, whereas the C1–C2–C3 angle is only 121°. Very likely the<br />

corresponding angle in the precursor acetylene 101 has a similar value [94].<br />

Another phane system with distorted triple bonds is the [2.2]paracyclophane/<br />

dehydroannulene hybrid 105 (Scheme 7.20) [95]. A prominent feature of the [14]<br />

annulene core of this hydrocarbon is the inward bending of the monoyne units<br />

389


390 7 Angle-strained Cycloalkynes<br />

Scheme 7.19<br />

by 4.5–11.8°, whereas the diyne moiety bends outwards by 8.1–11.9°. A similar<br />

structure was observed for the dehydroannulene 106, that corresponds to one deck<br />

of 105. Here the monoyne units are bent inwards by 3.9–11.5° and the diyne unit<br />

is bent outward by 8.6–11.2° [96].<br />

Finally, the [6.6]naphthalenophane 107 shows interesting structural features<br />

in its unsaturated bridges: the angle between the triple bond and the aromatic<br />

carbon atom to which it is bound is 173.8°, and that between it and the neighboring<br />

sp 2 -hybridized carbon atom amounts to 175.5°, their four atoms being<br />

arranged in a transoid fashion, not linear as for the simple model compounds in<br />

Scheme 7.1 [97].<br />

Pericyclynes [98], cyclic compounds in which acetylene units are separated from<br />

each other by a methylene group or a heteroatom, are another interesting class<br />

of polyacetylenes, since they could possess homoaromatic character. Several of<br />

Scheme 7.20


Scheme 7.21<br />

7.5 Cyclic Polyacetylenes<br />

these contain distorted acetylene units. For example (Scheme 7.21) in the thioether<br />

108 the C�C–C angles lie between 172.5° and 173.3° [99]; and in cyclopentayne<br />

109 the corresponding values for the isolated triple bonds range from 170.8° to<br />

178.8°, whereas in the butadiyne moiety the ‘outer’ angle (to the saturated carbon<br />

atom) amounts to 168.2° and the ‘inner’ enclosing three atoms of the triple bonds<br />

is 168.5° [99].<br />

Although cyclic polyacetylenes possessing three, 110 (Scheme 7.21) or more<br />

consecutive triple bonds apparently have not been prepared, the ultimate molecules<br />

in this series are the completely dehydrogenated cyclic polyacetylenes, the cyclo[n]<br />

carbons 111.<br />

There have been several reports on the preparation – or at least the detection – of<br />

this novel allotrope of carbon [100], and the topic has been reviewed several times<br />

[101]. Certainly, the lower members must possess (strongly) distorted acetylene<br />

units. However, because of the instability of the cyclocarbons no experimental<br />

evidence concerning their molecular structures is available. Mass spectrometric<br />

techniques are usually employed to detect them after their generation, for example<br />

by laser evaporation of graphite or from stable organic precursors [102]. To study<br />

this unique form of carbon by electronic and/or vibrational spectroscopy in the<br />

gas phase or under matrix isolation conditions would be extremely interesting<br />

[103]. It has been estimated that the deformation of the bond angle at each sphybridized<br />

carbon atom of cyclo-C 18 112 amounts to ca. 160° [104].<br />

With compound 112 we are ending our journey through the interesting world<br />

of deformed acetylenes that began with another cyclocarbon, the cyclocarbene 9<br />

(Scheme 7.2). Theoretical studies on ethynyl expanded prismanes are discussed<br />

in Section 2.3.3.2.<br />

391


392 7 Angle-strained Cycloalkynes<br />

7.6<br />

Spectroscopic Properties of Angle-strained Cycloalkynes<br />

Although spectroscopic data of selected strained cycloalkynes have been discussed<br />

several times above it seems reasonable to present these in a separate chapter<br />

in comparison to possibly discover any general trends. Needless to say, that this<br />

discussion suffers from the fact that the available spectroscopic information is<br />

not only incomplete but in the cases of the very reactive cycloalkynes impossible<br />

to obtain experimentally.<br />

IR and Raman spectra are available for a sizeable number of cyclohexynes to<br />

cyclononynes, most of them derivatives [12]; they indicate that a reduction in the<br />

C�C–C angle causes a shift to lower wave numbers. A linear correlation between<br />

absorption maxima and experimental and/or calculated C�C–C bond angles has<br />

been proposed; however, strong deviations from this correlation has also been<br />

noted [105]. To rationalize this observation it has been proposed that the hybridization<br />

is changed by the deformation from sp towards sp 2 . This would lead to<br />

a lower force constant for the C�C bond stretching vibration and hence a lower<br />

frequency of absorption [12]. In several cases two or three bands are registered<br />

in the region between 2100 and 2300 cm –1 , and this has been attributed to Fermi<br />

resonance with overtones or combination bands.<br />

Since the triple bond does not carry any hydrogen atom substituents, 1 H NMR<br />

data are of limited importance to probe the angle strain/deformation in small-ring<br />

cycloalkynes. However, 13 C NMR data indicate that increasing the ring strain by<br />

going to smaller ring sizes or introducing additional double bonds into the rings,<br />

shifts the 13 C signals to lower fields. A case in point is 3,3,7,7-tetramethylcycloheptyne<br />

73 which with � = 109.8 ppm for the acetylenic carbon atoms has not only<br />

the highest value for a simple cyclooctyne but is shifted downfield by 22.8 ppm in<br />

comparison to the unstrained reference compound 2,2,5,5-tetramethyl-3-hexyne<br />

[12]. Again, the observed shifts due to ring strain are rationalized as resulting<br />

from a changed hybridization of the acetylenic carbon atoms in the angle-strained<br />

cycloalkynes. In many of the NMR studies reported on these hydrocarbons the<br />

emphasis is on the dynamic behavior of these compounds (flexibility of the<br />

saturated polymethylene bridge).<br />

Hints concerning the orbital structure of angle-strained cycloalkynes may be<br />

derived from photoelectron (PE) and electron transmission (ET) spectra of these<br />

hydrocarbons. According to ab initio [106] and extended Hückel calculations [107]<br />

the energy of the HOMO is hardly effected by cis-bending of a triple bond up to<br />

ca. 30°. In contrast, the LUMO energy is decreased considerably in the cis-bent<br />

models. Experimental studies on various angle-strained cycloalkynes confirm these<br />

predictions. For seven-membered cycloalkynes such as 73 is has been observed<br />

that the otherwise degenerate �-orbitals are split into two PE bands [108, 109].


References<br />

1 C. D. Sherrill, E. F. C. Byrd, M. Head-<br />

Gordon, J. Chem. Phys. 2000, 113, 1447.<br />

2 A. R. Campanelli, A. Arcadi,<br />

A. Domenicano, F. Ramondo,<br />

I. Hargittai, J. Phys. Chem. A, 2006, 110,<br />

2045.<br />

3 S. Szafert, J. A. Gladysz, Chem. Rev.<br />

2003, 103, 4175.<br />

4 R. L. Hayes, E. Fattal, N. Govind,<br />

E. A. Carter, J. Am. Chem. Soc. 2001, 123,<br />

641.<br />

5 R. W. Wetmore, H. F. Schaefer, III,<br />

J. Chem. Phys. 1978, 69, 1648.<br />

6 D. Feller, D. A. Dixon, J. Phys. Chem. A,<br />

2000, 104, 3048.<br />

7 D. Sillars, R. I. Kaiser, N. Galland,<br />

Y. Hannachi, J. Phys. Chem. A, 2003, 107,<br />

5149.<br />

8 T. C. Dinadayalane, U. Deva Priyakumar,<br />

G. Narahari Sastry, J. Phys. Chem. A,<br />

2004, 108, 11433.<br />

9 Y. Amatatsu, M. Hosokawa, J. Phys.<br />

Chem. A, 2004, 108, 10238.<br />

10 F. R. Fronczek, M. S. Erickson, J. Chem.<br />

Cryst., 1995, 25, 737.<br />

11 I. Goldberg, Acta Crystallogr., Sect. B:<br />

Struct. Crystallogr. Cryst. Chem., 1975, 31,<br />

754.<br />

12 The most comprehensive compilations,<br />

including a wealth of physical<br />

(spectroscopic) data of angle-strained<br />

cycloalkynes, are by A. Krebs and<br />

J. Wilke, Top. Curr. Chem. 1983, 109,<br />

189 and H. Meier, Advances in Strain<br />

in Organic Chemistry, Vol. 1 (B. Halton,<br />

Ed.), JAI Press, London, 1991, pp. 215;<br />

for further reviews see: R. P. Johnson<br />

in J. F. Liebman, A. Greenberg (Eds.),<br />

Molecular Structure and Energetics,<br />

Vol. 3, Studies of Organic Molecules,<br />

VCH Publishers, Deerfield Beach, FL,<br />

1986, Chapter 3, pp. 85. The present<br />

review does not describe arynes<br />

(dehydroannulenes); for two leading<br />

modern references on the structure<br />

and new representatives of this class of<br />

compounds see: R. Warmuth, Angew.<br />

Chem. 1997, 109, 1406; Angew. Chem.<br />

Int. Ed. Engl. 1997, 36, 1347; A. Sygula,<br />

R. Sygula, P. W. Rabideau, Org. Lett.<br />

2005, 7, 4999; A. Sygula, R. Sygula,<br />

P. W. Rabideau, Org. Lett. 2006, 8, 5909.<br />

References<br />

13 G.Yu, Y. Ding, X. Huang, C. Sun, J. Phys.<br />

Chem. A, 2005, 109, 1594.<br />

14 R. A. Seburg, E. V. Patterson,<br />

J. F. Stanton, R. J. McMahon, J. Am.<br />

Chem. Soc., 1997, 119, 5847.<br />

15 P. Saxe, H. F. Schaefer, J. Am. Chem.<br />

Soc., 1980, 102, 3239; G. Fitzgerald,<br />

H. F. Schaefer, III, Israel J. Chem.,<br />

1983, 23, 93; M. Rubio, J. Stålring,<br />

A. Bernhardsson, R. Lindh, B. O. Roos,<br />

Theor. Chem. Acc. 2000, 105, 15 and<br />

references cited therein.<br />

16 H. P. Reisenauer, G. Maier, A. Riemann,<br />

R. W. Hoffmann, Angew. Chem. 1984, 96,<br />

596; Angew. Chem. Int. Ed. Engl. 1984,<br />

23, 641.<br />

17 D. J. DeFrees, A. D. McLean, Astrophys.<br />

J., 1986, 308, L31; H. Kanata,<br />

S. Yamamoto, S. Saito, Chem. Phys.<br />

Lett. 1987, 140, 221; D. L. Cooper,<br />

S. C. Murphy, Astrophys. J. 1988, 333,<br />

482; J. Takahashi, K. Yamashita, J. Chem.<br />

Phys. 1996, 104, 6613.<br />

18 Review: H. Hopf, Classics in Hydrocarbon<br />

Chemistry, Wiley-VCH, Weinheim, 2000,<br />

pp. 112.<br />

19 G. Maier, H. Pacl, H. P. Reisenauer,<br />

A. Meudt, R. Janoshek, J. Am. Chem.<br />

Soc. 1995, 117, 12712.<br />

20 D. L. Michalopoulos, M. E. Geusic,<br />

P. R. R. Langridge-Smith, R. E. Smalley,<br />

J. Chem. Phys. 1984, 80, 3556.<br />

21 L. Pandolfo, G. Paiaro, A. Somogyi,<br />

S. Catinella, P. Traldi, P. Rapid Commun.<br />

Mass Spectr., 1993, 7, 132.<br />

22 L. K. Montgomery, J. D. Roberts, J. Am.<br />

Chem. Soc., 1960, 82, 4750; G. Wittig,<br />

E. R. Wilson, Chem. Ber., 1965, 98, 451.<br />

For a recent review on highly unsaturated<br />

cyclobutane derivates see R. P. Johnson<br />

in S. Patai, Z. Rappoport (Eds.)<br />

Chemistry of Cyclobutanes, Vol. 1, Wiley<br />

and Sons, Chichester, 2005, pp. 589.<br />

23 G. Fitzgerald, P. Saxe, H. F. Schaefer, III,<br />

J. Am. Chem. Soc., 1983, 105, 690.<br />

24 J. C. Gilbert, St. Kirschner, Tetrahedron,<br />

1996, 52, 2279.<br />

25 H. A. Carlson, G. E. Quelch,<br />

H. F. Schaefer, III, J. Am. Chem. Soc.<br />

1992, 114, 5344.<br />

26 R. P. Johnson, K. J. Daoust, J. Am.<br />

Chem. Soc., 1995, 117, 362; H. Meier,<br />

393


394 7 Angle-strained Cycloalkynes<br />

M. Schmitt, Tetrahedron Lett., 1989, 30,<br />

5873.<br />

27 K. D. Baumgart, G. Szeimies,<br />

Tetrahedron Lett., 1984, 25, 737.<br />

28 S. Murata, T. Yamamoto, H. Tomioka,<br />

J. Am. Chem. Soc. 1993, 115, 4013.<br />

29 K. B. Wiberg, M. Marquez, J. Am. Chem.<br />

Soc., 1998, 120, 2932.<br />

30 S. L. Buchwald, R. B. Nielsen, Chem.<br />

Rev. 1988, 88, 1047.<br />

31 R. D. Adams, X. Qu, Synlett, 1996,<br />

6, 493 and earlier publications;<br />

cf. B. H. S. Thimmappa, Th. P. Fehiner,<br />

Chemtracts: Inorganic Chemistry, 1993, 5,<br />

173.<br />

32 For a recent review see J. C. Gilbert,<br />

St. Kirschner, In: Advances in Theoretically<br />

Interesting Molecules, Vol. 4 (B. Halton,<br />

Ed.), JAI Press, London, 1998, pp. 203.<br />

33 Review: R. F. C. Brown, Pyrolytic Methods<br />

in Organic Chemistry; Academic Press,<br />

New York, 1980, p. 124.<br />

34 L. Fitjer, S. Modaressi, Tetrahedron Lett.<br />

1983, 24, 5495; cf. L. Fitjer, U. Kliebisch,<br />

D. Wehle, S. Modaressi, Tetrahedron Lett.<br />

1982, 23, 1661.<br />

35 O. L. Chapman, J. Gano, P. R. West,<br />

M. Regitz, G. Maas, J. Am. Chem. Soc.<br />

1981, 103, 7033.<br />

36 For the calculation of the vibrational<br />

spectrum of singlet and triplet<br />

cyclopentyne see S. Olivella,<br />

M. A. Pericas, A. Riera, A. Sole, J. Am.<br />

Chem. Soc. 1986, 108, 6884.<br />

37 J. C. Gilbert, M. E. Baze, J. Am. Chem.<br />

Soc., 1984, 106, 1885; J. C. Gilbert,<br />

M. E. Baze, J. Am. Chem. Soc., 1983,<br />

105, 664; J. C. Gilbert, D.-R. Hou,<br />

J. W. Grimme, J. Org. Chem. 1999, 64,<br />

1529; M.-D. Su, J. Chin. Chem. Soc.<br />

(Taipei, Taiwan), 2005, 52, 599. For a<br />

DFT study of cycloadditions to strained<br />

cycloalkynes see St. M. Bachrach,<br />

J. C. Gilbert, D. W. Laird, J. Am. Chem.<br />

Soc. 2001, 123, 6706.<br />

38 J. M. Bolster, R. M. Kellog, J. Am. Chem.<br />

Soc. 1981, 103, 2868.<br />

39 M. A. Bennett, Pure & Appl. Chem. 1989,<br />

61, 1695; cf. S. L. Buchwald, R. T. Lum,<br />

R. A. Fisher, W. M. Davis, J. Am. Chem.<br />

Soc. 1989, 111, 9113.<br />

40 J. Nakayama, E. Ohshima, A. Ishii,<br />

M. Hoshino, J. Org. Chem. 1983, 48, 60<br />

and lit. quoted.<br />

41 P. G. Gassman, I. Gennick, J. Am. Chem.<br />

Soc. 1980, 102, 6863 and refs. cited.<br />

42 For a more detailed study in the<br />

presence of dihydropyrane as a trapping<br />

reagent see D. W. Laird, J. C. Gilbert,<br />

J. Am. Chem. Soc. 2001, 123, 6704.<br />

43 T. Kitamura, M. Kotani, T. Yokoyama,<br />

Y. Fujiwara, K. Hori, J. Org. Chem. 1999,<br />

64, 680.<br />

44 Review: T. Okuyama, M. Fujita, Acc.<br />

Chem. Res. 2005, 38, 679.<br />

45 N. L. Frank, K. K. Baldridge, J. S. Siegel,<br />

J. Am. Chem. Soc. 1995, 117, 2102.<br />

46 Review: K. Komatsu, Eur. J. Org. Chem.,<br />

1999, 1495.<br />

47 R. W. Hoffmann, Dehydrobenzene and<br />

Cycloalkynes, Academic Press, New<br />

York, 1967; H. G. Viehe, Chemistry of<br />

Acetylenes, Marcel Dekker, New York,<br />

1969; S. Patai (Ed.), The Chemistry of<br />

the Carbon–Carbon Triple Bond, Part<br />

I, John Wiley and Sons, Chichester,<br />

1980; S. Patai (Ed.), The Chemistry of<br />

the Carbon–Carbon Triple Bond, Part 2,<br />

John Wiley and Sons, Chichester, 1980;<br />

Houben-Weyl, Methoden der Organischen<br />

Chemie, Alkine, Di- und Polyine, Allene<br />

und Cumulene, Vol. V/2a, Georg Thieme<br />

Verlag, Stuttgart, 1981; S. Patai, Z. Rappoport<br />

(Eds.), The Chemistry of Triplebonded<br />

Functional Groups, Supplement C,<br />

John Wiley and Sons, Chichester, 1981;<br />

C. Wentrup, Reactive Molecules, Wiley-<br />

Interscience, New York, 1984.<br />

48 G. Wittig, U. Mayer, Chem. Ber. 1963, 96,<br />

342.<br />

49 G. Wittig, G. Harborth, Chem. Ber.,<br />

1944, 77, 306; cf. F. Scardiglia,<br />

J. D. Roberts, Tetrahedron, 1957, 1, 343.<br />

50 G. Wittig, A. Krebs, Chem. Ber. 1961, 94,<br />

3260; cf. G. Wittig, R. Pohlke, Chem. Ber.<br />

1961, 94, 3276.<br />

51 G. J. Baxter, R. F. C. Brown, Austr. J.<br />

Chem. 1978, 31, 327; K. L. Erickson,<br />

J. Wolinsky, J. Am. Chem. Soc. 1965, 87,<br />

1142.<br />

52 M. A. Bennett, G. B. Robertson,<br />

P. O. Whimp, T. Yoshida, J. Am. Chem.<br />

Soc. 1971, 93, 3797; G. B. Robertson,<br />

P. O. Whimp, J. Am. Chem. Soc. 1975,<br />

97, 1051.<br />

53 C. Wentrup. R. Blanch, H. Briehl,<br />

G. Gross, J. Am. Chem. Soc. 1988, 110,<br />

1874.


