27.12.2013 Views

Microscopic Interactions and Macroscopic Properties Final Report (1st

Microscopic Interactions and Macroscopic Properties Final Report (1st

Microscopic Interactions and Macroscopic Properties Final Report (1st

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Deutsche<br />

Forschungsgemeinschaft<br />

Macromolecular<br />

Systems:<br />

<strong>Microscopic</strong><br />

<strong>Interactions</strong> <strong>and</strong><br />

<strong>Macroscopic</strong> <strong>Properties</strong><br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


Deutsche<br />

Forschungsgemeinschaft<br />

Macromolecular Systems:<br />

<strong>Microscopic</strong> <strong>Interactions</strong><br />

<strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

<strong>Final</strong> report of the<br />

collaborative research centre 213,<br />

``Topospezifische Chemie und Toposelektive<br />

Spektroskopie von MakromolekÏlsystemen:<br />

Mikroskopische Wechselwirkung und<br />

Makroskopische Funktion'', 1984^1995<br />

Edited by<br />

HeinzHoffmann,MarkusSchwoerer<strong>and</strong><br />

Thomas Vogtmann<br />

Collaborative Research Centres


Deutsche Forschungsgemeinschaft<br />

Kennedyallee 40, D-53175 Bonn, Federal Republic of Germany<br />

Postal address: D-53175 Bonn<br />

Phone: ++49/228/885-1<br />

Telefax: ++49/228/885-2777<br />

E-Mail: (X.400): S = postmaster; P = dfg; A = d400; C = de<br />

E-Mail: (Internet RFC 822): postmaster@dfg.de<br />

Internet: http://www.dfg.de<br />

This book was carefully produced. Nevertheless, editors, authors <strong>and</strong> publisher do not warrant the information<br />

contained therein to be free of errors. Readers are advised to keep in mind that statements, data, illustrations,<br />

procedural details or other items may inadvertently be inaccurate.<br />

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

A catalogue record for this book is available from the British Library.<br />

Die Deutsche Bibliothek – CIP Cataloguing-in-Publication Data<br />

A catalogue record for this publication is available from Die Deutsche Bibliothek<br />

ISBN 3-527-27726-9<br />

© WILEY-VCH Verlag GmbH, D-69469 Weinheim (Federal Republic of Germany), 2000<br />

Printed on acid-free <strong>and</strong> chlorine-free paper<br />

All rights reserved (including those of translation into other languages). No part of this book may be reproduced<br />

in any form – by photoprinting, microfilm, or any other means – nor transmitted or translated into machine<br />

language without written permission from the publishers. Registered names, trademarks, etc. used in this book,<br />

even when not specifically marked as such, are not to be considered unprotected by law.<br />

Cover Design <strong>and</strong> Typography: Dieter Hüsken<br />

Composition: ProSatz Unger, D-69469 Weinheim<br />

Printing: betz-druck gmbh, D-64291 Darmstadt<br />

Bookbindung: J. Schäffer GmbH & Co. KG, D-67269 Grünstadt<br />

Printed in the Federal Republic of Germany


Contents<br />

Preface ...................................................<br />

List of Contributors ..........................................<br />

XV<br />

XVII<br />

I<br />

Mainly Solids<br />

1 Model Systems for Photoconductive Material ..................... 3<br />

Harald Meyer <strong>and</strong> Dietrich Haarer<br />

1.1 Introduction . . ............................................. 3<br />

1.2 Experimental techniques . . . ................................... 4<br />

1.3 Transport models ............................................ 5<br />

1.4 Results ................................................... 7<br />

1.4.1 Moleculary doped polymers . ................................... 7<br />

1.4.2 Side-chain polymers ......................................... 9<br />

1.4.3 Conjugated systems .......................................... 10<br />

1.5 Outlook .................................................. 12<br />

References . . . ............................................. 13<br />

2 Novel Photoconductive Polymers ............................... 15<br />

Jörg Bettenhausen <strong>and</strong> Peter Strohriegl<br />

2.1 Introduction . . ............................................. 15<br />

2.2 Liquid crystalline oxadiazoles <strong>and</strong> thiadiazoles ...................... 16<br />

2.2.1 The basic idea . ............................................. 16<br />

2.2.2 Monomer synthesis .......................................... 17<br />

2.2.3 Oligo- <strong>and</strong> polysiloxanes with pendant oxadiazole groups .............. 19<br />

2.2.4 Photoconductivity measurements ................................ 21<br />

2.3 Starburst oxadiazole compounds ................................. 22<br />

2.3.1 Motivation . . . ............................................. 22<br />

2.3.2 Synthesis of starburst oxadiazole compounds ....................... 24<br />

2.3.3 Thermal properties .......................................... 27<br />

References . . . ............................................. 30<br />

V


Contents<br />

3 Theoretical Aspects of Anomalous Diffusion in Complex Systems ...... 31<br />

Alex<strong>and</strong>er Blumen<br />

3.1 General aspects ............................................. 31<br />

3.2 Photoconductivity ........................................... 32<br />

3.3 The Matheron-de-Marsily model ................................ 36<br />

3.4 Polymer chains in MdM flow fields .............................. 38<br />

3.5 Conclusions . . ............................................. 42<br />

Acknowledgements .......................................... 42<br />

References . . . ............................................. 43<br />

4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation,<br />

Specific Heat <strong>and</strong> Thermal Conductivity in Polymers ............... 44<br />

Andreas Nittke, Michael Scherl, Pablo Esquinazi, Wolfgang Lorenz,<br />

Junyun Li, <strong>and</strong> Frank Pobell<br />

4.1 Introduction . . ............................................. 44<br />

4.2 Phenomenological theory for heat release . . ........................ 46<br />

4.2.1 Generalities . . . ............................................. 46<br />

4.2.2 The st<strong>and</strong>ard tunneling model with infinite cooling rate . .............. 47<br />

4.2.3 Influence of higher-order tunneling processes <strong>and</strong> a finite cooling rate . . . . 49<br />

4.2.4 The influence of a constant <strong>and</strong> thermally activated relaxation rate ....... 52<br />

4.3 Experimental details ......................................... 54<br />

4.4 Experimental results <strong>and</strong> discussion .............................. 55<br />

4.4.1 Specific heat <strong>and</strong> thermal conductivity ............................ 55<br />

4.4.2 Internal friction <strong>and</strong> sound velocity .............................. 58<br />

4.4.3 Heat release . . ............................................. 62<br />

4.5 Conclusions . . ............................................. 65<br />

Acknowledgements .......................................... 66<br />

References . . . ............................................. 66<br />

5 Spectral Diffusion due to Tunneling Processes at very low Temperatures 68<br />

Hans Maier, Karl-Peter Müller, Siegbert Jahn, <strong>and</strong> Dietrich Haarer<br />

5.1 Introduction . . ............................................. 68<br />

5.2 The optical cryostat .......................................... 69<br />

5.3 Theoretical considerations . . ................................... 71<br />

5.4 Temperature dependence . . . ................................... 73<br />

5.5 Time dependence ............................................ 74<br />

References . . . ............................................. 76<br />

6 Optically Induced Spectral Diffusion in Polymers Containing Water<br />

Molecules: A TLS Model System ............................... 78<br />

Klaus Barth, Dietrich Haarer, <strong>and</strong> Wolfgang Richter<br />

6.1 Introduction . . ............................................. 78<br />

6.2 Experimental setup for burning <strong>and</strong> detecting spectral holes ............ 79<br />

6.3 Reversible line broadening phenomena ............................ 80<br />

6.4 Induced spectral diffusion . . ................................... 83<br />

References . . . ............................................. 87<br />

VI


Contents<br />

7 Slave-Boson Approach to Strongly Correlated Electron Systems ....... 88<br />

Holger Fehske, Martin Deeg, <strong>and</strong> Helmut Büttner<br />

7.1 Introduction . . ............................................. 88<br />

7.2 Slave-boson theory for the t-t'-J model ............................ 90<br />

7.2.1 SU(2)-invariant slave-particle representation ........................ 90<br />

7.2.2 Functional integral formulation ................................. 93<br />

7.2.3 Saddle-point approximation . ................................... 95<br />

7.2.4 Magnetic phase diagram of the t-t'-J model ........................ 97<br />

7.3 Comparison with experiments .................................. 102<br />

7.3.1 Normal-state transport properties ................................ 102<br />

7.3.2 Magnetic correlations <strong>and</strong> spin dynamics . . ........................ 105<br />

7.3.3 Inelastic neutron scattering measurements . . ........................ 106<br />

7.4 Summary ................................................. 109<br />

References . . . ............................................. 110<br />

8 Non-Linear Excitations <strong>and</strong> the Electronic Structure of Conjugated<br />

Polymers ................................................. 113<br />

Klaus Fesser<br />

8.1 Introduction . . ............................................. 113<br />

8.2 Models ................................................... 114<br />

8.3 Disorder .................................................. 116<br />

8.4 Non-linear excitations ........................................ 117<br />

8.5 Perspective . . . ............................................. 119<br />

Acknowledgements .......................................... 120<br />

References . . . ............................................. 120<br />

9 Diacetylene Single Crystals ................................... 122<br />

Markus Schwoerer, Elmar Dormann, Thomas Vogtmann,<br />

<strong>and</strong> Andreas Feldner<br />

9.1 Introduction . . ............................................. 122<br />

9.2 Photopolymerization ......................................... 129<br />

9.2.1 Carbenes .................................................. 129<br />

9.2.2 Intermediate photoproducts . ................................... 131<br />

9.2.3 Electronic structure of dicarbenes ................................ 131<br />

9.2.3.1 Electron spin resonance of quintet states ( 5 DC n ) .................... 131<br />

9.2.3.2 ENDOR of quintet states . . . ................................... 136<br />

9.2.3.3 ESR <strong>and</strong> ENDOR of triplet dicarbenes 3 DC n ....................... 139<br />

9.2.4 Flash photolysis <strong>and</strong> reaction dynamics of diradicals .................. 141<br />

9.3 Holography . . . ............................................. 144<br />

9.3.1 Theory ................................................... 145<br />

9.3.2 Experimental setup .......................................... 147<br />

9.3.3 General characterization . . . ................................... 149<br />

9.3.4 Angular selectivity .......................................... 150<br />

9.3.5 Prepolymerized samples ....................................... 152<br />

9.3.6 Chain length, polymer profile, <strong>and</strong> grating profiles ................... 152<br />

9.3.7 Multrecording . ............................................. 154<br />

VII


Contents<br />

9.3.8 Holography . . . ............................................. 154<br />

9.4 Di-, pyro-, <strong>and</strong> ferroelectricity .................................. 155<br />

9.4.1 Dielectric properties of diacetylenes .............................. 156<br />

9.4.1.1 Correlation of polymer content <strong>and</strong> electric permittivity . .............. 156<br />

9.4.1.2 Application to topospecifically modified diacetylenes . . . .............. 158<br />

9.4.1.3 Additional applications ....................................... 159<br />

9.4.2 Pyroelectric diacetylenes . . . ................................... 159<br />

9.4.2.1 IPUDO ................................................... 159<br />

9.4.2.2 NP/4-MPU . . . ............................................. 160<br />

9.4.2.3 DNP/MNP . . . ............................................. 161<br />

9.4.2.4 Spurious piezo <strong>and</strong> pyroelectricity of diacetylenes .................... 161<br />

9.4.3 The ferroelectric diacetylene DNP ............................... 162<br />

9.4.4 Summary ................................................. 166<br />

9.5 Non-linear optical properties ................................... 167<br />

9.5.1 Aims of investigation ........................................ 167<br />

9.5.2 Experimental setup .......................................... 167<br />

9.5.3 Theoretical approaches ....................................... 170<br />

9.5.4 Sample preparation .......................................... 170<br />

9.5.5 Value <strong>and</strong> phase of the third order susceptibility w (3) .................. 170<br />

9.5.6 Relaxation of the singlet exciton ................................ 171<br />

9.5.7 The w (3) tensor components . ................................... 172<br />

9.5.8 Signal saturation ............................................ 173<br />

9.5.9 Spectral dispersion, phase, <strong>and</strong> relaxation of w (5) ..................... 174<br />

9.5.10 Conclusion . . . ............................................. 176<br />

References . . . ............................................. 177<br />

10 Matrix-Molecule Interaction in Dye-Doped Rare Gas Solids .......... 181<br />

Thomas Giering, Peter Geißinger, Wolfgang Richter, <strong>and</strong> Dietrich Haarer<br />