54 J. Tseng, M. L. McKee, Ph. B. Shevlin,<br />

J. Am. Chem. Soc. 1987, 109, 5474;<br />

S. Olivella, M. A. Pericas, A. Riera,<br />

A. Sole, J. Org. Chem., 1987, 52, 4160.<br />

55 W. Sander. O. L. Chapman, Angew.<br />

Chem. 1988, 100, 402; Angew. Chem. Int.<br />

Ed. Engl. 1988, 27, 398.<br />

56 Bicyclic cyclohexynes: H. Hart,<br />

K. Shahlai, J. Am. Chem. Soc. 1988, 110,<br />

7136; K. Komatsu, S. Aonuma, Y. Jinbu,<br />

R. Tsuji, C. Hirosawa, K. Takeuchi,<br />

J. Org. Chem. 1991, 56, 195.<br />

57 W. Ando, F. Hojo, S. Sekigawa,<br />

N. Nakayama, T. Shimizu, Organometallics,<br />

1992, 11, 1009; F. Hojo,<br />

S. Sekigawa, N. Nakayama, T. Shimizu,<br />

W. Ando, Organometallics, 1993, 12, 803.<br />

58 G. Wittig, J. Meske-Schüller, Liebigs Ann.<br />

Chem. 1968, 711, 65; cf. F. G. Willey,<br />

Angew. Chem. 1964, 76, 144, Angew.<br />

Chem. Int. Ed. Engl. 1964, 3, 138.<br />

Modern methods of preparation:<br />

M. Fujita, Y. Sakanishi, M. Nishii,<br />

T. Okuyama, J. Org. Chem. 2002, 67,<br />

8138; early molecular mechanics<br />

calculations: N. L. Allinger, A. Y. Meyer,<br />

Tetrahedron 1975, 31, 1807.<br />

59 A. Krebs, W. Cholcha, M. Müller, Th.<br />

Eicher, H. Pielartzik, H.-G. Schnöckel,<br />

Tetrahedron Lett. 1984, 25, 5027.<br />

60 A. Krebs, H. Kimling, Angew. Chem.<br />

1971, 83, 540; Angew. Chem. Int. Ed.<br />

Engl. 1971, 10, 509; cf. S. F. Karaev,<br />

A. Krebs, Tetrahedron Lett. 1973, 2853.<br />

61 A. Krebs, H. Kimling, Liebigs Ann. 1974,<br />

2074.<br />

62 J. Haase, A. Krebs, Zeitschr. Naturforsch.,<br />

Teil A, 1972, 27, 624.<br />

63 P. Schulte, G. Schmidt, C.-P. Kramer,<br />

A. Krebs, U. Behrens, J. Organomet.<br />

Chem. 1997, 530, 95; cf. F. Olbrich,<br />

G. Schmidt, U. Behrens, E. Weiss,<br />

J. Organomet. Chem. 1991, 418, 421.<br />

64 J. Wolinsky, K. Erickson, Chem. &<br />

Industry (London), 1964, 47, 1953;<br />

C. B. Reese, A. Shaw, J. Chem. Soc.,<br />

Chem. Commun. 1972, 787.<br />

65 Review: D. Heber, P. Rösner,<br />

W. Tochtermann, Eur. J. Org. Chem.<br />

2005, 4231.<br />

66 L. Brandsma, H. D. Verkruijsse,<br />

Synthesis, 1978, 290.<br />

67 H.-P. Bühl, H. Gugel, H. Kolshorn,<br />

H. Meier, Synthesis, 1978, 536.<br />

References<br />

For the preparation of derivatives of<br />

cyclo octyne using this protocol see<br />

H. Meier, H. Petersen, Synthesis, 1978,<br />

596.<br />

68 M. Traetteberg, W.Lüttke, R. Machinek,<br />

A. Krebs, H. J. Hohlt, J. Mol. Struct.<br />

1985, 128, 217; cf. J. Haase, A. Krebs,<br />

Z. Naturforsch. Teil A, 1971, 26, 1190;<br />

cf. E. Goldstein, B. Ma, J.-H. Lii,<br />

N. L. Allinger, J. Phys. Org. Chem. 1996,<br />

9, 191.<br />

69 G. Wittig, S. Fischer, Chem. Ber. 1972,<br />

105, 3542; M. A. Bennett, I. W. Boyd,<br />

J. Organomet. Chem., 1985, 290, 165;<br />

G. Gröger, U. Behrens, F. Olbrich,<br />

Organometallics, 2000, 19, 3354.<br />

70 H. Meier, M. Schmitt, Tetrahedron Lett.<br />

1989, 30, 5873.<br />

71 Review: H. Meier, N. Hanold,<br />

Th. Molz, H. J. Bissinger, H. Kolshorn,<br />

J. Zountsas, Tetrahedron, 1986, 42, 1711.<br />

72 H. Meier, Y. Dai, Tetrahedron Lett.,<br />

1993, 34, 5277; H. Meier, Y. Dai,<br />

H. Schumacher, H. Kolshorn,<br />

Chem. Ber. 1994, 127, 2035.<br />

73 A. T. Blomquist, R. E. Burge, Jr.,<br />

A. C. Sucsy, J. Am. Chem. Soc., 1952, 74,<br />

3636.<br />

74 V. Typke, J. Haase, A. Krebs, J. Mol.<br />

Struct. 1979, 56, 77.<br />

75 N. L. Allinger, A. Y. Meyer, Tetrahedron,<br />

1975, 31, 1807; cf. J. Sicher, M. Svoboda,<br />

J. Zavada, R. B. Turner, P. Goebel,<br />

Tetrahedron, 1966, 22, 659; F. A. L. Anet,<br />

I. Yavari, Tetrahedron, 1978, 34, 2879;<br />

R. B. Turner, A. D. Jarrett, P. Goebel,<br />

B. J. Mallon, J. Am. Chem. Soc. 1973, 95,<br />

790.<br />

76 H. Meier, H. Petersen, H. Kolshorn,<br />

Chem. Ber. 1980, 113, 2398.<br />

77 E. Kloster-Jensen, J. Wirz, Helv. Chim<br />

Acta, 1975, 58, 162.<br />

78 H. Detert, B. Rose, W. Mayer, H. Meier,<br />

Chem. Ber. 1994, 127, 1529.<br />

79 R. Destro, T. Pilati, M. Simonetta, J. Am.<br />

Chem. Soc. 1975, 97, 658.<br />

80 R. A. G. de Graaff, S. Gorter, C. Romers,<br />

H. N. C. Wong, F. Sondheimer, J. Chem.<br />

Soc. Perkin Trans. 2, 1981, 478.<br />

81 H. Sakurai, Y. Nakadeira, A. Hosomi,<br />

Y. Eriyama, C. Kabuto, J. Am. Chem. Soc.<br />

1983, 105, 3359.<br />

82 R. Gleiter, M. Karcher, R. Jahn, H. Irngartinger,<br />

Chem. Ber. 1988, 121, 735.<br />

395


396 7 Angle-strained Cycloalkynes<br />

83 R. Gleiter, St. Rittinger, H. Irngartinger,<br />

Chem. Ber. 1991, 124, 365; for 1-thia-6azacyclodeca-3,8-diynes<br />

see R. Gleiter,<br />

J. Ritter, H. Irngartinger, J. Lichtenthaler,<br />

Tetrahedron Lett. 1991, 32, 2883; for<br />

1,6-diazacyclodeca-3,8-diynes see<br />

J. Ritter, R. Gleiter, H. Irngartinger,<br />

Th. Oeser, J. Am. Chem. Soc. 1997, 119,<br />

10599; review: R. Gleiter, R. Merger in<br />

Modern Acetylene Chemistry, P. J. Stang,<br />

F. Diederich (Eds.), VCH, Weinheim<br />

1995, pp. 285; cf. R. Gleiter, Angew.<br />

Chem. 1992, 104, 29; Angew. Chem. Int.<br />

Ed. Engl. 1992, 31, 27.<br />

84 R. Gleiter, R. Merger, H. Irngartinger,<br />

J. Am. Chem. Soc. 1992, 114, 8927.<br />

85 R. Gleiter, M. Ramming, H. Weigl,<br />

V. Wolfart, H. Irngartinger, Th. Oeser,<br />

Liebigs Ann./Recueil, 1997, 1545.<br />

86 Review: R. Gleiter, D. B. Werz in<br />

Carbon Rich Compounds, M. M. Haley,<br />

R. R. Tykwinski (Eds.), Wiley-VCH,<br />

Weinheim 2006, pp. 295.<br />

87 Review: F. Sondheimer, Pure Appl.<br />

Chem. 1963, 7, 363; cf. M. Nakagawa in<br />

The Chemistry of Triple-bonded Functional<br />

Groups, S. Patai (ed.), Wiley-Interscience,<br />

New York, 1978, Vol. I, pp. 635.<br />

88 R. Gleiter, R. Merger, J. Chavez, T. Oeser,<br />

H. Irngartinger, H. Pritzkow, B. Nuber,<br />

Eur. J. Org. Chem. 1999, 2841; X-ray of<br />

1-aza-cyclotetradeca-3,5,10,12-tetrayne<br />

derivatives: E. M. Schmidt, R. Gleiter,<br />

F. Rominger, Chem. Eur. J. 2003, 9,<br />

1814 and E. M. Schmidt, R. Gleiter,<br />

F. Rominger, Eur. J. Org. Chem. 2004,<br />

2818.<br />

89 M. Kaftory, I. Agmaon, M. Ladika,<br />

P. J. Stang, J. Am. Chem. Soc. 1987,<br />

109, 782; K. N. Houk, L. T. Scott,<br />

N. G. Rondan, D. C. Spellmeyer,<br />

G. Reinhardt, J. L. Hyun, G. J. DeCicco,<br />

R. Weiss, M. H. M. Chen, L. S. Bass,<br />

J. Clardy, F. S. Jørgensen, T. A. Eaton,<br />

V. Sarkozi, C. M. Petit, L. Ng,<br />

K. D. Jordan, J. Am. Chem. Soc. 1985,<br />

107, 6556; A. G. Myers, S. D. Goldberg,<br />

Angew. Chem. 2000, 112, 2844; Angew.<br />

Chem. Int. Ed. Engl. 2000, 39, 2732.<br />

90 M. M. Haley, R. R. Tykwinski (Eds.),<br />

Carbon-Rich Compounds – from Molecules<br />

to Materials, Wiley-VCH, Weinheim,<br />

2006; V. Maraval, R. Chauvin, Chem.<br />

Rev. 2006, 106, 5317; E. L. Spitler,<br />

C. A. Johnson, II, M. M. Haley, Chem.<br />

Rev. 2006, 106, 5344.<br />

91 M. E. Maier, Synlett, 1995, 13.<br />

92 F. Mitzel, C. Boudon, J.-P. Gisselbrecht,<br />

J.-P. Seiler, M. Gross, F. Diederich, Helv.<br />

Chim. Acta, 2004, 87, 1130.<br />

93 M. Psiorz, H. Hopf, Angew. Chem. 1982,<br />

94, 639; Angew. Chem. Int. Ed. Engl. 1982,<br />

21, 640; cf. C. W. Chan, H. N. C. Wong,<br />

J. Am. Chem. Soc. 1988, 110, 462.<br />

94 K. Albrecht, D. C. R. Hockless,<br />

B. König, H. Neumann, M. A. Bennett,<br />

A. de Meijere, J. Chem. Soc. Chem.<br />

Commun., 1996, 543; cf. B. König,<br />

M. A. Bennett, A. de Meijere, Synlett,<br />

1994, 653.<br />

95 H. Hinrichs, A. K. Fischer, P. G. Jones.<br />

H. Hopf, M. M. Haley, Org. Lett. 2005,<br />

7, 3793; cf. H. Hinrichs, A. J. Boydston,<br />

P. G. Jones, K. Hess, R. Herges,<br />

M. M. Haley, H. Hopf, Chem. Eur. J.<br />

2006, 12, 7103.<br />

96 K. P. Baldwin, A. J. Matzger,<br />

D. A. Scheiman, C. A. Tessier,<br />

K. P. C. Vollhardt, W. J. Youngs, Synlett,<br />

1995, 125.<br />

97 H. Hopf, Chr. Werner, unpublished<br />

results.<br />

98 Reviews: L. T. Scott, M. J. Cooney<br />

in Modern Acetylene Chemistry<br />

(P. J. Stang, F. Diederich, Eds), VCH,<br />

Weinheim, 1995, pp. 321; A. de Meijere,<br />

S. I. Kozhushkov, Topics in Current<br />

Chemistry, Vol. 201, Springer, Berlin,<br />

1999, p. 1.<br />

99 L. T. Scott, private communication.<br />

100 Y. Rubin, M. Kahr, C. B. Knobler,<br />

F. Diederich, C. L. Wilkins, J. Am. Chem.<br />

Soc. 1991, 113, 495.<br />

101 F. Diederich in Modern Acetylene<br />

Chemistry (P. J. Stang, F. Diederich,<br />

Eds), VCH, Weinheim 1995, pp. 443;<br />

F. Diederich, L. Gobbi, Topics in Current<br />

Chemistry, Springer Verlag, Berlin,<br />

Vol. 201, 1999, pp. 43; K. Tahara, Y. Tobe,<br />

Chem. Rev. 2006, 106, 5274.<br />

102 F. Diederich, Y. Rubin, O. L. Chapman,<br />

N. S. Goroff, Helv. Chim. Acta, 1994, 77,<br />

1441; Y. Tobe, T. Fujii, H. Matsumoto,<br />

K. Naemura, Y. Achiba, T. Wakabayashi,<br />

J. Am. Chem. Soc. 1996, 118, 2758;<br />

Y. Tobe, T. Fujii, H. Matsumoto,<br />

K. Tsumuraya, D. Naguchi,<br />

N. Nakagawa, M. Sonoda, K. Naemura,


Y. Achiba, T. Wakabayashi, J. Am. Chem.<br />

Soc. 2000, 122, 1762.<br />

103 M. Grutter, M. Wyss, E. Riaplov, J. P.<br />

Maier, S. D. Peyerimhoff, M. Hanrath,<br />

J. Chem. Phys. 1999, 111, 7397.<br />

104 F. Diederich, Y. Rubin, C. B. Knobler,<br />

R. L. Whetten, K. E. Schriver,<br />

K. N. Houk, Y. Li, Science, 1989, 245,<br />

1088. Apparently cyclocarbons were<br />

first discussed in a publication in<br />

Nachr. Chem. Techn. 1984, 32, 379 in<br />

a (fictional) contribution by H. Musso<br />

and H. Hopf under alias names<br />

(S. Housmans, H. P. Honnef).<br />

References<br />

105 H. Meier, H. Petersen, H. Kolshorn,<br />

Chem. Ber. 1980, 113, 2398.<br />

106 N. G. Rondan, L. N. Domelsmith,<br />

K. N. Houk, A. T. Bowne, R. H. Levin,<br />

Tetrahedron Lett., 1979, 3237.<br />

107 D. M. Hoffmann, R. Hoffmann;<br />

C. R. Fiesel, J. Am. Chem. Soc., 1982,<br />

104, 3858.<br />

108 H. Schmidt, A. Schweig, A. Krebs,<br />

Tetrahedron Lett., 1974, 1471.<br />

109 G. Bieri, E. Heilbronner,<br />

E. Kloster-Jensen, A. Schmelzer,<br />

J. Wirz, Helv. Chim. Acta, 1974, 57,<br />

1265.<br />

397


8<br />

Molecules with Labile Bonds:<br />

Selected Annulenes and Bridged Homotropilidenes<br />

Richard V. Williams<br />

8.1<br />

Introduction<br />

Arguably, molecules with labile bonds are involved in any and every chemical<br />

reaction. Therefore, to be tractable, the scope of this chapter must be limited to<br />

a highly select range of molecules with especially labile bonds – some annulenes<br />

and the bridged homotropilidenes. The emphasis of this review is further refined<br />

to concentrate on the properties and consequences of the possession of labile<br />

bonds. In-keeping with the title ‘<strong>Strained</strong> <strong>Hydrocarbons</strong>’, heteroatoms are only<br />

included as modifiers of the parent hydrocarbon and systems in which they are the<br />

prime motivators are excluded. In general, metal containing compounds are not<br />

considered. The coverage throughout this chapter is limited to material published<br />

subsequent to the last major review in that particular area, with a brief overview of<br />

prior material sufficient to set the newer work in an appropriate context. Access<br />

to the older primary literature is provided, wherever possible, through citations<br />

to appropriate reviews.<br />

8.2<br />

Annulenes<br />

8.2.1<br />

Cyclobutadiene<br />

Interest in the simplest annulene, cyclobutadiene or [4]annulene (1), extends back<br />

for more than 130 years. Sondheimer introduced the term annulenes to describe<br />

‘the completely conjugated monocarbocyclic polyenes, the ring size being indicated<br />

by a number in brackets’ [1]. The salient areas of fascination with cyclobutadiene<br />

have been its structure, whether it has a square planar or rectangular, triplet or<br />

singlet ground state, and its (anti)aromaticity or lack thereof. Its preparation,<br />

chemistry and physical properties are summarized in several excellent reviews<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

399


400 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

[2–8]. In a very recent overview Bally presents the current state-of-the-art in the<br />

cyclobutadiene story as it relates to antiaromaticity [9]. Despite this long history,<br />

the first synthesis of a cyclobutadiene was only achieved in 1965. Petit et al.<br />

prepared 1 by ceric ion oxidation of the corresponding iron tricarbonyl complex<br />

2 [10]. Subsequently, many alternative syntheses of 1 involving the photolysis or<br />

thermolysis of strained precursors such as 3–7 have appeared [11]. 1 is an exceptionally<br />

reactive transient intermediate that can only be isolated in low-temperature<br />

matrices or even in a hemicarcerand at room temperatures [3–8] (see Chapter 10<br />

for discussion on reactions in “molecular flasks”). It dimerizes by an extremely<br />

rapid Diels–Alder self condensation even at cryogenic temperatures, but it can<br />

be trapped in situ to form derivatives with suitable reagents. Highly sterically<br />

encumbered derivatives, e.g. 8–13, are stable at ambient temperature and several<br />

of these compounds have had their X-ray structures determined [3–8].<br />

Cyclobutadienes provide prime examples of the synergy between experiment<br />

and theory. Much of the experimental work was and continues to be driven by<br />

theory, and theory has played a key role in the interpretation and rationalization of<br />

experimental results [3–9]. Coulson recognized that the Hückel molecular orbital<br />

theory predicted square planar (D 4h) triplet ground state of 1 would be subject<br />

to second order Jahn–Teller distortion which should result in a rectangular (D 2h)<br />

singlet ground state [12]. More advanced calculations incorporating electron<br />

correlation favor a rectangular singlet ground state for cyclobutadiene [13]. IR<br />

spectra from matrix isolated cyclobutadiene were initially interpreted as indicating<br />

a square geometry for 1 [17–19]. However, theory and experiment came into<br />

concert when Masamune et al. clearly demonstrated, using a Fourier transform<br />

IR spectrometer, that the IR spectrum of 1 does indeed correspond with a rectangular<br />

geometry. The earlier investigators were hampered by the low intensity<br />

of the signals and were misled by bands due to carbon dioxide (which complexes<br />

with the cyclobutadiene) produced in the photolytic preparation of 1 [4, 20]. In<br />

addition to the extensive discussion of the IR spectra of cyclobutadienes in the<br />

general reviews [4, 6, 7], several detailed analyses of the IR absorptions of cyclobutadienes<br />

have appeared [21–23]. Following the first IR investigations on 1, X-ray<br />

crystal structures confirming rectangular geometries for the stable substituted<br />

cyclobutadienes 8 and 10 became available [24, 25]. Subsequent X-ray structures<br />

determinations of 9, 11 [26] and 13 [27] all reveal a rectangular cyclobutadiene


8.2 Annulenes<br />

Figure 8.1 Several isolated cyclobutadiene derivatives with X-ray determined bond lengths.<br />

core. Examination of Figure 8.1, in which the X-ray determined bond lengths<br />

for 8–11 and 13 are shown, reveals that in 11 the cyclobutadiene core is almost<br />

square. Borden and Davidson rationalized this seeming anomaly by considering<br />

the reduction in steric strain (between the tertiary-butyl groups) upon lengthening<br />

the double bonds, and the reduction in energy upon concomitant shortening of<br />

the formal single bonds [28].<br />

Although the familiar term aromaticity defies a universally acceptable definition<br />

[29, 30, 31], it is generally used to describe completely conjugated cyclic systems<br />

in which (4n + 2) (where n is an integer) electrons are delocalized in a cyclic<br />

array. Such systems enjoy a ‘special stability’ (the aromatic molecule is of lower<br />

energy than a suitable non-aromatic model) due to this aromatic delocalization.<br />

Conversely, conjugated cyclic systems of 4n electrons do not enjoy significant<br />

special stability (non-aromatic) and may even be less stable (antiaromatic) than<br />

the appropriate model compound. It should be noted that geometric and magnetic<br />

criteria are also used in awarding the designation aromatic or anti aromatic to<br />

candidate compounds. It would be preferable if the terms, introduced by Garratt<br />

[1], diatropic (for systems having a diamagnetic ring current, equated with 4n + 2<br />

aromatics) and paratropic (for systems having a paramagnetic ring current,<br />

equated with 4n species which may or may not be destabilized by cyclic delocalization)<br />

were to be universally adopted. Just as benzene is the archetype aromatic<br />

molecule, cyclobutadiene was considered to be the prime example of an antiaromatic<br />

[5, 8, 32]. Recent work has increasingly tended to call this conclusion into<br />

question [9]. It should be noted, Bally and Masamune pointed out in their 1980<br />

review that the high energy of cyclobutadiene could result from several factors,<br />

not just its antiaromaticity, and that the reported degree of ‘negative resonance<br />

energies’ varied between 0 and more than 30 kcal mol –1 [4]. There is no doubt that<br />

1 exhibits properties associated with antiaromaticity and that it should be classed<br />

as antiaromatic [3–9]. However, Mo and Schleyer, in a very detailed analysis of<br />

the thermodynamics of (anti)aromaticity, show that the destabilization of cyclobutadiene<br />

by �-antiaromaticity is rather small. They attribute the majority of the<br />

classical thermodynamic instability of cyclobutadiene to result from strain and<br />

�–� and �–� repulsions [33].<br />

401


402 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

As the ground state of 1 is rectangular, then the two degenerate D 2h forms must<br />

lie on (at least) a double-minimum potential energy surface (PES). These forms<br />

could readily interconvert if the barrier(s) between them is sufficiently low. Balaban<br />

and F�rca�iu introduced the term automerization that they defined as being a particular<br />

case of isomerization in which both the molecular and structural formulae<br />

are conserved in the ‘reactant’ and ‘product’ [34]. Valence bond isomerization<br />

or valence tautomerization equally well describe such processes; however, the<br />

interconversions of 1a � 1b and the rectangular forms of other cyclobutadienes<br />

are most commonly referred to as automerizations. The first indication that such<br />

an automerization is facile came from the NMR spectra of the tri-tertiary-butyl<br />

derivatives 10 and 14. In each case the tertiary-butyl substituents at C2 and C4<br />

gave only one 1 H resonance. For 14 the C1 1 H resonance showed no line broadening<br />

down to 133 K and the C2/C4 13 C single resonance remained sharp to 88 K<br />

giving an estimated �G ‡ of less than 2.5 kcal mol –1 for automerization [35–37].<br />

Interestingly, the ring 1 H resonance in 14 occurs at 1.04 ppm higher field than for<br />

the C2 1 H resonance of cyclopentadiene. This significant upfield shift is taken as<br />

indicative of a paramagnetic ring current in 14 and of supporting its designation<br />

as antiaromatic [36]. The PES for the automerization of 1 (and some derivatives)<br />

has been extensively explored by calculation and there is general agreement that<br />

classically (vide infra) 1a automerizes to 1b through the square planar D 4h 1c [11].<br />

The calculated barrier heights have varied widely [11]. The current best level of<br />

theory (optimized multireference average quadratic coupled cluster, MR-AQCC,<br />

employing various extended correlation consistent basis sets) predicts a barrier<br />

height of 6.3 kcal mol –1 (electronic and zero-point energies) for the 1a � 1b automerization<br />

[16].<br />

Trapping experiments of specifically 1,2-dideuterated cyclobutadiene generated<br />

in solution convincingly demonstrate that cyclobutadiene has a rectangular D 2h<br />

ground state in solution as well as in low temperature matrices [38]. Repeating<br />

these trapping experiments at various temperatures and concentrations allowed<br />

the estimation of the activation barrier (�H ‡ ~1.6–10 kcal mol –1 ) for the automer-


8.3 Cyclooctatetraene<br />

ization [39]. This latter study also revealed a surprising large negative entropy of<br />

activation. Carpenter proposed that his negative entropy of activation could be<br />

rationalized by assuming the automerization of 1 proceeds mostly by heavy-atom<br />

tunneling [40]! This controversial proposal stimulated much research resulting<br />

in a consensus supporting tunneling in the automerization of 1 [41–43]. In an<br />

important experiment, Maier et al. found for the first time in a cyclobutadiene<br />

that the 13 C NMR spectra of the cyclobutadienes 15 are temperature dependent<br />

[44]. Coalescence temperatures were determined for the C2/C4 signals of 15b<br />

and 15c and for the quarternary carbons attached to C2/C4 of 15c. From these<br />

results 15a was estimated to have �G ‡ ~3.5 kcal mol –1 and for 15b and 15c, �G ‡<br />

was determined to be 4.5 and 5.8 kcal mol –1 respectively. These data demonstrate<br />

that the automerization of these substituted cyclobutadienes proceeds by a classical<br />

mechanism and does not involve tunneling.<br />

A further reflection of the lability of cyclobutadienes is their photochemical<br />

valence isomerization to the most strained cage compounds, tetrahedranes [6,<br />

45–48]. The first tetrahedrane, tetra-tertiary-butyltetrahedrane 16a, to be isolated<br />

and fully characterized (including X-ray structure determination) can be prepared<br />

by low-temperature irradiation of 11. Similarly, low-temperature irradiation of 12<br />

and 15a yield tetrahedranes 16b and 16c respectively.<br />

8.3<br />

Cyclooctatetraene<br />

Cyclooctatetraene ([8]annulene, COT) (17) was the next annulene to be isolated<br />

after benzene [49]. Contrary to the expectations of the time that it should exhibit<br />

similar properties (aromaticity) to benzene, it behaves as a simple polyolefin.<br />

There are several earlier and extensive reviews [5, 8, 50–54] of cyclooctatetraenes<br />

(COTs) and one recent general review of the annulenes which makes brief mention<br />

of COTs [55]. In 2001, Klärner reviewed the antiaromaticity of planar COT [56].<br />

Although 17 was originally prepared in 1911 by a low yielding multistep synthesis<br />

[49], it remained relatively unavailable until the advent of the Reppe et al. synthesis<br />

in 1948 [57]. With practical amounts of material then accessible, progress in the<br />

chemistry of COTs was rapid. The infrared and Raman spectra [58–60], and X-ray<br />

403


404 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

[61, 62] and electron diffraction [63, 64] determinations of COT revealed that it<br />

adopts a nonplanar equilibrium conformation. There was initially some dispute<br />

as to whether the structure is of D 2d (tub-shaped), D 4 (crown-shaped with alternating<br />

single and double bonds) or D 4d (crown-shaped with all equivalent bond<br />

lengths). Subsequent electron diffraction studies clearly established a D 2d tub<br />

geometry 17a for COT [65, 66].<br />

There are 21 isomeric (CH) 8 hydrocarbons possible, some of which can be<br />

accessed from COT and many of which lead to COT upon thermolysis or photolysis<br />

[67, 68]. In the context of this review the low temperature processes of valence<br />

isomerization via bond shifting (BS, 17a � 17c), ring inversion (RI, 17a � 17d)<br />

and valence isomerizations (VI, 17a � 18) to bicyclo[4.2.0]octa-2,4,7-triene 18 and<br />

to semibullvalene 19 via high temperature equilibration are of most importance.<br />

It was early recognized that as a consequence of COT’s nonplanar equilibrium<br />

geometry, mono- and appropriately polysubstituted derivatives would be chiral<br />

and that racemization may ensue should the nonplanar ring undergo inversion<br />

with a sufficiently low activation barrier, e.g. 20a � 20d [69, 70]. Similarly, racemization<br />

can also result from the higher energy BS [52–54]. In 1962 Mislow and<br />

Perlmutter reported the first isolation of an optically active COT, the diacid 21 [71].<br />

From the rate of racemization of 21, they estimated the RI activation energy to<br />

be 27 kcal mol –1 . The substituents hinder planarization which doubtless results<br />

in a higher RI activation energy than that for simpler COTs and permitted the<br />

isolation of each enantiomer.<br />

Also in 1962 Anet reported an elegant NMR experiment using the splittings<br />

in the 13 C satellites of the single 1 H resonance of 17 to estimate the barrier for<br />

BS (�G ‡ ~13.7 kcal mol –1 ) in COT, 17a � 17c [72]. The BS activation energy was<br />

later refined by Luz and Meiboom (10.9 kcal mol –1 ) from their dynamic NMR<br />

studies on partially perdeutero (all hydrogens of some molecules replaced with


8.4 Bond Shifting, Ring Inversion and Antiaromaticity<br />

deuterium but not for every molecule in the macroscopic sample) COT in nematic<br />

solvents [73], and again by Naor and Luz in a new analysis of Luz and Meiboom’s<br />

[73] data obtained for C 8H 8 COT in nematic solvents (�H ‡ = 10.0 kcal mol –1 ) [74].<br />

Shortly after studying BS in COT, Anet group in another ingenious NMR experiment<br />

determined the activation parameters for both BS (�G ‡ = 17.1 kcal mol –1 at<br />

–2 °C) and RI (�G ‡ = 14.7 kcal mol –1 at –2 °C) in the alcohol 22 [75]. Of perhaps<br />

greater significance, they postulated that BS involves a planar fully delocalized<br />

bond equalized transition state (TS) 23 while RI proceeds through a planar bond<br />

alternating TS 24. This notion of planar bond equalized or alternating transition<br />

states (TSs) persists today, at least for the parent and simple substituted COTs (vide<br />

infra). Oth surveyed his own results and those of others studying the dynamics<br />

of COTs [76].<br />

Subsequent to these pioneering studies, many syntheses for specifically functionalized<br />

COTs were developed and the BS and RI for a multitude of these COTs<br />

were examined [5, 8, 50–54, 68]. In addition to NMR determinations of the BS and<br />

RI activation parameters, Paquette, in particular, extracted the RI data from detailed<br />

analyses of the complex kinetics for the loss of optical activity in appropriate COTs<br />

and their NMR determined BS values [52–54]. It was found that increasing substitution<br />

of the COT nucleus tended to slow the rates of both BS and RI.<br />

8.4<br />

Bond Shifting, Ring Inversion and Antiaromaticity<br />

The proposal of Anet group that the TS for COT BS is the planar D 8h and for COT<br />

RI the planar D 4h species 17e and 17f respectively [75], appeared to offer the perfect<br />

method of evaluating antiaromaticity in (planar) COT. Assuming BS and RI go<br />

through 17e/f, then the difference in activation barriers for BS and RI (��G ‡ ,<br />

��H ‡ , etc.) directly yields the (classical) resonance energy for the antiaromatic<br />

17e [52]. However, the current view is that both D 8h and D 4h COT are paratropic<br />

and antiaromatic [56, 77]. While 17f appears to be universally accepted as the RI<br />

TS, the veracity of 17e as the BS TS has been questioned. Dewar et al. considered<br />

the experimental ��G ‡ s for a range of COTs (typically 2–4 kcal mol –1 ) to be far<br />

too small compared with their calculated differences in the heats of formation<br />

for 17e/f (13.9, MINDO/2, and 15.4, � approximation, kcal mol –1 ) for 17e to be<br />

the BS TS [78]. They suggested a crown-shaped species similar to 17b may be the<br />

BS TS. Based on symmetry and qualitative energy considerations, Ermer et al.<br />

concluded that neither 17e nor a crown TS were likely and that a flattened saddle<br />

405


406 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

of D 2d symmetry 17g was the best choice for the TS [79]. Paquette group initially<br />

supported 17e as the BS TS [80], but, following extensive additional studies, later<br />

concluded for certain highly substituted or annelated (cyclooctatetraenophanes)<br />

COTs that a planar TS is untenable for the observed rapid BSs [54]. He suggested<br />

that a lower energy pathway would result from pseudorotation of the COT nucleus<br />

to give a flattened saddle TS (analogous to 17g). He left unanswered the question<br />

of the BS TS in the parent COT 17.<br />

Hrovat and Borden, using CASSCF calculations, provided an answer to this<br />

question [81]. They clearly demonstrated that 17e and 17f are the TSs for the<br />

parent COT and that the CASSCF/6-31G* activation BS and RI barriers (including<br />

zero-point correction) calculated at the optimized CASSCF/3-21G* geometry<br />

(CASSCF/6-31G*//CASSCF/3-21G*) are 14.7 and 10.6 kcal mol –1 , respectively.<br />

These barriers (and the difference between them, 4.1 kcal mol –1 ) are in excellent<br />

agreement with experiments. The results from several more recent calculations<br />

at higher levels of theory all support Borden and Hrovat’s earlier results [82–84].<br />

Garavelli et al. optimized 17e and 17f using CASSCF/6-31G* and found activation<br />

barriers of 14.5 and 10.9 kcal mol –1 (including zero-point correction) [84]. Incorporation<br />

of dynamic electron correlation (CASPT2) results in 17e and 17f becoming<br />

essentially degenerate and after zero-point correction 17e is 1.9 kcal mol –1 lower<br />

in energy than 17f results in the authors claim to be within the error bars for<br />

CASPT2. It is of interest to note that the singlet D 8h TS 17e, just as the singlet D 4h<br />

TS 1c for cyclobutadiene automerization, violates Hund’s rule [81].<br />

In concurrence with an earlier CASSCF/6-31G* study of the RI and BS processes<br />

by Castaño et al. [83], Garavelli et al. also propose that there is a bifurcation point<br />

(BP) of D 4h symmetry between the two TSs, 17e and 17f and similarly 17e and 17f �,<br />

and that BS avoids passage through the lower lying 17f/17f �, instead proceeding directly<br />

from the BP to the D 2d COT ground state (Figure 8.2). The path followed from<br />

17e toward 17f to yield D 2d COT is illustrated by the dashed lines in Figure 8.2.


8.4 Bond Shifting, Ring Inversion and Antiaromaticity<br />

Figure 8.2 Bond shifting (BS), ring inversion (RI) showing the proposed bifurcation point (BP)<br />

between 17e and 17f.<br />

The COT PES has also been probed by means of transition state spectroscopy<br />

[85, 86]. Linberger, Borden and coworkers used photodetachment of an electron<br />

from the COT radical anion to generate and study the planar TSs 17e/f by means<br />

of photo electron spectroscopy [87]. They confirmed that 17e and 17f are separated<br />

in energy by 3–5 kcal mol –1 and that, in violation of Hund’s rule, the D 8h singlet<br />

is 12.1 kcal mol –1 lower in energy than the D 8h triplet [87, 88]. Zewail et al. also<br />

generated the D 4h 17f by photodetachment of the COT radical anion and examined<br />

the TS dynamics for RI using pump/probe femtosecond laser pulses and timedependent<br />

mass spectrometric analysis [89].<br />

It has been suggested that heavy atom tunneling is involved in the COT BS, just<br />

as in cyclobutadiene automerization [82, 90]. The small atomic displacements and<br />

activation barrier and the negative entropy of activation for BS are supportive of<br />

the participation of tunneling in BS.<br />

Through suitable substitution the activation barrier for BS can be raised to such<br />

a level as to allow the isolation of ‘shelf-stable’ bond shift isomers [52, 54]. For<br />

example, the annelated derivatives 25a and 25b were the first bond shift isomers<br />

to be isolated and the tetramethyl COTs 26a/26b followed shortly after.<br />

Similarly Ermer et al. predicted, from geometric considerations, and Trindle later<br />

confirmed, from DFT and ab initio calculations, that annelation of COT with small<br />

rings or bicycles would favor a planar conformation of the COT nucleus [79, 91].<br />

407


408 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

The X-ray structure of perfluorotetracyclobuteno-COT (27a) has a planar COT<br />

nucleus with ~D 4h symmetry resulting from the small ring annelation increasing<br />

the internal COT angle from 120° to 135° [92]. The endocyclic (to the cyclobutene)<br />

bonds (en) are on average 0.072 Å shorter than the exocyclic bonds (ex). Baldridge<br />

and Siegel calculated (B3PW91/DZ(2d,p)) that 27a is 17.2 kcal mol –1 more stable<br />

than 27b, while for the corresponding hydrocarbons 28b is 2.3 kcal mol –1 more<br />

stable than 28a [93]. They attribute this reversal in tautomeric stability to the<br />

electronic perturbations induced by perfluorination. They also predicted that<br />

tetrakisbicyclohexeno-COT should be planar with 29b favored by 32.7 kcal mol –1<br />

over 29a. Matsuura and Komatsu confirmed these predictions by the synthesis<br />

of 29b [94]. Considering the 1 H chemical shift of the bridgehead bicyclohexeno<br />

protons, the nucleus independent chemical shifts (NICS) value and the magnetic<br />

susceptibility exaltation of 29, they concluded that 29 is less antiaromatic than the<br />

parent D 4h COT. They attributed this decreased antiaromaticity to the electronic<br />

interaction between the bicyclohexeno-annelating groups and the COT nucleus.<br />

Planar COTs are antiaromatic [56], with the ‘D 4h’ conformer 2–4 kcal mol –1<br />

lower in energy than the ‘D 8h’ in the parent COT and simple derivatives. However,<br />

just as for cyclobutadiene the antiaromatic destabilization is only small and is<br />

certainly not responsible for the nonplanar D 2d tub ground state [77]. Departure<br />

from planarity can be understood in terms of Jahn–Teller distortion from the D 8h<br />

to the D 4h geometry which puckers to the D 2d tub principally to relieve angle strain.<br />

Pseudo-Jahn–Teller coupling has also been invoked to account for the puckering of<br />

the D 4h COT [95]. The Hückel rule (4n + 2 = aromatic, 4n = ‘antiaromatic’) strictly<br />

only applies to planar monocyclic fully conjugated systems. However, Haddon<br />

showed using his �-orbital axis vector (POAV) analysis that conjugation is hardly<br />

interrupted even when the �-system is significantly distorted from planarity, as<br />

in, for example, the various bridged [10]annulenes which are strongly aromatic<br />

despite their considerable nonplanarity [96–98]. Jenneskens et al. determined from<br />

the analysis of current density maps that D 4h COT has a large paramagnetic ring<br />

current and that even as the ring puckers toward the D 2d tub a paramagnetic ring<br />

current is maintained for ~80% of the geometry change [99, 100].


8.5<br />

Valence Isomerization<br />

8.5 Valence Isomerization<br />

At ambient temperature COT is in rapid equilibrium with bicyclo[4.2.0]octa-<br />

2,4,7-triene 18 formed by thermally allowed 6-� disrotatory electrocyclization. At<br />

100 °C ~0.01% of 18 is present in the equilibrium mixture [50, 51, 67, 68, 101,<br />

102]. In agreement with Huisgen and Mietzsch’s experimental determination<br />

of the 17 � 18 activation barrier (28.1 kcal mol –1 ) [101], Gravelli et al. calculated<br />

(CASPT2, zero point corrected) this barrier to be 26.9 kcal mol –1 (Figure 8.3) [84].<br />

Due to its ground state tub conformation, no diene unit is available in COT to<br />

participate in Diels–Alder cycloadditions. However, 18 is an active participant in<br />

Diels–Alder chemistry [50, 51]. Polymethylation and bridging of the COT nucleus<br />

can increase the thermodynamic stability of the bicyclic tautomers. Such stabilization<br />

is apparent for the bridged 30a/b, octamethyl-COT and for tetramethyl 26a<br />

and pentamethyl 31a, but not their BS isomers 26b and 31b [50]. Bicyclo[4.2.0]<br />

octa-2,4,7-triene 18 and its derivatives are doubtless involved in the rich high<br />

temperature chemistry of the COTs [67].<br />

Paquette and coworkers showed that thermolysis of semibullvalene 19 and<br />

some of its derivatives yielded COTs [103]. A later kinetic study of the 19 � 17<br />

equi librium found an activation energy of 39.8 kcal mol –1 with 17 more stable<br />

than 19 (�H 0 = 2.4 kcal mol –1 ) and ~2.7% of 19 at 300 °C [104]. Recent CASPT2<br />

calculations are in remarkably good agreement with these experimental data [84],<br />

and in accord with earlier CASSCF calculations [105], predict a single step reaction<br />

passing through a TS of C 2 symmetry and a bifurcation point from which the two<br />

semibullvalene tautomers result (Figure 8.3).<br />

The position and ease of attaining the semibullvalene � COT thermal equilibrium<br />

is highly dependent on substitution [106, 107]. In general, it is more<br />

common for semibullvalenes to function as synthetic precursors of specifically<br />

substituted COTs than for them to be prepared by this thermal isomerization.<br />

The complex photochemistry of COTs is summarized in these referenced<br />

reviews and articles [51, 67, 84, 108, 109]. A practical synthesis of semibullvalene<br />

19 results from the gas phase (vibrationally hot) [84] photolysis of COT [109].<br />

409


410 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

Figure 8.3 Equilibrium between 17, 18 and 19.<br />

Another (CH) 8 isomer important in the synthesis of semibullvalene is barrelene<br />

32 [110]. Initial interest in barrelene centered on Hine’s proposal that inter action<br />

of its 6-� electrons may lead to aromatic character [111]. Zimmerman, Paufler<br />

et al. were the first to prepare 32 and the discovery of semibullvalene resulted<br />

from the Zimmerman group’s later investigations of its properties [110, 112,<br />

113]. Zimmerman concluded that barrelene is a delocalized system, but, due to<br />

the enforced Möbius �-orbital overlap, it is not stabilized by this delocalization<br />

having the same �-energy as three isolated ethylenes and hence is non-aromatic<br />

[113, 114]. Photoelectron spectroscopy studies confirmed the delocalization and<br />

also suggested a degree of through-bond (hyperconjugative) interaction [115].<br />

The acetone sensitized photolysis of barrelene not only marked the discovery of<br />

semibullvalene but also the discovery of the di-�-methane rearrangement [116].<br />

8.6<br />

Ions Derived from COT<br />

Based on the known ease of reduction of COT, Katz carried out a detailed study of<br />

the COT dianion 33 [117, 118]. He concluded that 33 is planar and aromatic. The<br />

reduction of COTs by both dissolving metals and electrochemically affirms these<br />

conclusions [50, 51, 55]. Despite intensive efforts and its anticipated aromaticity,<br />

the COT dication 34 remains unknown. A few substituted COT dications, e.g., 35a<br />

and 35b, have been prepared at low temperature and deemed to be planar aromatics<br />

[119]. Komatsu and coworkers prepared the bicyclo[2.2.2]octeno annelated COT


8.7 The Higher Annulenes<br />

dication 36 which is stable at room temperature, has a tub conformation and<br />

undergoes RI with �G ‡ ~10.8 kcal mol –1 [120, 121].<br />

Protonation of COT or reaction with some other electrophiles yields the homotropylium<br />

cation 37 or one of its derivatives [50, 51]. Homotropylium cations are the<br />

most extensively studied and best established homoaromatics [122, 123]. The idea<br />

of homoaromaticity, first proposed by Roberts [124] and Doering and coworkers<br />

[125] in 1956, was generalized 3 years later by Winstein [126] who actually introduced<br />

the term ‘homoaromaticity’. If the cyclic conjugation of an aromatic<br />

system is interrupted by one or more saturated units (often a CH 2 group) and the<br />

resulting molecule still enjoys ‘special stability’, then it is termed homoaromatic<br />

[122, 123, 127–134]. Thus a molecule will be considered homoaromatic if one<br />

(or more) through space interactions complete the cyclic delocalization of 4n + 2<br />

electrons and that this through space interaction is energy lowering. Winstein<br />

termed homoaromatic systems with one interruption to the cyclic conjugation<br />

monohomoaromatic and those with multiple interruptions, bis-, tris-, and tetrahomoaromatic<br />

etc. [127, 128].<br />

8.7<br />

The Higher Annulenes<br />

Two recent reviews [55, 135] and earlier articles [1, 5] provide extensive coverage<br />

of the higher annulenes. The focus of interest in these annulenes continues to<br />

be the study of their aromaticity, their structures and their complex conformational<br />

and configurational isomerism. All annulenes are known between [4]- and<br />

[30]annulene except for [26]- and [28]annulenes. Most of the higher ([10]- and larger<br />

for the purposes of this review) annulenes are examples of systems with labile<br />

bonds and exist in a wide variety of configurations and conformations [55, 135].<br />

In addition to reviewing the dynamics of COTs, Oth also reviewed the dynamics<br />

of [12]-, [14]-, [16]- and [18]annulenes as determined by variable-temperature NMR<br />

studies [76]. It has long been held that: (1) the 4n + 2 annulenes are diatropic with<br />

bond-equalized structures while the singlet planar 4n annulenes are paratropic<br />

and bond alternating [30]; (2) as the annulene ring size increases, bond length<br />

alternation will set in. The position of onset of bond alternation was debated<br />

with a mild consensus in favor of about [20]annulene [55, 135]. Schleyer et al.<br />

have challenged this uneasy truce by asserting both [14]- and [18]annulene have<br />

significantly bond alternating structures [136]. X-ray structures for [18]annulene<br />

determined at 80 K [137] and 111 K [138] indicate a bond equalized D 6h geometry.<br />

411


412 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

Schleyer et al. suggest that these structures are incorrect as a consequence of<br />

insoluble disorder problems [136]. In a commentary on the Schleyer et al. paper,<br />

Ermer concludes that if there is indeed disorder, it may be possible to solve these<br />

problems experimentally [139]. He also pointed out that the Schleyer et al. data<br />

apply to the gas and solution phases and postulated that crystal packing forces<br />

may favor the D 6h geometry.<br />

Modeling [10]annulene computationally has proved to be difficult. Castro,<br />

Karney et al. summarize previous work in this area and examine the interconversion<br />

of various [10]annulene isomers [140]. They support, from their BH&HYLP<br />

and coupled cluster calculations, Masamune’s [141] NMR based structural assignments<br />

for the two crystalline compounds mono-trans 38 and all-cis 39 isolated by<br />

his group and his proposed equilibration of 38 with heart-shaped 40.<br />

As already discussed for cyclobutadiene and COT, the higher annulenes can<br />

also undergo bond shift isomerization. This bond shifting along with the conformational<br />

and other configurational equilibria [1, 55, 76] of [12]-, [14]- and<br />

[16]annulenes were re-examined computationally by Castro and Karney et al.<br />

[142–144]. They concluded that the bond shifting, 41 � 42, 43 � 44 and 45 � 46,<br />

proceeds through a Möbius transition state (antiaromatic for the [14]- and aromatic<br />

for the [12]- and [16]annulenes). The recently revived interest in and study of<br />

Möbius systems is reviewed by Herges [145].<br />

Bridged derivatives are known for most of the higher annulenes and these<br />

derivatives are discussed in the two most recent reviews on annulenes [55, 135].<br />

In the majority of these annulenes the bridging group is a methano group, e.g.,<br />

the methano-bridged [10]annulene (47 and 48) and [14]annulene (49 and 50).<br />

At ambient temperature, 1,6-methano[10]annulene 47 is in equilibrium with a<br />

very small amount of the isomeric norcaradiene 51 [146–152]. Substitution on


8.8 Bridged Homotropilidenes<br />

C11 tends to increase the stability of the norcaradienic form and, reaffirming the<br />

homoaromatic transannular (C1,C6) interaction [123], it has even been suggested<br />

that the dimethyl compound 52 is nonclassical (a barrierless flat PES linking<br />

52a � 52b) [153].<br />

Dimethyldihydropyrene 53, another bridged [14]annulene, has long served as<br />

an excellent NMR probe for aromaticity [154, 155] and is now being studied for its<br />

potential applications as a photoswitch [156]. The dark green 53 is photo-bleached<br />

by visible light to give the colorless cyclophanediene 54 which undergoes ready<br />

thermal, although forbidden, electrocyclization to return 53. A recent computational<br />

investigation (UB3LYP/6-31G*) revealed the transition state for this 6-�<br />

conrotatory electrocyclization is biradical-like accounting for the low thermal<br />

activation barrier [157].<br />

8.8<br />

Bridged Homotropilidenes<br />

The trivial name for 1,3,5-cycloheptatriene 55 is tropilidene and hence homotropilidene<br />

is bicyclo[5.1.0]octa-2,5-diene 56. Doering reasoned that the Cope<br />

rearrangement in homotropilidene, via the cis conformer 56b, would be greatly<br />

accelerated compared with the parent 1,5-hexadiene due to the ring strain induced<br />

upon cyclopropanation and to the restricted conformational mobility of the vinyl<br />

groups [158]. The homotropilidene degenerate Cope rearrangement proved to<br />

be exceptionally rapid (�H ‡ 21.4 kcal mol –1 less than for 1,5-hexadiene), which<br />

lead Doering to introduce a new term ‘fluctional’ now commonly fluxional [159,<br />

160]. He defined fluxional (structure), which refers to a dynamic system and is<br />

413


414 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

not to be confused with resonance hybrid, as ‘an organic molecule which must be<br />

described as the mean between two identical structures’. The next logical development<br />

in designing molecules capable of ever more rapid Cope rearrangement<br />

was to restrict conformational mobility even further and to more closely mimic<br />

the required boat conformation for the Cope TS by imposing a bridge between<br />

C4,C8 as in 57 [158–160].<br />

Again expectations were met; the bridged homotropilidenes, bullvalene 58<br />

[161], barbaralane 59 [162], barbaralone 60 [162] and semibullvalene 19 [110], all<br />

undergo degenerate Cope rearrangement much more rapidly than 56. Remarkably,<br />

by a series of Cope rearrangements every hydrogen and every carbon atom<br />

in bullvalene become equivalent. The order of reactivity is 19 > 59 > 60 > 58.<br />

Thus began the quest for a neutral bishomoaromatic bridged homotropilidine<br />

[123, 132, 163, 164].<br />

It was suggested long ago that the semibullvalene nucleus is the system which<br />

most closely approaches the elusive goal of neutral homoaromaticity [165]. Semibullvalene<br />

19 was first prepared by Zimmerman and Grunewald in 1966 [110].<br />

With the low field NMR spectrometer available to them at that time the degenerate<br />

Cope rearrangement 19a � 19b could not be frozen out. Later, Anet group using a<br />

251 MHz ( 1 H) NMR spectrometer were able to achieve temperatures lower than<br />

the coalescence temperature and consequently determined the activation barrier<br />

for this Cope rearrangement (�G ‡ 5.5 kcal mol –1 , �H ‡ 4.8 kcal mol –1 ) [166]. The<br />

value for the activation barrier was later refined using 13 C dynamic NMR data<br />

(�G ‡ 6.2 kcal mol –1 , �H ‡ 5.2 kcal mol –1 ) [167]. It is generally accepted that the<br />

C 2v symmetric homoaromatic species 19c is the transition state for this process<br />

[163, 164].<br />

Calculations by Dewar and Lo [168] and Hoffmann and Stohrer [169] suggested<br />

that substitution of the semibullvalene nucleus with electron withdrawing groups<br />

at the 2, 4, 6, and 8 positions and electron donating groups at the 1 and 5 positions<br />

would lead to a decrease in the activation energy for the Cope rearrangement<br />

through stabilization of the TS. Many syntheses of substituted semibullvalenes<br />

and barbaralanes were developed, and in agreement with Dewar’s and Hoffmann’s<br />

predictions the activation barrier was lowered [163, 164]. However, no example has<br />

yet been prepared in which this ‘barrier’ becomes negative resulting in the delocalized<br />

species, analogous to 19c, becoming the ground state [123, 163, 164].


8.9 Recent Developments<br />

In an alternative approach to eliminating the barrier to Cope rearrangement in<br />

semibullvalenes, small ring annelation was calculated to destabilize the localized<br />

forms and to favor the delocalized form, e.g., 61, 62, 63 and 64 [123, 132, 163,<br />

164]. Again no neutral homoaromatic ground state bridged homotropilidene has<br />

yet been characterized experimentally.<br />

8.9<br />

Recent Developments<br />

Bullvallene. This was used as a starting material in the synthesis of triaza- and tri thia-<br />

[3]-peristylanes 65 [170] and 66 [171] and also the bullvalene trisepoxide 67 [172].<br />

Barbaralane. Reiher and Kirchner and Kirchner and Sebastiani computationally<br />

examined the tetraphosphabarbaralane 68, and briefly the corresponding tetraazaanalog<br />

and the parent 59 [173, 174]. They conclude that 68 is homoaromatic with<br />

a C 2v ground state which is not biradicaloid. Using molecular dynamics simulations<br />

they demonstrate that at finite temperatures (as opposed to the usual 0 K)<br />

68 still prefers a C 2v ground state and that distortion (in a Cope rearrangement<br />

sense) to degenerate C s forms is a vibrational mode (no energy barrier) and not a<br />

reaction. Using molecular dynamics modeling they present 3-dimensional NICS<br />

analysis confirming the homoaromatic nature of 68. By (computationally) linking<br />

together two or more barbaralanes, e.g. 69 and 70, Tantillo et al. arrived at a new<br />

class of compounds – �-polyacenes [175]. These extended barbaralanes were<br />

computed to have completely delocalized bishomoaromatic singlet ground states<br />

and again are predicted not to be biradicaloid. The authors originally envisaged<br />

these extended barbaralanes as potential sigmatropic shiftamers in the sense of<br />

69a � 69b [175, 176].<br />

415


416 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

Semibullvalene. In addition to semibullvalenes, many of the studies considered here<br />

also discuss barbaralanes. Some semibullvalenes and barbaralanes, with extremely<br />

low activation barriers to the Cope rearrangement, are thermochromic and solvatochromic<br />

[177]. The observed solvatochromism is attributed to the changing concentration<br />

of the delocalized homoaromatic species. Time-dependent density functional<br />

theory (TD-DFT) calculations confirm that the long-wavelength absorptions,<br />

responsible for the color of these compounds, arise solely from the delocalized<br />

species. For the semibullvalene 71 and barbaralanes 72 and 73, increasing solvent<br />

dipolarity (dipolar solvents possess a permanent dipole moment whereas polar<br />

solvents are characterized by a significant dielectric constant [178]) increases the<br />

concentration and correspondingly the thermodynamic stability of the delocalized<br />

forms. The stabilization of the delocalized form by dipolar solvents is so successful<br />

with 72 that it goes from a localized (in most solvents) to a delocalized homoaromatic<br />

solvate ground state in the highly dipolar solvent N,N�-dimethylpropylene<br />

urea [177]. In complete contrast, the bisanhydride 64 appears to follow the opposite<br />

trend with the delocalized form increasing in concentration with decreasing solvent<br />

polarity, perhaps supporting the notion that 64 is homoaromatic in the gas phase<br />

[164, 179]. Various self-consistent reaction field (SCRF) methods were used to<br />

model the effect of solvent on these bridged homotropilidenes [177]. While the<br />

results of these calculations indicated small selective solvent stabilizations of the<br />

delocalized over localized forms for 71–73, the calculated solvent effects on 64 were<br />

negligible. A gas phase electron diffraction structure determination of 64 resulted<br />

in three models, a delocalized homoaromatic structure, a localized structure or a<br />

2 : 1 mix of localized:delocalized forms, none of these possibilities could be ruled<br />

out as they all fit the diffraction data [179].


8.9 Recent Developments<br />

Confirming suggestions in an earlier review [163], calculations on 19, 58, 59<br />

and dihydrobullvalene 74 reveal that semibullvalene rearranges most rapidly of<br />

the group as it enjoys the greatest relief of strain in going from the localized form<br />

to the homoaromatic TS and that the degree of aromaticity in the TS is greatest<br />

for barbaralane (and least for semibullvalene!) [180].<br />

Recently Brown, Henze and Borden used unrestricted DFT, CASSCF and<br />

CASPT2 to reinvestigate parent semibullvalene 19 and some derivatives (61 and<br />

75) along with the new system, epoxide 76, with a view to determining if the C 2v<br />

symmetric species were necessarily homoaromatic [181]. They concluded that<br />

the C 2v geometries of 19, 75 and 76 enjoy homoaromatic stabilization while that<br />

of 61 does not and that these C 2v species are the ground states for 61, 75 and 76.<br />

Both 61 and 76 were calculated to possess significant triplet character and 61 was<br />

predicted to have a triplet ground state.<br />

Calculations on a series of substituted semibullvalenes 71, 77–82 revealed that<br />

the B3LYP/6-31G* method tends to overestimate the stability of the delocalized<br />

homoaromatic species by up to about 3 kcal mol –1 compared with experimental<br />

values determined in solution [182]. In agreement with experiment, two phenyl<br />

and two cyano groups 80 were found to be more effective in stabilizing the delocalized<br />

form than four phenyl 82 or four cyano 81 substituents. Derivatives 80,<br />

81 and 82 were all predicted to have homoaromatic ground states. Subsequent<br />

(TD) B3LYP/6-31G* calculations on semibullvalenes 19, 71, 78, 79, 80 and 83<br />

and barbaralanes 59, 72, 73, 84, 85 and 86 reaffirmed that the delocalized species<br />

are responsible for the long wavelength electronic absorptions observed in the<br />

visible region for some of these compounds [183]. The position of the calculated<br />

and observed IR bands are in good agreement. 80 was predicted to be blue and<br />

to be the best target for synthesis of a neutral bishomoaromatic.<br />

Sauer et al. published full experimental details for their ingenious one-pot<br />

syntheses of a wide variety of 1,5-dimethyl-3,7-disubstitued semibullvalenes<br />

417


418 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

Scheme 8.1<br />

(Scheme 8.1) [163, 184], which prompted Zhou and Birney to computationally<br />

examine the fragmentation of the proposed intermediate diazo compounds 87 to<br />

semibullvalenes 88 [185]. Zhou and Birney concluded that the fragmentation led<br />

to a bifurcation point and from there to the localized semibullvalenes and thus<br />

avoiding the Cope TS for these semibullvalenes. In another one-pot synthesis, Xi<br />

group developed a copper (I) mediated approach to octasubstituted semibullvalenes<br />

from 1,4-diiodobutadienes (Scheme 8.2) [186]. Thermolysis of the resulting<br />

octasubstituted semibullvalenes provides octasubstituted COTs.<br />

Scheme 8.2<br />

In an extension of Gompper’s modification of the Saunders’ isotopic perturbation<br />

method [187, 188], Quast, Williams et al. studied, by variable-temperature<br />

13 C NMR spectroscopy in a variety of solvents, the nondegenerate Cope rearrangement<br />

of a series of unsymmetrically substituted semibullvalenes (89–96)<br />

[189]. They developed a new treatment in which the effects of the substituents on<br />

chemical shift were specifically accounted for and determined the thermodynamic<br />

quantities for these skewed equilibria. The Cope equilibrium could only be frozen<br />

out for 89 and it was determined that in 89 the preferred valence tautomer is the<br />

one with the ethyl group on the cyclopropane unit.


8.9 Conclusions<br />

Jiao and coworkers used the isolobal relationship of a boron carbonyl (BCO)<br />

moiety to a CH group in their design of potentially homoaromatic semibullvalenes,<br />

barbaralanes and bullvalenes [190]. Their calculations (B3P86/6-311+G**)<br />

indicate that 97 and 98 should be homoaromatic. The other BCO ‘replaced’ bridged<br />

homotropilidines under investigations showed reduced barriers to their Cope<br />

rearrangement compared to their respective hydrocarbons, but still possessed a<br />

localized ground state.<br />

8.9<br />

Conclusions<br />

The rapid development and current affordability of computer hardware and<br />

software has resulted in an unprecedented ability to carry out ever more sophisticated<br />

calculations on increasingly bigger molecules. These developments have<br />

had a major impact in the area of molecules with labile bonds. In all the systems<br />

discussed in this chapter, theory has played a leading role in identifying synthetic<br />

targets and important experimental studies and the rationalization and interpretation<br />

of experimental results. Through this synergy between experiment and<br />

theory, the complex properties of cyclobutadiene and cyclooctatetraene are now<br />

well understood. Enticing new results resurrect old questions about the annulenes:<br />

Where does natural bond alternation begin and just how (anti)aromatic are<br />

they? The location of Möbius transition states for annulene isomerization nicely<br />

addresses the long-standing and vexing question of why these formal cis � trans<br />

isomerizations are so facile. The cyclophanediene � dihydropyrene equilibrium<br />

is now better understood and predictions regarding substituent effects on the<br />

barrier to this electrocyclization are being tested experimentally. Many barbaralanes<br />

and semibullvalenes with miniscule experimental barriers to their Cope<br />

rearrangement are now known. The thermochromism and solvatochromism of<br />

these bridged homotropilidenes has been thoroughly investigated and even a<br />

semibullvalene solvate with a homoaromatic ground state found. Again, through<br />

calculations, several ground state bishomoaromatic semibullvalenes have been<br />

identified. Now all that remains is the synthesis and experimental verification of<br />

the first neutral bishomoaromatic!<br />

419


420 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

References<br />

1 Sondheimer, F. Acc. Chem. Res. 1972, 5,<br />

81.<br />

2 Cava, M. P.; Mitchell, M. J. Cyclobutadiene<br />

and Related Compounds; Academic<br />

Press: New York, 1967.<br />

3 Maier, G. Angew. Chem. Int. Ed. 1974, 13,<br />

425.<br />

4 Bally, T.; Masamune, S. Tetrahedron<br />

1980, 36, 343.<br />

5 Balaban, A. T.; Banciu, M.; Ciorba, V.<br />

Annulenes, Benzo-, Hetero-, Homo-<br />

Derivative and their Valence Isomers,<br />

Vol. 1; CRC Press, Inc.: Boca Raton,<br />

1987.<br />

6 Maier, G. Angew. Chem. Int. Ed. 1988, 27,<br />

309.<br />

7 Arnold, B. R.; Michl, J. In Kinetics and<br />

Spectroscopy of Carbenes and Biradicals;<br />

Platz, M. S., Ed.; Plenum Press: New<br />

York, 1990, p. 1.<br />

8 Minkin, V. I.; Glukhovtsev, M. N.;<br />

Simkin, B. Y. Aromaticity and Antiaromaticity<br />

electronic and Structural<br />

Aspects; John Wiley & Sons, Inc., New<br />

York, 1994.<br />

9 Bally, T. Angew. Chem. Int. Ed. 2006, 45,<br />

6616.<br />

10 Watts, L.; Fitzpatrick, J. D.; Petit, R.<br />

J. Am. Chem. Soc. 1965, 87, 3253.<br />

11 See references 2–7 and references cited<br />

therein.<br />

12 Coulson, C. A. Chem. Soc. (London)<br />

Spec. Publ. 1958, No. 12, 85.<br />

13 The results from earlier calculations<br />

are summarized in references 2–8 and<br />

12. For more recent calculations see<br />

references [14–16] and references cited<br />

therein.<br />

14 Carsky, P.; Bartlett, R. J.; Fitzgerald, G.;<br />

Noga, J.; Spirko, V. J. Chem. Phys. 1988,<br />

89, 3008.<br />

15 Cremer, D.; Kraka, E.; Joo, H.;<br />

Stearns, J. A.; Zwier, T. S. PCCP 2006, 8,<br />

5304.<br />

16 Eckert-Maksic, M.; Vazdar, M.;<br />

Barbatti, M.; Lischka, H.; Maksic, Z. B.<br />

J. Chem. Phys. 2006, 125, 064310.<br />

17 Lin, C. Y.; Krantz, A. J. Chem. Soc.,<br />

Chem. Commun. 1972, 1111.<br />

18 Chapman, O. L.; McIntosh, C. L.;<br />

Pacansky, J. J. Am. Chem. Soc. 1973, 95,<br />

614.<br />

19 Krantz, A.; Lin, C. Y.; Newton, M. D.<br />

J. Am. Chem. Soc. 1973, 95, 2744.<br />

20 Masamune, S.; Souto-Bachiller, F. A.;<br />

Machiguchi, T.; Bertie, J. E. J. Am. Chem.<br />

Soc. 1978, 100, 4889.<br />

21 Hess, B. A., Jr.; Schaad, L. J.; Carsky, P.;<br />

Zahradnik, R. Chem. Rev. 1986, 86, 709.<br />

22 Redington, R. L. J. Chem. Phys. 1998,<br />

109, 10781.<br />

23 Hess, B. A., Jr.; Schaad, L. J. Mol. Phys.<br />

2000, 98, 1107.<br />

24 Irngartinger, H.; Rodewald, H. Angew.<br />

Chem. Int. Ed. 1974, 13, 740.<br />

25 Delbaere, L. T. J.; James, M. N. G.;<br />

Nakamura, N.; Masamune, S. J. Am.<br />

Chem. Soc. 1975, 97, 1973.<br />

26 The results for 9 and 11 are discussed in<br />

reference [6].<br />

27 Sekiguchi, A.; Tanaka, M.; Matsuo, T.;<br />

Watanabe, H. Angew. Chem. Int. Ed.<br />

2001, 40, 1675.<br />

28 Borden, W. T.; Davidson, E. R.<br />

J. Am. Chem. Soc. 1980, 102, 7958.<br />

29 For an excellent discussion of the<br />

problems associated with defining<br />

aromaticity see reference 8.<br />

30 Schleyer, P. v. R.; Jiao, H. Pure Appl.<br />

Chem. 1996, 68, 209.<br />

31 See also the Thematic Issue of Chemical<br />

Reviews, May 2001, 101, Number 5,<br />

Guest Editor Schleyer, P. v. R., on<br />

Aromaticity and the multitude of<br />

references cited therein.<br />

32 Wiberg, K. B. Chem. Rev. 2001, 101, 1317.<br />

33 Mo, Y.; Schleyer, P. v. R. Chem. Eur. J.<br />

2006, 12, 2009.<br />

34 Balaban, A. T.; F�rca�iu, D. J. Am. Chem.<br />

Soc. 1967, 89, 1958.<br />

35 Maier, G.; Alzerreca, A. Angew. Chem.<br />

Int. Ed. 1973, 12, 1015.<br />

36 Masamune, S.; Nakamura, N.; Suda, M.;<br />

Ona, H. J. Am. Chem. Soc. 1973, 95,<br />

8481.<br />

37 Maier, G.; Schaefer, U.; Sauer, W.;<br />

Hartan, H.; Matusch, R.; Oth, J. F. M.<br />

Tetrahedron Lett. 1978, 1837.<br />

38 Whitman, D. W.; Carpenter, B. K.<br />

J. Am. Chem. Soc. 1980, 102, 4272.<br />

39 Whitman, D. W.; Carpenter, B. K.<br />

J. Am. Chem. Soc. 1982, 104, 6473.<br />

40 Carpenter, B. K. J. Am. Chem. Soc. 1983,<br />

105, 1700.