10.1 Introduction . . ............................................. 181<br />

10.2 Stochastic theory ............................................ 183<br />

10.3 Rare gases ................................................. 186<br />

10.4 Experimental . . ............................................. 187<br />

10.5 Inhomogeneous absorption lines ................................. 188<br />

10.6 Pressure effects ............................................. 190<br />

10.7 Rare gas mixtures ........................................... 191<br />

10.8 Summary ................................................. 194<br />

Acknowledgements .......................................... 195<br />

References . . . ............................................. 195<br />

II<br />

Mainly Micelles, Polymers, <strong>and</strong> Liquid Crystals<br />

11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant<br />

Solutions ................................................. 199<br />

Heinz Hoffmann<br />

11.1 General behaviour of surfactants ................................ 199<br />

VIII


Contents<br />

11.2 From globular micelles towards bilayers . . . ........................ 200<br />

11.3 Viscoelastic solutions with entangled rods . ........................ 202<br />

11.3.1 General behaviour ........................................... 202<br />

11.3.2 Viscoelastic systems ......................................... 205<br />

11.3.3 Mechanisms for the different scaling behaviour ..................... 209<br />

11.4 Viscoelastic solutions with multilamellar vesicles .................... 211<br />

11.4.1 The conditions for the existence of vesicles ........................ 211<br />

11.4.2 Freeze fracture electron microscopy .............................. 212<br />

11.4.3 Rheological properties ........................................ 213<br />

11.4.4 Model for the shear modulus ................................... 217<br />

11.5 Ringing gels . . ............................................. 220<br />

11.5.1 Introduction . . ............................................. 220<br />

11.5.2 The aminoxide system ........................................ 221<br />

11.5.3 The bis-(2-ethylhexyl)sulfosuccinate system ........................ 224<br />

11.5.4 PEO-PPO-PEO block copolymers ................................ 226<br />

11.6 Lyotropic mesophases ........................................ 227<br />

11.6.1 Introduction . . ............................................. 227<br />

11.6.2 Nematic phases <strong>and</strong> their properties .............................. 228<br />

11.6.3 Cholesteric phases <strong>and</strong> their properties ............................ 232<br />

11.6.4 Vesicle phases <strong>and</strong> L 3 phases ................................... 233<br />

11.7 Shear induced phenomena . . ................................... 236<br />

11.7.1 General ................................................... 236<br />

11.7.2 Under what conditions do we find drag-reducing surfactants? ........... 236<br />

11.8 SANS measurements on micellar systems . . ........................ 239<br />

11.9 A new rheometer ............................................ 243<br />

References . . . ............................................. 247<br />

12 Photophysics of J Aggregates .................................. 251<br />

Hermann Pschierer, Hauke Wendt, <strong>and</strong> Josef Friedrich<br />

12.1 Introduction . . ............................................. 251<br />

12.2 Basic aspects of pressure <strong>and</strong> electric field phenomena in hole burning<br />

spectroscopy of J aggregates ................................... 252<br />

12.3 Experimental . . ............................................. 253<br />

12.4 Results ................................................... 254<br />

12.5 Discussion . . . ............................................. 256<br />

12.5.1 Pressure phenomena ......................................... 256<br />

12.5.2 Electric field-induced phenomena ............................... 258<br />

Acknowledgements .......................................... 258<br />

References . . . ............................................. 259<br />

13 Convection Instabilities in Nematic Liquid Crystals ................ 260<br />

Lorenz Kramer <strong>and</strong> Werner Pesch<br />

13.1 Introduction . . ............................................. 260<br />

13.2 Basic equations <strong>and</strong> instability mechanisms ........................ 264<br />

13.2.1 The director equation ......................................... 264<br />

13.2.2 The velocity field ........................................... 266<br />

IX


Contents<br />

13.2.3 Electroconvection ........................................... 267<br />

13.2.3.1 The st<strong>and</strong>ard model .......................................... 267<br />

13.2.3.2 The weak electrolyte model . ................................... 268<br />

13.2.4 Rayleigh-Bénard convection . ................................... 269<br />

13.3 Theoretical analysis .......................................... 269<br />

13.4 Rayleigh-Bénard convection . ................................... 275<br />

13.5 Electrohydrodynamic convection ................................ 278<br />

13.5.1 Linear theory <strong>and</strong> type of bifurcation ............................. 278<br />

13.5.2 Results of Ginzburg-L<strong>and</strong>au equation ............................. 279<br />

13.5.3 Beyond the Ginzburg-L<strong>and</strong>au equation ............................ 281<br />

13.5.3.1 Experimental results ......................................... 281<br />

13.5.3.2 Theoretical results <strong>and</strong> discussion ................................ 282<br />

13.6 Concluding remarks .......................................... 286<br />

Acknowledgements .......................................... 288<br />

Note added . . . ............................................. 289<br />

References . . . ............................................. 290<br />

14 Preparation <strong>and</strong> <strong>Properties</strong> of Ionic <strong>and</strong> Surface Modified<br />

Micronetworks ............................................. 295<br />

Michael Mirke, Ralf Grottenmüller, <strong>and</strong> Manfred Schmidt<br />

14.1 Introduction . . ............................................. 295<br />

14.2 Polymerization in normal microemulsion . . ........................ 295<br />

14.2.1 Mechanism <strong>and</strong> size control . ................................... 295<br />

14.2.2 Surface functionalization of microgels ............................ 298<br />

14.3 Polymerization in inverse microemulsion . . ........................ 299<br />

14.3.1 Preparation of ionic microgels .................................. 299<br />

14.3.2 <strong>Properties</strong> of ionic microgels <strong>and</strong> interparticle interaction .............. 299<br />

14.4 Conclusion <strong>and</strong> relevance to future work . . ........................ 303<br />

References . . . ............................................. 304<br />

15 Ferrocene-Containing Polymers ................................ 305<br />

Oskar Nuyken,Volker Burkhardt, Thomas Pöhlmann, Max Herberhold,<br />

Fred Jochen Litterst, <strong>and</strong> Christian Hübsch<br />

15.1 Introduction . . ............................................. 305<br />

15.2 Addition polymers ........................................... 306<br />

15.2.1 Radical polymerization ....................................... 306<br />

15.2.2 Radical copolymerization . . . ................................... 308<br />

15.2.3 Anionic polymerization of VFc ................................. 308<br />

15.2.3.1 Living polymerization ........................................ 309<br />

15.2.3.2 Block copolymers ........................................... 312<br />

15.2.4 Polymeranalogeous reactions ................................... 314<br />

15.3 Polymers with ferrocene units in the main chain ..................... 315<br />

15.3.1 Polycondensation ............................................ 315<br />

15.3.2 Polymers by addition of dithiols to diolefins ........................ 316<br />

15.3.2.1 Radical reaction ............................................ 316<br />

15.3.2.2 Base catalyzed reactions . . . ................................... 316<br />

X


Contents<br />

15.3.2.3 Acid catalyzed reactions . . . ................................... 318<br />

15.3.3 1,1'-dimercapto-ferrocene as initiator ............................. 319<br />

15.3.4 Reductive coupling .......................................... 320<br />

15.4 Mößbauer studies of polymers containing ferrocene .................. 320<br />

References . . . ............................................. 323<br />

16 Transfer of Vibrational Energy in Dye-Doped Polymers ............. 325<br />

Johannes Baier, Thomas Dahinten, <strong>and</strong> Alois Seilmeier<br />

16.1 Introduction . . ............................................. 325<br />

16.2 Experimental . . ............................................. 326<br />

16.3 Results <strong>and</strong> discussion ........................................ 327<br />

16.4 Summary ................................................. 332<br />

References . . . ............................................. 332<br />

17 Picosecond Laser Induced Photophysical Processes of Thiophene<br />

Oligomers ................................................ 333<br />

Dieter Grebner, Matthias Helbig, <strong>and</strong> Sabine Rentsch<br />

17.1 Introduction . . ............................................. 333<br />

17.2 Experimental . . ............................................. 334<br />

17.3 Spectroscopic properties of oligothiophenes ........................ 337<br />

17.4 Results ................................................... 337<br />

17.4.1 Picosecond-transient spectra of oligothiophenes in solution ............. 337<br />

17.4.2 Time behaviour of transient spectra .............................. 339<br />

17.4.3 Size dependence of spectroscopic properties of oligothiophenes ......... 341<br />

17.4.4 Size dependence of the kinetic behaviour of oligothiophenes ............ 342<br />

17.5 Discussion . . . ............................................. 343<br />

References . . . ............................................. 343<br />

18 Topospecific Chemistry at Surfaces ............................. 344<br />

Hans Ludwig Krauss<br />

18.1 Introduction . . ............................................. 344<br />

18.1.1 The problem . . ............................................. 344<br />

18.1.2 Preparative <strong>and</strong> analytical methods ............................... 344<br />

18.1.3 Industrial applications ........................................ 345<br />

18.1.4 The st<strong>and</strong>ard procedures of the Phillips process ..................... 345<br />

18.1.5 Earlier work . . ............................................. 346<br />

18.2 The support . . . ............................................. 346<br />

18.2.1 Unmodified silica ........................................... 346<br />

18.2.2 Modified silica ............................................. 347<br />

18.2.3 Others .................................................... 348<br />

18.3 Transition metal surface compounds .............................. 348<br />

18.3.1 The metals . . . ............................................. 348<br />

18.3.2 Impregnation <strong>and</strong> activation . ................................... 349<br />

18.4 Coordinatively unsaturated sites ................................. 350<br />

18.4.1 Reduction of saturated surface compounds . ........................ 350<br />

18.4.2 Elimination of lig<strong>and</strong>s ........................................ 352<br />

XI


Contents<br />

18.5 Physical properties of the coordinatively unsaturated sites .............. 353<br />

18.5.1 Topologically different sites . ................................... 353<br />

18.5.2 Optical <strong>and</strong> magnetic properties ................................. 353<br />

18.6 Chemical properties of the coordinatively unsaturated sites ............. 355<br />

18.6.1 Survey of catalytic reactions ................................... 355<br />

18.6.2 Olefin polymerization ........................................ 356<br />

18.6.3 Other catalytic reactions . . . ................................... 360<br />

18.7 Deactivation . . ............................................. 361<br />

18.8 Summary <strong>and</strong> outlook ........................................ 362<br />

References . . . ............................................. 363<br />

III<br />

Biopolymers<br />

19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers 369<br />

Stefan Limmer, Günther Ott, <strong>and</strong> Mathias Sprinzl<br />

19.1 Introduction . . ............................................. 369<br />

19.2 Site-specific NMR spectroscopy of chemically synthesized RNA duplexes . . 370<br />

19.2.1 Stability of tRNA-derived acceptor stem duplexes .................... 371<br />

19.2.2 Manganese ion binding sites at RNA duplexes ...................... 374<br />

19.2.3 Structural determination of short RNA duplexes by 2D NMR spectroscopy . 377<br />

19.2.4 NMR derived model of the tRNA Ala acceptor arm ................... 379<br />

19.2.5 Chemical shifts <strong>and</strong> scalar coupling as an indicator of RNA structure in the<br />

vicinity of a G-U pair ........................................ 380<br />

19.2.6 Structure of aminoacyl-tRNA <strong>and</strong> transacylation of the aminoacyl residue . . 382<br />

19.3 Structure of elongation factor Tu ................................ 383<br />

19.3.1 Sequence of Thermus thermophilus EF-Tu.......................... 384<br />

19.3.2 Crystallization, X-ray analysis, <strong>and</strong> the tertiary structure . .............. 386<br />

19.3.3 Nucleotide binding <strong>and</strong> GTPase reaction . . ........................ 387<br />

19.3.4 Mechanism of GTP induced conformational change of EF-Tu ........... 388<br />

19.3.5 Aminoacyl-tRNA in complex with EF-Tu 7 GTP ..................... 389<br />

19.3.6<br />

1 H NMR of yeast Phe-tRNA Phe EF-Tu7GTP complex . . .............. 391<br />

19.3.7<br />

13 C NMR studies of the Val-tRNA Val EF-Tu7GTP ternary complex ....... 393<br />

19.3.8 Role of EF-Tu in complex with aminoacyl-tRNA ..................... 395<br />

19.3.9 EF-Tu interaction with EF-Ts ................................... 395<br />

19. 3.10 Site-directed mutagenesis of EF-Tu ............................... 396<br />

19.4 Summary <strong>and</strong> conclusions . . ................................... 397<br />

Acknowledgement ........................................... 397<br />

References . . . ............................................. 398<br />

20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers ......... 401<br />

Alex<strong>and</strong>er Wokaun<br />

20.1 Introduction . . ............................................. 401<br />

20.2 Diffusion in surfactant systems .................................. 402<br />

20.2.1 Structural characteristics of micellar solutions, cubic phases, <strong>and</strong> multilamellar<br />

vesicles from NMR self-diffusion measurements .............. 402<br />

XII


Contents<br />

20.2.2 Probing of mobilities in multilamellar vesicles by forced Rayleigh scattering 405<br />

20.2.3 Dimensionality of diffusion in lyotropic mesophases from fluorescence<br />

quenching ................................................. 409<br />

20.2.4 Summary of results .......................................... 412<br />

20.3 Vibrational spectroscopy <strong>and</strong> conformational analysis of oligonucleotides . . 413<br />

20.3.1 Spectroscopic characterization of right <strong>and</strong> left-helical forms of a hexadecanucleotide<br />

duplex ........................................ 413<br />

20.3.2 SERS spectra of deoxyribonucleotides ............................ 415<br />

20.3.3 Studies of chromophore-DNA interaction by vibrational spectroscopy . . . . . 417<br />

20.3.4 Summary of results .......................................... 418<br />

20.4 Related projects carried out within the framework of the Collaborative<br />

Research Centre ............................................ 419<br />

20.5 Remarks <strong>and</strong> acknowledgements ................................ 421<br />

References . . . ............................................. 422<br />

21 Energy Transport by Lattice Solitons in a-Helical Proteins ........... 424<br />

Franz-Georg Mertens, Dieter Hochstrasser, <strong>and</strong> Helmut Büttner<br />

21.1 Introduction . . ............................................. 424<br />

21.2 The model ................................................. 426<br />

21.3 Quasicontinuum approximation ................................. 428<br />

21.4 Velocity range for the quasicontinuum approach ..................... 431<br />

21.5 Solitary waves for realistic parameter values ........................ 432<br />

21.6 Iterative method <strong>and</strong> stability ................................... 434<br />

21.7 Conclusion . . . ............................................. 437<br />

References . . . ............................................. 438<br />

IV<br />

Appendix<br />

22 Documentation of the Collaborative Research Centre 213 ............ 443<br />

Markus Schwoerer <strong>and</strong> Heinz Hoffmann<br />

22.1 List of Members ............................................ 443<br />

22.2 Heads of Projects (Teilprojektleiter) .............................. 444<br />

22.2.1 Projektbereich A: Gemeinsame Einrichtungen ...................... 444<br />

22.2.2 Projektbereich B: Festkörper ................................... 444<br />

22.2.3 Projektbereich C: Funktionale Systeme – Mizellen, Oberflächen<br />

und Polymere . ............................................. 445<br />

22.2.4 Projektbereich D: Biopolymere ................................. 446<br />

22.3 Guests .................................................... 447<br />

22.4 Co-workers . . . ............................................. 450<br />

22.5 International Cooperation . . . ................................... 457<br />

22.6 Funding .................................................. 459<br />

XIII


Preface<br />

At the end of 1983, about eight years after the inauguration of the University of Bayreuth<br />

the Deutsche Forschungsgemeinschaft (DFG) agreed to establish the Collaborative Research<br />

Centre 213, TOPOMAC, for the promotion of basic research in chemistry <strong>and</strong> physics of<br />

macromolecular systems. The title of TOPOMAC in its full length, “Topospezifische Chemie<br />

und Toposelektive Spektroskopie von Makromolekülsystemen: Mikroskopische Wechselwirkung<br />

und Makroskopische Funktion” expessed the intention of the original applicants:<br />

a productive cooperation between physicists, chemists <strong>and</strong> biochemists across the mutual<br />

borders of their original research fields. Until the end of 1995 TOPOMAC was supported by<br />

the Deutsche Forschungsgemeinschaft with 28 MDM.<br />

The present book documents the achievements of TOPOMAC. It is not a minute addition<br />

of all the results which have been published in periodical journals but rather a survey of<br />

important research fields of members of TOPOMAC. The articles have been written towards<br />

or after the end of the support period as both, review <strong>and</strong> original publications.<br />

The book covers the fields of:<br />

. Electronic, photoelectric, thermal, dielectric, optical <strong>and</strong> magnetic properties of macromolecular<br />

solids (polymers <strong>and</strong> polymer crystals),<br />

. Micellar structures, J-aggregates, liquid crystals, µ-gels, ferrocene-containing polymers<br />

<strong>and</strong> topospecific chemistry at surfaces <strong>and</strong> in single crystals, <strong>and</strong><br />

. Biopolymers <strong>and</strong> surfactant systems as studied by site directed spectroscopy <strong>and</strong> site directed<br />

chemistry <strong>and</strong> also by the theory of energy transport.<br />

During the period of its support TOPOMAC had 30 members (Teilprojektleiter). Only<br />

ten of them have been members for the entire period, mainly because twenty times a call<br />

from other universities or research institutions reached one of the members of the Collaborative<br />