41 See for example, references 7, 14, 22, 35<br />

and 37 and references cited therein.<br />

42 Orendt, A. M.; Arnold, B. R.;<br />

Radziszewski, J. G.; Facelli, J. C.;<br />

Malsch, K. D.; Strub, H.; Grant, D. M.;<br />

Michl, J. J. Am. Chem. Soc. 1988, 110,<br />

2648.<br />

43 Zuev, P. S.; Sheridan, R. S.; Albu, T. V.;<br />

Truhlar, D. G.; Hrovat, D. A.;<br />

Borden, W. T. Science 2003, 299, 867.<br />

44 Maier, G.; Wolf, R.; Kalinowski, H. O.<br />

Angew. Chem. Int. Ed. 1992, 31, 738.<br />

45 Maier, G.; Neudert, J.; Wolf, O. Angew.<br />

Chem. Int. Ed. 2001, 40, 1674.<br />

46 Maier, G.; Neudert, J.; Wolf, O.;<br />

Pappusch, D.; Sekiguchi, A.; Tanaka, M.;<br />

Matsuo, T. J. Am. Chem. Soc. 2002, 124,<br />

13819.<br />

47 Tanaka, M.; Sekiguchi, A. Angew. Chem.<br />

Int. Ed. 2005, 44, 5821.<br />

48 Nemirowski, A.; Reisenauer, H. P.;<br />

Schreiner, P. R. Chem. Eur. J. 2006, 12,<br />

7411.<br />

49 Willstätter, R.; Waser, E. Berichte der<br />

Deutschen Chemischen Gesellschaft 1911,<br />

44, 3423.<br />

50 Paquette, L. A. Tetrahedron 1975, 31,<br />

2855.<br />

51 Fray, G. I.; Saxton, R. G. The Chemistry<br />

of Cyclooctatetraene and Its Derivatives;<br />

Cambridge University Press: Cambridge,<br />

1978.<br />

52 Paquette, L. A. Pure Appl. Chem. 1982,<br />

54, 987.<br />

53 Paquette, L. A. Advances in Theoretically<br />

Interesting Molecules 1992, 2, 1.<br />

54 Paquette, L. A. Acc. Chem. Res. 1993, 26,<br />

57.<br />

55 Spitler, E. L.; Johnson, C. A., II;<br />

Haley, M. M. Chem. Rev. 2006, 106, 5344.<br />

56 Klärner, F.-G. Angew. Chem. Int. Ed.<br />

2001, 40, 3977.<br />

57 Reppe, W.; Schlichting, O.; Meister, H.<br />

Ann. 1948, 560, 93.<br />

58 Lippincott, E. R.; Lord, R. C., Jr.<br />

J. Am. Chem. Soc. 1946, 68, 1868.<br />

59 Flett, M. S. C.; Cave, W. T.; Vago, E. E.;<br />

Thompson, H. W. Nature (London) 1947,<br />

159, 739.<br />

60 Lippincott, E. R.; Lord, R. C., Jr.;<br />

McDonald, R. S. J. Chem. Phys. 1948, 16,<br />

548.<br />

61 Kaufman, H. S.; Fankuchen, I.; Mark, H.<br />

J. Chem. Phys. 1947, 15, 414.<br />

References<br />

62 Kaufman, H. S.; Fankuchen, I.; Mark, H.<br />

Nature (London) 1948, 161, 165.<br />

63 Bastiansen, O.; Hassel, O.; Langseth, A.<br />

Nature (London) 1947, 160, 128.<br />

64 Bastiansen, O.; Hassel, O. Acta Chem.<br />

Scand. 1949, 3, 209.<br />

65 Karle, I. L. J. Chem. Phys. 1952, 20, 65.<br />

66 Bastiansen, O.; Hedberg, L.; Hedberg, K.<br />

J. Chem. Phys. 1957, 27, 1311.<br />

67 Hassenrück, K.; Martin, H. D.; Walsh, R.<br />

Chem. Rev. 1989, 89, 1125.<br />

68 Gajewski, J. J. Hydrocarbon Thermal<br />

Isomerizations, Second Edn.; Elsevier<br />

Academic Press: Amsterdam, 2004.<br />

69 Cope, A. C.; Kinter, M. R. J. Am. Chem.<br />

Soc. 1951, 73, 3424.<br />

70 Cope, A. C.; Burg, M.; Fenton, S. W.<br />

J. Am. Chem. Soc. 1952, 74, 173.<br />

71 Mislow, K.; Perlmutter, H. D. J. Am.<br />

Chem. Soc. 1962, 84, 3591.<br />

72 Anet, F. A. L. J. Am. Chem. Soc. 1962, 84,<br />

671.<br />

73 Luz, Z.; Meiboom, S. J. Chem. Phys.<br />

1973, 59, 1077.<br />

74 Naor, R.; Luz, Z. J. Chem. Phys. 1982, 76,<br />

5662.<br />

75 Anet, F. A. L.; Bourn, A. J. R.; Lin, Y. S.<br />

J. Am. Chem. Soc. 1964, 86, 3576.<br />

76 Oth, J. F. M. Pure Appl. Chem. 1971, 25,<br />

573.<br />

77 Wannere, C. S.; Moran, D.; Allinger,<br />

N. L.; Hess, B. A., Jr.; Schaad, L. J.;<br />

Schleyer, P. v. R. Org. Lett. 2003, 5, 2983.<br />

78 Dewar, M. J. S.; Harget, A. J.;<br />

Haselbach, E. J. Am. Chem. Soc. 1969,<br />

91, 7521.<br />

79 Ermer, O.; Klaerner, F. G.; Wette, M.<br />

J. Am. Chem. Soc. 1986, 108, 4908.<br />

80 Gardlik, J. M.; Paquette, L. A.; Gleiter, R.<br />

J. Am. Chem. Soc. 1979, 101, 1617.<br />

81 Hrovat, D. A.; Borden, W. T.<br />

J. Am. Chem. Soc. 1992, 114, 5879.<br />

82 Andres, J. L.; Castano, O.; Morreale, A.;<br />

Palmeiro, R.; Gomperts, R. J. Chem.<br />

Phys. 1998, 108, 203.<br />

83 Castano, O.; Palmeiro, R.; Frutos, L. M.;<br />

Luisandres, J. J. Comput. Chem. 2002, 23,<br />

732.<br />

84 Garavelli, M.; Bernardi, F.; Cembran, A.;<br />

Castano, O.; Frutos, L. M.; Merchan, M.;<br />

Olivucci, M. J. Am. Chem. Soc. 2002,<br />

124, 13770.<br />

85 Polanyi, J. C.; Zewail, A. H. Acc. Chem.<br />

Res. 1995, 28, 119.<br />

421


422 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

86 Neumark, D. M. Science (Washington,<br />

D.C.) 1996, 272, 1446.<br />

87 Wenthold, P. G.; Hrovat, D. A.;<br />

Borden, W. T.; Lineberger, W. C. Science<br />

(Washington, D.C.) 1996, 272, 1456.<br />

88 Wenthold, P. G.; Lineberger, W. C. Acc.<br />

Chem. Res. 1999, 32, 597.<br />

89 Paik, D. H.; Yang, D.-S.; Lee, I. R.;<br />

Zewail, A. H. Angew. Chem. Int. Ed.<br />

2004, 43, 2830.<br />

90 Dewar, M. J. S.; Merz, K. M., Jr. J. Phys.<br />

Chem. 1985, 89, 4739.<br />

91 Trindle, C. Int. J. Quantum Chem 1998,<br />

67, 367.<br />

92 Einstein, F. W. B.; Willis, A. C.;<br />

Cullen, W. R.; Soulen, R. L. J. Chem.<br />

Soc., Chem. Commun. 1981, 526.<br />

93 Baldridge, K. K.; Siegel, J. S. J. Am.<br />

Chem. Soc. 2001, 123, 1755.<br />

94 Matsuura, A.; Komatsu, K. J. Am. Chem.<br />

Soc. 2001, 123, 1768.<br />

95 Blancafort, L.; Bearpark, M. J.;<br />

Robb, M. A. Mol. Phys. 2006, 104, 2007.<br />

96 Haddon, R. C. J. Am. Chem. Soc. 1986,<br />

108, 2837.<br />

97 Haddon, R. C.; Scott, L. T. Pure Appl.<br />

Chem. 1986, 58, 137.<br />

98 Haddon, R. C. Acc. Chem. Res. 1988, 21,<br />

243.<br />

99 Fowler, P. W.; Havenith, R. W. A.;<br />

Jenneskens, L. W.; Soncini, A.; Steiner, E.<br />

Angew. Chem. Int. Ed. 2002, 41, 1558.<br />

100 Havenith, R. W. A.; Fowler, P. W.;<br />

Jenneskens, L. W. Org. Lett. 2006, 8, 1255.<br />

101 Huisgen, R.; Mietzsch, F. Angew. Chem.<br />

Int. Ed. 1964, 3, 83.<br />

102 Huisgen, R.; Mietzsch, F.; Boche, G.;<br />

Seidl, H. Chem. Soc. Spec. Publ. 1965,<br />

19, 3.<br />

103 Paquette, L. A.; Russell, R. K.;<br />

Wingard, R. E., Jr. Tetrahedron Lett.<br />

1973, 1713.<br />

104 Martin, H. D.; Urbanek, T.; Walsh, R.<br />

J. Am. Chem. Soc. 1985, 107, 5532.<br />

105 Castano, O.; Frutos, L.-M.; Palmeiro, R.;<br />

Notario, R.; Andres, J.-L.; Gomperts, R.;<br />

Blancafort, L.; Robb, M. A. Angew. Chem.<br />

Int. Ed. 2000, 39, 2095.<br />

106 Quast, H.; Heubes, M.; Dietz, T.;<br />

Witzel, A.; Boenke, M.; Roth, W.<br />

Eur. J. Org. Chem. 1999, 813.<br />

107 Vij, A.; Palmer, J. L.; Chauhan, K.;<br />

Williams, R. V. J. Chem. Crystallogr. 2001,<br />

30, 621.<br />

108 Scott, L. T.; Jones, M., Jr. Chem. Rev.<br />

1972, 72, 181.<br />

109 Turro, N. J.; Liu, J.-M.;<br />

Zimmerman, H. E.; Factor, R. E. J. Org.<br />

Chem. 1980, 45, 3511.<br />

110 Zimmerman, H. E.; Grunewald, G. L.<br />

J. Am. Chem. Soc. 1966, 88, 183.<br />

111 Hine, J.; Brown, J. A.; Zalkow, L. H.;<br />

Gardner, W. E.; Hine, M. J. Am. Chem.<br />

Soc. 1955, 77, 594.<br />

112 Zimmerman, H. E.; Paufler, R. M.<br />

J. Am. Chem. Soc. 1960, 82, 1514.<br />

113 Zimmerman, H. E.; Grunewald, G. L.;<br />

Paufler, R. M.; Sherwin, M. A. J. Am.<br />

Chem. Soc. 1969, 91, 2330.<br />

114 Zimmerman, H. E. Acc. Chem. Res.<br />

1971, 4, 272.<br />

115 Haselbach, E.; Heilbronner, E.;<br />

Schroeder, G. Helv. Chim. Acta 1971, 54,<br />

153.<br />

116 Zimmerman, H. E.; Armesto, D. Chem.<br />

Rev. 1996, 96, 3065.<br />

117 Katz, T. J. J. Am. Chem. Soc. 1960, 82,<br />

3784.<br />

118 Katz, T. J. J. Am. Chem. Soc. 1960, 82,<br />

3785.<br />

119 Olah, G. A.; Staral, J. S.; Liang, G.;<br />

Paquette, L. A.; Melega, W. P.; Carmody,<br />

M. J. J. Am. Chem. Soc. 1977, 99, 3349.<br />

120 Nishinaga, T.; Komatsu, K.; Sugita, N.<br />

J. Chem. Soc., Chem. Commun. 1994,<br />

2319.<br />

121 Komatsu, K.; Nishinaga, T. Synlett 2005,<br />

187.<br />

122 Childs, R. F. Acc. Chem. Res. 1984, 17,<br />

347.<br />

123 Williams, R. V. Chem. Rev. 2001, 101,<br />

1185.<br />

124 Applequist, D. E.; Roberts, J. D.<br />

J. Am. Chem. Soc. 1956, 78, 4012.<br />

125 Doering, W. v. E.; Laber, G.;<br />

Vonderwahl, R.; Chamberlain, N. F.;<br />

Williams, R. B. J. Am. Chem. Soc. 1956,<br />

78, 5448.<br />

126 Winstein, S. J. Am. Chem. Soc. 1959, 81,<br />

6524.<br />

127 Winstein, S. Chem. soc. Spec. Publ. 1967,<br />

21, 5.<br />

128 Winstein, S. Q. Rev. Chem. Soc. 1969, 23,<br />

965.<br />

129 Story, P. R. In Carbonium Ions, Vol. III,<br />

Olah, G. A., Schleyer, P. v. R. (Eds.);<br />

Wiley-Interscience: New York, 1972,<br />

p. 283.


130 Warner, P. M. In Topics in Nonbenzenoid<br />

Aromatic Chemistry, Vol. 2, Nozoe, T.,<br />

Breslow, R., Hafner, K., Ito, S., Murata, I.<br />

(Eds.); Hirokawa Publishing Co.: Tokyo,<br />

1977, p. 283.<br />

131 Paquette, L. A. Angew. Chem. Int. Ed.<br />

1978, 17, 106.<br />

132 Williams, R. V.; Kurtz, H. A. Adv. Phys.<br />

Org. Chem. 1994, 29, 273.<br />

133 Childs, R. F.; Cremer, D.; Elia, G.<br />

Chemistry of the Cyclopropyl Group 1995,<br />

2, 411.<br />

134 Cremer, D.; Childs, R. F.; Kraka, E.<br />

Chemistry of the Cyclopropyl Group 1995,<br />

2, 339.<br />

135 Kennedy, R. D.; Lloyd, D.; McNab, H.<br />

J. Chem. Soc., Perkin Trans. 1 2002, 1601.<br />

136 Wannere, C. S.; Sattelmeyer, K. W.;<br />

Schaefer, H. F., III; Schleyer, P. v. R.<br />

Angew. Chem. Int. Ed. 2004, 43, 4200.<br />

137 Bregman, J.; Hirshfeld, F. L.;<br />

Rabinovich, D.; Schmidt, G. M. J.<br />

Acta Cryst. 1965, 19, 227.<br />

138 Gorter, S.; Rutten-Keulemans, E.;<br />

Krever, M.; Romers, C.;<br />

Cruickshank, D. W. J. Acta Crystallogr.,<br />

Sect. B: Struct. Sci 1995, B51, 1036.<br />

139 Ermer, O. Helv. Chim. Acta 2005, 88,<br />

2262.<br />

140 Castro, C.; Karney, W. L.; McShane;<br />

M., C.; Pemberton; P., R. J. Org. Chem.<br />

2006, 71, 3001.<br />

141 Masamune, S.; Darby, N. Acc. Chem. Res.<br />

1972, 5, 272.<br />

142 Castro, C.; Karney, W. L.; Valencia, M. A.;<br />

Vu, C. M. H.; Pemberton, R. P. J. Am.<br />

Chem. Soc. 2005, 127, 9704.<br />

143 Pemberton, R. P.; McShane, C. M.;<br />

Castro, C.; Karney, W. L. J. Am. Chem.<br />

Soc. 2006, 128, 16692.<br />

144 Moll, J. F.; Pemberton, R. P.;<br />

Gutierrez, M. G.; Castro, C.;<br />

Karney, W. L. J. Am. Chem. Soc. 2007,<br />

129, 274.<br />

145 Herges, R. Chem. Rev. 2006, 106, 4820.<br />

146 Vogel, E. Pure Appl. Chem. 1969, 20, 237.<br />

147 Vogel, E. Proceedings of the Robert A.<br />

Welch Foundation Conference on Chemical<br />

Research 1969, 12, 215.<br />

148 Vogel, E. Isr. J. Chem. 1980, 20, 215.<br />

149 Cremer, D.; Dick, B. Angew. Chem. 1982,<br />

94, 877.<br />

150 Farnell, L.; Radom, L. J. Am. Chem. Soc.<br />

1982, 104, 7650.<br />

References<br />

151 Vogel, E. Pure Appl. Chem. 1982, 54,<br />

1015.<br />

152 Vogel, E.; Scholl, T.; Lex, J.;<br />

Hohlneicher, G. Angew. Chem. 1982, 94,<br />

878.<br />

153 Dorn, H. C.; Yannoni, C. S.;<br />

Limbach, H.-H.; Vogel, E. J. Phys. Chem.<br />

1994, 98, 11628.<br />

154 Mitchell, R. H. Chem. Rev. 2001, 101,<br />

1301.<br />

155 Williams, R. V.; Armantrout, J. R.;<br />

Twamley, B.; Mitchell, R. H.; Ward, T. R.;<br />

Bandyopadhyay, S. J. Am. Chem. Soc.<br />

2002, 124, 13495.<br />

156 Mitchell, R. H. Eur. J. Org. Chem. 1999,<br />

2695.<br />

157 Williams, R. V.; Edwards, W. D.;<br />

Mitchell, R. H.; Robinson, S. G. J. Am.<br />

Chem. Soc. 2005, 127, 16207.<br />

158 Doering, W. v. E. Proc. Robert A. Welch<br />

Found. Conf. Chem. Res. 1990, 34, 334.<br />

159 Doering, W. v. E.; Roth, W. R.<br />

Tetrahedron 1963, 19, 715.<br />

160 Doering, W. v. E.; Roth, W. R. Angew.<br />

Chem. Int. Ed. 1963, 2, 115.<br />

161 Schröder, G. Angew. Chem. Int. Ed. 1963,<br />

2, 481.<br />

162 Doering, W. v. E.; Ferrier, B. M.;<br />

Fossel, E. T.; Hartenstein, J. H.;<br />

Jones, M., Jr.; Klumpp, G. W.;<br />

Rubin, R. M.; Saunders, M. Tetrahedron<br />

1967, 13, 3943.<br />

163 Williams, R. V. Advances in Theoretically<br />

Interesting Molecules 1998, 4, 157.<br />

164 Williams, R. V. Eur. J. Org. Chem. 2001,<br />

227.<br />

165 Paquette, L. A.; Liao, C. C.; Burson, R. L.;<br />

Wingard, R. E., Jr.; Shih, C. N.; Fayos, J.;<br />

Clardy, J. J. Am. Chem. Soc. 1977, 99,<br />

6935.<br />

166 Cheng, A. K.; Anet, F. A. L.; Mioduski, J.;<br />

Meinwald, J. J. Am. Chem. Soc. 1974, 96,<br />

2887.<br />

167 Moskau, D.; Aydin, R.; Leber, W.;<br />

Guenther, H.; Quast, H.; Martin, H. D.;<br />

Hassenrueck, K.; Grohmann, K.;<br />

Miller, L. S. Chem. Ber. 1989, 122, 925.<br />

168 Dewar, M. J. S.; Lo, D. H. J. Am. Chem.<br />

Soc. 1971, 93, 7201.<br />

169 Hoffmann, R.; Stohrer, W. D.<br />

J. Am. Chem. Soc. 1971, 93, 6941.<br />

170 Mehta, G.; Vidya, R.; Sharma, P. K.;<br />

Jemmis, E. D. Tetrahedron Lett. 2000, 41,<br />

2999.<br />

423


424 8 Molecules with Labile Bonds: Selected Annulenes and Bridged Homotropilidenes<br />

171 Mehta, G.; Gagliardini, V.;<br />

Schaefer, C.; Gleiter, R. Org. Lett.<br />

2004, 6, 617.<br />

172 Liang, S.; Lee, C.-H.; Kozhushkov, S. I.;<br />

Yufit, D. S.; Howard, J. A. K.; Meindl, K.;<br />

Ruehl, S.; Yamamoto, C.; Okamoto, Y.;<br />

Schreiner, P. R.; Rinderspacher, B. C.;<br />

De Meijere, A. Chem. Eur. J. 2005, 11,<br />

2012.<br />

173 Reiher, M.; Kirchner, B. Angew. Chem.<br />

Int. Ed. 2002, 41, 3429.<br />

174 Kirchner, B.; Sebastiani, D. J. Phys.<br />

Chem. A 2004, 108, 11728.<br />

175 Tantillo, D. J.; Hoffmann, R.;<br />

Houk, K. N.; Warner, P. M.;<br />

Brown, E. C.; Henze, D. K. J. Am. Chem.<br />

Soc. 2004, 126, 4256.<br />

176 Tantillo, D. J.; Hoffmann, R. Acc. Chem.<br />

Res. 2006, 39, 477.<br />

177 Seefelder, M.; Heubes, M.; Quast, H.;<br />

Edwards, W. D.; Armantrout, J. R.;<br />

Williams, R. V.; Cramer, C. J.;<br />

Goren, A. C.; Hrovat, D. A.;<br />

Borden, W. T. J. Org. Chem. 2005, 70,<br />

3437.<br />

178 Reichardt, C. Solvents and Solvent<br />

Effects in Organic Chemistry; 3rd edn.;<br />

VCH: Weinheim, Germany, 2002.<br />

179 Samdal, S.; Richardson, A. D.;<br />

Hedberg, K.; Gadgil, V. R.; Meyer, M. M.;<br />

Williams, R. V. Helv. Chim. Acta 2003,<br />

86, 1741.<br />

180 Hrovat, D. A.; Brown, E. C.;<br />

Williams, R. V.; Quast, H.; Borden, W. T.<br />

J. Org. Chem. 2005, 70, 2627.<br />

181 Brown, E. C.; Henze, D. K.; Borden,<br />

W. T. J. Am. Chem. Soc. 2002, 124, 14977.<br />

182 Hrovat, D. A.; Williams, R. V.;<br />

Goren, A. C.; Borden, W. T. J. Comput.<br />

Chem. 2001, 22, 1565.<br />

183 Goren, A. C.; Hrovat, D. A.;<br />

Seefelder, M.; Quast, H.; Borden, W. T.<br />

J. Am. Chem. Soc. 2002, 124, 3469.<br />

184 Sauer, J.; Bauerlein, P.; Ebenbeck, W.;<br />

Schuster, J.; Sellner, I.; Sichert, H.;<br />

Stimmelmayr, H. Eur. J. Org. Chem.<br />

2002, 791.<br />

185 Zhou, C.; Birney, D. M. Org. Lett. 2002,<br />

4, 3279.<br />

186 Wang, C.; Yuan, J.; Li, G.; Wang, Z.;<br />

Zhang, S.; Xi, Z. J. Am. Chem. Soc. 2006,<br />

128, 4564.<br />

187 Gompper, R.; Schwarzensteiner, M. L.;<br />

Wagner, H. U. Tetrahedron Lett. 1985, 26,<br />

611.<br />

188 Saunders, M.; Jimenez-Vazquez, H. A.<br />

Chem. Rev. 1991, 91, 375.<br />

189 Heubes, M.; Dietz, T.; Quast, H.;<br />

Seefelder, M.; Witzel, A.; Gadgil, V. R.;<br />

Williams, R. V. J. Org. Chem. 2001, 66,<br />

1949.<br />

190 Wu, H.-S.; Jiao, H.; Wang, Z.-X.;<br />

Schleyer, P. v. R. J. Am. Chem. Soc. 2003,<br />

125, 10524.


9<br />

Molecules with Nonstandard Topological Properties:<br />

Centrohexaindane, Kuratowski’s Cyclophane and<br />

Other Graph-theoretically Nonplanar Molecules<br />

Dietmar Kuck<br />

9.1<br />

Introduction<br />

It may be suspected that students of organic chemistry believe, for probably quite<br />

varying periods of time, in the perfect tetrahedral coordination of the saturated<br />

carbon atom. Colloquial mentioning that an organic molecule, say, methanol,<br />

contains a ‘tetrahedral carbon’ incorrectly implies a sphere of a regular tetrahedron.<br />

Van’t Hoff and Le Bel’s model [1] has just been too appealing. The propensity to<br />

oversimplify seems to be a major driving force but also a means for our understanding<br />

chemical structures. Simplicity comes and goes with symmetry and our<br />

individual views of chemical beauty change over the years. What is more beautiful:<br />

ideal shape or a (slightly) pertubated appearance [2]?<br />

In fact, there are only a few perfectly tetrahedral molecules in organic chemistry.<br />

Methane, the tetrahalomethanes and also neopentane belong to this group, albeit<br />

only the central carbon atom has the perfectly tetrahedral coordination, that is,<br />

its valence orbitals are subject to perfect sp 3 -hybridization. The central carbons<br />

in tetraethylmethane [3] and tetrabenzylmethane [4] do not, nor do many more<br />

‘organic’ carbon atoms [5].<br />

9.1.1<br />

Is All This Trivial?<br />

Another insight that comes late, or later, in the process of learning about organic<br />

chemistry concerns structural topology, rather than three-dimensional (3-D) topography,<br />

of organic molecules. It frequently comes as a challenge to students to<br />

present the 3-D structural formula of a molecule in the plane of a sheet of paper.<br />

Strychnine is a clear example of a complex polycyclic organic structure which can<br />

be projected so that we ‘see’ its seven rings put one alongside the other in apparently<br />

two dimensions. However, there are naturally occurring and also non-natural<br />

organic compounds which cannot be drawn in the plane with one ring beside the<br />

other. Strychnine is too simple a structure! This is documented by several reports<br />

on topologically nonplanar organic molecules [6–14].<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

425


426 9 Molecules with Nonstandard Topological Properties<br />

Ask students for an organic compound the structure of which cannot be drawn<br />

in the plane without crossing at least two bonds. It is very likely that they won’t<br />

find any but they would spontaneously suggest either C 60, or dodecahedrane (or<br />

one of the other two Platonic hydrocarbons) – or diamond, which, notwithstanding<br />

the carats involved, may be considered a set of ultimate organic molecules given<br />

its hydrogen-terminated surface [8, 15].<br />

In fact, all of the known or otherwise conceivable globular molecules, such<br />

as the fullerenes [16] (discussed in Chapter 5), the dodecahedranes [17] (briefly<br />

presented in Section 2.4.1.6), superphane [18, 19] (mentioned in Section 4.2),<br />

and their synthesis precursors, or the various carcerands [20], are ‘topologically<br />

planar’. Even ‘oligomeric’ fullerenes [21] are � since their structures can be represented<br />

as ‘planar graphs’, that is, as polycyclic arrangements of rings that are<br />

multiply fused side-by-side in the plane. It is not possible to draw topologically,<br />

or graph-theoretically nonplanar (‘gt-nonplanar’), molecules in such a simple way<br />

(Figure 9.1). At least one crossing of a pair of bonds cannot be avoided.<br />

The smallest conceivable gt-nonplanar’carbon-based structure bearing, at the<br />

same time, a carbon atom with perfect tetrahedral coordination would be the<br />

carbon cluster C 5 1. However, C 5 isomers are known to be linear and, of course,<br />

extremely difficult to generate experimentally [22]. A somewhat more realistic<br />

analog of 1 is the next-higher homolog bearing a perfectly ‘tetrahedral’ carbon<br />

atom, viz. ‘centrohexaquadrane’ (C 11 H 12 , 2). This hypothetical hydrocarbon was<br />

discussed in 1981 in an inspiring conceptual article concerning the aggregation<br />

of rings about a common carbon center [23]. The first experimentally realistic<br />

gt-nonplanar hydrocarbon bearing a central carbon atom with ideal tetrahedral<br />

symmetry is centrohexaquinane (C 17 H 24 , 3). This molecule has been calculated to<br />

have overall T molecular symmetry [23, 24] and represents, owing to the almost<br />

perfect geometrical fit of the cyclopentane units with the central neopentane core,<br />

a low-strain ‘polycyclane’ [23]. More specifically, it may be considered the prototypical<br />

member of the unusual family of ‘centropolycyclanes’, that is, polycyclic<br />

hydrocarbons bearing several (up to six) rings mutually fused about the four C–C<br />

bonds of neopentane [23]. Notwithstanding its geometrically familiar constituents,<br />

including the unique, perfectly sp 3 -hybridized central carbon atom (and 16<br />

nonperfect cases), centrohexaquinane (3) has remained a hypothetical molecule<br />

to date [25–29]. The same holds true for its highest unsaturated congener, centrohexaquinacene<br />

(4, Figure 9.2), which also has low internal strain but has been<br />

calculated to have a perfect T d -symmetrical molecular structure [24].<br />

Figure 9.1 The three simplest topologically nonplanar carbon-based K 5 molecules are<br />

all hypothetical: tetrahedral cluster C 5 1; centrohexaquadrane 2 [23]; centrohexaquinane 3.