Research Centre 213. Fourteen of them followed this call <strong>and</strong> left the Collaborative Research<br />

Centre 213. Less than two years after the end of the period of support some of the remaining<br />

former members of Collaborative Research Centre 213 together with young <strong>and</strong><br />

new faculty members began to continue the formal cooperation between chemists <strong>and</strong> physicists<br />

in the field of macromolecular research. Their actual cooperation in the meantime<br />

never had been terminated.<br />

The editors would like to express the sincere thanks of the members of TOPOMAC to<br />

the foreign guests of TOPOMAC, to the Deutsche Forschungsgemeinschaft, to the University<br />

XV


Preface<br />

of Bayreuth <strong>and</strong> also to the Freistaat Bayern. Our foreign guests stayed for long or short periods<br />

between one year <strong>and</strong> one day. They have contributed in an essential <strong>and</strong> special manner<br />

to the success on our research fields. They also strongly intensified the national <strong>and</strong> international<br />

scientific relations of both, the members <strong>and</strong> the research students of TOPO-<br />

MAC. The cooperation of the speakers of TOPOMAC with Dr. Funk from the Deutsche Forschungsgemeinschaft<br />

office throughout the entire period of support was excellent. The help<br />

of the former president of the university, Dr. K. D. Wolff <strong>and</strong> his chancellor, W. P. Hentschel<br />

as well as the continuous support by the late Ministerialrat G. Grote <strong>and</strong> by J. Großkreutz<br />

from the Bayerisches Staatsministerium für Unterricht, Kultus, Wissenschaft und Kunst has<br />

been an essential stimulus for the scientific members of TOPOMAC. Last but not least we<br />

thank Doris Buntkowski for her faithful <strong>and</strong> reliable work as our secretary.<br />

The Editors<br />

XVI


List of Contributors<br />

Alex<strong>and</strong>er Blumen<br />

Theoretische Polymerphysik<br />

Universität Freiburg<br />

Herrmann-Herder-Straße 3<br />

79104 Freiburg<br />

Elmar Dormann<br />

Physikalisches Institut<br />

Universität Karlsruhe<br />

Engesserstr. 7<br />

67131 Karlsruhe<br />

Pablo Esquinazi<br />

Institut für Experimentelle Physik II<br />

Universität Leipzig<br />

Linnestraße 5<br />

04103 Leipzig<br />

Klaus Fesser<br />

Fachbereich Physik<br />

Universität Greifswald<br />

Domstraße 10a<br />

17489 Greifswald<br />

Josef Friedrich<br />

Lehrstuhl für Physik<br />

Technische Universität München<br />

85350 Freising-Weihenstephan<br />

Holger Fehske<br />

Theoretische Physik I<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Dietrich Haarer<br />

Experimentalphysik IV<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Heinz Hoffmann<br />

Physikalische Chemie I<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Lorenz Kramer<br />

Theoretische Physik II<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Hans Ludwig Krauss<br />

Heunischstraße 5 b<br />

96049 Bamberg<br />

Franz G. Mertens<br />

Theoretische Physik<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Oskar Nuyken<br />

Lehrstuhl für Makromolekulare Stoffe<br />

Technische Universität München<br />

Lichtenbergstraße 4<br />

85747 München<br />

XVII


List of Contributors<br />

Sabine Rentsch<br />

Institut für Optik und Quantenelektronik<br />

Friedrich Schiller Universität Jena<br />

Max-Wien-Platz 1<br />

07743 Jena<br />

Wolfgang Richter<br />

Experimentalphysik IV<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Paul Rösch<br />

Biopolymere<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Manfred Schmidt<br />

Institut für Physikalische Chemie<br />

Universität Mainz<br />

Welder-Weg 11<br />

55099 Mainz<br />

Markus Schwoerer<br />

Experimentalphysik II<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Alois Seilmeier<br />

Physikalisches Institut<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Mathias Sprinzl<br />

Biochemie<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Peter Strohriegel<br />

Makromolekulare Chemie I<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Thomas Vogtmann<br />

Experimentalphysik II<br />

Universität Bayreuth<br />

Universitätsstraße 30<br />

95447 Bayreuth<br />

Alex<strong>and</strong>er Wokaun<br />

Bereich F5<br />

Paul Scherrer Institut<br />

CH-5232 Villingen<br />

XVIII


I<br />

Mainly Solids<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


1 Model Systems for Photoconductive Materials<br />

Harald Meyer <strong>and</strong> Dietrich Haarer<br />

1.1 Introduction<br />

The effect of photoconductivity, i. e. the increase of the electrical conductivity of a material<br />

upon illumination with light of suitable photon energy, was first discovered in selenium by<br />

W. Smith in 1873.<br />

A typical application for these materials is the xerographic process [2, 3], which was<br />

developed by C. F. Carlson in 1942 [1]. Here, a photoconducting film on top of a grounded<br />

electrode is homogeneously charged by a corona discharge. Typical field strengths are up to<br />

10 6 V/m. In a second step the image of the original document is projected onto the film. At<br />

the areas where the film is illuminated charge carriers are generated in the photoconducting<br />

film. One species traverses the film <strong>and</strong> recombines at the grounded electrode whereas the<br />

oppositely charged species neutralizes the surface charges. With this step the original image<br />

is transferred into an electrostatic image on the photoconductor film. Subsequently, small toner<br />

particles are deposited on the photoconductor. They stick to the charged regions <strong>and</strong><br />

thus generate a real image. In the next step, this image is transferred onto paper <strong>and</strong> fixed<br />

by thermally fusing the toner particles on the paper.<br />

The process for a laser printer is similar except for the fact that the image is written<br />

onto the photoconductor directly by a laser or a diode array. If the toner particles are fused<br />

directly onto the photoconductor instead of transferring them onto paper the photoconductor<br />

can be used as an offset printing master [4].<br />

Potential systems for commercial use have to meet several requirements:<br />

a) sensitivity in the visible region of the spectrum;<br />

b) good charge transport properties, i. e. charge carrier mobility in excess of 10 –7 cm 2 /Vs<br />

[3] <strong>and</strong> minor trapping effects;<br />

c) good mechanical <strong>and</strong> dielectrical properties;<br />

d) excellent film forming properties <strong>and</strong> possibility of manufacturing defect free large area<br />

films;<br />

e) mechanical flexibility for the use in small sized desktop devices.<br />

The last two requirements cannot be met neither by organic nor inorganic crystalline<br />

materials. Therefore both, organic <strong>and</strong> inorganic amorphous photoconductors, have been de-<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

3


1 Model Systems for Photoconductive Materials<br />

veloped. Inorganic materials like e. g. a-Se, As 2 Se 3 or alloys of Se <strong>and</strong> Te show good transport<br />

properties with mobilities in the range of 0.1 cm 2 /Vs at room temperature [7], <strong>and</strong> high<br />

sensitivity for visible light together with moderate mechanical <strong>and</strong> dielectric properties. One<br />

of their most important disadvantages compared to organic systems is that most of these materials<br />

are highly toxic.<br />

In the past 30 years, since H. Hoegl discovered photoconductivity in the organic polymer<br />

poly(N-vinylcarbazole) (PVK) (Fig. 1.2) [5, 6], organic materials have almost completely<br />

replaced their inorganic counterparts, although the effective mobility in these systems is<br />

typically 5 orders of magnitude lower as compared to a-Se.<br />

Organic photoconductors can be regarded as large or medium b<strong>and</strong>gap semiconductors.<br />

A typical value for commercially applied systems is 3.5 eV for carbazole derivatives<br />

(Section 1.4). This inherent lack of sensitivity in the visible region has been overcome by<br />

using charge transfer systems [5] or multilayer systems with additional charge generation<br />

layers. For a review see e. g. Ref. [3]. On the other h<strong>and</strong>, the large b<strong>and</strong>gap of organic materials<br />

virtually eliminates the influence of thermally generated charge carriers <strong>and</strong> thus improves<br />

the dielectric properties as compared to low b<strong>and</strong>gap materials.<br />

Therefore the main target of the works, which will be described in the following, was<br />

to identify the key factors which limit the charge carrier mobilities in organic systems <strong>and</strong><br />

to develop new high mobility materials.<br />

1.2 Experimental techniques<br />

Besides the xerographic discharge method [3, 10], the time-of-flight (TOF) technique is generally<br />

used to study the charge carrier transport of thin organic films. Here, the photoconducting<br />

film with a typical thickness of 10 mm is s<strong>and</strong>wiched between two electrodes (Fig. 1.1). Electron-hole<br />

pairs are generated by the energy hn of a strongly absorbed laser pulse, which is irradiated<br />

through one of the semitransparent electrodes. For most organic materials the charge<br />

carrier generation process can be described by an Onsager model, either in its one or three dimensional<br />

form [11–13].<br />

The wavelength of the laser is chosen to ensure that the penetration depth is considerably<br />

less than the sample thickness. Thus the charge carriers are generated close to the illuminated<br />

surface. Under the influence of an externally applied electrical field the electronhole<br />

pairs are separated <strong>and</strong> one species, depending on the polarity of the external field, immediately<br />

recombines at the illuminated electrode. The opposite charged species drifts<br />

through the sample, thus giving rise to a time dependent photocurrent I p (t).<br />

For a quantitative analysis the knowledge about the electrical field inside the sample<br />

is necessary. This can be achieved by performing the measurements in the small signal limit,<br />

i. e. disturbations due to space charge effects are negligible <strong>and</strong> the electric field in the bulk<br />

is determined by the externally applied field. It turns out, that this condition is fulfilled as<br />

long as the generated photocharge Q tot is less than 10 % of the charge Q = CU, which is<br />

4


1.3 Transport models<br />

Figure 1.1: Principle setup for TOF experiments.<br />

stored on the electrodes. With typical experimental parameters (sample thickness d =10mm,<br />

sample area A =4cm 2 ,10 21 monomer units per cm 3 , voltage U = 300 V, sample capacitance<br />

C = 1 nF) the above criterion leads to an upper limit for the photocharges of Q tot


1 Model Systems for Photoconductive Materials<br />

neglected, or in other words, the detrapping always occurs to states at e d . The traps can be<br />

caused by static disorder as well as by self trapping due to polaronic effects or both.<br />

Mathematically equivalent with the MT model is the Continuous Time R<strong>and</strong>om Walk<br />

(CTRW) model [47–49], where the exponential trap distribution density g (e) in the MT<br />

model,<br />

<br />

"<br />

g…"† /exp ; …2†<br />

k B T 0<br />

corresponds to the well-known algebraic distribution of hopping times in the CTRW model<br />

[45],<br />

…t† /t 1 ; …3†<br />

where the disorder parameter a in Eq. 3 corresponds to the dimensionless temperature<br />

(a = T/T 0 ).<br />

A variety of experimental data [16, 25, 26, 33, 34] can be described by the empirical<br />

formula,<br />

<br />

ef f ˆ 0 exp<br />

p<br />

" 0 E<br />

k B T ef f<br />

<br />

; …4†<br />

first proposed by Gill [33], where e 0 is the zero field activation energy, E the electrical field<br />

<strong>and</strong> b the Poole-Frenkel factor. This equation describes the thermal activated release from<br />

localized traps where the activation energy is lowered by an external field according to the<br />

Poole-Frenkel effect [37].<br />

The effective temperature T eff is related to the physical temperature T by<br />

1<br />

ˆ 1<br />

T ef f T<br />

1<br />

T 0<br />

:<br />

…5†<br />

Initially, the characteristic temperature T 0 simply was an empirical parameter. In<br />

Section 1.4, however, we shall see that in certain cases this parameter can be interpreted microscopically.<br />

An alternative approach [28, 50–54] is based on the assumption that the density of<br />

states can be modelled by a Gaussian distribution. Charge carrier transport occurs via direct<br />

hopping between the localized sites. In general, the differences between the two models are<br />

too small to be detected experimentally. Since our data can be quantitatively explained<br />

within the framework of multiple trapping we will restrict the discussion to this model.<br />

6


1.4 Results<br />

1.4 Results<br />

As stated in Section 1.1, organic photoconductors can be regarded as large or medium b<strong>and</strong>gap<br />

semiconductors. In contrast to inorganic crystalline semiconductors the electronic coupling between<br />

adjacent sites is weak. Therefore the electronic states in these materials are primarily of<br />

molecular character <strong>and</strong> the charge carrier transport has to be regarded as an incoherent hopping<br />

process between two neighbouring sites. For the following discussion organic photoconductors<br />

will be divided into three main groups namely molecularly doped polymers (Section<br />

1.4.1), side-chain polymers (Section 1.4.2.), <strong>and</strong> conjugated systems (Section 1.4.3).<br />

1.4.1 Molecularly doped polymers<br />

In molecularly doped polymers (MDPs) the transport molecules are molecularly dispersed in<br />

an inert matrix. Due to crystallization the maximum concentration of chromophores, which<br />

can be achieved, is typically in the range of 50 mol%.<br />

Typical chemical compounds include oxadiazole derivatives [14], pyrazolines [15–<br />

18], hydrazones [19–22], carbazole derivatives [23–26], triphenylmethane (TPM) derivatives<br />

[27, 28], triphenylamine (TPA) derivatives [29, 30], <strong>and</strong> TAPC [31], which can be regarded<br />

as a dimer of TPA. The charge carrier mobilities at room temperature are typically<br />

in the range from 10 –6 cm 2 /Vs for N-isopropylcarbazole [25] to 10 –4 cm 2 /Vs for p-diethylaminobenzaldehyde<br />

diphenyl hydrazone (DEH) [22].<br />

In order to study the effect of chemical substitutions of the respective transport molecule<br />

N-isopropylcarbazole (NIPC) <strong>and</strong> derivatives thereof, with electron donor as well as acceptor<br />

substituents (Fig. 1.2), have been investigated in a polycarbonate host [26].<br />

As compared by 3,6-dibromo-N-isopropylcarbazole (DBr-NIPC) the effective hole<br />

mobility of 3,6-dimethoxy-N-isopropylcarbazole (DMO-NIPC) is slightly decreased. The<br />

reason for this behaviour is the fact that the cation, which is relevant for the transport process,<br />

is stabilized by electron donating substituents like the methoxy group, as can be seen<br />

from the shift of the ionisation potentials [26]. This leads to stronger localization of the<br />

charge within the aromatic rings as compared to derivatives with moderate electron acceptors<br />

like bromine. Therefore the spatial overlap of the electronic states of adjacent molecules,<br />

which are involved in the transport process, is reduced.<br />

With strong acceptors like the nitro groups in 3,6-dinitro-N-isopropylcarbazole (DN-<br />