9.2<br />

Topologically Nonplanar Graphs and Molecular Motifs<br />

9.2.1<br />

The Centrohexaquinacene Core<br />

9.2 Topologically Nonplanar Graphs and Molecular Motifs<br />

What is the beauty of the centrohexaquinane motif, and of centrohexaquinacene<br />

4, in particular? Most of their properties become evident when regarded with the<br />

eyes of an organic chemist and analyzed with a touch of mathematics.<br />

1. It represents a relatively large organic structure centered about a truly sp 3 -<br />

hybridized carbon atom.<br />

2. The tetrahedral symmetry of the central carbon atom is extended into the 3-D<br />

space.<br />

3. Centrohexaquinacene 4 contains six perfectly planar cyclopentene rings.<br />

4. Each of these six rings is bisected by one of the Cartesian axes and the central<br />

carbon atom shared by all rings is placed in the origin of this Cartesian coordinate<br />

system.<br />

5. Thus, placed at one of the six tips of an octahedron, each of the six double<br />

bonds of centrohexaquinacene 4 points out into one of the six directions of<br />

the 3-D space (�x, �y, �z) [28, 29].<br />

6. The organic substructures of 3 and 4 are manifold: They contain three mutually<br />

fused spiro[4.4]nonane units (and thus represent ‘superspiro’ arrangements),<br />

four intermingled [3.3.3]propellanes as well as four intermingled triquinacenetype<br />

entities, and finally, three partially intercrossing [5.5.5.5]fenestrane units<br />

[23, 28].<br />

7. In these various substructures, the geometrical abnormalities and the corresponding<br />

symmetries are largely preserved. In turn, all deviations from<br />

perfect geometries and minor symmetries in the substructures are completely<br />

cancelled in the complete centrohexacyclic frameworks of 3 and 4, in particular.<br />

Figure 9.2 The hypothetical centrohexaquinacene 4 and an ‘exploded’ view illustrating the<br />

ortho gonal orientation of the six cyclopentene rings of 4 in the Cartesian coordinate system<br />

[28, 29].<br />

427


428<br />

9 Molecules with Nonstandard Topological Properties<br />

8. Stepwise dismantling of 3 and 4, representing the largest congeners, by<br />

removing one C 2 (ethano or etheno) bridge after the other leads to all of the<br />

lower centropolyquinanes and centropolyquinacenes, respectively.<br />

9. The centrohexacyclic structures of 3 and 4 give rise to another special molecular<br />

topography: All surfaces of these C 17 molecular cores are concave when<br />

regarding the cyclopentene rings of 4 as strictly two-dimensional units.<br />

10. To complete this survey of abnormalities, it is repeated here that the molecular<br />

topology of 3 and 4 is nonplanar: The constitution of the 17 carbon atoms corresponds<br />

to the complete graph K 5, a nonplanar graph comprising five centers<br />

all of which are pairwise interconnected. This special abnormality will be<br />

discussed in detail below.<br />

9.2.2<br />

The Nonplanar Graphs K 5 and K 3,3 and Some Molecular Representatives<br />

A key publication on graph theory appeared in 1930 in Fundamenta Mathematicae,<br />

written by the Polish mathematician C. Kuratowski [30]. In this article entitled ‘Sur<br />

le problème des courbes gauches en topologie’, Kuratowski showed that all topologically<br />

nonplanar graphs can be reduced to two graphs, K 5 and K 3,3 , both of which<br />

can be sketched by pencil-and-paper drawings of a simple tetrahedron. These two<br />

sketches are depicted by a modern computer-aided ‘drawing’ in Figure 9.3.<br />

Although the two graphs K 5 and K 3,3 have a common origin, they are fundamentally<br />

different. In the case of the graph K 5 , the central point of the tetrahedron is<br />

added and connected to the four corner points. The whole set of ten lines (edges<br />

of the graph) connecting five tetravalent points (the vertices of the graph) comprise<br />

the ‘complete graph’ K 5 . Each of the five vertices is connected to all of the other<br />

ones. The graph K 3,3 , also nonplanar, is somewhat less simple and elegant. In<br />

this case, two opposite edges of the tetrahedron are divided by two additional<br />

points, which are mutually connected. Coloring these additional points and their<br />

neighboring points at the tetrahedron’s tips in opposite ways, as depicted in<br />

Figure 9.3, yields two sets of three vertices of the ‘bipartite’ K 3,3 graph. Notably,<br />

the three trivalent vertices of each set are not interconnected, but each of them is<br />

associated with each vertex of the other set. Thus, the graph K 3,3 comprises two<br />

‘isolated’ sets of three vertices and a total of nine edges.<br />

Figure 9.3 The sketches of the K 5 graph (a) and the K 3,3 graph (b), as reproduced from<br />

Kuratowski’s publication of 1930 [30] using modern computer-aided graphics.


9.2 Topologically Nonplanar Graphs and Molecular Motifs<br />

Figure 9.4 The complete graph K 5 in four different representations (see text).<br />

The graph K 5 is often represented as a regular pentagon with ‘exhaustive’<br />

interconnection of the tips (Figure 9.4a). The impossibility of projecting it into<br />

the plane without any mutual crossing of edges is most simply illustrated in<br />

Figure 9.4b. Slight rearrangement in the plane gives a filled square (Figure 9.4c)<br />

with the same topological properties, and distortion into the third dimension<br />

generates the tetrahedron with its central vertex being topographically different but<br />

topologically equivalent to the other four (Figure 9.4d). This recalls Kuratowski’s<br />

original representation of the K 5 graph, as depicted in Figure 9.3a.<br />

The graph K 3,3 is often represented in a rectangular form, as depicted in<br />

Figure 9.5a. Maximum efforts to separate the edges again leave one unavoidable<br />

crossing, as most simply shown in Figure 9.5b. Rearrangement of points, as<br />

discussed for the graph K 5, yields the (unusual) rectangular representation of<br />

Figure 9.5c, which can easily be converted into the regular hexagonal form of<br />

Figure 9.5d. Of course, this representation is still topologically nonplanar, as shown<br />

in Figure 9.5e. Finally, out-of-plane distortion generates a spatial impression of<br />

the interconnection of the six vertices of the K 3,3 graph. Kuratowski’s tetrahedronbased<br />

definition (Figure 9.3b) can be most easily recognized from the illustrations<br />

in Figure 9.5d and 9.5e, in which the extra diametrical connectivity is represented<br />

by the two middle vertices and their common edge.<br />

Figure 9.5 The complete graph K 3,3 in six different representations (see text).<br />

429


430<br />

9 Molecules with Nonstandard Topological Properties<br />

Figure 9.6 Centrohexaindane 5, a T d -symmetrical K 5 -hydrocarbon and the hexabenzo analog of<br />

3 and 4, and Kuratowski’s cyclophane 6, a formally D 2d -symmetrical K 3,3 -polyether. Note that the<br />

front and the back naphthalene-2,7-diyl bridges have been simplified by a two extra long bonds<br />

(O–O).<br />

The number and variety of gt-nonplanar organic compounds has increased<br />

recently, and the examples are collected and discussed in several reports, including<br />

quite recent ones [8–14, 28, 29, 31–34]. Instead of commenting on the various sorts<br />

of such nevertheless unusual structures, two of the most representative cases are<br />

presented in more detail below. To date, the most versatile among these is centrohexaindane<br />

(5, Figure 9.6) [35–37], the hexabenzo analog of hypothetical parent<br />

K 5 hydrocarbons, centrohexaquinane 3 and centrohexaquinacene 4. This highest<br />

member of the centropolyindane family can be made in gram amounts [37] and<br />

a variety of derivatives has become accessible [9, 28, 29]. The most representative<br />

K 3,3 counterpart is considered to be Kuratowski’s cyclophane 6 [38], a macrocyclic<br />

polyether, although a number of other topologically nonplanar organic compounds<br />

of the K 3,3 topology were known previously [8, 10, 11, 39–41]. The synthesis of<br />

Kuratowski’s cyclophane was designed as a directed construction of a nonplanar<br />

organic compound of the K 3,3 type and, as such, it appears to represent a unique<br />

example of a non-natural polycyclic structure.<br />

9.3<br />

Centrohexaindane<br />

Centrohexaindane 5 is a colorless, high melting solid (m.p. > 400 °C). Its solubility<br />

in various solvents is surprisingly high and single crystals can be grown from<br />

some of them, including para-xylene and triethylamine [29]. The fact that this<br />

‘heavy’ C 41 H 24 hydrocarbon does not behave as a stone-like solid is attributed to<br />

the peculiarity that it is one of the very organic molecules that, disregarding the<br />

rims of the benzene rings, have exclusively concave faces. Thus, solvent molecules<br />

like triethylamine tend to fill the four ‘hollow’ tribenzotriquinacene cavities. In<br />

fact, a strongly negative electrostatic potential has been calculated for the concave<br />

side of the tribenzotriquinacene skeleton [42].


9.3 Centrohexaindane<br />

Figure 9.7 The structure of centrohexaindane 5, as deduced from X-ray structure analysis<br />

of a single crystal containing one molecule of triethylamine per 5 (C 41 H 24 �NEt 3 ) [28, 29, 43].<br />

Views of the tribenzotriquinacene cavity (stick model, left) and of the triptindane (propellane)<br />

axis (space-filling model, right).<br />

The molecular structure of centrohexaindane as determined by X-ray diffraction<br />

of single crystals grown from triethylamine solutions is shown in Figure 9.7 [43].<br />

All three intermingled [2,2�]spirobiindane units are strictly linear and the two<br />

benzene rings of each are mutually orientated at right angles. The four tribenzotriquinacene<br />

entities are strictly C 3v -symmetrical; not only are the planes of their<br />

three indane wings orientated at 120° but the long axes of the indane wings are<br />

strictly orthogonal to each other. The three fenestrindane units comprised in 5 are<br />

held in a perfect D 2d -symmetrical conformation. The six C–C–C bond angles at<br />

the central carbon atom of 5 are identical, within the limits of experimental error,<br />

with the ideal tetrahedral angle (109° 28�) expected for the four C–C bonds at an<br />

undisturbed sp 3 -hybridized atom. From all this follows the overall T d -symmetry<br />

of the ground state of centrohexaindane, which is only very slightly affected by<br />

the crystal packing.<br />

9.3.1<br />

Centrohexaindane and Structural Regularities of the Centropolyindane Family<br />

As the highest member of the regular centropolyindanes [28, 29, 44], centrohexaindane<br />

(5) contains all of the lower members of this family as substructures. Thus,<br />

removal of just one indane unit leads to centropentaindane 7 [45], another very rigid<br />

polycyclic system bearing five benzene rings stretching practically ortho gonally<br />

into five of the six directions of the 3-D space. Removal of another indane unit gives<br />

rise to either fenestrindane 8 [46], an interesting congener of the widely studied<br />

all-cis-[5.5.5.5]fenestranes [5, 28], or to a trifuso-centrotetraindane 9 [47] which, like<br />

5 and 7, represents a rigid molecular framework. Among the higher centropolyindanes,<br />

fenestrindane 8 is particular since it is conformationally flexible, as are<br />

two of the three regular centrotriindanes comprised in the framework of centrohexaindane,<br />

viz. 11 [46b, 48] and 12 [49, 50]. The only centrotriindanes having a<br />

conformationally rigid framework are the tribenzotriquinacenes, e.g. 10 [48, 51].<br />

In fact, the three indane wings of the latter centrotriindane have been shown to<br />

stretch into the 3-D space almost perfectly at right angles [28, 29].<br />

431


432<br />

9 Molecules with Nonstandard Topological Properties<br />

Figure 9.8 Centrohexaindane 5 and the lower centropolyindanes 7�12. The highly systematic<br />

structural and geometrical relationship between the centropolyindanes is reflected by the nearly<br />

constant � max value of their UV/Vis absorption and the systematic incremental 13 C chemical<br />

shifts of the central carbon atoms. a The � max value of 12 was taken from [49b].<br />

Centrohexaindane has such a very close structural and geometrical relationship<br />

with the various lower centropolyindanes, that a comparison of some spectroscopic<br />

properties is informative. The UV/Vis spectrum of centrohexaindane reveals that<br />

the six �-electron systems of the mutually fused indane units do not interact, in<br />

agreement with the diphenylmethane-type junction of the aromatic rings and the<br />

orthogonal arrangement in space [28, 29]. Thus, the absorption at � max = 276.5 nm<br />

(� = 5800) in n-heptane is roughly four times as strong as the corresponding lowenergy<br />

band in indane, and there is only a very minor bathochromic shift compared<br />

with the latter hydrocarbon (� max = 273.0 nm, � = 1500) [48]. With the decreasing<br />

number of indane units attached to the neopentane core, the extinction coefficient<br />

decreases steadily. By contrast, the � max values of the centropolyindanes were found<br />

to reflect a characteristic feature of the indane units annealed to the neopentane


9.3 Centrohexaindane<br />

core, namely their conformational geometry. In centrohexaindane 5 and in all other<br />

congeners containing at least one tribenzotriquinacene unit (cf. 10), the indane<br />

wings are forced into a planar conformation and their � max values are identical,<br />

viz. 276.0�276.5 nm. Of course, this also holds true for the tribenzotriquinacenes<br />

themselves, e.g. 10. Fenestrindane 8 and the other two centrotriindanes, 11 and<br />

12, and indane itself have conformationally flexible frameworks [28]. Their cyclopentane<br />

rings are allowed to partially move out-of-plane giving an envelope form<br />

even in fenestrindane, and the UV/Vis spectra of all of these indane hydrocarbons<br />

exhibit � max values in the range of 273.0�274.0 nm (Figure 9.8).<br />

In view of the fact that the central carbon atom of centrohexaindane (5) is a<br />

perfectly ‘aliphatic’, sp 3 -hybridzed atom, the resonance of the 13 C nucleus at this<br />

particular position is also noteworthy. It appears at �(C centro ) = 95.0 with particularly<br />

low relative intensity, as expected for such a nucleus embedded in scaffold<br />

of sixteen (!) other quaternary carbon atoms. As measured under standard conditions<br />

at 600 MHz, the intensity ratio falls considerably short of the expected<br />

value, i.e. [C centro ] : [C � ] � 0.25. The chemical shift of the central carbon nucleus<br />

in centrohexaindane decreases systematically when one or more indane wings are<br />

removed stepwise from the neopentane core. Thus, the value of �(C centro ) = 83.2<br />

was found for centropentaindane 7, whereas fenestrindane 8 has �(C centro ) = 71.9<br />

and the isomeric trifusocentrotetraindane 9 has �(C centro ) = 70.9. Thus, as illustrated<br />

in Figure 9.8, stepwise removal of the ortho-phenylene bridges from<br />

centrohexa indane gives rise to a constant upfield shift of 10�11.5 ppm. Similarly<br />

to the UV/Vis data, this finding reflects the highly systematic aufbau principle<br />

of the centropolyindane family, based on largely non-strained indane units and<br />

culminating in the T d -symmetrical centrohexaindane.<br />

9.3.2<br />

Syntheses of Centrohexaindane<br />

Centrohexaindane 5 can be prepared along three independent synthesis routes and<br />

in gram amounts. The first of these involves 11 steps starting from 1,3-indanedione,<br />

a commercially available building block. It involves the synthesis of several<br />

benzoannelated [5.5.5.6]- and [5.5.5.5]fenestranes, including fenestrindane 8, and<br />

affords centrohexaindane in an overall yield of ca. 12%. One of the variants of the<br />

‘fenestrane route’ to centrohexaindane is shown in Scheme 9.1 [35, 37, 46].<br />

Two-fold Michael addition of 1,3-indanedione 13 to dibenzylideneacetone 14<br />

yields the trans-diphenylspirotriketone 15 [52, 53]. Protection of the sterically accessible<br />

cyclohexanone functionality followed by reduction of the other two carbonyl<br />

groups gives the dispirane 16. Originally, the trans orientation in 15 and 16 of the<br />

phenyl groups appeared to be particularly favorable to construct the all-cis-[5.5.5.6]<br />

fenestrane framework of ketone 17 by bicyclodehydration [44b] with concomitant<br />

hydrolysis of the acetal group and, in fact, the yield of this conversion is excellent.<br />

However, the cis-diphenyl stereoisomers of 15 and 16 were found to undergo the<br />

two-fold cyclodehydration as well, giving rise to the strained cis,cis,cis,trans-[5.5.5.6]<br />

fenestrane skeleton [54]. Subsequent ring contraction of fenestranone 17 via<br />

433


434<br />

9 Molecules with Nonstandard Topological Properties<br />

Scheme 9.1 The synthesis of centrohexaindane 5 along the fenestrane route.<br />

the corresponding �,��-dibromoketone, followed by Favorskii rearrangement,<br />

yields the all-cis-tribenzo[5.5.5.5]fenestrane carboxylic acid 18. Decarboxylation<br />

of the latter compound suffers from mediocre efficiency but cleanly affords the<br />

tribenzo[5.5.5.5]fenestrene 19, which undergoes benzoannelation at its double<br />

bond by use of Raasch’s reagent (tetrachlorothiophene-S,S-dioxide) [55] and<br />

subsequent dehalogenation/aromatization. Noteworthily, fenestrindane 8, the<br />

all-cis-tetrabenzo[5.5.5.5]fenestrane, is a highly versatile parent centropolyindane<br />

which has been functionalized in various ways at both its four bridgehead positions<br />

and at the eight outer positions of the arene periphery [56, 57]. In the two final<br />

steps, fenestrindane is converted into its tetrabromo derivative 20, a remarkably<br />

distorted fenestrane bearing significantly flattened geometry at the central carbon<br />

atom [�(C–C–C) = 121.5°] [28, 56, 58] and subsequent condensation with two<br />

molecules of benzene under Lewis acid catalysis yields centrohexaindane [35].<br />

Although the majority of the steps of this multistep synthesis have good or even<br />

very good yields, the fenestrane route to centrohexaindane is rather lengthy and<br />

its efficiency is limited. On the other hand, the eventual incorporation of two<br />

molecules of benzene across the open angles of the fenestrane framework offers<br />

valuable advantages in view of the construction of substituted centrohexaindanes<br />

(see below).


9.3 Centrohexaindane<br />

Scheme 9.2 The synthesis of centrohexaindane 5 along the propellane route (via 22 and 23)<br />

and along the broken fenestrane route (via 11 and 7).<br />

Two much shorter and more efficient syntheses of centrohexaindane have<br />

been developed later and are illustrated in Scheme 9.2. The shortest access to 5<br />

starts again from 1,3-indanedione 13 which, in the first step, is converted into its<br />

2,2-dibenzyl derivative 21 and then to the [3.3.3]propellane ketone 22. Two-step<br />

oxidation of the latter compound affords the highly versatile 1,3,3�-triketone 23,<br />

called ‘triptindanetrione’ because of its relation to the parent centrotriindane<br />

12, dubbed ‘triptindane’ (Figure 9.8) [49]. Two further steps including three-fold<br />

addition of the elements of benzene (from phenyllithium and by hydrolysis) and<br />

subsequent three-fold cyclodehydration of the intermediate, highly crowded propellanetriol<br />

completes the six-step ‘propellane route’ to centrohexaindane 5 [36, 37]. It<br />

is remarkable that a total of ten new C–C bonds are formed by this sequence with<br />

an overall yield of 25%. Since three adjacent indane wings are attached in the last<br />

two steps, the propellane route represents a valuable alternative to the fenestrane<br />

route in case of the synthesis of substituted centrohexaindanes (see below).<br />

The third synthesis of centrohexaindane also involves 2,2-dibenzyl-1,3-indanedione<br />

21. This time, however, the diketone is converted into a ‘broken fenestrane’,<br />

viz. the C 2 -symmetrical centrotriindane 11, by reduction to the corresponding<br />

trans-1,3-indanediol and subsequent acid-catalyzed two-fold cyclo dehydration<br />

(Scheme 9.2) [44b]. Careful bromination of the benzhydrylic and benzylic positions<br />

of 11 and subsequent incorporation of two further benzene units furnish centropentaindane<br />

7, the second-highest centropolyindane, in strikingly high yield<br />

[45]. In the final two steps, the bromination/condensation sequence is repeated,<br />

435


436<br />

9 Molecules with Nonstandard Topological Properties<br />

involving the highly reactive bridgehead dibromo derivative of 7 and leading to<br />

centrohexaindane 5 in an overall yield of 40%. To date, this seven-step ‘broken<br />

fenestrane route’ has not been used for the synthesis of derivatives of centrohexaindane<br />

but it offers the possibility of incorporating one single arene unit in<br />

a directed way [37].<br />

9.3.3<br />

Multiply-functionalized Centrohexaindanes<br />

In fact, the directed synthesis of centrohexaindanes bearing various patterns of<br />

substitution at the six poles of the polycyclic framework appears to be a promising<br />

goal. The spatial orientation of the six arene units is well defined, so unusual<br />

and interesting supramolecular aggregation, in particular in the solid state, may<br />

be envisioned. However, only a few centrohexaindanes bearing a pinpointed set<br />

of functional groups have been made to date, and they are shown in Figure 9.9.<br />

Nitration of the parent centrohexaindane 5 under optimized conditions gives<br />

a mixture of four hexanitro derivatives 24a�d, all of which bearing each nitro<br />

group at one of the six peripheral C–C edges [29, 59]. Two of these constitutional<br />

isomers have C 1 molecular symmetry and the other two are C 3 symmetrical.<br />

Thus, four racemates are formed in ratios (ca. 3 : 3 : 1 : 1) that correspond to the<br />

ratios expected for the case of statistical attack of the nitrating reagent. The four<br />

racemates can be separated from each other and have been fully characterized<br />

[59]. Conversion of the original mixture to the as yet hypothetical twelve-fold<br />

nitrated centrohexaindane and use of the latter to generate the corresponding<br />

T d -symmetrical dodeca aminocentrohexaindane represents a similarly challenging<br />

and promising goal.<br />

At variance from the a posteriori functionalization of the parent centrohexaindane,<br />

the a priori incorporation of substituents turned out to be useful in several<br />

cases. Following the propellane route, the C s -symmetrical dimethyl and tetramethyl<br />

derivatives 25 and 26 were synthesized [60]. Remarkably, these centrohexaindanes<br />

contain the substituents at the strongly sterically-shielded ortho positions; in the<br />

case of the tetramethyl derivative 26, each of the methyl groups points into one of<br />

the four sterically-shielded propellane cavities of the centrohexaindane framework.<br />

The tetramethoxycentrohexaindanes 27 and 28 represent two examples of several<br />

electron-rich derivatives of 5 that have become accessible along the fenestrane<br />

route [61]. The four-fold bridgehead bromination of fenestrindane (8 � 20,<br />

cf. Scheme 9.1) can easily be used to prepare the corresponding bridgehead<br />

tetraalcohol, which in turn can be condensed under Brønsted acid catalysis with<br />

various methoxybenzenes, e.g. with veratrole and hydroquinone dimethyl ether in<br />

the cases of 27 and 28, respectively. Again, the facile incorporation of four methoxy<br />

groups into the sterically-shielded inner arene positions of 28 is remarkable. Thus,<br />

the 1 H NMR spectrum of 28 clearly reflects the close neighborhood of the four<br />

methoxy groups to the eight ortho protons at the adjacent benzene rings, which<br />

resonate at particularly low field (� = 8.22 ppm, compared with � = 7.74 ppm in<br />

the case of 27) [59, 61].


9.3 Centrohexaindane<br />

Figure 9.9 Selected derivatives of centrohexaindane synthesized either by a posteriori functionalization<br />

of 5 (24a�d) or by a priori incorporation of the functional groups (25�29, see text).<br />

437


438<br />

9 Molecules with Nonstandard Topological Properties<br />

The only twelve-fold functionalized derivative of 5 reported so far is the T dsymmetrical<br />

dodecamethoxycentrohexaindane 29 [62]. In this case, the propellane<br />

route proved to be viable, albeit with some inevitable restrictions and losses of<br />

yields in the last steps. As mentioned above, the construction of centrohexaindanes<br />

bearing functional groups at geometrically well-definable points is considered a<br />

great challenge.<br />

9.4<br />

K 5 versus K 3,3 Molecules<br />

Compared with the K 3,3 hydrocarbons and the heterocyclic organic molecules with<br />

K 3,3 topology presented in the next section, most of the gt-nonplanar counterparts<br />

of the K 5-type are remarkably simple. However, as mentioned above, the synthesis<br />

of the smallest prototypical hydrocarbon that conceivably should be accessible,<br />

centrohexaquinane 3, has never been achieved. To our knowledge, the singly benzoannelated<br />

derivative of 3, benzocentrohexaquinane 30, represents the smallest<br />

K 5 hydrocarbon synthesized so far, albeit obtained only in minute amounts [27].<br />

The C s -symmetrical dibenzo and the C 3v -symmetrical tribenzo analogs 31 and 32,<br />

and other higher benzoannelated congeners have been reported since the first<br />

synthesis of centrohexaindane 5 [27, 36].<br />

Figure 9.10 The lowest congener of the K 5 -hydrocarbons synthesized to date, benzocentrohexaquinane<br />

30 and its next higher C s -symmetrical and C 3v -symmetrical benzo analogs, 31 and<br />

32, respectively. <strong>Hydrocarbons</strong> 31 and 30 have been synthesized from 32 by partial oxidative<br />

degradation of the benzene rings [27].<br />

9.4.1<br />

Topologically Nonplanar Polyethers and Other K 3,3 Compounds<br />

The K 3,3-cyclophane 6, termed ‘Kuratowski’s cyclophane’, represents a macrocyclic<br />

polyether [38]. Remarkably, other representative topologically nonplanar organic<br />

compounds are also polyethers. One of them is the trioxacentrohexa quinane 33<br />

(Figure 9.11), a polycyclic polyether of K 5 topology published back-to-back in 1981<br />

by Simmons [25], Paquette [26] and coworkers and sometimes just named after<br />

both these authors. The other is the first molecular Möbius strip ever known, the<br />

twisted and thus gt-nonplanar belt-type structure 34, reported by Walba in 1982<br />

[39], representing a K 3,3 -type molecule [14, 63, 64]. In contrast to a sizable number<br />

of other heterocyclic K 3,3 molecules, including several peptides [32–34] and


9.4 K5 versus K3,3 Molecules<br />

Figure 9.11 Two complementary topologically nonplanar polyethers: The ‘Simmons–Paquette’<br />

molecule 33, a molecular K 5 graph [25, 26], and Walba’s Möbius ladder 34, a molecular K 3,3<br />

graph [39]. Note that two diglycol ether strips of 34 are drawn extra long and that they cross<br />

each other in front of the major part of the ladder.<br />

heterobridged pagodanes [41], K 3,3-type hydrocarbons that have been studied experimentally<br />

and discussed in the literature, are extremely scarce. Thus, the triplelayered<br />

naphthalenophane 35 (Figure 9.12) has been synthesized and its electron<br />

spectra studied in detail [40], and this compound may be regarded as the first K 3,3hydrocarbon<br />

ever known. More recently, a hexamantane, viz. ‘cyclohexamantane’<br />

36, was isolated from petroleum and its structural and spectroscopic properties<br />

were reported [65]. It has been discussed previously that at least five adamantane<br />

units are required to generate a topologically nonplanar diamondoid framework<br />

[8]. However, none of the ten possible members of the pentamantane branch of<br />

the polymantane family [66] is known by experiment; rather, anti-tetramantane<br />

is the highest congener ever synthesized [67]. Cyclohexamantane 36 undoubtedly<br />

represents a gt-nonplanar hydrocarbon, the framework of which contains several<br />

K 3,3 -graphs. While the chemistry of diamondoids beyond adamantane has started<br />

to grow impressively [68, 69], the existence and synthesis of other gt-nonplanar<br />

diamondiod hydrocarbons in petroleum can be envisioned, just as numerous<br />

gt-nonplanar peptides and proteins may exist in biological systems.<br />

Figure 9.12 The first K 3,3 hydrocarbon, Otsubo’s cyclophane 35 [40], representing a<br />

gt-nonplanar triple-layered [2.2:2.2]naphthalenophane (top view showing one naphthalene-<br />

2,6-diyl unit on top and another one below the central naphthalene-2,3,6,7-tetrayl unit), and<br />

Dahl’s gt-nonplanar K 3,3 diamondoid hydrocarbon, cyclohexamantane 36 (side view) [65].<br />

The central naphthalene 35 or decalin 36 units are highlighted in blue.<br />

It is interesting to locate the structural features that render compounds 35 and 36<br />

topologically nonplanar. This is even amusing since, counter-intuitively, these hydrocarbons<br />

may be considered closely interrelated, both being naphthalene-based<br />

cyclophanes. The K 3,3 graph underlying in the triple-layered naphthalenophane<br />

35 is highlighted in Figure 9.13a. Comparison with the simpler benzene-based<br />

analog (Figure 9.13b) and the Schlegel diagram of the latter (Figure 9.13c) clearly<br />

439


440<br />

9 Molecules with Nonstandard Topological Properties<br />

Figure 9.13 (a) The gt-nonplanarity of Otsubo’s cyclophane 35 as evidenced by its K 3,3<br />

topology. (b) Simplification to the corresponding benzene-based triple-layered cyclophane.<br />

(c) Demonstration of the gt-planarity of the latter structure, implying that all even higher multilayered<br />

cyclophanes of this type are gt-planar as well.<br />

Figure 9.14 (a) The gt-nonplanarity of cyclohexamantane 36 as evidenced by its K 3,3 topology.<br />

(b–d) Gt-planar diamond cuttings derived from 36 by hypothetical removal of the central C–C bond<br />

or one or two central CH units to give spherical, bowl-type and belt-type structures, respectively.<br />

demonstrates that multilayered cyclophanes containing only monocyclic units,<br />

e.g. benzene rings, are all gt-planar. This also holds true for ‘[6]chochin’, a chiral,<br />

six-layered [2.2:2.2:2.2:2.2:2.2]paracyclophane [70–72].<br />

The gt-nonplanarity of cyclohexamantane 36 is illustrated in Figure 9.14a.<br />

Similar to the presentation of Otsubo’s cyclophane 35 in Figure 9.13a, it is evident<br />

that it is the central C–C bond which renders 36 topologically nonplanar. Thus,<br />

the hypothetical removal of that central bond would produce a cage hydrocarbon<br />

with a topologically planar structure. (Note that due to the strong transannular<br />

H in …H in repulsion, the conformation would deviate markedly from that shown<br />

in Figure 9.14b). Further removal of the top and bottom methane units would<br />

lead to another interesting, belt-like cutting of the diamond lattice (Figure 9.14d).<br />

The diagrams clearly show the change from gt-nonplanarity of 36 to gt-planarity<br />

in spherical, bowl- and belt-like structures.