NIPC) the ionisation potential is already larger than for the host matrix polycarbonate. In<br />

this case even the matrix acts as a trap for holes, thus preventing efficient charge carrier<br />

transport [26].<br />

Since the effective mobility is determined by the spatial overlap of the transport states,<br />

the concentration dependence of the effective mobility can be described by<br />

<br />

ef f / r 2 exp<br />

<br />

2r<br />

; …6†<br />

r 0<br />

7


1 Model Systems for Photoconductive Materials<br />

N<br />

H 3 CO<br />

N<br />

N<br />

N<br />

N<br />

C 2 H 5<br />

R<br />

R<br />

BTA<br />

R: H: NIPC<br />

Br: DBr-NIPC<br />

OCH 3 : DMO-NIPC<br />

NO 2 : DN-NIPC<br />

Si<br />

Si<br />

O<br />

CH 3<br />

PMPS<br />

n<br />

Ca n<br />

CH CH 2<br />

PVK<br />

(CH 2 ) n<br />

Ca<br />

n<br />

R<br />

Polysiloxan<br />

R<br />

n<br />

Ca:<br />

N<br />

R:<br />

O<br />

DPOP-PPV<br />

Figure 1.2: Chemical structure of photoconducting materials. Abbreviations see text.<br />

H<br />

PPV<br />

where r is the mean distance between adjacent transport molecules <strong>and</strong> r 0 the wave function<br />

decay parameter. For carbazole derivatives r 0 is typically in the range 1.3–1.6 Å, depending<br />

on the substituted groups [26].<br />

In recent years, also percolation models have been successfully applied to experimental<br />

data [55] as an alternative approach to model the transport properties of doped disordered<br />

systems. For polycarbonate doped with a derivative of benztriazole (BTA) (Fig. 1.2) it has<br />

been shown that the concentration dependence of the effective mobility can be described at<br />

low concentrations by [55]<br />

ef f ˆ 0 …p p c † t …7†<br />

for p>p c , where p is the concentration of transport molecules, p c the percolation threshold,<br />

<strong>and</strong> t the critical exponent. The experimental values (p c = 0.095 <strong>and</strong> t = 2.46) are consistent<br />

with 3D-continuum percolation calculations [56].<br />

Since this system shows good transport properties over a wide range of concentrations,<br />

the influence of extrinsic traps could be studied [57]. Here, a sample containing 25 wt% of<br />

BTA in a polycarbonate host was doped with small amounts of the organic dye astrazone<br />

orange (AO). It turned out that the effective mobility is reduced by a factor of 2 when the<br />

8


1.4 Results<br />

concentration of AO is increased to 0.5 wt%. This finding becomes also important when<br />

thinking of new applications like organic electroluminescent devices. One way to increase<br />

the efficiency in these devices is to molecularly disperse the luminescent moiety into a<br />

charge transport material. But there is a trade-off between efficient trapping of the charge<br />

carriers – which is improved by higher concentrations of the luminescent groups – <strong>and</strong> concentration<br />

quenching of the fluorescence, which can be avoided by keeping the concentration<br />

as low as possible. The above described measurements [57] show that organic dyes can be<br />

very efficient traps for charge carriers. Therefore the concentration of the luminescent moiety<br />

can be as low as 2% or less [61], thus avoiding concentration quenching without loosing<br />

high trapping efficiency.<br />

1.4.2 Side-chain polymers<br />

In side-chain polymers, where the chromophore is covalently bonded to a polymer main<br />

chain, the concentration of transport molecules can be increased as compared to MDPs without<br />

causing crystallization. Because the underlying physical mechanism of charge carrier<br />

transport (nearest neighbour hopping between weakly coupled, localized states) is the same<br />

for both MDPs <strong>and</strong> side-chain polymers, the transport properties are qualitatively similar.<br />

This finding offers the opportunity to tailor the thermal <strong>and</strong> mechanical properties of<br />

the photoconductor without seriously affecting the transport properties. This can be achieved<br />

either by variation of the spacer length between the chromophore <strong>and</strong> the backbone or by variation<br />

of the backbone itself. A typical example for this group is PVK [12, 32, 34, 35]. Here,<br />

the transition from dispersive to non-dispersive transport could be observed [34]. Due to the<br />

fact, that the TOF curves have been measured over 10 decades in time, it was possible by<br />

means of a numerical inverse Laplace transform [34] to calculate from the measured photocurrent<br />

the trapping rate distribution, which is a measure for the density of localized states.<br />

In polysiloxane derivatives with pendant carbazolyl groups [58, 59] (Fig. 1.2) the effect<br />

of a variation of the spacer length was investigated. Starting from a spacer length of 3 or 5<br />

methylene units up to 6 or 11 groups, the glass transition temperature drops from its initial value<br />

of 51 8Cor78Cdownto–58Cor–458C for the C 11 spacer. It turns out, that the zero field<br />

activation energy e act for the hole transport remains unchanged <strong>and</strong> is the same for NIPC or<br />

PVK, e act = 0.51 eV [58]. This indicates, that the activation energy in side-chain polymers<br />

with pendant carbazolyl groups is dominated by intrinsic properties of the carbazolyl moiety<br />

<strong>and</strong> is not much affected by other factors like the morphology of the polymer [58].<br />

However, the absolute values for m eff differ by roughly one order of magnitude, where the<br />

compound with the shortest spacer shows an effective mobility in the range of 10 –6 cm 2 /Vs<br />

(T = 300 K, E =3710 5 V/cm). For the materials with a spacer length of 5 <strong>and</strong> 6 methylene<br />

units the mobility is lower by one order of magnitude <strong>and</strong> therefore comparable to the mobility<br />

in NIPC or PVK. A striking feature, when comparing NIPC with PVK, is the fact that<br />

polycarbonate doped with 20 wt% of NIPC shows the same hole mobility at room temperature<br />

as compared to PVK where the carbazole concentration amounts to 86 wt%. Obviously,<br />

the covalent attachment of the chromophore to a stiff backbone, like PVK with its glass<br />

transition temperature of 227 8C, induces a mutual orientation of adjacent pendant groups,<br />

9


1 Model Systems for Photoconductive Materials<br />

which reduces the electronic coupling between the two sites. Therefore it is desirable to<br />

induce a well-defined amount of flexibility in both, the spacer <strong>and</strong> the backbone, to allow<br />

for mutual reorientation of the chromophores. For a polysiloxane backbone with carbazolyl<br />

transport groups the optimum spacer length n is equal to 3.<br />

1.4.3 Conjugated systems<br />

In contrast to the two main groups described above the charge carrier transport in conjugated<br />

systems occurs via the polymer backbone. Various substituents are used to tailor the<br />

mechanical properties <strong>and</strong> the processibility. For instance, the chain can be a skeleton with<br />

quasi-s-conjugation like in polysilanes [36, 38–42] or polygermylenes [39].<br />

The second group of main chain polymers contains a p-conjugated backbone like<br />

polyacetylene, polythiophene, polypyrrole, polyphenylene, poly(phenylenevinylene) (PPV),<br />

<strong>and</strong> their derivatives. For a review see [43, 44].<br />

Both types of conjugated systems have been investigated. In the case of poly(methylphenyl<br />

silane) (PMPS), a quasi-s-conjugated polymer (Fig. 1.2), non-dispersive transport<br />

has been found down to a temperature of 250 K [60]. For lower temperatures, the transport<br />

is dispersive, which indicates that the transport in these materials is controlled by traps.<br />

Compared to the materials described above, the depths of the relevant traps are much lower.<br />

The zero field activation energy turns out to be as low as 0.37 eV as compared to 0.5 eV for<br />

materials containing carbazole. This gives a hole mobility at room temperature of roughly<br />

10 –3 cm 2 /Vs, which is 3 orders of magnitude higher than for the materials described above.<br />

The high mobility <strong>and</strong> the transparency in the visible region makes PMPS also useful for<br />

applications in multilayer electroluminescent devices [61–63].<br />

Besides the s-bonded PMPS, we investigated oligomeric conjugated compounds based<br />

on carbazole [64]. First experiments with a trimer of a carbazole derivative showed that this<br />

material is soluble in most of the common organic solvents. It forms a low molecular weight<br />

glass <strong>and</strong> shows no tendency to crystallize even after several months. In this material the<br />

p-system has been extended over three monomer units, which results in remarkably high<br />

hole mobilities as compared to the monomer model compound NIPC. First TOF experiments<br />

show an increase in mobility by roughly two orders of magnitude as compared to unconjugated<br />

carbazole systems, which reaches 2 7 10 –4 cm 2 /Vs [64].<br />

The second group of p-conjugated polymers contains one p z -orbital per carbon atom<br />

which st<strong>and</strong>s perpendicular to the plane of the s-bonded skeleton of the main chain. The interaction<br />

between these orbitals leads to electronic states, which are delocalized – at least<br />

partly – along the conjugated chain. As compared to the intrachain coupling, the intrachain<br />

coupling is lower by typically one or two orders of magnitude [66] thus leading to a large<br />

anisotropy of the electronic <strong>and</strong> optical properties. These materials can be regarded as basically<br />

one-dimensional organic semiconductors with b<strong>and</strong>gaps in the visible region of the<br />

spectrum, e. g. 1.5 eV (trans-(CH) x ) [43], 2.0 eV (polythiophene), 2.5 eV (polyphenylene vinylene),<br />

<strong>and</strong> 3.0 eV [poly(para-phenylene)] [67].<br />

We investigated the charge transport properties of the p-conjugated polymer poly<br />

(para-phenylenevinylene) <strong>and</strong> its soluble substituted derivative DPOP-PPV (Fig. 1.2). In<br />

10


1.4 Results<br />

contrast to e. g. (CH) x , PPV can be prepared in undoped form. If oxygen is carefully heated<br />

out, PPV exhibits excellent film forming properties as well as thermal stability under ambient<br />

conditions. In recent years, PPV has drawn additional attention due to its use in organic<br />

electroluminescent devices [45].<br />

In the following we will show that the charge carrier transport in these materials can<br />

be described by the conventional models mentioned before, which have been developed for<br />

the characterization of disordered molecular systems.<br />

Based on the MT model a method has been developed to calculate the distribution of<br />

capture rates from the measured photocurrent [68, 69]. In contrast to Ref. [34], the method<br />

is based on a Fourier transform technique with better numerical stability as compared to the<br />

inverse Laplace transform used in Ref. [34].<br />

With this approach the capture rate density o c is given by<br />

! c …" ! †k B T ˆ 2<br />

p<br />

Q tot sin <br />

t mic<br />

I…!†<br />

<br />

!<br />

! ; …8†<br />

where Q tot is the total photogenerated charge, t mic the microscopic transit time, <strong>and</strong> I (o) the<br />

Fourier transform of the measured photocurrent, <strong>and</strong> f the phase shift (Eq. 9). For DPOP-<br />

PPV, t mic = 7.5610 –10 s for the given experimental conditions.<br />

<br />

ˆ tan 1 Im I…!†<br />

Re I…!†<br />

…9†<br />

The trap depth e o is related to the frequency o by the expression<br />

<br />

! ˆ r 0 exp<br />

<br />

" !<br />

; …10†<br />

k B T<br />

where r 0 is the attempt-to-escape frequency (r 0 =10 10 s –1 for DPOP-PPV).<br />

In DPOP-PPV the effective mobility is field dependent <strong>and</strong> thermally activated according<br />

to Gill’s formula (Eq. 4) [33] with a characteristic temperature T 0 = 465 K (Fig. 1.3).<br />

The calculated trap density g (e) is exponential down to a pronounced cutoff energy e d ,<br />

<br />

g…"† /exp<br />

<br />

"<br />

: …11†<br />

k B T 0<br />

This cutoff energy corresponds exactly to the measured activation energy of the effective<br />

mobility. Furthermore, the initially simply empirical parameter T 0 in Gill’s formula<br />

(Eq. 4) could be correlated with the decay constant k B T 0 of the trap distribution.<br />

By calculating the total number of trapping events it can be seen that the typical distance<br />

which a charge carrier travels between two trapping events is of the order of 4 Å. This<br />

value is comparable to the intrachain distance <strong>and</strong> indicates that the transport in these materials<br />

can be best described by conventional hopping between closely neighbouring sites. No<br />

evidence for b<strong>and</strong>like transport has been found.<br />

11


µ eff<br />

[cm 2 /Vs]<br />

1 Model Systems for Photoconductive Materials<br />

10 -2<br />

10 -3<br />

10 -4<br />

350 K 300 K 270 K<br />

230 K<br />

500 kV/cm<br />

350 kV/cm<br />

200 kV/cm<br />

75 kV/cm<br />

0kV/cm<br />

10 -5<br />

10 -6<br />

10 -7<br />

10 2<br />

10 0<br />

T 0<br />

= 465K<br />

10 -2<br />

10 -4<br />

10 -6<br />

10 -8<br />

10 -10<br />

0 2 4 6 8<br />

10 -8<br />

2.5 3.0 3.5 4.0 4.5<br />

1/T [1000/K]<br />

Figure 1.3: large frame: effective mobility m eff of DPOP-PPV as a function of temperature for some<br />

electrical fields, data points for zero field were extrapolated from field dependent measurements; inset:<br />

Arrhenius fits for different electrical fields. For details see text. Data from [68, 69].<br />

For the case of PPV the dispersion of the photocurrent is too large for transit time determination.<br />

It can be concluded that even at room temperature the release times from the<br />

deepest traps, which control the transport properties, are in the range of 1 s or less. This corresponds<br />

to an effective mobility of less than 10 –8 cm 2 /Vs [69]. Based on TSC measurements<br />

from different groups [70], we conclude that this behaviour is primarily caused by the<br />

existence of grain boundaries.<br />

It is reasonable to assume that in conjugated polymers with a rigid backbone the preferred<br />

orientation of the main chain will be parallel to the film surface. Therefore, the<br />

charge carrier transport through the film has to occur perpendicular to the backbone, i. e.<br />

perpendicular to the direction with high intrinsic mobility, which would mean a principle<br />

limit for this class of materials for this kind of applications.<br />

1.5 Outlook<br />

As described above, disordered organic materials have been developed with effective hole<br />

mobilities of up to 10 –3 cm 2 /Vs together with good mechanical <strong>and</strong> dielectric properties.<br />

However it seems that for disordered systems a considerable improvement of this value will<br />

12


References<br />

be difficult to achieve, because for all investigated material groups the charge carrier transport<br />

is limited by the localized nature of the electronic states <strong>and</strong> by the hopping mechanism<br />

of the transport process.<br />

Based on recent other works [8, 9], we think that for developing even higher mobility<br />

materials it is essential to improve the structural order of the material. This may open new<br />

application fields, e. g. the use of active organic materials in semiconductor devices. One<br />

promising way is the use of highly ordered liquid crystals. With these materials, however,<br />

charge carrier mobilities up to 0.1 cm 2 /Vs or even more are realistic, which would make<br />

them comparable to single crystalline systems [65].<br />

References<br />

1. C. F. Carlson: US Pat. 2297691 (1942)<br />

2. D. Haarer: Angew. Makromol. Chem., 183, 197 (1990)<br />

3. D.M. Pai, B.E. Springett: Rev. Mod. Phys., 65(1), 163 (1993)<br />

4. K.-W. Klüpfel, M. Tomanck, F. Endermann: German Patent No. DE-B-11 45184 (1963)<br />

5. H. Hoegl, O. Süs, W. Neugebauer, Kalle AG: German Patent No. DBP 1068115. Chem. Abstr., 55,<br />

20742 a (1961)<br />

6. H. Hoegl: J. Phys. Chem., 69(3), 755 (1965)<br />

7. G. Pfister: Phys. Rev. Lett., 36(5), 271 (1976)<br />

8. D. Adam, F. Closs, T. Frey, D. Funhoff, D. Haarer, H. Ringsdorf, P. Schuhmacher, K. Siemensmeyer:<br />

Phys. Rev. Lett., 70(4), 457 (1993)<br />

9. D. Adam, P. Schuhmacher, J. Simmerer, L. Häussling, K. Siemensmeyer, K.H. Etzbach, H. Ringsdorf,<br />

D. Haarer: Nature, 371, 141 (1994)<br />

10. H. Tung Li, P.J. Regensburger: J. Appl. Phys., 34, 1730 (1963)<br />

11. L. Onsager: Phys. Rev. B, 54, 554 (1938)<br />

12. H. Kaul, D. Haarer: Ber. Bunsenges. Phys. Chem., 91, 845 (1987)<br />

13. M. Gailberger, H. Bässler: Phys. Rev. B, 44(16), 8643 (1991)<br />

14. Y. Kanemitsu, Y. Sugimoto: Phys. Rev. B, 16(21), 14 182 (1992)<br />

15. Y. Sano, K. Kato, M. Yokoyama, Y. Shirota: H. Mikawa, Mol. Cryst. Liq. Cryst. 36(1–2), 137<br />

(1976)<br />

16. A. Peled, L.B. Schein, D. Glatz: Phys. Rev. B, 41(15), 10835 (1990)<br />

17. Y. Kanemitsu, H. Funada, Y. Masumoto: J. Appl. Phys., 71(1), 300 (1992)<br />

18. H. Bässler, P.M. Borsenberger: Chem. Phys., 177(3), 763 (1993)<br />

19. P.N. S<strong>and</strong>a, T. Takamori, D.B. Dove: J. Appl. Phys., 64(3), 1229 (1988)<br />

20. J.X. Mack, L.B. Schein, A. Peled: Phys. Rev. B, 39, 7500 (1989)<br />

21. Y. Kanemitsu, H. Funada, S. Imamura: J. Appl. Phys., 67(9), 4152 (1990)<br />

22. P.M. Borsenberger, L.T. Pautmeier, H. Bässler: Phys. Rev. B, 46(19), 12145 (1992)<br />

23. J. Mort, G. Pfister, S. Grammatica: Solid State Commun., 18, 693 (1976)<br />

24. P.M. Borsenberger, L.E. Contois, A.I. Ateya: J. Appl. Phys., 50(2), 914 (1979)<br />

25. S.J. Santos Lemus, J. Hirsch: Philos. Mag. B, 53(1), 25 (1986)<br />

26. D. Haarer, H. Meyer, P. Strohriegl, D. Naegele: Makromol. Chem., 192(3), 617 (1991)<br />

27. D.M. Pai, J.F. Yanus, M. Stolka, D. Renfer, W.W. Limburg: Philos. Mag., 48(6), 505 (1983)<br />

28. H. Bässler: Philos. Mag., 50(3), 347 (1984)<br />

29. G. Pfister: Phys. Rev. B, 16(8), 3676 (1977)<br />

13


1 Model Systems for Photoconductive Materials<br />

30. P.M. Borsenberger: J. Appl. Phys., 68(12), 6263 (1990)<br />

31. P.M. Borsenberger L. Pautmeier, R. Richert, H. Bässler: J. Chem. Phys., 94(12), 8276 (1991)<br />

32. J.M. Pearson, M. Stolka: Polymer Monographs, Vol. 6: Poly(N-vinylcarbazole), Gordon <strong>and</strong> Breach<br />

Sci. Publ., New York (1981)<br />

33. W.D. Gill: J. Appl. Phys., 43, 5033 (1972)<br />

34. E. Müller-Horsche, D. Haarer, H. Scher: Phys. Rev. B, 35, 1273 (1987)<br />

35. F.C. Bos, D.M. Burl<strong>and</strong>: Phys. Rev. Lett., 58, 152 (1987)<br />

36. R.G. Kepler, J.M. Zeigler, L.A. Harrah, S.R. Kurtz: Phys. Rev. B, 35(6), 2818 (1987)<br />

37. J. Frenekl: Phys. Rev., 54, 647 (1938)<br />

38. K. Shimakawa, T. Okada, O. Imagawa: J. Non-Cryst. Solids, 114(1), 345 (1989)<br />

39. M. Abkowitz, M. Stolka: Solid State Commun., 78(4), 269 (1991)<br />

40. H. Kaul: PhD thesis, Universität Bayreuth (1991)<br />

41. E. Brynda, S. Nespurek, W. Schnabel: Chem. Phys., 175(2–3), 459 (1993)<br />

42. V. Cimrova, S. Nespurek, R. Kuzel, W. Schnabel: Synth. Met., 67(1–3), 103 (1994)<br />

43. A.J. Heeger, S. Kivelson, J.R. Schrieffer, W.-P. Su: Rev. Mod. Phys., 60(3), 781 (1988)<br />

44. J. L. Brédas, R. Silbey (eds.): Conjugated Polymers, Kluwer Academic Publishers, Dordrecht<br />

(1991)<br />

45. J.H. Burroughes, D.D.C. Bradley, A.R. Brown, R.N. Marks, K. Mackay, R.H. Friend, P.L. Burns,<br />

A.B. Holmes: Nature, 347(6293), 539 (1990)<br />

46. F.W. Schmidlin: Phys. Rev. B, 16(6), 2362 (1977)<br />

47. H. Scher, E.W. Montroll: Phys. Rev. B, 12(6), 2455 (1975)<br />

48. J. Nool<strong>and</strong>i: Phys. Rev. B, 16(10), 4474 (1977)<br />

49. G. Pfister, H. Scher, Adv. in Phys, 27(5), 747 (1978)<br />

50. P.M. Borsenberger, L. Pautmeier, H. Bässler: J. Chem. Phys., 94(8), 5447 (1991)<br />

51. G. Schönherr, H. Bässler, M. Silver: Philos. Mag. B, 44(3), 369 (1981)<br />

52. H. Bässler, G. Schönherr, M. Abkowitz, D.M. Pai: Phys. Rev. B, 26(6), 3105 (1982)<br />

53. L. Pautmeier, R. Richert, H. Bässler: Philos. Mag. Lett., 59(6), 325 (1989)<br />

54. L. Pautmeier, R. Richert, H. Bässler: Synth. Met., 37, 271 (1990)<br />

55. H. Domes, R. Leyrer, D. Haarer, A. Blumen: Phys. Rev. B, 36(8), 4522 (1987)<br />

56. D. Stauffer: Introduction to Percolation Theory, Taylor <strong>and</strong> Francis, London (1985)<br />

57. H. Domes: PhD thesis, Universität Bayreuth (1988)<br />

58. H. Domes, R. Fischer, D. Haarer, R. Strohriegl: Makromol. Chem., 190, 165 (1989)<br />

59. H. Schnörer, H. Domes, A. Blumen, D. Haarer: Philos. Mag. Lett., 58(2), 101 (1988)<br />

60. H. Kaul: PhD thesis, Universität Bayreuth (1991)<br />

61. H. Suzuki, H. Meyer, J. Simmerer, J. Yang, D. Haarer: Adv. Mater., 5, 743 (1993)<br />

62. H. Suzuki, H. Meyer, S. Hoshino, D. Haarer: J. Appl. Phys. in press (1995)<br />

63. J. Kido, K. Nagai, Y. Okamoto, T. Skotheim: Appl. Phys. Lett., 59, 2760 (1991)<br />

64. C. Beginn, J.V. Grazulevicius, P. Strohriegl,J. Simmerer, D. Haarer: Macromol. Chem. Phys., 195,<br />

2353 (1994)<br />

65. N. Karl: in: K. Sumino (ed.): Defect Control in Semiconductors, Elsevier Science Publishers,<br />

North Holl<strong>and</strong> (1990)<br />

66. P. Gomes da Costa, R.G. D<strong>and</strong>rea, E.M. Conwell: Phys. Rev. B, 47(4), 1800 (1993)<br />

67. G. Leising, K. Pichler, F. Stelzer: in: H. Kuzmany, M. Mehring, S. Roth, (eds.): Springer Series in<br />

Solid-State Sciences, Vol. 91: Electronic <strong>Properties</strong> of Conjugated Polymers III, Springer, Berlin<br />

(1989)<br />

68. H. Meyer: PhD thesis, Universität Bayreuth (1994)<br />

69. H. Meyer, D. Haarer, H. Naarmann, H.H. Hörhold: Phys. Rev. B, in press (1995)<br />

70. M. Onoda, D.H. Park, K. Yoshino: J. Phys. (London), Condens. Matter, 1(1), 113 (1989)<br />

14


2 Novel Photoconductive Polymers<br />

Jörg Bettenhausen <strong>and</strong> Peter Strohriegl<br />

2.1 Introduction<br />

Electrophotography is the only area in which the conductivity of sophisticated organic materials<br />

<strong>and</strong> polymers is exploited in a large scale industrial process today. Photoconductors are<br />

characterized by an increase of electrical conductivity upon irradiation. According to this definition<br />

photoconductive materials are insulators in the dark <strong>and</strong> become semiconductors if<br />

illuminated. In contrast to electrically conductive compounds photoconductors do not contain<br />

free carriers of charge. In photoconductors these carriers are generated by the action of<br />

light.<br />

The discovery of photoconductivity dates back to 1873 when W. Smith found the effect<br />

in selenium. Based on this discovery C. F. Carlson developed the principles of the xerographic<br />

process already in 1938.<br />

Photoconductivity in organic polymers was first discovered in 1957 by H. Hoegl, who<br />

found that poly(N-vinylcarbazole) (PVK) <strong>and</strong> charge transfer complexes of PVK with electron<br />

acceptors like 2,4,7-trinitrofluorenone act as photoconductors [1].<br />

Besides the application of photoconductive polymers in photocopiers these materials<br />

are also widely used in laser printers in the last years. The third area in which photoconductors<br />

are applied is the manufacturing of electrophotographic printing plates.<br />

The organic photoconductors used in practice are based on two types of systems. The<br />

first one are polymers in which the photoconductive moiety is part of the polymer, for example<br />

a pendant or in-chain group. The second group involves low molecular weight compounds<br />

imbedded in a polymer matrix. These so-called moleculary doped polymers are<br />

widely used today.<br />

One interesting class of photoconductive materials are oxadiazoles. It is known for a<br />

long time that derivatives of 1,3,4-oxadiazole are good photoconductors. Compound 1 for<br />

example is described in patents <strong>and</strong> was frequently applied in photocopiers [2].<br />

1 2<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

15


2 Novel Photoconductive Polymers<br />

The oxadiazole 1 is a hole transport material since the electron withdrawing effect of<br />

the oxadiazole group with three electronegative heteroatoms is overcompensated by two<br />

electron donating amino groups.<br />

Within the last years, oxadiazoles like 2-(biphenyl)-5-(4-tert.butylphenyl)-1,3,4-<br />

oxadiazole (PBD) 2 have been frequently applied in organic light emitting diodes [3]. Here<br />

the electron withdrawing oxadiazole unit dominates the electronic properties <strong>and</strong> the oxadiazole<br />

compounds act as electron injection <strong>and</strong> transport layers. Furthermore, 2,5-diphenyloxadiazoles<br />

have been used as building blocks in thermostable polymers <strong>and</strong> they are highly<br />

fluorescent as well.<br />

In this paper the synthesis <strong>and</strong> characterization of a number of novel low molecular<br />

weight oxadiazole derivatives <strong>and</strong> polymers is described. The compounds show different molecular<br />

shapes, e. g. rod-like <strong>and</strong> star-shaped structures have been realized.<br />

In addition to the compounds described here, a series of different main chain <strong>and</strong> side<br />

group polymers with oxadiazole moieties are presently synthesized in our research group [4].<br />

2.2 Liquid crystalline oxadiazoles <strong>and</strong> thiadiazoles<br />

2.2.1 The basic idea<br />

The aim of this work is the preparation of photoconductive compounds with high carrier mobilities.<br />

There are a number of indications that the carrier mobility, i. e. the speed of the<br />

charge particles within the sample, depends on the geometric arrangement of the photoconductive<br />

moieties.<br />

At room temperature for instance, the mobility in single crystals of aromatic compounds<br />

like anthracene or perylene is very high, i. e. in the range of 10 –1 cm 2 /Vs [5]. In<br />

amorphous polymers the carrier mobilities are orders of magnitude smaller <strong>and</strong> typical values<br />

are in the range of 10 –6 <strong>and</strong> 10 –8 cm 2 /Vs [6].<br />

So we were interested to investigate if the higher order in the liquid crystalline state<br />

leads to higher mobilities in comparison to less ordered amorphous solids. In liquid crystalline<br />

polymers a macroscopic orientation of the photoconductive groups in the mesophase<br />

can be achieved by means of electric or magnetic fields. The orientation can be frozen in by<br />

lowering the temperature below the glass transition. By this, materials with a degree of order<br />

between a perfect single crystal <strong>and</strong> totally disordered amorphous polymers are obtained.<br />

With this in mind we started the synthesis of a series of liquid crystalline model compounds<br />

<strong>and</strong> polymers in which the side groups possess both mesogenic <strong>and</strong> photoconductive<br />

properties.<br />

16


2.2 Liquid crystalline oxadiazoles <strong>and</strong> thiadiazoles<br />

2.2.2 Monomer synthesis<br />

In the past years we have synthesized a variety of compounds with oxadiazole <strong>and</strong> thiadiazole<br />

moieties [7]:<br />

3 X=S 4 X=O<br />

The rod-like thiadiazole derivative 3 is liquid crystalline <strong>and</strong> shows a smectic as well<br />

as a nematic LC phase with the following phase behaviour: c 103 s a 180 n 203 i. An acrylate<br />

monomer was prepared by hydroboration of the terminal double bond <strong>and</strong> subsequent<br />

esterification with acryloyl chloride <strong>and</strong> then polymerized. The resulting polymer 5 exhibits<br />

a mesophase which has not yet been identified.<br />

5<br />

Unfortunately the clearing point of the polymer is at 246 8C, where the material starts<br />

to decompose. Therefore orientation in an electric field was not possible.<br />

In contrast to the thiadiazole the oxadiazole 4 is neither liquid crystalline as monomer<br />

nor as polymer. The reason for the lack of mesogenic properties is that the substitution of<br />

the sulphur by oxygen introduces a bend into the molecule which prevents the formation of<br />

a LC phase.<br />

Recently liquid crystalline oxadiazoles (Scheme 2.1) have been described [8]. Here<br />

the oxadiazole is coupled to at least two benzene or cyclohexane rings. Thereby an increase<br />

of the mesophase stability is achieved.<br />

Scheme 2.1: Liquid crystalline oxadiazole compounds [8].<br />

17


2 Novel Photoconductive Polymers<br />

6 7<br />

8<br />

9<br />

10-13<br />

10 R=Ph-OC 6<br />

H 13<br />

12 R = Ph-N(CH 3<br />

) 2<br />

11 R=C 7<br />

H 15<br />

13 R = Ph-N(CH 3<br />

)C 6<br />

H 13<br />

Scheme 2.2: Monomer synthesis.<br />

Our aim was to synthesize a series of different oxadiazole monomers which are functionalized<br />

for the preparation of polymers. Scheme 2.2 shows the synthesis of the monomeric<br />