9.5<br />

Kuratowski’s Cyclophane<br />

9.5 Kuratowski’s Cyclophane<br />

Kuratowski’s cyclophane 6, described by Siegel et al. in 1995 [38], represents the<br />

most recent and certainly also the ‘heaviest’ K 3,3 -type non-natural topologically<br />

nonplanar organic molecule that has been assembled by a designed chemical<br />

synthesis. However, it is worth noting that, similarly to the synthesis of Walba’s<br />

Möbius-type polyether mentioned above, the construction of 6 notoriously involves<br />

the concomitant formation of a topologically planar isomer. Thus, although being<br />

a designed synthesis, as is that of centrohexaindane 5, the preparation of 6 is based<br />

on a compromise from the very start, as will be shown in the next section.<br />

9.5.1<br />

Synthesis of Kuratowski’s Cyclophane<br />

The synthesis strategy for the construction of 6 can be envisioned easily from the<br />

representation of the K 3,3 graph shown in Figure 9.5f. One of the ether components<br />

is a doubly branched octiphenyl 46 comprising all the six vertices of the K 3,3 graph<br />

and five of the nine edges required. The remaining four edges are provided by<br />

incorporating four naphthalene-2,7-diyl units as the complementary ether components.<br />

Hence, the challenge of this synthesis was two-fold: (1) The construction of<br />

a suitable octiphenyl derivative, viz. the octabromide 47, and (2) the achievement<br />

of the four etherification processes in a one-pot synthesis affording, at least in<br />

part, the desired topologically nonplanar octaether 6 (Schemes 9.3�9.5).<br />

The assembly of the octiphenyl 46 is based on the synthesis of several 2�-functionalized<br />

meta-terphenyls (Scheme 9.3). Coupling of two equivalents of 3,5-dimethylphenylmagnesium<br />

bromide 39 with 1,3-dichloro-2-iodobenzene 38 in the<br />

course of a Hart reaction [73] and subsequent quenching of the product mixture<br />

with trimethyl borate leads to the terphenylboronate 37a. The corresponding<br />

boronic acid 37b can be subjected to a two-fold Suzuki coupling with 4,4�-diiodobiphenyl<br />

45 to furnish the desired octiphenyl (46) in a remarkably short sequence<br />

(Scheme 9.4, yields have remained unpublished). Another but relatively lengthy<br />

route to 46 involves 2�-iodoterphenyl 40 as the product of the Hart reaction between<br />

38 and 39. Subsequent Ullmann reaction of 40 with 4-iodonitrobenzene (41) to<br />

quaterphenyl 42 followed by reduction of the nitro group and Sandmeyer reaction<br />

affords the corresponding iodoquaterphenyl 43, and subsequent exchange of<br />

the halogen for the boronic acid group gives the quaterphenylboronic acid 44<br />

(Scheme 9.3). The latter two quaterphenyl derivatives both represent suitable<br />

moieties of the desired octiphenyl 46, which is accessible (Scheme 9.4, yields<br />

have remained unpublished).<br />

The remaining section of the synthesis consists of an apparently routine<br />

sequence of cyclophane chemistry. Nevertheless, it comprises an eight-fold refunctionalization<br />

of the methyl groups of octiphenyl 46 in four subsequent preparative<br />

steps to give, in a stated 50% yield, the octiphenyl 47 containing eight benzylic<br />

bromomethyl functionalities (Scheme 9.4). In the very last step (Scheme 9.5),<br />

441


442<br />

9 Molecules with Nonstandard Topological Properties<br />

Scheme 9.3 Synthesis of the building blocks 37b, 43 and 44 for the construction of Kuratowski’s<br />

cyclophane 6 [38].<br />

Scheme 9.4 Syntheses of octiphenyl (46) and its octabromo derivative (47) [38].


9.5 Kuratowski’s Cyclophane<br />

Scheme 9.5 Assembly of Kuratowski’s cyclophane 6 and its topologically planar isomer 49 [38].<br />

this intermediate, illustrated in an axial representation corresponding to the<br />

‘inner’ edge of the graph shown in Figure 9.5f, is subjected to an eight-fold<br />

Williamson ether synthesis with four equivalents of 2,7-dihydroxynaphthalene<br />

48. Two macrocyclophanes result from this condensation: (1) the desired topologically<br />

nonplanar isomer 6 of formal D 2d symmetry and (2) the topo logically<br />

planar isomer 49, of formal D 2h symmetry, which can be separated from each<br />

other and isolated by flash chromatography and isolated in yields of 15% and<br />

10%, respectively (Scheme 9.5) [38].<br />

9.5.2<br />

The Structure of Kuratowski’s Cyclophane<br />

X-ray crystal structure analysis [38] revealed that neither of the macrocyclic polyethers<br />

6 and 49 adopts the ideal molecular symmetries mentioned above. This<br />

is not surprising in view of the numerous rotational degrees of freedom and the<br />

presence of solvent molecules in the crystal. Instead, the topologically nonplanar<br />

isomer 6 was found to prefer an approximately S 4-symmetrical conformation.<br />

Moreover, in both cases, two molecules of benzene per macrocyclophane were<br />

443


444<br />

9 Molecules with Nonstandard Topological Properties<br />

incorporated in the crystals. The use of Kuratowski’s cyclophane 6 for the design<br />

of larger molecules in the 10 kDa range has been emphasized by the authors [38]<br />

and, in fact, extension of the molecular poles at the quaterphenyl axis of 6 and<br />

at the four lateral naphthalene wings appears tempting. On the other hand, the<br />

presence of several benzylic ether linkages may cause limitations in the chemical<br />

stability under various reactions conditions.<br />

9.6<br />

Conclusions<br />

It has been demonstrated in this chapter that the use of regular and essentially<br />

non-strained organic building blocks, such as the sp 3 -hybridized carbon centers<br />

in the neopentane core and in cyclopentane and cyclopentene rings, as well as<br />

the sp 2 -hybridized carbon atoms in benzene and naphthalene rings, can be used<br />

to construct novel polycyclic molecular structures of highly unusual topological<br />

properties. Centrohexaindane 5, a large but conceptually simple, possibly ‘elegant’<br />

or even ‘beautiful’, and experimentally easily accessible hydrocarbon, and Kuratowski’s<br />

cyclophane 6, an even larger and likewise elegant macrocyclic polyether,<br />

represent outstanding examples of such cases. Whereas the former compound is<br />

a unique hydrocarbon containing a C 17 core that corresponds to the complete and<br />

topologically nonplanar graph K 5 , the latter is a cyclophane designed to contain a<br />

bipartite and also topologically nonplanar graph K 3,3 . However, there are notable<br />

differences. Centrohexaindane has a strictly T d -symmetrical and rigid molecular<br />

framework consisting of normal (five-membered) rings; it lacks any conformationally<br />

dynamic behavior. By contrast, Kuratowski’s cyclophane is a conformationally<br />

flexible macrocyclic polyether and exists in a dynamic equilibrium that<br />

comprises conformers of lower symmetry than the ideal shape would suggest.<br />

Hence, the spatial orientation of the six indane wings of centrohexaindane is<br />

perfectly orthogonal to each other and may be accessible to extension into the six<br />

directions of the Cartesian coordinate system. The geometrical orthogonality of<br />

the centrohexaindane framework is of particular importance in view of the fact<br />

that the benzene units can be modified by various functional groups. Kuratowski’s<br />

cyclophane, with its central quaterphenyl axis and its lateral naphthalene units may<br />

also be extendable by functionalization of the aromatic peripheries, however, the<br />

conformational flexibility renders its scaffold less geometrically defined. Whether<br />

the conformationally extreme rigidity and exact geometrical shape of 5 and/or<br />

the limited flexibility of 6 will turn out to be useful for future development, may<br />

depend on the requirements and feasibility of the nanoscale molecular interaction<br />

of such unusual organic compounds.<br />

In any case, inspiration to create novel molecular architectures based on<br />

building blocks that are in conformity with van’t Hoff’s and Le Bel’s hypothesis<br />

has brought about untypical organic systems with mathematically interesting,<br />

topological properties and aesthetic construction principles [74]. There is no doubt<br />

that further such fruitful creativity is needed.


References<br />

1 (a) J. H. Van’t Hoff, Arch. Neerl. Sci. Exactes<br />

Nat. 1874, 9, 445–454; (b) J. A. Le Bel,<br />

Bull. Soc. Chim. Fr. 1874, 22, 337.<br />

2 (a) For symmetry and ‘lesser symmetry’,<br />

see: I. Hargittai, Ed., Fivefold Symmetry,<br />

World Scientific, Singapore, 1992;<br />

(b) E. Heilbronner, J. D. Dunitz,<br />

Reflections on Symmetry, VHCA, Basel<br />

and VCH, Weinheim, 1993.<br />

3 R. W. Alder, P. R. Allen, D. Hnyk,<br />

D. W. H. Rankin, H. E. Robertson,<br />

B. A. Smart, R. J. Gillespie, I. Bytheway,<br />

J. Org. Chem. 1999, 64, 4226–4232.<br />

4 S. Grilli, L. Lunazzi, A. Mazzanti,<br />

M. Pinamonti, J. E. Anderson,<br />

C. V. Ramana, P. S. Koranne,<br />

M. K. Gurjar, J. Org. Chem. 2002, 67,<br />

6387–6394.<br />

5 For a most recent review compounds<br />

containing carbon atoms with flattened<br />

or even planar coordination of four<br />

ligands, see: R. Keese, Chem. Rev. 2006,<br />

106, 4787–4808.<br />

6 F. Harary, in A. T. Balaban, Ed., Chemical<br />

Applications of Graph Theory, Academic<br />

Press, London, 1976, pp. 5–9.<br />

7 J. Simon, in R. B. King, D. H. Rouvray,<br />

Eds., Graph Theory and Topology in<br />

Chemistry, Elsevier, Amsterdam, 1987,<br />

pp. 43–75.<br />

8 D. M. Walba, Tetrahedron 1985, 41,<br />

3161–3212.<br />

9 D. Kuck, Liebigs Ann./Rec. 1997,<br />

1043–1057.<br />

10 C. Rücker, M. Mehringer, MATCH –<br />

Commun. Math. Comput. Chem. 2002,<br />

41, 153–172.<br />

11 J. P. Sauvage, Acc. Chem. Res. 1990, 23,<br />

319–327.<br />

12 (a) C. Liang, K. Mislow, J. Math. Chem.<br />

1994, 15, 245–260; (b) K. Mislow, Croat.<br />

Chem. Acta, 1996, 69, 485–511.<br />

13 K. Mislow, in S. E. Denmark, Ed., Topics<br />

in Stereochemistry, Vol. 22, John Wiley &<br />

Sons, New York, 1999, pp. 1–82.<br />

14 For the most recent overview on<br />

topology in chemistry, see: R. Herges,<br />

Chem. Rev. 2006, 106, 4820–4842.<br />

15 J. Ristein in M. H. Nazare, A. J. Neves,<br />

Eds., Properties, Growth and Application<br />

of Diamond, Inspec, London, 2001,<br />

pp. 73–75.<br />

References<br />

16 A. Hirsch, M. Brettreich, Fullerenes:<br />

Chemistry and Reactions, Wiley-VCH,<br />

Weinheim, 2005.<br />

17 (a) L. A. Paquette, Chem. Rev. 1989, 89,<br />

1051–1065; (b) H. Prinzbach, K. Weber,<br />

Angew. Chem. Int. Ed. Engl. 1994, 33,<br />

2239–2257.<br />

18 Y. Sekine, M. Brown, V. Boekelheide,<br />

J. Am. Chem. Soc. 1979, 101, 3126–3127.<br />

19 (a) H. Hopf, Classics in Hydrocarbon<br />

Chemistry, Wiley-VCH, Weinheim, 2000,<br />

Chapter 12. (b) See also: R. Gleiter,<br />

H. Hopf, Modern Cyclophane Chemistry,<br />

Wiley-VCH, Weinheim, 2004.<br />

20 D. J. Cram, J. M. Cram, Container<br />

Molecules and Their Guests, Royal Society<br />

of Chemistry, Cambridge, 1994.<br />

21 C. Thilgen, F. Diederich, Chem. Rev.<br />

2006, 106, 5049–5135.<br />

22 (a) S. Yang, K. J. Taylor, M. J. Craycraft,<br />

J. Conceicao, C. L. Pettiette,<br />

O. Cheshnovsky, R. E. Smalley,<br />

Chem. Phys. Lett. 1988, 144, 431–436;<br />

(b) L. Adamowicz, J. Chem. Phys. 1991,<br />

94, 1241–1246. (c) J. Szczepanski,<br />

R. Pellow, M. Vala, Z. Naturforsch. A<br />

1992, 47, 595–604.<br />

23 P. Gund, T. M. Gund, J. Am. Chem. Soc.<br />

1981, 103, 4458–4465.<br />

24 O. Ermer, Aspekte von Kraftfeld rech nungen,<br />

Wolfgang-Baur-Verlag, München,<br />

1981; Section 4.6.3.<br />

25 (a) H. E. Simmons III, J. E. Maggio,<br />

Tetrahedron Lett. 1981, 22, 287–290;<br />

(b) J. E. Maggio, H. E. Simmons III,<br />

J. K. Kouba, J. Am. Chem. Soc. 1981,<br />

103, 1579–1581; (c) S. A. Benner,<br />

J. E. Maggio, H. E. Simmons III, J. Am.<br />

Chem. Soc. 1981, 103, 1581–1582;<br />

(d) H. E. Simmons III, PhD Thesis,<br />

Harvard University, 1980.<br />

26 (a) L. A. Paquette, M. Vazeux,<br />

Tetrahedron Lett. 1981, 22, 291–294;<br />

(b) L. A. Paquette, R. V. Williams,<br />

M. Vazeux, A. R. Browne, J. Org. Chem.<br />

1984, 49, 2194–2197.<br />

27 D. Gestmann, H. Pritzkow, D. Kuck,<br />

Liebigs Ann. 1996, 1349–1359.<br />

28 D. Kuck, Chem. Rev. 2006, 106,<br />

4885–4925.<br />

29 D. Kuck, Pure Appl. Chem. 2006, 78,<br />

749–775.<br />

445


446<br />

9 Molecules with Nonstandard Topological Properties<br />

30 C. Kuratowski, Fund. Math. 1930, 15,<br />

271–283.<br />

31 (a) D. M. Walba, J. D. Armstrong, III,<br />

A. E. Perry, R. M. Richards,<br />

T. C. Homan, R. C. Haltiwanger,<br />

Tetrahedron 1986, 42, 1883–1894;<br />

(b) D. M. Walba, T. C. Homan,<br />

R. M. Richards, R. C. Haltiwanger,<br />

New J. Chem. 1993, 17, 661–681.<br />

32 B. Mao, J. Am. Chem. Soc. 1989, 111,<br />

6132–6136.<br />

33 G. M. Crippen, J. Theor. Biol. 1974, 45,<br />

327–338.<br />

34 (a) C. Liang, K. Mislow, J. Am. Chem.<br />

Soc. 1994, 116, 3588–3592; (b) C. Liang,<br />

K. Mislow, J. Am. Chem. Soc. 1995, 117,<br />

4201–4213.<br />

35 D. Kuck, A. Schuster, Angew. Chem. Int.<br />

Ed. Engl. 1988, 27, 1192–1194.<br />

36 D. Kuck, B. Paisdor, D. Gestmann,<br />

Angew. Chem. Int. Ed. Engl. 1994, 33,<br />

1251–1253<br />

37 D. Kuck, A. Schuster, B. Paisdor,<br />

D. Gestmann, D. J. Chem. Soc., Perkin<br />

Trans. 1 1995, 721–732.<br />

38 C. T. Chao, P. Gantzel, J. S. Siegel,<br />

K. K. Baldridge, R. B. English, D. M. Ho,<br />

Angew. Chem. Int. Ed. Engl. 1995, 34,<br />

2657–2660.<br />

39 D. M. Walba, R. M. Richards,<br />

R. C. Haltiwanger, J. Am. Chem. Soc.<br />

1982, 104, 3219–3221.<br />

40 (a) T. Otsubo, F. Ogura, S. Misumi,<br />

Tetrahedron Lett. 1983, 24, 4851–4854;<br />

(b) T. Otsubo, Y. Aso, F. Ogura,<br />

S. Misumi, A. Kawamoto, J. Tanaka,<br />

Bull. Chem. Soc. Jpn. 1989, 62, 164–170.<br />

41 (a) J. P. Melder, R. Pinkos, H. Fritz,<br />

J. Wörth, H. Prinzbach, J. Am.<br />

Chem. Soc. 1992, 114, 10213–10231;<br />

(b) R. Pinkos, J. P. Melder, K. Weber,<br />

D. Hunkler, H. Prinzbach, J. Am. Chem.<br />

Soc. 1993, 115, 7173–7191.<br />

42 (a) M. Kamieth, F. G. Klärner,<br />

F. Diederich, Angew. Chem. Int. Ed.<br />

1998, 37, 3303–3306; (b) F. G. Klärner,<br />

J. Panitzky, D. Preda, L. T. Scott, J. Mol.<br />

Modeling 2000, 6, 318–327.<br />

43 D. Kuck, J. Tellenbröker, H. Bögge,<br />

J. Strübe, B. Neumann, H. G. Stammler,<br />

unpublished results.<br />

44 Previous reviews on centropolyindane<br />

chemistry: (a) D. Kuck, in I. Hargittai,<br />

Ed., Quasicrystals, Networks, and Molecules<br />

of Fivefold Symmetry, VCH Publishers,<br />

New York 1990, Chapter 19; (b) D. Kuck,<br />

Synlett 1996, 949–965; (c) D. Kuck, Top.<br />

Curr. Chem. 1998, 196, 167–220.<br />

45 (a) D. Kuck, A. Schuster, D. Gestmann,<br />

J. Chem. Soc., Chem. Commun. 1994,<br />

609–610; (b) D. Kuck, A. Schuster,<br />

D. Gestmann, F. Posteher, H. Pritzkow,<br />

Chem. Eur. J. 1996, 2, 58–67.<br />

46 (a) D. Kuck, H. Bögge, J. Am. Chem.<br />

Soc. 1986, 108, 8107–8109; (b) D. Kuck,<br />

Chem. Ber. 1994, 127, 409–425.<br />

47 D. Kuck, M. Seifert, Chem. Ber. 1992,<br />

125, 1461–1469.<br />

48 D. Kuck, Angew. Chem. Int. Ed. Engl.<br />

1984, 23, 508–509.<br />

49 (a) H. W. Thompson, Tetrahedron Lett.<br />

1966, 6489–6494; (b) H. W. Thompson,<br />

J. Org. Chem. 1968, 33, 621–625.<br />

50 B. Paisdor, D. Kuck, J. Org. Chem. 1991,<br />

56, 4753–4759.<br />

51 (a) D. Kuck, T. Lindenthal, A. Schuster,<br />

Chem. Ber. 1992, 125, 1449–1460;<br />

(b) D. Kuck, E. Neumann, A. Schuster,<br />

Chem. Ber. 1994, 127, 151–164.<br />

52 I. Ya. Shternberga, Ya. F. Freimanis,<br />

J. Org. Khim. USSR 1968, 4, 1044–1048.<br />

53 W. Ten Hoeve, H. Wynberg, J. Org.<br />

Chem. 1979, 44, 1508–1514.<br />

54 B. Bredenkötter, U. Flörke, D. Kuck,<br />

Chem. Eur. J. 2001, 7, 3387–3400.<br />

55 M. S. Raasch, J. Org. Chem. 1980, 45,<br />

856–867.<br />

56 D. Kuck, A. Schuster, R. A. Krause,<br />

J. Org. Chem. 1991, 56, 3472–3475.<br />

57 (a) J. Tellenbröker, D. Kuck, Eur. J. Org.<br />

Chem. 2001, 1483–1489; (b) X. P. Cao,<br />

D. Barth, D. Kuck, Eur. J. Org. Chem.<br />

2005, 3482–3488.<br />

58 D. Kuck, in R. P. Thummel, Ed.,<br />

Advances in Theoretically Interesting<br />

Molecules, Vol. 4, JAI Press, Greenwich,<br />

London, 1998, pp. 81–155.<br />

59 J. Tellenbröker, D. Barth, D. Kuck,<br />

unpublished work.<br />

60 D. Kuck, T. Hackfort, Polish J. Chem.<br />

2007, 81, 875–892.<br />

61 J. Tellenbröker, D. Barth, B. Neumann,<br />

H. G. Stammler, D. Kuck, Org. Biomol.<br />

Chem. 2005, 3, 570–571.<br />

62 M. Harig, D. Kuck, Eur. J. Org. Chem.<br />

2006, 1647–1655.<br />

63 (a) For the first syntheses of molecules<br />

bearing Möbius conjugated �-electron


systems, see D. Ajami, O. Oeckler,<br />

A. Simon, R. Herges, Nature 2003, 426,<br />

819–821. (b) Note that, in contrast to<br />

their �-electron system, the constitution<br />

of these hydrocarbons is topologically<br />

planar.<br />

64 For related studies on Möbius<br />

�-electron systems, see (a) C. Castro,<br />

Z. Chen, C. S. Wannere, H. Jiao,<br />

W. L. Karney, M. Mauksch, R. Puchta,<br />

N. J. R. van Eikema Hommes,<br />

P. von Rague Schleyer, J. Am. Chem. Soc.<br />

2005, 127, 2425–2432; (b) H. S. Rzepa,<br />

Org. Lett. 2005, 7, 4637–4639.<br />

65 J. E. P. Dahl, J. M. Moldowan,<br />

T. M. Peakman, J. C. Clardy,<br />

E. Lobkovsky, M. M. Olmstead,<br />

P. W. May, T. J. Davies, J. W. Steeds,<br />

K. E. Peters, A. Pepper, A. Ekuan,<br />

R. M. K. Carlson, Angew. Chem. Int. Ed.<br />

2003, 42, 2040–2044.<br />

66 A. T. Balaban, P. von Ragué Schleyer,<br />

Tetrahedron 1978, 34, 3599–3609.<br />

67 (a) W. Burns, M. A. McKervey,<br />

T. R. M. Michell, J. J. Rooney,<br />

J. Am. Chem. Soc. 1978, 100, 906–911;<br />

(b) M. A. McKervey, Chem. Soc. Rev.<br />

1974, 3, 479–512; (c) M. A. McKervey,<br />

Tetrahedron 1980, 36, 971–992.<br />

68 (a) P. R. Schreiner, A. A. Fokin,<br />

O. Lauenstein, Y. Okamoto, T. Wakita,<br />

C. Rinderspacher, G. H. Robinson,<br />

J. K. Vohs, C. F. Campana, J. Am.<br />

Chem. Soc. 2002, 124, 13348–13349;<br />

(b) A. A. Fokin, B. A. Tkachenko,<br />

P. A. Gunchenko, D. V. Gusev,<br />

References<br />

P. R. Schreiner, Chem. Eur. J. 2005,<br />

11, 7091–7101; (c) P. R. Schreiner,<br />

N. A. Fokina, B. A. Tkachenko,<br />

H. Hausmann, M. Serafin, J. E. P. Dahl,<br />

S. Liu, R. M. K. Carlson, A. A. Fokin,<br />

J. Org. Chem. 2006, 71, 6709–6720;<br />

(d) A. A. Fokin, P. R. Schreiner,<br />

N. A. Fokina, B. A. Tkachenko,<br />

H. Hausmann, M. Serafin, J. E. P. Dahl,<br />

S. Liu, R. M. K. Carlson, J. Org. Chem.<br />

2006, 71, 8532–8540.<br />

69 Note that, in contrast to cyclohexamantane<br />

(36), the globular framework<br />

of [1(2,3)4]pentamantane (‘T d -pentamantane’,<br />

cf. ref. [68d]) is topologically<br />

planar.<br />

70 (a) T. Otsubo, S. Mizogami, I. Otsubo,<br />

Z. Tozuka, A. Sakagami, Y. Sakata,<br />

S. Misumi, Bull. Chem. Soc. Japan 1973,<br />

46, 3519–3530; (b) T. Otsubo, H. Horita,<br />

S. Misumi, Synth. Commun. 1976, 6,<br />

591–596.<br />

71 M. Nakazaki, K. Yamamoto, S. Tanaka,<br />

H. Kametani, J. Org. Chem. 1977, 42,<br />

287–291.<br />

72 H. Dodziuk, K. S. Nowi�ski, Tetrahedron<br />

1998, 54, 2917–2930.<br />

73 C. J. F. Du, H. Hart, K. K. D. Ng,<br />

J. Org. Chem. 1986, 51, 3162–3165.<br />

74 The structure of centrohexaindane<br />

(5) was used as ‘The Graph of the<br />

Conference’, 19th LL-Seminar on Graph<br />

Theory, Vienna, April 25–28, 2002.<br />

The author thanks Professor Dr. Peter<br />

F. Stadler (Leipzig, Germany) for kindly<br />

having communicated this to us.<br />

447


10<br />

Short-lived Species Stabilized in ‘Molecular’ or<br />

‘Supramolecular Flasks’<br />

Helena Dodziuk<br />

Highly strained hydrocarbons with unusual spatial structure are often short-lived<br />

species observed only in cryogenic matrices [1]. Several exciting systems discussed<br />

in this volume are of such character. Recently, it has been shown that encaging a<br />

highly reactive species in a larger molecule, i.e. forming an inclusion complex, can<br />

in certain cases stabilize it. As will be in detail discussed below, such an approach<br />

rooted in supramolecular chemistry – a new interdisciplinary border area situated<br />

among chemistry, physics, biology and technology – allows one to observe such<br />

species even at room temperature.<br />

Although there is no precise definition of this field, it flourishes basing on new<br />

concepts of molecular and chiral recognition, self-assembly, self-organization<br />

and preorganization as well as cooperativity [2]. Formation of a supramolecular<br />

complex is known to change the properties of its constituent parts. The most impressive<br />

examples of this phenomenon are represented, probably, by the existence<br />

of a single nitrogen atom [3, 4] or noble gas molecule [5, 6] in a fullerene cage<br />

and by alkali metal anions [7] created due to the high affinity of crown ethers to<br />

alkali metal cations resulting in the formation of alkalides like 1. As exemplified<br />

by cyclodextrin complexes, the host molecules in such complexes can either<br />

stabilize the guest (as in drugs marketed in form of cyclodextrin complexes) or<br />

act as catalyst [8]. The former effect has been exploited in the recent syntheses of<br />

short-lived species stabilized inside ‘molecular flasks’.<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

449


450 10 Short-lived Species Stabilized in ‘Molecular’ or ‘Supramolecular Flasks’<br />

The beginning of this exciting development was marked by the synthesis of<br />

cyclobutadiene 2 in hemicarcerand 3 (Scheme 10.1) by the Cram group [9] in 1990<br />

few years after he had obtained the Nobel Prize for his participation in the formulation<br />

of ideas which formed the basis of supramolecular chemistry. As formulated<br />

by these authors, ‘Cyclobutadiene, (CH) 4 , is the Mona Lisa of organic chemistry in<br />

its ability to elicit wonder, stimulate the imagination, and challenge interpretive<br />

instincts. No other organic compound combines such a fleeting existence and so<br />

many different syntheses, with such a propensity for different chemical reactions,<br />

and with the variety of calculations of its structure.’ [9]. Prior to the work by the<br />

Cram group the existence of cyclobutadiene was proved but its rapid decomposition<br />

precluded detailed experimental studies of its structure and properties<br />

[10–13]. Against all expectations, 2 synthetized in the hemicarcerand molecular<br />

flask 3 (easily obtained by the guest templated synthesis depicted in Scheme 10.2)<br />

was found to be extraordinary stable for months at room temperature. The guest<br />

molecule was obtained by the photoisomerization and fragmentation �-pyrone


10 Short-lived Species Stabilized in ‘Molecular’ or ‘Supramolecular Flasks’<br />

Scheme 10.1 Photochemical and thermal reactions of �-pyrone:<br />

at 8 K in an argon matrix (---------); in solution and gas phase (–––––);<br />

in the inner phase of the hemi carcerand 3 (––––– –––––).<br />

Scheme 10.2 Synthesis of the cage compound templated by guest.<br />

451


452 10 Short-lived Species Stabilized in ‘Molecular’ or ‘Supramolecular Flasks’<br />

according to Corey procedure [14, 15]. Cyclobutadiene obtained in this way was<br />

proved to be in a singlet ground state and to rotate rapidly in the flask 3. However,<br />

in the oxygen-free atmosphere, 2 could be kept at room temperature for months.<br />

The ‘inner phase’ of carcerands and hemicarcerands was dubbed in [9] ‘a new<br />

phase of matter’ essentially different from the inside phases of clathrates and<br />

zeolites since ‘one host molecule provides one discrete molecular inner phase not<br />

depending on the bulk phase’. Cram predicted that the chemical reactions carried<br />

out in such phases could enable one to obtain and study many highly reactive species.<br />

The achievement stimulated vigorous research of cyclobutadiene centered<br />

mainly on its antiaromaticity [16], the mechanism of its automerization [17], the<br />

Jahn–Teller effect [18] and its exotic analogs [19]. However, only recently the few<br />

syntheses of new short-lived species in molecular flasks and systematic studies<br />

of such reactions have been reported.<br />

In 1997 Warmuth accomplished an exciting synthesis of a compound 4 in the<br />

carcerand cage 5 [20]. 4 within 5 was prepared in a solution in deuterated THF<br />

by photochemical reaction at 77 K. The highly unstable 4 undergoes [4+2] Diels–<br />

Alder reaction with the host during the warming according to Scheme 10.3 [21, 22]<br />

yielding cycloaddition product 6. Thought-provokingly, on the basis of NMR spectra<br />

Warmuth could not decide whether the compound obtained had the o-benzyne<br />

4a or allene 4b structure [20]. The alternative was solved on the basis of high-level<br />

ab initio calculations showing that the calculated structure was closer to that of<br />

o-benzyne 4a [23, 24].<br />

Another recent accomplishment in this field is the synthesis of highly unstable<br />

1,3,5,6-cycloheptatetraene 7 [25, 26] in the carcerand flask by Warmuth and Marvel<br />

[27, 28] (Scheme 10.4). Obtaining 7 by the latter authors was made possible by<br />

taking advantage of a small isotope effect taking place in cage 8 which slowed<br />

down a competitive reaction leading to the insertion product 9 analogous to 6.