1,3,4-oxadiazoles. The oxadiazole moiety is formed by a cyclisation reaction of the<br />

diacylhydrazine derivatives 8 with phosphorous oxychloride. The unsymmetrical bishydrazides<br />

8 are prepared by treatment of benzoyl chlorides 6 with the appropriate monohydrazides<br />

7. The last reaction step is an esterification with 4-(5-hexenyloxy)benzoyl chloride.<br />

The esterification reaction is essential in this case because of two reasons. First the rod-like<br />

18


2.2 Liquid crystalline oxadiazoles <strong>and</strong> thiadiazoles<br />

mesogen is formed, <strong>and</strong> second a functional group for the preparation of LC polymers is introduced.<br />

Additionally the oxadiazole monomer 14 was synthesized, in which the ester group is<br />

replaced by a biphenyl unit. By this it is possible to investigate the influence of the ester<br />

function on the photoconductive properties of the liquid crystalline compounds.<br />

14<br />

The liquid crystalline behaviour <strong>and</strong> the phase transitions of the monomeric oxadiazole<br />

derivatives 10–14 were determined by DSC <strong>and</strong> polarizing microscopy. In Tab. 2.1 the<br />

phase transition temperatures are summarized:<br />

Table 2.1: Phase behaviour of the monomeric oxadiazoles 10–14<br />

Compound<br />

Transition temperature in 8C<br />

10 k 141 n 167 i<br />

11 k 90 s a 116 i<br />

12 k 162 n 179 i<br />

13 k 127 (s a 111) i<br />

14 k 72 s a 77 i<br />

Except compound 13 all monomers are enantiotropic liquid crystalline <strong>and</strong> show nematic<br />

or smectic mesophases. Only derivative 13 shows a monotropic s a phase.<br />

2.2.3 Oligo <strong>and</strong> polysiloxanes with pendant oxadiazole groups<br />

The next step was the synthesis of liquid crystalline polysiloxanes with oxadiazole groups in<br />

the mesogenic unit. The polymers were prepared by a polymeranalogous reaction of the<br />

monomers 10 <strong>and</strong> 11 with poly(hydrogenmethylsiloxane) 15.<br />

Both polymers are liquid crystalline. The transition into the isotropic phase takes place<br />

at about 200 8C <strong>and</strong> therefore at a much higher temperature than in the monomers. The high<br />

clearing temperature makes the orientation of the polymers difficult which is preferably carried<br />

out near the clearing point. Up to now it was not possible to prepare well-defined polymers<br />

with the monomers 12–14. A possible reason is the inactivation of the Pt catalyst by<br />

the amino groups.<br />

If the cyclic tetrasiloxane 18, which is commercially available in high purity, is used<br />

instead of 15, well-defined monodisperse model compounds are obtained [9]. These com-<br />

19


2 Novel Photoconductive Polymers<br />

10, 11 15<br />

16 R=Ph-OC 6<br />

H 13<br />

17 R=C 7<br />

H 15<br />

Scheme 2.3: Synthesis of polysiloxanes with pendant oxadiazole groups.<br />

pounds can be highly purified, for example by preparative gel permeation chromatography.<br />

This is very important for photoconductors, because it is well-known that even traces of impurities<br />

may reduce the carrier mobility.<br />

18<br />

So the monomeric compounds 10 <strong>and</strong> 12–14 have been reacted with 18 to yield the<br />

following tetrameric derivatives:<br />

19 R=Ph-OC 6<br />

H 13<br />

,n=1 21 R = Ph-N(CH 3<br />

)C 6<br />

H 13<br />

,n=<br />

20 R = Ph-N(CH 3<br />

) 2<br />

,n=1 22 R = Ph-N(CH 3<br />

)C 6<br />

H 13<br />

,n=<br />

Scheme 2.4: Tetrasiloxanes with pendant oxadiazole groups.<br />

20


2.2 Liquid crystalline oxadiazoles <strong>and</strong> thiadiazoles<br />

All cyclosiloxanes are liquid crystalline <strong>and</strong> show several advantages compared to the<br />

polymers. So the cyclic siloxanes form stable glasses but their clearing points are much<br />

lower. For example compounds 21 <strong>and</strong> 22 possess clearing temperatures at only 102 8C.<br />

The phase behaviour of the tetramers are listed in Tab. 2.2.<br />

Table 2.2: Phase behaviour of the tetrameric siloxanes 19–22.<br />

Comp.<br />

Transition temperature in 8C<br />

19 g 54<br />

a)<br />

m 1 108<br />

a)<br />

m 2 117 s c 155 n 185 i<br />

20 g 67 s a 138 n 181 i<br />

21 g 34 k 66 n 102 i<br />

22 g 38 s a 82 n 102 i<br />

a) Mesophase not identified<br />

2.2.4 Photoconductivity measurements<br />

For the characterization of the novel photoconductive compounds two different experimental<br />

techniques have been used. Some measurements were made by BASF AG using a steady-state<br />

method. These experiments allow a quick characterization, but in contrast to the time-of-flight<br />

(TOF) method no statements about transient photocurrents <strong>and</strong> carrier mobilities are possible.<br />

The TOF measurements were carried out in the group of D. Haarer, Universität Bayreuth.<br />

The tetrameric cyclosiloxanes are very suitable compounds for physical investigations.<br />

Because of the low viscosity the preparation of samples is much easier than in the case of<br />

the polymers. Additionally the dark currents of the tetramers are very low.<br />

All the compounds are photoconductive, as shown by the steady-state method. For example,<br />

the measurements of the thiadiazole 3 showed a distinct rise of the photocurrent at<br />

the transition from the crystalline to the mesophase, as shown in Fig. 2.1. The decrease of<br />

Figure 2.1: Temperature dependence of the photocurrent of the thiadiazole 3.<br />

21


2 Novel Photoconductive Polymers<br />

the current at the transition from the smectic to the nematic phase is directly correlated with<br />

the loss of order. These results show that the basic idea of this work is correct.<br />

Nevertheless, we were not able to carry out TOF measurements with calamitic monomers,<br />

because the dark currents in the liquid crystalline phases were too high. As mentioned<br />

before, such measurements are possible with the cyclic tetramers. The transient photocurrents<br />

of the tetrasiloxanes 19, 21, 22 illustrate that the carrier transport is totally dispersive,<br />

i. e. dominated by deep traps in which the charge carriers are captured. This is a typical behaviour<br />

of amorphous polymers. No transit time could be detected <strong>and</strong> no statements about<br />

the carrier mobility can be made for the rod-like mesogens.<br />

In contrast it was shown in the past years that discotic liquid crystals like hexapentyloxytriphenylene<br />

(HPT) 23 have carrier mobilities up to 10 –3 cm 2 /Vs in the liquid crystalline<br />

D ho phase [10]. Here, the disc-shaped molecules are ideally stacked above each other<br />

<strong>and</strong> therewith allow a fast carrier transport. The discotic hexathioether 24 with a highly ordered<br />

helical columnar (H) phase exhibits mobilities up to 10 –1 cm 2 /Vs, which are almost as<br />

high as in organic single crystals [11].<br />

6<br />

6<br />

23 24<br />

One important observation during the photoconductivity measurements of the tetramer<br />

19 was, that it showed almost the same photocurrent for holes <strong>and</strong> electrons. This was interesting<br />

because of the lack of electron transporting materials <strong>and</strong> led us synthesize a variety<br />

of starburst oxadiazole compounds, which are described in the next Chapter.<br />

2.3 Starburst oxadiazole compounds<br />

2.3.1 Motivation<br />

The synthesis of novel materials with high carrier mobilities is one of the major goals in the<br />

field of photoconductive polymers. Within the last years different approaches have been pursued<br />

to reach this goal. First, the photoconductive properties of conjugated polymers like<br />

poly(phenylenevinylene) <strong>and</strong> poly(methyl phenylsilane) have been investigated [12]. Another<br />

approach are liquid crystals which are the topic of the first Chapter. The third way to realize<br />

22


2.3 Starburst oxadiazole compounds<br />

the goal are glasses of large extended aromatic amines. The best investigated representatives<br />

of this class of compounds are N,N'-diphenyl-N,N'-bis(3-methylphenyl)-[1,1'-biphenyl]-4,4'-<br />

diamine (TPD) 25 from Xerox [13] <strong>and</strong> 1,1-bis(di-4-tolyl-aminophenyl)cyclohexane (TAPC)<br />

26 from Kodak [14].<br />

Both compounds are derivatives of triphenylamine, a well-known photoconductor. If<br />

thin films of TPD <strong>and</strong> TAPC are prepared by vacuum evaporation both compounds form<br />

metastable glasses. In such glasses carrier mobilities up to 10 –3 cm 2 /Vs for TPD [13] <strong>and</strong><br />

10 –2 cm 2 /Vs for TAPC [14] have been reported. But both TAPC <strong>and</strong> TPD glasses are metastable<br />

<strong>and</strong> have a strong tendency to crystallize. If the molecules are imbedded in a polymer<br />

matrix, e. g. polycarbonate or polystyrene, morphologically stable materials are formed, but<br />

the mobilities decrease drastically [13].<br />

25 26<br />

Several attempts have been made to overcome the problems with metastable TPD <strong>and</strong><br />

TAPC glasses. Recently we described the synthesis of 3,6-bis[(9-hexyl-3-carbazolyl)ethynyl]-9-hexylcarbazole<br />

27, a trimeric model compound of poly(carbazolylene ethynylene)<br />

[15]. This material shows a glass transition temperature of 41 8C <strong>and</strong> forms glasses which<br />

are stable for more than a year. Mobilities up to 10 –4 cm 2 /Vs (E =67 10 5 V/cm, T =308C)<br />

have been measured by the time-of-flight technique.<br />

27<br />

Another interesting approach are starburst compounds with high glass transition temperatures.<br />

So 4,4',4@-tris(N-carbazolyl)triphenylamine, which has been published recently<br />

[16], shows a glass transition at 151 8C <strong>and</strong> forms a stable glass.<br />

Beginning with the work of Thomalia [17] in 1986, starburst molecules (dendrimers)<br />

have achieved enormous interest within the last years. Dendrimers are highly branched regular<br />

molecules, which are usually prepared by stepwise reactions. In many cases the behaviour<br />

of dendritic macromolecules are different from that of linear polymers, e. g. the former<br />

23


2 Novel Photoconductive Polymers<br />

usually show enhanced solubility. The reasons for these differences are the unique three-dimensional<br />

structure <strong>and</strong> the large number of chain ends in dendrimers.<br />

Two different methods have been developed for the stepwise synthesis of starburst<br />

dendrimers:<br />

a) the divergent approach, where the synthesis starts from a core molecule with two or<br />

more reactive groups;<br />

b) the convergent approach, in which the synthesis starts at the outer sphere of the dendrimer.<br />

In the divergent approach, the reaction of the core molecule with two or more reagents<br />

containing at least two protected branching sites is followed by removal of the protecting<br />

groups <strong>and</strong> subsequent reaction of the liberated reactive groups which leads to the starburst<br />

molecule of the <strong>1st</strong> generation. The process is repeated until the desired size is reached. In<br />

the convergent approach the synthesis starts at what will become the outer surface of the<br />

dendrimer. Step by step large dendrimer arms are prepared <strong>and</strong> finally the completed arms<br />

are coupled to the core. Both methods produce well-defined large dendritic molecules whose<br />

structures are manifolds of the building blocks. They allow structural as well as functional<br />

group variation.<br />

On the other h<strong>and</strong>, hyperbranched polymers can be synthesized in a one-step reaction<br />

using highly functionalized monomers of the type A x B, where x is 2 or larger. This method<br />

does not yield polymers of such a well-defined structure, but has the advantage to provide<br />

rapidly large quantities of material. The control of the degree of branching is difficult <strong>and</strong><br />

mainly depends on statistics, steric effects, <strong>and</strong> the reactivity of the functional groups.<br />

The aim of our work is the synthesis of starburst compounds with oxadiazole moieties.<br />

In contrast to the triphenylamine <strong>and</strong> carbazole derivatives oxadiazoles are strong electron<br />

acceptors. Therefore their transport characteristics can be switched from hole transport materials<br />

like 2,5-(4-diethylaminophenyl)-oxadiazole 1, in which the electron withdrawing effect<br />

of the oxadiazole ring is overcompensated by the two electron donating amino groups,<br />

to electron transport materials like 2-(biphenyl)-5-(4-tert.butylphenyl)-oxadiazole (PBD) 2.<br />

Recently, Saito demonstrated that oxadiazoles in a polycarbonate matrix show electron transport<br />

with mobilities up to 10 –5 cm 2 /Vs [18]. Such electron transport materials are attractive<br />

for the use in copiers <strong>and</strong> in organic light-emitting diodes.<br />

2.3.2 Synthesis of starburst oxadiazole compounds<br />

Starburst oxadiazole compounds have been mentioned for the first time in a thermodynamic<br />

study of the structure-property-relationship in low molecular weight organic glasses [19,<br />

20]. Upon rapid cooling these compounds form glasses with T g s between 77 <strong>and</strong> 169 8C.<br />

We have used three different approaches for the preparation of starburst oxadiazole<br />

compounds, which are schematically shown in Fig. 2.2.<br />

The methods A <strong>and</strong> B can be compared to the divergent synthesis of dendrimers.<br />

Method A starts from a central core <strong>and</strong> involves the reaction of the acid chloride groups<br />

<strong>and</strong> a subsequent cyclisation with phosphorous oxychloride to the oxadiazole ring (Chap-<br />

24


O<br />

O<br />

Cl<br />

C<br />

C<br />

Cl<br />

O<br />

C<br />

Cl<br />

Method A<br />

Method B<br />

O<br />

H 2N NH C<br />

O<br />

O<br />

O<br />

O<br />

C<br />

NH<br />

NH<br />

C<br />

C<br />

NH NH C<br />

O<br />

O<br />

C<br />

NH<br />

NH<br />

C<br />

C<br />

N N<br />

N NH<br />

N<br />

N<br />

O<br />

N N O<br />

O<br />

N<br />

N<br />

=core<br />

= shell<br />

Method C<br />

X<br />

Z<br />

+<br />

Y<br />

X<br />

X<br />

Z<br />

Z<br />

X = OH, C CH Y = Br, F Z = O , C C<br />

=core<br />

=shell<br />

Figure 2.2: The different approaches to starburst oxadiazoles.<br />

25


2 Novel Photoconductive Polymers<br />

ter 1). The main difficulty in this case is that three functional groups must react in one step.<br />

Partially incomplete cyclisation causes problems, because the products contain one or two<br />

bishydrazides <strong>and</strong> are difficult to separate from the target compounds, which have three oxadiazole<br />

groups. Therefore we recently developed a totally different route (method B). The<br />

compounds 31 a–f were prepared by the reaction of benzene tricarbonyl chloride 30 with<br />

the tetrazoles 29. The latter were synthesized from benzonitriles 28 with potassium azide<br />

(Scheme 2.5). With the loss of nitrogen the tetrazole ring is transferred to the oxadiazoles<br />

[21]. This gives an easy access to unsymmetrically substituted oxadiazoles which we used<br />

for the first time in the synthesis of starburst oxadiazole compounds.<br />

In the third approach (method C) which is similar to the convergent synthesis a preformed<br />

oxadiazole precursor is prepared by stepwise synthesis <strong>and</strong> finally coupled to the<br />

core molecule.<br />

R 1<br />

R 2 CN + NaN 3<br />

NH 4 Cl<br />

R 1<br />

R 2<br />

R 3<br />

R 3<br />

28 29<br />

C<br />

N<br />

N<br />

NH<br />

N<br />

Cl<br />

O<br />

C<br />

O<br />

C<br />

Cl<br />

O<br />

C<br />

Cl<br />

+ 3<br />

R 2<br />

R 1<br />

R 3<br />

C<br />

N<br />

N<br />

NH<br />

N<br />

31a-f<br />

pyridine<br />

2h, reflux<br />

31a-f<br />

30 29<br />

Scheme 2.5: Synthesis of starburst oxadiazoles via tetrazole intermediates.<br />

For the preparation of the novel starburst molecules, different core molecules have<br />

been used: 1,3,5-benzene tricarboxylic acid chloride, 1,3,5-tris(4-benzoyl)benzene, 1,3,5-<br />

triethynylbenzene, <strong>and</strong> 4,4',4@trihydroxytriphenylamine.<br />

Except benzenetricarboxylic acid which is commercially available, all core molecules<br />

were synthesized following well-known literature procedures [22, 23].<br />

The structures of the starburst oxadiazole compounds are summarized in Tab. 2.3.<br />

31 c <strong>and</strong> 31g have been described before [19, 20], but no synthetisis procedure was given.<br />

The oxadiazoles 31a–h have been prepared according to method B (Fig. 2.3), starting<br />

from benzene tricarboxylic acid chloride.<br />

Compound 31g with tert.-butyl substituents was synthesized by method A, too. So a<br />

comparison of methods A <strong>and</strong> B is possible. The tetrazole route (method B) exhibits several<br />

advantages compared to the ring closure with dehydrating agents (method A). One advantage<br />

is the short reaction time. The reaction is finished within 2 hours whereas heating for<br />

1–3 days is necessary if the ring closure is carried out with phosphorous oxychloride. Even<br />

more important is the facile work up procedure. In the case of the dehydration with POCl 3<br />

column chromatography <strong>and</strong> subsequent recrystallization is necessary to purify the product<br />

which is finally obtained in 33% yield. In contrast only one filtration on a short silica gel<br />

26


2.3 Starburst oxadiazole compounds<br />

Table 2.3: Structures of the starburst oxadiazole compounds 31–34.<br />

Comp. Core a) Shell a)<br />

31<br />

N<br />

O<br />

N<br />

R 1<br />

R 3<br />

R 2<br />

31a R 1 =R 3 =H,R 2 =CH 3<br />

31b R 2 =R 3 =CH 3 ,R 1 =H<br />

31c R 1 =R 3 =CF 3 ,R 2 =H<br />

31d R 1 =R 3 =H,R 2 =C 2 H 5<br />

31e R 1 =R 3 =H,R 2 =OC 2 H 5<br />

31f R 1 =R 3 =H,R 2 = CH(CH 3 ) 2<br />

31g R 1 =R 3 =H,R 2 = C(CH 3 ) 3<br />

31h R 1 =R 3 =H,R 2 = N(C 2 H 5 ) 2<br />

N N<br />

32 R<br />

R = C(CH 3 ) 3<br />

O<br />

C<br />

C<br />

C C<br />

N N<br />

33 R<br />

R = C(CH 3 ) 3<br />

C<br />

C<br />

O<br />

N N<br />

34 N<br />

O<br />

R R = C(CH 3 ) 3<br />

O<br />

a) cf. Figure 2.2.<br />

column is sufficient for the purification of 31 g, prepared by the tetrazole route. In this case<br />

the yield of the pure product is 69%. The starburst oxadiazole 32 has been prepared by<br />

method B too, whereas for 33 <strong>and</strong> 34 method C (Fig. 2.2) was used. The identity of all compounds<br />

was confirmed by NMR <strong>and</strong> FTIR spectroscopy.<br />

2.3.3 Thermal properties<br />

The thermal behaviour of the starburst molecules has been investigated by differential scanning<br />

calorimetry <strong>and</strong> thermogravimetric measurements.<br />

27


2 Novel Photoconductive Polymers<br />

Table 2.4: Thermal properties of the starburst compounds 31–34.<br />

a)<br />

a)<br />

Compound Molecular mass T g T m<br />

[g/mol] [8C] [8C] [8C]<br />

31a 553 – 334 348<br />

31b 595 – 333 358<br />

31c 919 – 250 336<br />

31d 595 – 276 360<br />

31e 643 108 259 337<br />

31f 637 97 225 349<br />

31g 679 142 257 366<br />

31h 724 128 299 329<br />

32 907 165 297 409<br />

33 979 – 320 356<br />

34 1122 137 – 391<br />

a) 3rd heating, determined by DSC with 20 K/min<br />

b) onset of decomposition in nitrogen, thermogravimetric measurement with 10 K/min<br />

T dec<br />

b)<br />

Among the compounds 31 a–h with a small benzene core both crystalline <strong>and</strong> glass<br />

forming materials exist. So 31 a–d with small methyl, ethyl, or trifluoromethyl substituents<br />

show only melting points in the DSC experiment with cooling rates of 20 K/min. Naito <strong>and</strong><br />

Miura reported that it is possible to obtain glasses even with small methyl or trifluoromethyl<br />

substituents if the compounds are rapidly cooled with liquid nitrogen [19]. In contrast, the<br />

compounds 31 e with ethoxy, 31f with iso-propyl, 31g with tert-butyl substituents, <strong>and</strong> 31h<br />

with diethylamino groups form glasses upon cooling with 20 K/min in the DSC equipment.<br />

When the amorphous samples are heated again the glass transition appears first but on<br />

further heating the samples start to recrystallize <strong>and</strong> consequently show a melting point at<br />

higher temperatures. Compound 32 with the triphenylbenzene core behaves similar like<br />

31e–h. The most stable glasses are formed by 34 with the triphenylamine core. The DSC<br />

diagram of the novel glass forming oxadiazole compound is shown below. Upon both, heating<br />

<strong>and</strong> cooling, only a glass transition is observed at 137 8C (Fig. 2.3). In our DSC experi-<br />

Figure 2.3: DSC scan of the starburst oxadiazole compound 34, 3rd heating <strong>and</strong> cooling, with 20 K/min.<br />

28


2.3 Starburst oxadiazole compounds<br />

ments we never found any evidence for crystallization of 34. Even in the first heating no<br />

melting point is observed up to 350 8C. Consequently, transparent amorphous films are obtained<br />

from the starburst oxadiazole compound 34 by solvent casting.<br />

The thermal behavior of compound 33 is somewhat different. No reproducible DSC<br />

scans are obtained in subsequent heating-cooling cycles. We attribute this to thermal crosslinking<br />

of the triple bonds [24].<br />

The thermal stability has been monitored by thermogravimetric measurements. In<br />

most cases the onset of decomposition is in the range from 330–370 8C. The oxadiazole<br />

compound 32 with a triphenyl benzene core shows a somewhat higher thermal stability up<br />

to 410 8C.<br />

(CH 3 ) 3 C<br />

N<br />

O<br />

N<br />

O<br />

C(CH 3 ) 3<br />

N<br />

N O<br />

O<br />

N<br />

O<br />

N<br />

O<br />

N<br />

C(CH 3 ) 3<br />

34<br />

The starburst oxadiazole compounds are now being tested as electron injection <strong>and</strong><br />

transport layer in organic LEDs <strong>and</strong> as photoconductors. First tests of two layer LEDs with<br />

PPV show that the novel materials possess properties comparable to 2 but have the great advantage<br />

to show no recrystallization if thin films were made by spin-coating. We will report<br />

on these measurements in the near future.<br />

29


2 Novel Photoconductive Polymers<br />

References<br />

1. H. Hoegl, O. Süs, W. Neugebauer, Kalle AG: DBP 1068115, Chem. Abstr., 55, 20742 a (1961),<br />

H. Hoegl: J. Phys. Chem., 69, 755 (1965)<br />

2. W. Wiedemann: Chem.-Ztg., 106, 275 (1982) <strong>and</strong> references therein<br />

3. A.R. Brown, D.D.C. Bradley, J.H. Burroughes, R.H. Friend, N.C. Green-ham, P.L. Burn, A.B.<br />

Holmes, A. Kraft: Appl. Phys. Lett., 61, 2793 (1992)<br />

4. E. Buchwald, M. Meier, S. Karg, W. Rieß, M. Schwoerer, P. Pösch, H.-W. Schmidt, P. Strohriegl:<br />

Adv. Mater., 7, 839 (1995), M. Greczmiel, P. Pösch, H.-W. Schmidt, P. Strohriegl, E. Buchwald,<br />

M. Meier, W. Rieß, M. Schwoerer: Makromol. Symposia, 102, 371 (1996)<br />

5. W. Warta, R. Stehle, N. Karl: Appl. Phys., A36, 163 (1985)<br />

6. P.M. Borsenberger, D.S. Weiss: in: Organic Photoreceptors For Imaging Systems, M. Dekker, New<br />

York, (1993)<br />

7. R.G. Müller: PhD thesis, Bayreuth, (1992)<br />

8. D. Girdziunaite, C. Tschierske, E. Novotna, H. Kresse, A. Hetzheim: Liq. Cryst., 10, 397 (1991)<br />

9. F.H. Kreuzer, D. Andrejewski, W. Haas, N. Häberle, G. Riepl, P. Spes: Mol. Cryst. Liq. Cryst.,<br />

199, 345 (1991)<br />

10. D. Adam, F. Closs, T. Frey, D. Funhoff, D. Haarer, H. Ringsdorf, P. Schuhmacher, K. Siemensmeyer:<br />

Phys. Rev. Lett., 70, 457 (1993)<br />

11. D. Adam, P. Schuhmacher, J. Simmerer, L. Häussling, K. Siemensmeyer, K.H. Etzbach, H. Ringsdorf,<br />

D. Haarer: Nature, 371, 141 (1994)<br />

12. M. Gailberger, H. Bässler: Phys. Rev. B, 44, 8643 (1991), M.A. Abkowitz, M.J. Rice, M. Stolka:<br />

Phil. Mag. B, 61, 25 (1990)<br />

13. M. Stolka, J.F. Yanus, D.M. Pai: J. Phys. Chem., 88, 4707 (1984)<br />

14. P.M. Borsenberger, L. Pautmeier, R. Richert, H. Bässler: J. Phys. Chem., 94, 8276 (1991)<br />

15. C. Beginn, J.V. Grazulevicius, P. Strohriegl, J. Simmerer, D. Haarer: Macromol. Chem. Phys., 195,<br />

2353 (1994)<br />

16. Y. Kuwabara, H. Ogawa, H. Inoda, N. Noma, Y. Shirota: Adv. Mater., 6, 677 (1994)<br />

17. D.A. Tomalia, H. Baker, J. Dewald, M. Hall, G. Kallos, S. Martin, J. Roeck, J. Ryder, P. Smith:<br />

Macromolecules, 19, 2466 (1986)<br />

18. H. Tokuhisa, M. Era, T. Tsutsui, S. Saito: Appl. Phys. Lett., 66, 3433 (1995)<br />

19. K. Naito, A. Miura: J. Phys. Chem., 97, 6240,(1993)<br />

20. S. Egusa, Y. Watanabe: EP 553950 A2 to Toshiba Corp.<br />

21. R. Huisgen, J. Sauer, H.J. Sturm, J.H. Markgraf: Chem. Ber., 93, 2106 (1960)<br />

22. E. Weber, M. Hecker, E. Koepp, W. Orlia, M. Czugler, I. Csöregh: J. Chem. Soc., Perkin Trans. II,<br />

1251 (1988)<br />

23. S.J.G. Linkletter, G.A. Pearson, R.I. Walter: J. Am. Chem. Soc., 99, 5269 (1977)<br />

24. S.-C. Lin, E.M. Pearce: in: High-Performance Thermosets – Chemistry, <strong>Properties</strong>, Applications,<br />

Carl Hanser Verlag, Munich, 137 (1994)<br />

30


3 Theoretical Aspects of Anomalous Diffusion<br />

in Complex Systems<br />

Alex<strong>and</strong>er Blumen<br />

3.1 General aspects<br />

Many relaxation phenomena in the solid phase depend on diffusion, which is usually treated<br />

in the framework of statistical methods. Regular diffusion, known as Brownian motion, is<br />

characterized by a linear increase of the mean-squared displacement with time. On the other<br />

h<strong>and</strong>, for a whole series of phenomena this simple relation does not hold; their temporal<br />

evolution of the mean-squared displacement is non-linear <strong>and</strong> thus obeys at long times:<br />

<br />

r 2 <br />

…t† t<br />

<br />

…1†<br />

with g 0 1. Relation 1 is referred to as anomalous diffusion. In the case that g < 1 one denotes<br />

the behaviour as subdiffusive. A subdiffusive pattern of motion often results from disorder<br />

[1–4]. One has to note, however, that the asymptotic law, Eq. 1, emerges only when<br />

the disorder influences the motion on all scales. In the case that g > 1 the motion is termed<br />

superdiffusive. A classical example for superdiffusive behaviour is furnished by the motion<br />

of particles in a turbulent flow. In this paper we focus on several models for anomalous diffusion<br />

which involve polymeric systems.<br />

Now, the mean-squared displacement is a basic characteristic feature for the motion<br />

but, as an averaged quantity, it can provide only restricted information about the basic microscopic<br />

mechanisms involved. In several works we have also studied the propagator P(r,t),<br />

the probability to be at r at time t having started at the origin at t = 0. Space limitations prevent<br />

us from going into details. Here we focus on Ar 2 (t)S <strong>and</strong> refer to Klafter et al. <strong>and</strong> Zumofen<br />

et al. [4, 5] for in-depth analyses of the propagator.<br />

From the beginning, the role of time-dependent aspects in disordered media has to be<br />

emphasized. In usual r<strong>and</strong>om walk problems it is often assumed that the disorder is<br />

quenched, so that the dynamics evolves over a static substrate, i. e. the geometrical or the energetical<br />

disorder is frozen in. However, it is also possible that the dynamical processes are<br />

directly under the influence of temporal (possibly medium induced) fluctuations.<br />