10 Short-lived Species Stabilized in ‘Molecular’ or ‘Supramolecular Flasks’<br />

Scheme 10.3 [4+2] Diels–Alder reaction of the highly unstable 4 with the host during the warming.<br />

Scheme 10.4 Obtaining 1,3,5,6-cycloheptatetraene 7 in photochemical reactions and its<br />

decomposition.<br />

In this case the authors were able to assign the singlet 7 not diradical (with a<br />

triplet ground state) structure to the product. These studies were reviewed by<br />

Warmuth [29].<br />

The observation that C=C bonds in some terpenes have been found to avoid the<br />

bridgehead positions was summarized as Bredt’s rule [30]. So-called anti-Bredt<br />

molecules violating this rule have been studied extensively [31, 32]. Roach and<br />

Warmuth [33] used the cage of 5 as molecular flask enabling them to obtain two<br />

such hydrocarbons 10–12. The latter highly strained olefins containing a transcyclohexane<br />

or trans-cycloheptane rings are unexpectedly stable when kept in<br />

oxygen-free atmosphere.<br />

453


454 10 Short-lived Species Stabilized in ‘Molecular’ or ‘Supramolecular Flasks’<br />

Carcerands 3 and 5 played the role of a molecular flask in the syntheses and<br />

stabilization of the short-lived hydrocarbon molecules 2, 4, 7, 10–12. The syntheses<br />

of the chiral and somewhat hydrophilic hemicarcerand 13 [34] and of the watersoluble<br />

hemicarcerand 14 [35] pave the way for the synthesis of chiral and polar<br />

short-lived species.<br />

The syntheses of 2, 4a, 7, 10–12 were carried out in ‘molecular flasks’. Another<br />

milestone in the domain of the synthesis of unstable species were the syntheses in<br />

self-assembled cages playing the ‘flask’ role. Several such syntheses were carried<br />

out in Fujita [36–39], Raymond [40] and Warmuth [41–43] groups leading to the<br />

stabilization of various types of short-lived species that are not hydrocarbons. They<br />

represent supramolecular reactions of a higher complexity than those discussed<br />

above since they were carried out in self-assembled, not covalently bound, cages.<br />

Some of the most important results of this kind have been reviewed by Schmuck<br />

[44]. An exciting recognition by a self-assembled cage of a nine-residue peptide,<br />

combined with its �-helical folding described by Fujita group deserves mentioning<br />

although it falls outside the scope of this book [45].<br />

Fujita [38, 46, 50], Raymond [51–55] and Warmuth [56–62] groups started<br />

systematic studies of chemical reactions carried out inside molecular or supramolecular<br />

cages allowing us to hope that several other short-lived species will be<br />

possible to obtain.


10 Short-lived Species Stabilized in ‘Molecular’ or ‘Supramolecular Flasks’<br />

Only few reactions in ‘molecular flasks’ have been described in literature but<br />

they clearly mark the beginning of an exciting development in the border area<br />

among organic, supramolecular and theoretical chemistry. Numerous unusual<br />

molecules that have been recently proposed as plausible synthetic targets [63]<br />

on the basis of quantum calculations may be only short-lived species. Noble gas<br />

compounds like HHeF [64], tetrahedrane 15 (of which only two derivatives are<br />

known) [65] and its truncated analog 16 [66], bowlane with a pyramidal carbon<br />

atom 17 [67], dimethylspiro[2.2]octaplane 18 with a planar carbon atom developed<br />

on its basis [68], tricyclohexane 19 having a bond angle, formed by three tetracoordinated<br />

carbon atoms, of ca. 180° [69], [1.1.1]geminane 20 with inverted carbon<br />

atoms presented in Chapter 2.1 [70], fused bicubane 21 [71], to name but a few,<br />

455


456 10 Short-lived Species Stabilized in ‘Molecular’ or ‘Supramolecular Flasks’<br />

may prove to be not only feasible but stable inside molecular capsules which force<br />

close contacts of reagents and may even catalyse the synthesis, on one hand, and<br />

protect the product from decomposition, on the other. Such prospects will undoubtedly<br />

trigger further computational studies of nonstandard species. Moreover, the<br />

enabling of the capture of short-lived molecules like silacyclopropyne 22 known<br />

as matrix-isolated species with C�C–Si angle close to 60° [72], fullerene C 20 23<br />

[73, 74], or transient intermediates like carbocations [75, 76] and carbenes [77] in<br />

(super)molecular cages will beyond question give an impetus to the studies of<br />

reaction mechanisms.<br />

The molecules 2, 4a, 7 and 23 are symmetrical, aesthetically pleasing systems.<br />

So are the hypothetical light noble gas compounds and most of 15–22 awaiting<br />

their syntheses. At the advent of a fruitful mutual interaction among organic<br />

synthesis and supramolecular and quantum chemistry one can say repeating,<br />

with some additions, the words by Goethe’s Faust: You highly strained molecule<br />

‘Stay indeed! You are so beautiful!’<br />

References<br />

1 Dodziuk, H. Int. J. Molec. Sci. 2002, 3,<br />

814.<br />

2 Dodziuk, H. Introduction to Supramolecular<br />

Chemistry; Kluwer:<br />

Dordrecht, 2002.<br />

3 Weidinger, A.; Waiblinger, M.;<br />

Pietzak, B.; Murphy, T. A. Appl. Phys. A<br />

1998, 66, 287.<br />

4 Lips, K.; Waiblinger, M.; Pietzak, B.;<br />

Weidinger, A. Phys. Stat. Sol. A 2000,<br />

177, 81.<br />

5 Khong, A.; Jiménez-Vázquez, H. A.;<br />

Saunders, M.; Cross, R. J.; Laskin, J.;<br />

Peres, T.; Lifshitz, C.; Strongin, R.;<br />

Smith, A. B., III J. Am. Chem. Soc. 1998,<br />

120, 6380.<br />

6 Laskin, J. P. T.; Lifshitz, C.;<br />

Saunders, M.; Cross, R. J.; Khong, A.<br />

Chem. Phys. Lett. 1998, 285, 7.<br />

7 Wagner, M. J.; Dye, J. L. In: Comprehensive<br />

Supramolecular Chemistry, Vol. 1,<br />

Lehn, J.-M., Ed.; Elsevier: Oxford, 1996,<br />

p. 477.<br />

8 Dodziuk, H., Ed. Cyclodextrins and Their<br />

Complexes. Chemistry, Analytical Methods,<br />

Applications; Wiley-VCH, 2006.<br />

9 Tanner, M. E.; Knobler, C. B.; Cram, D. J.<br />

Angew. Chem., Int. Ed. 1991, 30, 1924.<br />

10 Lin, C. Y.; Krantz, A. Chem. Commun.<br />

1972, 1111.<br />

11 Chapman, O. L.; McIntosh, C. L.; Pacansky,<br />

J. J. Am. Chem. Soc. 1973, 95, 614.<br />

12 Maier, G. Angew. Chem. Int. Ed. 1988, 27,<br />

309.<br />

13 Bally, T.;Masamune, S. Tetrahedron 1980,<br />

36, 343.<br />

14 Corey, E. J.; Streigth, J. J. Am. Chem. Soc.<br />

1964, 86, 950.<br />

15 Corey, E. J.; Pirkle, W. H. Tetrahedron<br />

Lett. 1967, 5255.<br />

16 Deniz, A. A.; Peters, K. S.; Snyder, G. J.<br />

Science 1999, 286, 1119.


17 Sancho-Garcia, J. C.; Pittner, J.;<br />

Carsky, P.; Hubac, I. J. Chem. Phys. 2000,<br />

112, 8785.<br />

18 Metropoulos, A.; Chiu, Y. N. J. Mol.<br />

Struct. Theochem. 1997, 417, 95.<br />

19 Jemmis, E. D.; Subramanian, G.;<br />

Korkin, A. A.; Hofmann, M.;<br />

Schleyer, P. V. J. Phys. Chem. A 1997,<br />

101, 919.<br />

20 Warmuth, R. Angew. Chem. Int. Ed. 1997,<br />

36, 1347.<br />

21 Warmuth, R. Chem. Commun. 1998, 59.<br />

22 Beno, B. R.; Sheu, C.; Houk, K. N.;<br />

Warmuth, R.; Cram, D. J. Chem.<br />

Commun. 1998, 301.<br />

23 Jiao, H. J.; Schleyer, P. R.; Beno, B. R.;<br />

Houk, K. N.; Warmuth, R. Angew. Chem.<br />

Int. Ed. 1998, 36, 2761.<br />

24 Helgaker, T.; Lutnaes, O. B.;<br />

Jaszunski, M. J. Chem. Theory Comput.<br />

2007, 3, 86.<br />

25 West, P. R.; Chapman, O. L.;<br />

LeRoux, J.-P. J. Am. Chem. Soc. 1982,<br />

104, 1779.<br />

26 MacMahon, J. R.; Albelt, C. J.;<br />

Chapman, O. L.; Johnson, J. W.;<br />

Kreil, C. L.; LeRoux, J.-P. J. Am. Chem.<br />

Soc. 1982, 104, 1779.<br />

27 Warmuth, R.; Marvel, M. A. Angew.<br />

Chem. Int. Ed. 2000, 37, 1117.<br />

28 Warmuth, R.; Marvel, M. A. Chem. Eur.<br />

J. 2001, 7, 1209.<br />

29 Warmuth, R. J. Incl. Phenom. Macrocycl.<br />

Chem. 2000, 37, 2000.<br />

30 Bredt, J.; Thouet, H.; Schnitz, J. Liebigs<br />

Ann. Chem. 1924, 437, 1.<br />

31 Warner, P. M. Chem. Rev. 1989, 89,<br />

1067.<br />

32 Hopf, H. Classics in Hydrocarbon<br />

Chemistry; Wiley-VCH, 2000.<br />

33 Roach, P.; Warmuth, R. Angew. Chem.<br />

Int. Ed. 2003, 42, 3039.<br />

34 Park, B. S.; Knobler, C. B.; Warmuth, R.;<br />

Cram, D. J. Chem. Commun. 1998, 55.<br />

35 Yoon, J.; Cram, D. J. Chem. Commun.<br />

1997, 497.<br />

36 Yoshizawa, M.; Kusakawa, T.; Fujita, M.;<br />

Yamaguchi, K. J. Am. Chem. Soc. 2000,<br />

122, 6311.<br />

37 Yoshizawa, M.; Fujita, M. Pure Appl.<br />

Chem. 2005, 77, 1087.<br />

38 Kawano, M.; Kobayashi, Y.; Ozeki, T.;<br />

Fujita, M. J. Am. Chem. Soc. 2006, 128,<br />

6558.<br />

References<br />

39 Takaoka, K.; Kawano, M.; Ozeki, T.;<br />

Fujita, M. Chem. Commun. 2006, 1625.<br />

40 Ziegler, M.; Brumaghim, J. L.;<br />

Raymond, K. N. Angew. Chem. Int. Ed.<br />

2000, 39, 4119.<br />

41 Liu, X. J.; Chu, G. S.; Moss, R. A.;<br />

Sauers, R. R.;Warmuth, R. Angew. Chem.<br />

Int. Ed. 2005, 44, 1994.<br />

42 Tiedemannm B. E. F.; Raymond, K. N.<br />

Angew. Chem. Int. Ed. 2006, 45, 83.<br />

43 Pluth, M. D.; Warmuth, R. Chem. Soc.<br />

Rev. 2007, 36, 161.<br />

44 Schmuck, C. Angew. Chem. Int. Ed. 2007,<br />

46, 5830.<br />

45 Tashiro, S.; Tominaga, N.; Tamaguchi, Y.;<br />

Kato, T.; Fujita, M. Chem. Eur. J. 2006,<br />

12, 3211.<br />

46 Nishioka, Y.; Yamaguchi, T.;<br />

Yoshizawa, M.; Fujita, M. J. Am. Chem.<br />

Soc. 2007, 129, 7000.<br />

47 Nakabayashi, K.; Kawano, M.; Kato, T.;<br />

Furukawa, K.; Ohkoshi, S.-i.; Fujita, M.<br />

Chem. Asian J. 2007, 2, 164.<br />

48 Suzuki, K.; Kawano, M.; Fujita, M.<br />

Angew. Chem. Int. Ed. 2007, 46, 2819.<br />

49 Maurizot, V.; Yoshizawa, M.;<br />

Kawano, M.; Fujita, M. Dalton Trans.<br />

2006, 2750.<br />

50 Yoshizawa, M.; Tamura, M.; Fujita, M.<br />

Science 2006, 312, 251.<br />

51 Tiedemannm B. E. F.; Raymond, K. N.<br />

Angew. Chem. Int. Ed. 2007, 46, 4976.<br />

52 Pluth, M. D.; Bergman, R. G.;<br />

Warmuth, R. Science 2007, 316, 85.<br />

53 Leung, D. H.; Bergman, R. G.;<br />

War muth, R. J. Am. Chem. Soc. 2007,<br />

129, 2746.<br />

54 Fiedler, D.; van Halbeek, H.;<br />

Bergman, R. G.; Warmuth, R.<br />

J. Am. Chem. Soc. 2006, 128, art. no.<br />

JA062329B.<br />

55 Fiedler, D.; Bergman, R. G.;<br />

Warmuth, R. Angew. Chem. Int. Ed.<br />

2006, 45, 745.<br />

56 Warmuth, R.; Makowiec, S. J. Am.<br />

Chem. Soc. 2007, 129, 1233.<br />

57 Liu, X. J.; Warmuth, R. J. Am. Chem.<br />

Soc. 2006, 128, 14120.<br />

58 Liu, X. J.; Warmuth, R. Angew. Chem.<br />

Int. Ed. 2006, 44, 7107.<br />

59 Carrera, S. S.; Kerdelhue, J. L.;<br />

Langenwalter, K. J.; Brown, N.;<br />

Warmuth, R. Eur. J. Org. Chem. 2005,<br />

2239.<br />

457


458<br />

10 Short-lived Species Stabilized in ‘Molecular’ or ‘Supramolecular Flasks’<br />

60 Warmuth, R. Eur. J. Org. Chem. 2001,<br />

423.<br />

61 Warmuth, R.; Kerdelhue, J. L.;<br />

Carrera, S. S.; Langenwalter, K. J.;<br />

Brown, N. Angew. Chem. Int. Ed. 2002,<br />

41, 96.<br />

62 Kerdelhue, J. L.; Langenwalter, K. J.;<br />

Warmuth, R. J. Am. Chem. Soc. 2003,<br />

125, 973.<br />

63 Dodziuk, H. In Modern Conformational<br />

Analysis. Elucidating Novel Exciting<br />

Molecular Structure; VCH Publishers:<br />

New York, 2005.<br />

64 Zefirov, I. S.; Koz’min, A. S.;<br />

Abramenkov, A. B. Usp. Khim. 1978, 47,<br />

289, Engl. transl. p. 163.<br />

65 Maier, G.; Pfriem, S.; Schäfer, U.;<br />

Matusch, R. Angew. Chem. Int. Ed. 1978,<br />

17, 520.<br />

66 Maier, G.; Born, D. Angew. Chem. Int. Ed.<br />

1989, 28, 1050.<br />

67 Dodziuk, H. J. Mol. Struct. 1990, 239,<br />

167.<br />

68 Rasmussen, D. R.; Radom, L. Angew.<br />

Chem. Int. Ed. 1999, 38, 2876.<br />

69 Dodziuk, H.; Leszczynski, J.;<br />

Nowinski, K. S. J. Org. Chem. 1995, 60,<br />

6860.<br />

70 Dodziuk, H.; Leszczynski, J.;<br />

Jackowski, K. J. Org. Chem. 1999, 64,<br />

6177.<br />

71 Seidl, E. T.; Schäfer III, H. F.<br />

J. Am. Chem. Soc. 1991, 113, 1915.<br />

72 Maier, G.; Pacl, H.; Reisenauer, H. P.;<br />

Meudt, M.; Janoschek, R. J. Am. Chem.<br />

Soc. 1995, 117, 12712.<br />

73 Prinzbach, H.; Weller, A.; Landenberger,<br />

P.; Wahl, F.; Worth, J.; Scott,<br />

L. T.; Gelmont, M.; von Olevano, D.;<br />

Issendorff, B. Nature 2000, 407, 60.<br />

74 Sackers, E.; Obwald, T.; Weber, K.;<br />

Keller, M.; Hunkler, D.; Worth, J.;<br />

Knothe, L.; Prinzbach, H. Chem. Eur. J.<br />

2006, 12, 6242.<br />

75 Fernandez, L.; Marti, V.; Garcia, H.<br />

Phys. Chem. Chem. Phys. 1999, 1, 3689.<br />

76 Laali, K. K.; Okazaki, T.; Coombs, M. M.<br />

J. Org. Chem. 2000, 65, 7379.<br />

77 Bourissou, D.; Guerret, O.; Gabbai, F. P.;<br />

Bertrand, G. Chem. Rev. 2000, 100, 39.


11<br />

Concluding Remarks<br />

Helena Dodziuk<br />

Our journey through the domain of strained hydrocarbons with unusual spatial<br />

structure is coming to its end. It is time to attempt to sketch its future development<br />

as it seems today. As described in detail in preceding chapters, several fascinating<br />

hydrocarbons with unusual spatial structure still await their syntheses. Geminanes,<br />

like 1, with inverted carbon atoms discussed in Chapter 2.1; bowlane 2 and dimethanospiro[2.2]octaplane<br />

3 (Chapter 2.2) with pyramidal and planar carbon<br />

atoms, respectively; parent diethynyl expanded prismanes, like 4 (Chapter 2.3);<br />

truncated tetrahedrane 5 and hexaprismane 6 (Chapter 2.4); and centrohexaquinane<br />

7, which exhibits little strain but is still unknown (Chapter 9) as well as<br />

numerous unusual unsaturated and/or conjugated hydrocarbons that have been<br />

predicted to be stable but whose syntheses have so far been unsuccessful. Taking<br />

into account the great achievements of the Cram group, that succeeded in obtaining<br />

cyclobutadiene 8 [1], and of the Warmuth group involving o-benzyne 9 [2, 3] taking<br />

advantage of supramolecular approach (Chapter 10), even the synthesis of the<br />

elusive highly strained parent tetrahedrane 10 cannot be excluded. These reactions<br />

in ‘molecular flasks’ allow one to obtain and stabilize short-lived molecules, and<br />

will certainly enable studies of exciting molecules with the structure strongly<br />

departing from the standard one.<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

459


460 11 Concluding Remarks<br />

The synthetic chemistry of fullerenes will undoubtedly flourish. One can expect<br />

that it will develop in four directions:<br />

1. In addition to a total synthesis of the parent fullerene, those of more specific<br />

C 60 and higher fullerene derivatives will be vigorously pursued, driven not only<br />

by pure interest but also by the prospects of marketable applications.<br />

2. Syntheses of larger fullerene cages will be indispensable for applications of<br />

endohedral fullerene complexes.<br />

3. With the same purpose in mind, the syntheses of ‘opened’ fullerenes with a<br />

hole and their chemical closing, is discussed in Chapter 5.5. In this respect,<br />

the author is very interested to discover whether Murata group [4] will succeed<br />

in closing C 70 with two hydrogen molecules inside (discussed in Section 5.5),<br />

as this had been calculated to be less stable than the complex involving only<br />

one H 2 molecule inside the cage [5].<br />

4. Today ‘in’ isomers of hydrogenated fullerenes are hypothetical moieties but,<br />

taking into account their unusual structure and predicted interesting properties<br />

(for example, the exciting possibility of reactions between substituents inside<br />

the cage discussed in Chapter 2.4), they could become a hot topic when the<br />

appropriate synthetic methods will be developed.<br />

One can also expect numerous synthetic investigations in the field of carbon<br />

nanotubes although in this area aggregate formation, called synkinesis by Fuhrhop<br />

and Köning [6], will be more actively carried on than the standard synthesis.<br />

Physicochemical studies, interesting per se and forming the basis of applications<br />

of strained hydrocarbons as well as model calculations of these systems<br />

will thrive.<br />

As discussed in detail in Chapter 1.2, the applications are not (and should not<br />

be) the major driving force for studying strained hydrocarbons. However, eventually<br />

many strained hydrocarbons will be used as drug carriers, parts of molecular<br />

(or supramolecular) devices, etc.<br />

And last, but not least, unusual hydrocarbons or reagents from which they are<br />

made will be the subject of rapidly developing supramolecular chemistry. Endohedral<br />

fullerene complexes as well as nested fullerenes and multiwalled carbon<br />

nanotubes on the one hand and gorgeous supramolecular cube (crystallized from<br />

water solution of 11) obtained by the Kuck group in their quest for centropolyindanes<br />

[7] illustrate this point.


References<br />

I hope that this book has shown that the domain of strained hydrocarbons is<br />

an interdisciplinary field in which synthetic, physical and theoretical chemists are<br />

actively involved. Driven by the interest of numerous specialists from diversified<br />

areas the field will undoubtedly flourish.<br />

References<br />

1 Tanner, M. E.; Knobler, C. B.; Cram, D. J.<br />

Angew. Chem., Int. Ed. 1991, 30, 1924.<br />

2 Warmuth, R. Angew. Chem. Int. Ed. 1997,<br />

36, 1347.<br />

3 Jiao, H. J.; Schleyer, P. V.; Beno, B. R.;<br />

Houk, K. N.; Warmuth, R. Angew. Chem.<br />

Int. Ed. 1998, 36, 2761.<br />

4 Koichi, K.; Murata, M. M. S.; Murata, Y.<br />

In: 230th ACS National Meeting 2005<br />

2005, p ORGN-338.<br />

5 Dodziuk, H. Chem. Phys. Lett. 2005, 410,<br />

39.<br />

6 Fuhrhop, J.-H.; Koening, J. Membranes<br />

and Molecular Assemblies. The Synkinetic<br />

Approach; The Royal Society: Cambridge,<br />

1994.<br />

7 Kuck, D. Chem. Rev. 2006, 106, 4885.<br />

461


Index<br />

a<br />

acenaphthene skeleton 78<br />

acenaphthene-5,6-diyl bis(diarylmethylium)<br />

78<br />

acene<br />

– twisted 23<br />

[e,l]acephenanthrylene 184<br />

acetylene 26<br />

actuator 363<br />

adamantene 103<br />

adamantylideneadamantane 108<br />

– derivative 108<br />

alkaplane 48<br />

alkenes<br />

– acyclic 106<br />

– bridgehead 103 f.<br />

– computational data 114<br />

– distorted 103 ff., 112<br />

– nonplanar 103 ff.<br />

– planar 112 ff.<br />

– pyramidalized 112 ff.<br />

alkylidenecycloproparene 182<br />

alkynes 75 f.<br />

all-cis-tribenzo[5.5.5.5]fenestrane 434<br />

allene<br />

– bicyclic 25<br />

– � �-bond<br />

deformation 123<br />

– cyclic 123<br />

– eight-membered ring 134<br />

– four- and fi ve-membered ring 124<br />

– polycyclic 134 f.<br />

– seven-membered ring 131<br />

– strain estimate 123<br />

– strained cyclic 122 ff.<br />

amylose 351<br />

annulene 399 ff.<br />

– higher 411 ff.<br />

[4]annulene 399 ff.<br />

[8]annulene 403 ff.<br />

[10]annulene 411 f.<br />

methano-bridged 412<br />

[12]annulene 411 f.<br />

[14]annulene 411 f.<br />

[16]annulene 411 f.<br />

[18]annulene 411<br />

antiaromaticity 399 ff.<br />

�-antiaromaticity 401<br />

�-antiaromaticity 50<br />

antibacterial activity 313<br />

armchair nanotube 337 f.<br />

aromatic character 164<br />

aromatic ring<br />

– aromatic character 153 f.<br />

– distorted 164<br />

aromatics<br />

– alkylated 148<br />

– strained 147 ff.<br />

aromaticity 401<br />

�-aromaticity 50<br />

asterane 52 f.<br />

– 3-asterane 53<br />

– 4-asterane 53<br />

automerization 402 f.<br />

azacycloheptatetraene 135<br />

aza[60]fullerene 220<br />

b<br />

B3LYP (Becke three-parameter exchange<br />

functional coupled with Lee-Yang-Parr<br />

correction) 290<br />

back-clump 75<br />

barbaralane 414 f.<br />

barbaralone 414<br />

barrelene 410<br />

battery 365<br />

benzdiyne 178<br />

benzene<br />

– non-standard 147 ff.<br />

benzene rings<br />

– nontypical spatial structure 22<br />

<strong>Strained</strong> <strong>Hydrocarbons</strong>: Beyond the van’t Hoff and Le Bel Hypothesis. Edited by Helena Dodziuk<br />

Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

ISBN: 978-3-527-31767-7<br />

463


464<br />

Index<br />

benzocentrohexaquinane 438<br />

benzophane 157<br />

benzo[a]phenanthrene 148<br />

m-benzyne 17<br />

[m][n]betweenanene 21<br />

1,1�-bi(adamantyl) 87<br />

bi[4]asterane 58<br />

bi(bicyclo[1.1.0]butyl) 83<br />

bi(bicyclo[1.1.1]pentyl) 83<br />

– derivative 88<br />

1,1�-bi(bicyclo[1.1.1]pentyl) 87<br />

bi(cubyl) 83<br />

bicycloalkane 91<br />

bicyclo[n.n.n]alkane 91<br />

bicycloalkene 107<br />

– hydrogenation enthalpy 107<br />

– multiply unsaturated 109<br />

bicyclobutane<br />

– inverted geometry 33<br />

bicyclo[2.2.1]hept-2-en-5-yne 382<br />

bicyclo[3.3.0]oct-1(5)-ene 113<br />

bicyclo[4.4.1]undecapentaene 109<br />

bicyclo[4.4.2]dodecapentaene 109<br />

bicyclo[5.1.0]octa-2,5-diene 413<br />

bicyclo[5.3.1]undecapentaene 109<br />

bicyclo[5.3.1]undeca-1,3,5(11)-triene 110<br />

bifurcation point 410<br />

1,1�-bi(homocubyl) 88<br />

Bingel reaction 214 ff.<br />

bio-related material 365<br />

bipentaprismane 58<br />

bi[n]prismane 57<br />

bis(acetoxymethyl)tricyclo[2.1.0.0 2,5 ]pentan-<br />

3-one 84<br />

bisaddition<br />

– fullerene 213<br />

bisallene<br />

– cyclic 136<br />

bishomoaromatic 419<br />

bis(triphenylmethanol) 75<br />

bi(tetrahedranyl) 83<br />

1,1�-bi(tricyclo[3.1.0.0 2,6 ]hexyl) 88 f.<br />

1,1�-bi(tricyclo[4.1.0.0 2,7 ]heptyl) 88<br />

bond shifting 405<br />

bond<br />

– nonplanar C=C 112, 144<br />

– nonplanar C–C 44, 68 f.<br />

– shifting 405<br />

– ultralong C=C 70<br />

– ultrashort C=C 82<br />

bowlane 17, 46, 455<br />

– dimer 48<br />

bulk-heterojunction 305<br />

bullvalene 414 f.<br />

butatriene<br />

– � �-bond<br />

deformation 137<br />

– cyclic 137 ff.<br />

– fi ve- to nine-membered ring 137<br />

– strain estimate 137<br />

9-t-butylanthracene 148<br />

(2-t-butylcubyl)cubane 88<br />

c<br />

C60 , see also fullerene<br />

– derivative for optical limiting 299<br />

– derivative for photovoltaic application<br />

304<br />

– fi ve-membered ring-fused 222<br />

– four-membered ring-fused 221<br />

– six-membered ring-fused 223<br />

– terminated oligo( p-phenylene ethynylene)<br />

(OPE) hybrid compound 302<br />

– three-membered ring-fused 218<br />

C60-OPE hybrid 302<br />

C 60 (X) n isomer<br />

– steric strain 236 ff.<br />

C70 295<br />

C–C bond<br />

– elongated 74<br />

– expandability of ultralong bond 79<br />

– stiff molecular frame 72<br />

– super-ultralong 78<br />

– ultralong 70 ff.<br />

– ultrashort 82 ff.<br />

– ultrashort endocyclic bridging 83 ff.<br />

– ultrashort exocyclic intercage 86<br />

C(sp 2 )-C(sp 2 ) bond<br />

– fullerene cage 232<br />

C=C bond<br />

– nonplanar 112 ff.<br />

– stretch 116<br />

C–C(Me) bond<br />

– ultrashort 91<br />

C2v-[15]triangulene 19<br />

C10H10 saturated cage 63<br />

C12H12 saturated cage 63<br />

C60H6 293<br />

C60H18 294<br />

C60H36 294 f.<br />

C60H60 – all-out isomer 66<br />

– atom numbering 66<br />

– in isomers 66<br />

– ten-in isomer 66<br />

C 60 (OH) n 313<br />

C 60 -tetrathiafulvalene (TTF) dyad 302<br />

C70H2 298<br />

2–<br />

C70 298


C70H8 296<br />

(CH) 2n cage structure 59 ff.<br />

cage structure<br />

– nonplanar C=C bond 112 ff.<br />

carbene 456<br />

carbocation 456<br />

carbohelicene 166<br />

carbon atom<br />

– inverted 43, 58<br />

– planar 44 ff.<br />

– pyramidal 44 ff.<br />

carbon nanotube (CNT) 335 ff.<br />

– applications 356 ff.<br />

– armchair 337<br />

– aromaticity 345<br />

– bamboo-like structure 344<br />

– cap structure 343<br />

– chiral 338<br />

– cycloaddition 353<br />

– electrical conductivity 339<br />

– fl uorination 353<br />

– functionalization 347 ff.<br />

– HRTEM 341<br />

– hydrogenation 352<br />

– multi-walled (MWNT) 342 ff.<br />

– properties 357<br />

– resonance structure 346<br />

– safety 360<br />

– semiconductor-type 365<br />

– single-walled (SWNT) 335 ff.<br />

– structure 342 f.<br />

– symmetry 338<br />

– toxicity 360<br />

– zigzag 337<br />

carboxyfullerene 312<br />

carcerand cage 452 ff.<br />

catalysis<br />

– electrocatalytic activity of fullerenes<br />

270<br />

catenane 7<br />

centrohexaindane 430 ff.<br />

– multiply functionalized 436<br />

centrohexaquadrane 426<br />

centrohexaquinane 426<br />

centrohexaquinacene 427<br />

centropentaindane 432 ff.<br />

centropolycyclane 426<br />

centropolyindane 431<br />

trifuso-centrotetraindane 431 f.<br />

centrotriindane 432<br />

chemical sensor 365<br />

chiral nanotube 338<br />

[6]chochin 440<br />

[n]circulene 149<br />

[7]circulene 149<br />

Clar resonance structure 345 f.<br />

clathrate 451<br />

clump 75<br />

chochin 24<br />

computational chemistry 15<br />

conductive composite 364 ff.<br />

corannulene 25, 149<br />

coronene 149<br />

coupled cage compound 86<br />

– ultrashort exocyclic intercage C–C<br />

bond 86<br />

cubane 8 f., 50, 59 ff.<br />

– highly strained derivative 13<br />

– oligomer 56<br />

cubene 55, 62<br />

cubylcubane 56<br />

cumulene 25, 122<br />

– cyclic 136<br />

cuneane 59 ff.<br />

cyclic system<br />

– conjugated 401<br />

cyclization<br />

– Hopf 127<br />

– Myers-Saito type 131<br />

cycloaddition 353 f.<br />

– carbon nanotube 353 f.<br />

cycloalkene 107<br />

cycloalkyne<br />

– angle-strained 375 ff.<br />

– distorted 375<br />

– spectroscopic property 392<br />

cyclobutadiene 399 ff., 450<br />

– bond length 401<br />

1,2-cyclobutadiene 124<br />

cyclobutyne 378<br />

cyclodecyne 386<br />

cyclodextrin 449<br />

1,2-cycloheptadiene 131<br />

trans-cycloheptane 453<br />

cycloheptatetraene 132<br />

– 1,2,4,6-cycloheptatetraene 132<br />

– 1,3,5,6-cycloheptatetraene 453<br />

– incarterated 134<br />

1,2,4,5- cycloheptatriene 133<br />

1,3,5-cycloheptatriene 413<br />

cycloheptatrienylidene 188<br />

cycloheptyne 384<br />

1,2-cyclohexadiene 125<br />

cyclohexamantane 439 f.<br />

cyclohexane<br />

– planar ring 67 ff.<br />

– trans 454<br />

–<br />

trispirocyclopropanated 19<br />

Index<br />

465


466<br />

Index<br />

1,2,3-cyclohexatriene 138<br />

1,2,4-cyclohexatriene 127 f.<br />

E-cyclohexene 103<br />

cyclohexyne 382<br />

cyclononyne 386<br />

1,2-cyclooctadiene 135<br />

cyclooctatetraene (COT) 403 ff.<br />

– annelated 406<br />

– bond shifting 407<br />

– conformation 404<br />

– ion 410 f.<br />

– ring inversion 407<br />

[2.2]cyclooctatetraenophane 22<br />

E-cyclooctene 103 ff.<br />

Z-cyclooctene 107<br />

cyclooctyne 385 f.<br />

1,2-cyclopentadiene 124 f.<br />

cyclopentyne 379<br />

cyclophane 12, 23, 67, 109, 150 ff.<br />

– distorted 153<br />

– K3,3<br />

438<br />

– Kuratowski 430 ff., 441 ff.<br />

– naphthalene-based 439<br />

– NMR characteristics 165<br />

[n]cyclophane 154<br />

[m,n]cyclophane 151<br />

[2,2]cyclophane 151<br />

[2.2.2](1,3,5)cyclophane 165<br />

[26 ](1,2,3,4,5,6)cyclophane-1-ene 24<br />

cyclopropabenzene 192<br />

cyclopropabenzenyl anion 191<br />

cyclopropannulene 178<br />

cycloproparene 176 ff.<br />

– charge transfer (CT) complexation 189<br />

– heteroatom 187<br />

cyclopropyne 376<br />

cyclopropynylidene 376<br />

d<br />

decaprismane C20H20 50<br />

Td-1,2-dehydro-5,7-adamantanediyl<br />

dication 16<br />

1,2-dehydrocubane 55<br />

m-dehydrocubane 17<br />

dehydroprismane 55<br />

1,5-dehydroquadricyclane 17<br />

dehydrotriprismane 55<br />

density functional theory (DFT) method 15<br />

density of state (DOS) 243 f.<br />

density of unoccupied state (DUOS) 243 f.<br />

diademane 60 ff.<br />

di-p-benzhexaphyrin 23<br />

dibenzylation<br />

C<br />

2– 298<br />

– 70<br />

di-t-butylbenzene 148<br />

1,2-di-t-butylethylene 108<br />

(Z)-1,2-di-t-butylethylene<br />

(2,2,5,5-tetramethylhex-3-ene) 106<br />

dichloro[3.2.1]propellane 34<br />

9,9�-didehydroanthracene (DDA) 113<br />

Diels-Alder reaction<br />

– fullerene 213<br />

diiodonaphthocyclobutene 71<br />

5,6-dilithioacenaphthene 78<br />

dimethanospiro[2.2]octaplane 48<br />

N,N�-dimethylaniline 268<br />

1,1-dimethyl-2,3-bis-(t-butyl)cyclobutane<br />

106<br />

1,2-dimethyl-2,3-bis-(t-butyl)<br />

methylcyclopropane 106<br />

2,3-dimethyl-2-butene 108<br />

1,1-dimethyl-2-neopentylcyclopropane<br />

106<br />

dimethylspiro[2.2]octaplane 455<br />

E-1-(2,2-dimethyl-1-tetralinylidene)-<br />

2,2-dimethyltetralin 109<br />

1,5-dimethyltricyclo[2.1.0.0 2,5 ]pentan-<br />

3-one 84<br />

2,4-dimethyltricyclo[2.1.0.0 2,5 ]pentane 85<br />

3,6-dimethoxycyclopropa[b]naphthalene<br />

181<br />

9,9�-diphenyl-9,9�-bifl uorenyl 75<br />

trans-diphenylspirotriketone 433<br />

DNA photocleavage 310<br />

dodecadiyne 388<br />

dodecahedrane 12, 60<br />

– C20H20<br />

64<br />

dodecahedron 20<br />

Doering-Moore-Skattebøl rearrangement<br />

125 ff.<br />

double bond<br />

– distorted 21<br />

drop casting method 274<br />

e<br />

electrochemistry 259 f.<br />

electron emitter 366<br />

electron energy loss spectroscopy<br />

(EELS) 243 ff.<br />

electron-mediating property 272<br />

electronic part 363<br />

enyne<br />

– Diels-Alder reaction 126<br />

– photorearrangement 124<br />

ethylene<br />

– most distorted 111<br />

– t-butyl-substituted<br />

104 ff.<br />

Z-ethylene 104


f<br />

syn-fenchilydenefenchane 104<br />

fenestrane 17, 44 ff.<br />

[4.4.4.4]fenestrane 44<br />

all-cis-[5.5.5.5]fenestrane 431<br />

all-cis-[5.5.5.6]fenestrane 433<br />

fenestranone 434<br />

fenestrindane 431<br />

ferrocene 268<br />

ferrocene (Fc/Fc + ) system 262<br />

ferrocene-porphyrin-fullerene<br />

construct 269<br />

fi lm preparation method 274<br />

fullerene 6 ff., 205 ff.<br />

– absorption spectrum 280<br />

– � �-alanine<br />

derivative 313<br />

– alkylation 217<br />

– application 299 ff.<br />

– binding energy 290<br />

– biological application 310 ff.<br />

– bond 209, 232<br />

– C20<br />

11, 206<br />

– C60<br />

, see also C60 6 f., 205 ff.<br />

– C80<br />

207<br />

– cage 232<br />

– cyclic voltammetry 269 f.<br />

– cycloaddition 218<br />

– cyclobutane-annulated 221<br />

– � �-cyclodextrine<br />

subunit 311<br />

– density of state (DOS) 243 f.<br />

– density of unoccupied state (DUOS)<br />

243 f.<br />

– Diels-Alder adduct 223<br />

– DNA photocleavage 310<br />

– electrocatalytic activity 270<br />

– electron affi nity 260<br />

– electronic property 259<br />

– electronic structure 246<br />

– endohedral 282 ff.<br />

– fi lm 274 ff.<br />

– fi lm property 277<br />

– fullerene interaction 278<br />

– functionalization 263<br />

– guest substrate 288<br />

– higher 297<br />

– host 287<br />

– hydrogenated 291<br />

– hydrogenation 216<br />

– intra-/intermolecular association 263<br />

– in-out isomerism 64<br />

– ionization potential 260<br />

– isomer 212, 292 ff.<br />

– nanostructured fi lm 276<br />

– non-IPR stability 232<br />

Index<br />

– nonplanar steric strain 226 ff.<br />

– nonplanar steric strain parameter 229<br />

– nuclear magnetic resonance (NMR)<br />

250 ff.<br />

– open cage 284 f.<br />

– orbital picture 243 ff.<br />

– oxidation 215<br />

– perhydrogenated 60<br />

– physicochemical property 225<br />

– POAV ( �-orbital axis vector) analysis<br />

226 ff.<br />

– polyhydroxylated 313<br />

– porphyrin hybrid 312<br />

– property 255<br />

– radical addition 211<br />

– reaction 210<br />

– reactivity of hydrogenated fullerene<br />

297<br />

– reduction 215<br />

– reduction potential 261<br />

– reversible molecular incorporation<br />

and ejection 285<br />

– Schlegel diagram 226 f.<br />

– siloxane group 303<br />

– single crystal X-ray structure 225<br />

– sol-gel glasses 300 ff.<br />

– structure 291<br />

– superconductivity 281<br />

– synthesis 291<br />

– tris-malonic acid 312<br />

– UV/Vis absorption spectrum 247 f.<br />

– vibrational spectra 239 ff.<br />

– water interaction 278<br />

[60]fullerene 208<br />

– photovoltaic application 304<br />

fullerene aggregate 273<br />

fullerene-ferrocene dyad 268<br />

fullerene-OPE hybrid 303<br />

fullerene-porphyrin hybrid 312<br />

fullerene-TTF dyad 303<br />

fullerenol (C60 (OH) n ) 313<br />

fullerodendrimer 301<br />

fulleroid 220<br />

fulleropyrrolidinium salt 313<br />

fulvalene 111<br />

functionalization 347 ff.<br />

– covalent side-wall 352<br />

– endohedral 355<br />

– non-covalent 349 f.<br />

g<br />

graph<br />

– nonplanar 428<br />

graphene 337<br />

467


468<br />

Index<br />

h<br />

H2@C60 284 ff.<br />

He@C60 282 ff.<br />

helicene 166<br />

– asymmetric synthesis 171<br />

– nonracemic 171<br />

– physicochemical property 173<br />

– structure 172<br />

[4]helicene 148<br />

[5]helicene 168 ff.<br />

[6]helicene 166 ff.<br />

[7]helicene 168 f.<br />

[14]helicene 172<br />

helvetane 64<br />

hemicarcerand 450 ff.<br />

heptacyclo[6.4.0.0 2,4 .0 3,7 .0 5,12 .0 6,10 .0 9,11 ]<br />

dodecane 63<br />

heterophane 157<br />

heterohelicene 166<br />

heterojunction 305 ff.<br />

hexaanilino[60]fullerene 217<br />

hexa-t-butylbenzene 22<br />

hexaene 25<br />

hexagon-hexagon-pentagon junction<br />

(HHP) 225<br />

hexahelicene 166<br />

hexahydrosuperphane 67<br />

hexakis(trimethylsilyl) derivative 22<br />

hexanitro[60]fullerene 217<br />

hexaphenylethane (HPE) 73<br />

– acenaphthalene-type 80<br />

– bis(10-methylspiroacridine)-type 81<br />

– clumped derivative 74<br />

– cross-clumped 75<br />

– derivative with super-ultralong C–C<br />

bond 78<br />

– dihydropyracylene-type 80<br />

– dispiro 75<br />

hexaprismane 67<br />

homoaromaticity 109, 411<br />

homotropilidene 413<br />

bridged 413 ff.<br />

homotropylium cation 411<br />

Hopf cyclization chemistry 128<br />

Hückel rule 408<br />

hydrocarbon<br />

– bicyclic 20<br />

– distorted saturated 33 ff.<br />

– highly strained natural compound 13<br />

– K3,3<br />

438<br />

– K5<br />

438<br />

– saturated 18, 33 ff.<br />

– short-lived 449 ff.<br />

– strained, see strained hydrocarbon<br />

– unusual spatial structure 5<br />

hydrogenation enthalpy<br />

bicycloalkene 107<br />

i<br />

in-out isomerism 59<br />

– perhydrogenated fullerene C60H60<br />

64<br />

indium tin oxide (ITO) 304<br />

inversed photoemission spectroscopy<br />

(IPES) 243 f.<br />

inverted geometry 33<br />

isobenzene 127<br />

isodesmic equation 15<br />

isolated-pentagon rule (IPR) 228<br />

isomorph<br />

– conformational 79<br />

israelane 64<br />

k<br />

K3,3 molecule 438<br />

K5 molecule 438<br />

Knight shift<br />

– isotropic 252<br />

Korringa relation 258<br />

Kuratowski<br />

– cyclophane 430 ff., 441 ff.<br />

– graph 428<br />

l<br />

ladderane 12<br />

Langmuir-Blodgett (LB) method 274 ff.<br />

Langmuir-Schäffer (LS) method 274 ff.<br />

low-friction surface 364<br />

m<br />

magic angle spinning (MAS) 256<br />

membrane 364<br />

[n]metacyclophane 154 ff.<br />

[1]metacyclophane 150<br />

[2.2]metacyclophane 24<br />

[2,2]metaparacyclophane 152<br />

[5]metacyclophane 110, 150 f.<br />

[6]metacyclophane 110, 151<br />

metallacyclopenta-2,3,4-triene 139<br />

metallacyclopropabenzene 187<br />

methane<br />

– planar 45<br />

– tetrahedral 45<br />

1,5-methano[10]annulene 109<br />

1,6-methano[10]annulene 109, 412<br />

methanofullerene 220<br />

methano[70]fullerene 265<br />

1-(3-methoxycarbonyl)propyl-1-1-phenyl-[6,6]<br />

methanofullerene ([60]PCBM) 306


in-methylcyclophane 91<br />

6-methylene-1,2,4-cyclohexatriene 131<br />

N-methylphenothiazine 268<br />

Mills-Nixon effect 190<br />

Möbius strip 23<br />

molecular fl ask 449 ff.<br />

molecule<br />

– nonplanar 425 ff.<br />

monobenzyl C70H2 298<br />

multiporphyrin dendrimer 271<br />

n<br />

N@C 60 289<br />

nanokid 11<br />

nanoputane 11<br />

nanotechnology 6<br />

nanotube, see also carbon nanotube<br />

335 ff.<br />

nanotube cap 340<br />

naphthalene 148<br />

naphthalene-1,8-diyl bis(diarylmethylium)<br />

75<br />

naphthalenophane 439<br />

[6.6]naphthalenophane 390<br />

naphthocyclobutene derivative 71<br />

Ne@C 60 287<br />

near-fi eld microscopy probe 366<br />

nitrocubane 52<br />

NMR (Nuclear Magnetic Resonanz)<br />

38 f., 109, 151, 250 ff.<br />

norcaradiene 412<br />

o<br />

octabisvalene 59 ff.<br />

octacyclopropylcubane 12 ff.<br />

octahedrane 12<br />

– highly strained derivative 13<br />

octaplane 47<br />

octiphenyl 442<br />

– octabromo derivative 442<br />

olefi n strain energy (OSE) 55<br />

olympiadane 7<br />

OPE-C60 hybrid (oligophenyleneethynylene-C60<br />

) 309<br />

optical limiting (OL) 299<br />

OPV-C60 hybrid (oligophenylenevinylene-C60<br />

) 308 f.<br />

organofullerene 267<br />

– electron donor moieties 267<br />

orthogonene 17<br />

oxa[3.2.1]propellane 34<br />

p<br />

�–� interaction<br />

– carbon nanotube 345 ff.<br />

– concave–convex 25<br />

P3HT (poly(3-hexylthiophene)) 307<br />

paddlane 44 ff.<br />

[1.1.1.1]paddlane 42<br />

padogane 60<br />

– heterobridged 439<br />

para-cyclophane 24<br />

[n]paracyclophane 154 ff.<br />

[m.n]paracyclophane 161<br />

[0.0]paracyclophane 161<br />

[1.n]paracyclophane 161<br />

[1.1]paracyclophane 161 f.<br />

[2.2]paracyclophane 110, 161 f.<br />

[2.2]paracyclophane/dehydroannulene<br />

hybrid 389<br />

[2.2:2.2:2.2:2.2:2.2]paracyclophane 440<br />

[4]paracyclophane 159<br />

[4.4]paracyclophane 163<br />

[6]paracyclophane 158<br />

[10]paracyclophane 155<br />

[60]PCBM 306<br />

pentaene 25<br />

pentamantane 439<br />

pentaprismane 50 ff., 60 ff.<br />

percubylcubane 56<br />

perfl uorotetracyclobutenocyclooctatetraene<br />

408<br />

[5]pericyclyne<br />

– permethylated 26<br />

[6]pericyclyne<br />

– permethylated 26<br />

(Ph3P) 2Pt complex 119 f.<br />

phenylcarbene<br />

hemicarceplexed 134<br />

photocyclodehydrogenation 169<br />

photodehydrocyclisation 168<br />

photodynamic therapy (PDT) 310<br />

photorearrangement<br />

– enyne 124<br />

photovoltaic cell 309<br />

photovoltaic conversion 308<br />

phthalocyanine 304, 351<br />

POAV (�-orbital axis vector) analysis<br />

226 ff., 408<br />

polyacetylene<br />

– cyclic 387 f.<br />

polyaniline (PANI) 351<br />

polycubane 56<br />

polycyclane 426<br />

Index<br />

469


470<br />

Index<br />

polyether<br />

– topologically nonplanar 438<br />

polyhydroxyfullerene 212<br />

polymantane 439<br />

polyprismane 56<br />

poly[n]prismane 57<br />

porphyrin 351<br />

porphyrin-fullerene-dyad (P-C60 ) 312<br />

– metalated 312<br />

Prato reaction 213 ff.<br />

prismane 16, 49<br />

– C2nH2n<br />

prismane 49<br />

– expanded 52 ff.<br />

– ethynyl-expanded 54<br />

– fused 57<br />

[n]-prismane 64<br />

[6]prismane 64<br />

[7]prismane 64<br />

pristine fullerene 278<br />

propellane<br />

– small-ring 35 f.<br />

[k.l.m]propellane 33 f.<br />

– inverted geometry 33<br />

[k.1.1]propellane 43<br />

[1.1.1]propellane 9 ff., 34 ff.<br />

– precursor 42<br />

– preparation and reactivity 38 ff.<br />

[2.1.1]propellane<br />

– preparation and reactivity 40 f.<br />

[2.2.1]propellane<br />

– preparation and reactivity 40 f.<br />

[2.2.2]propellane 35<br />

[4.1.1]propellane 35<br />

propella[34 ]prismane 62<br />

pyramidalization 114 ff.<br />

pyramidane 17<br />

– derivative 16<br />

pyrene derivative 350<br />

r<br />

ratchet phase 256<br />

reverse saturable absorption (RSA) 300<br />

ring<br />

– inversion 405<br />

– nonplanar C=C bond 112 ff.<br />

rotator phase 256<br />

rotaxane system 6<br />

s<br />

scavenger activity 313<br />

semibullvalene 404 ff., 414 ff.<br />

– octasubstituted 418<br />

short-lived species<br />

– stabilization in molecular fl ask 449 ff.<br />

silacyclopropyne 377, 456<br />

sol-gel glasses 300 f.<br />

solar cell 305 ff.<br />

– all-polymer 307<br />

spin coating method 274<br />

spiropentane 18<br />

spiro[3.3]pentane 48<br />

[n]staffane 42<br />

stilbene<br />

– cyclic 108 f.<br />

– t-butyl-substituted<br />

108 f.<br />

strain energy 14, 104 f.<br />

strain estimate 123, 137<br />

strained hydrocarbon 12<br />

– computation 12<br />

structural composite 363<br />

supercapacity 365<br />

superconductivity 8<br />

– fullerene 281<br />

supercubane 56<br />

supramolecular fl ask 449 ff.<br />

surfactant 275<br />

t<br />

1,1,2,2-tetraarylacenaphthene derivative<br />

75<br />

tetraasterane 19<br />

tetracyclo[3.1.0.0 1,3 .0 3,5 ]hexane 17<br />

tetra-tertiary-butyltetrahedrane 403<br />

tetrahedral coordination 425 f.<br />

tetrahedrane 12, 61 ff., 403, 455<br />

– bis-tetrahedrane<br />

17<br />

tetrakisbicyclohexenocyclooctatetraene<br />

408<br />

tetrakis-t-butylethene 21<br />

tetralin 150<br />

anti-tetramantane 439<br />

tetramethyl cyclooctatetraene 407<br />

tetraphenylnaphthocyclobutene 73<br />

7,7,8,8-tetraphenyl-o-quinodimethane 73<br />

tetraphosphabarbaralane 415<br />

tetrathiafulvalene (TTF) 268<br />

3,3,7,7-tetramethylcycloheptyne 384<br />

4-thia-3,3,5,5-tetramethylcyclopentyne<br />

379<br />

topography 425 ff.<br />

topology 425 ff.<br />

triangulane 19<br />

all-cis-tribenzo[5.5.5.5]fenestrane 434<br />

tribenzo[5.5.5.5]fenestrene 434<br />

tri(bicyclo[1.1.1]pentyl) derivative 88<br />

tri-t-butylethylene 106<br />

1,2,3-tri-t-butylnaphthalene 22<br />

tricyclo[2.1.0.0 1,3 ]pentane 17


tricyclo[2.1.0.0 2,5 ]pentan-3-one 83 ff.<br />

tricyclo[2.1.0.0 2,5 ]pentane 83 f.<br />

– ultrashort endocyclic bridging C–C<br />

bond 83 ff.<br />

tricyclo[3.1.0 1,3 .0 3,5 ]hexane 18<br />

tricyclo[3.3.n.0 3,7 ]alk-3(7)-ene 113 ff.<br />

tricyclo[3.3.0.0 3,7 ]oct-1(7)-ene 119<br />

tricyclo[3.3.9.0 3,7 ]non-3(7)-ene 117<br />

tricyclo[3.3.10.0 3,7 ]dec-3(7)-ene 117<br />

tricyclo[3.3.11.0 3,7 ]undec-3(7)-ene 115<br />

[4.1.0.0 1,6 ]tricycloheptane 18<br />

tricyclo[4.2.2.2 2,5 ]dodeca-1,5-diene 21<br />

triene 25<br />

trifuso-centrotetraindane 431 f.<br />

trioxacentrohexaquinane 438<br />

triprismane 50, 62<br />

tri[n]prismane 55<br />

triprismene 55<br />

tris-�-homobenzene 69<br />

– cis and trans 69<br />

tropolidene 413<br />

u<br />

ultraviolet photoemission spectroscopy<br />

(UPS) 243 ff.<br />

umbrella confi guration 58<br />

UV/Vis spectra 432<br />

UV/Vis spectra of fullerenes 246 ff.<br />

v<br />

valence isomerization 404 ff.<br />

w<br />

windowpane, see fenestrane<br />

Index<br />

x<br />

X-ray of fullerenes 225 ff.<br />

X-ray absorption spectroscopy (XAS) 243<br />

xylene 148<br />

z<br />

zeolite 452<br />

zero-point vibrational energy (ZPVE) 15<br />

zigzag nanotube 337<br />

zirconacyclohepta-2,4,5,6-tetraene 139<br />

471

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!