First, in Section 3.2 we concentrate on photoconductivity, whose canonical description<br />

involves the continuous time r<strong>and</strong>om walk (CTRW) approach [1, 4, 6–8]. Basic ingredients<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

31


3 Theoretical Aspects of Anomalous Diffusion in Complex Systems<br />

in CTRWs are waiting time distributions with long time-tails. For photoconductivity one<br />

usually has g 1 shows up.<br />

An interesting dynamical feature is observed when the object to move has itself an internal<br />

structure; that is the case, for instance, for a polymer. In this case, because of the connections<br />

which exist between different segments of the polymer, the motion gets a very rich<br />

structure. The details of this situation will be discussed in Section 3.4.<br />

<strong>Final</strong>ly, we close in Section 3.5 with some conclusions.<br />

3.2 Photoconductivity<br />

A major field of application of scaling ideas in the time domain is photoconductivity [1–3,<br />

7, 12]; here the use of polymeric photoconductors in printing <strong>and</strong> copying represents one of<br />

the most sophisticated applications for organic materials. Such materials are nowadays<br />

superior to their inorganic counterparts. Now, the issue at stake in amorphous photoconductors<br />

is the appearance of dispersive transport, as contrasted to the familiar diffusive Gaussian<br />

behaviour. One observes experimentally that the motion of the charge carriers becomes<br />

slower <strong>and</strong> slower with the passage of time, a situation mirrored by Eq. 1 with g t T one finds<br />

I t 1 : …4†<br />

For instance measurements on polysiloxanes with pendant carbazole groups [13] follow<br />

Eqs. 3 <strong>and</strong> 4 with g = 0.58 very well over four decades in time, see below. In fact this is<br />

32


3.2 Photoconductivity<br />

typical for g significantly lower than one. However, one should note that the behaviour of<br />

I (t) is complex, since for g > 1 in the long-time limit the transport behaviour is non-dispersive.<br />

This is related to the existence of a finite mean waiting-time t = g/(g – 1). It follows<br />

that around g = 1 there is a crossover from dispersive to non-dispersive behaviour [13], as is<br />

also confirmed experimentally.<br />

From the preceding discussion it is obvious that the transport of charge carriers is<br />

much influenced by the waiting-time distributions (WTD) between hops. Physically the<br />

WTD arise from the local disorder in the sample. For large disorder the WTD are not exponential<br />

but decay much more slowly. In line with the preceding arguments one takes for the<br />

WTD expressions c (t) which behave at long times algebraically [7, 12]<br />

…t† t 1 ; …5†<br />

where 0 < g < 1. This choice for c (t) reproduces Eqs. 2–4, see below. Equation 5 is not valid<br />

near the time-origin. Therefore in calculations one prefers to work with functions welldefined<br />

for t 6 0. A suitable choice is for instance the series<br />

…t† ˆ1 a<br />

a<br />

X 1<br />

nˆ1<br />

a n b n exp … b n t† …6†<br />

with a < 1 <strong>and</strong> b < a –1 [8]. This function is everywhere continuous <strong>and</strong> finite, <strong>and</strong> for purely<br />

imaginary t it turns into the Weierstrass function. One verifies readily that for large t Eq. 6<br />

obeys Eq. 5 <strong>and</strong> that the corresponding g in Eq. 5 is g =lna/ln b. Equation 6 is very useful,<br />

since it allows to vary g freely by a judicious choice of a <strong>and</strong> b. Another choice for c (t)<br />

which is continuous for t 6 0is<br />

…t† ˆ…1 ‡ t† :<br />

…7†<br />

Here again the long-time behaviour follows Eq. 5.<br />

We now recall the basic ingredients of r<strong>and</strong>om walks in continuous time, the so-called<br />

CTRW [6–8]. Let c (r, t) be the probability distribution of making a step of length r in the<br />

time interval t to t +dt. The total transition probability in this time interval is<br />

…t† ˆX<br />

r<br />

…r;t†:<br />

…8†<br />

Furthermore the survival probability at the initial site is<br />

…t† ˆ1<br />

R t<br />

0<br />

…† d;<br />

…9†<br />

so that, switching to the Laplace space (t ? u), one has<br />

…u† ˆ‰1 …u†Š=u: …10†<br />

33


3 Theoretical Aspects of Anomalous Diffusion in Complex Systems<br />

The probability density Z(r,t) of just arriving at r in the time interval t to t +dt obeys<br />

the iterative relation<br />

…r;t†ˆP<br />

r 0<br />

R t<br />

0<br />

…r 0 ;† …r r 0 ;t † d ‡ …t† r;0 ; …11†<br />

in which the initial condition of starting at t = 0 from r = 0 is incorporated. One then has<br />

for the probability P (r,t), that the particle is at r at time t,<br />

P …r;t†ˆRt<br />

…r;t † …† d: …12†<br />

0<br />

Now P (r, t) also obeys an iterative relation:<br />

P …r;t†ˆX<br />

r 0<br />

R t<br />

0<br />

P …r 0 ;† …r r 0 ;t † d ‡ …t† r;0 …13†<br />

as may be seen either by inspection or, more formally, by using Eqs. 11 <strong>and</strong> 12. Clearly, a<br />

description of such convolutions, Eqs. 11–13, is more compact in Fourier-Laplace space.<br />

For P (k, u) one has from Eq. 13:<br />

P …k;u†ˆP …k;u† …k;u†‡ …u†<br />

…14†<br />

with the immediate solution<br />

P …k;u†ˆ …u†=‰ 1 …k;u† Š ˆ 1 …u†<br />

u<br />

1<br />

1 …k;u† : …15†<br />

The last expression generalizes the usual diffusion relation to r<strong>and</strong>om media, in which<br />

spatial <strong>and</strong> temporal aspects are coupled through c (k,u). The analysis is much simplified if<br />

such aspects decouple, which is the case for instance when the disorder is mainly energetic<br />

<strong>and</strong> the r<strong>and</strong>om walker moves over a rather regular lattice. In the decoupled case:<br />

…r;t†ˆ …r† …t† :<br />

…16†<br />

From this it follows immediately that also c (k,u) =l(k) c (u) is decoupled. In the decoupled<br />

scheme P (k, u) takes the form<br />

P …k;u†ˆ1 …u†<br />

u<br />

1<br />

1 …k† …u† ; …17†<br />

where l (k) = P qp(q)e –ik7q is the structure function of the infinite lattice <strong>and</strong> p (q) denotes<br />

the probability that a step extends over the distance q.<br />

34


3.2 Photoconductivity<br />

Coupling is very important for superlinear behaviour. Thus for the c (r,t) given by<br />

Eq. 16 the mean-squared displacement is either divergent or increases sublinearly or at most<br />

linearly in time. In order to obtain finite Ar 2 (t)S with a superlinear temporal behaviour,<br />

coupled c (r,t) forms have to be used [4]. An example is the WTD<br />

…r;t†ˆAr …r t †; …18†<br />

in which the d-function couples r <strong>and</strong> t. This WTD leads to Lévy walks.<br />

Let us now focus on the mean-squared displacement. Evidently, one has<br />

<br />

r 2 R<br />

…t† ˆ r 2 P …r;t† dr ˆ rk 2 P …k;t† j kˆ0<br />

…19†<br />

from which, in the decoupled scheme, using Eqs. 5 <strong>and</strong> 17 it follows that in the absence of<br />

any bias Eq. 2 is fulfilled. The procedure is more readily examplified by calculating the current<br />

from the mean displacement of the carrier Ar(t)S in a biasing field [8]. Setting L –1 for<br />

the inverse Laplace transform it follows:<br />

P<br />

r…t† ˆ rP …r;t†ˆir k P …k;t† kˆ0 ˆ L 1 … ir k P …k;u† j kˆ0 †<br />

r<br />

<br />

ˆ L 1<br />

@P…k;u†<br />

@<br />

<br />

ir k …k† j kˆ0<br />

ˆ1<br />

<br />

ˆ L 1 …u†<br />

u ‰ 1 …u† Š<br />

<br />

<br />

q ;<br />

where in the last line we work in the decoupled scheme <strong>and</strong> AqS = P q qp (q) is the mean displacement<br />

per hop. In the presence of a bias AqS 0 0. Thus, setting |AqS| = 1, we have for the<br />

current I (t) in an infinite lattice of any dimension,<br />

I …t† ˆ d <br />

r…t† <br />

<br />

dt ˆ L 1 …u†<br />

: …21†<br />

1 …u†<br />

For a finite chain of N sites Eq. 21 takes in the Laplace space the form [3, 13]:<br />

I…u† ˆXN<br />

nˆ1<br />

‰ …u† Š n ˆ<br />

…u†<br />

1 …u†<br />

…20†<br />

<br />

1 ‰ …u† Š N : …22†<br />

It is precisely this function which was used in Refs. [13, 14] to analyse, together with<br />

the WTD Eq. 6, the time-of-flight currents of photoconductive carriers in polysiloxanes with<br />

pendant carbazole groups. In this work one achieved with g = 0.58 a good agreement with<br />

the experimental findings. The short time behaviour indeed obeys Eq. 3, whereas at long<br />

times the form of Eq. 4 is reproduced fairly well.<br />

35


3 Theoretical Aspects of Anomalous Diffusion in Complex Systems<br />

3.3 The Matheron-de-Marsily model<br />

As mentioned in the introduction, superdiffusive (enhanced) behaviour is often found in turbulent<br />

flows. In this Section we adopt a picture different from CTRW. We focus on the influence<br />

of r<strong>and</strong>omly distributed biasing external fields <strong>and</strong> follow the description of Oshanin<br />

<strong>and</strong> Blumen [11]. We take a three-dimensional (3D) solvent for which the flow is parallel to<br />

the Y-axis. The direction <strong>and</strong> magnitude of the flow depend only on the X coordinate of the<br />

position vector [9] so that V Y , the non-vanishing component of the flow field, obeys:<br />

V Y …X;Y;Z†ˆVX ‰ Š: …23†<br />

Here furthermore V[X] is a r<strong>and</strong>om function of X. Geometrically the system consists<br />

of parallel layers perpendicular to the X-axis. In each layer the value of V Y is constant but<br />

varies from layer to layer [9].<br />

We assume the r<strong>and</strong>om function V[X] of Eq. 23 to be Gaussian with zero mean,<br />

AV[X]S = 0, <strong>and</strong> with the covariance<br />

<br />

VX ‰ 1 ŠVX ‰ 2 Š ˆ … j X1 X 2 j†: …24†<br />

Here the brackets denote configurational averages, which are conveniently expressed<br />

through Fourier integrals,<br />

… jX 1 X 2 j† ˆ R1 1<br />

dwQ…w† exp ‰ iw…X 1 X 2 † Š: …25†<br />

Now many possibilities for Q(w) can be envisaged. For simplicity we take here only a<br />

flat spectrum, Q(w) =W/2p, as in the original Matheron-de-Marsily (MdM) model [9] in<br />

which the flows are delta-correlated,<br />

… jX 1 X 2 j† ˆ W …X 1 X 2 †: …26†<br />

We start from the Langevin dynamics of a single spherical bead subject to the MdM flow.<br />

This allows us to display enhanced (superlinear) diffusion in a simple situation [9, 10, 22]. The<br />

study of the dynamics of Rouse polymers in such flows is deferred to the next Section.<br />

Let R(t) be the position of the center of mass of the bead at time t, <strong>and</strong> we assume<br />

that R(0) = 0. The components X(t), Y(t) <strong>and</strong> Z(t) ofR(t) obey the following Langevin equations<br />

m d2 X<br />

dt 2 ˆ dX<br />

dt ‡ f X…t† ;<br />

m d2 Z<br />

dt 2 ˆ dZ<br />

dt ‡ f Z…t† ;<br />

…27†<br />

…28†<br />

36


3.3 The Matheron-de-Marsily model<br />

<br />

m d2 Y<br />

dt 2 ˆ dY<br />

dt<br />

<br />

VX ‰ Š<br />

‡ f Y …t† : …29†<br />

Here m denotes the mass of the bead <strong>and</strong> z the friction constant. The terms f X (t), f Y (t),<br />

<strong>and</strong> f Z (t) give the r<strong>and</strong>om (thermal-noise) forces exerted on the bead by the solvent molecules.<br />

These forces are Gaussian, with the moments<br />

f i …t† ˆ0<br />

…30†<br />

<strong>and</strong><br />

f i …t† f j …t 0 †ˆ2T i;j …t t 0 † ; …31†<br />

where i,j B {X,Y,Z}. The dash st<strong>and</strong>s for thermal averaging, d i,j is the Kronecker-delta <strong>and</strong><br />

the temperature T is measured in units of the Boltzmann constant k B .<br />

Conventionally the acceleration terms in Eqs. 27–29 are neglected, since they are<br />

small relative to the other terms [15]. This leads to<br />

dX<br />

dt ˆ f X…t† ;<br />

dZ<br />

dt ˆ f Z…t† ;<br />

dY<br />

dt ˆ V‰X…t†Š ‡ f Y…t† :<br />

…32†<br />

…33†<br />

…34†<br />

Note that in Eq. 34 the X <strong>and</strong> Y coordinates are coupled. Equations 32 <strong>and</strong> 33 are<br />

readily solved,<br />

X…t† ˆ 1 Rt<br />

df X …† ;<br />

Z…t† ˆ 1 Rt<br />

df Z …† :<br />

0<br />

0<br />

…35†<br />

…36†<br />

Thus the bead undergoes a conventional diffusive motion between the layers (along<br />

the X-axis) <strong>and</strong> in the Z direction. One sees it readily by evaluating, say:<br />

X 2 …t† ˆ 2Rt R<br />

d t<br />

1 d 2 f X … 1 † f X … 2 †ˆ2 1 T Rt R<br />

d t<br />

1 d 2 … 1 2 †ˆ2…T=† t; …37†<br />

0 0<br />

0 0<br />

so that X 2 …t† ˆ2D 1 t, where D 1 = T/z is the diffusion coefficient of a single bead.<br />

37


3 Theoretical Aspects of Anomalous Diffusion in Complex Systems<br />

A similar procedure can be performed with respect to the solution of Eq. 34,<br />

Y…t† ˆRt<br />

0<br />

dV‰ X…† Š‡ 1 Rt<br />

df Y …† :<br />

0<br />

…38†<br />

The averaging involves now both, the thermal noise <strong>and</strong> the configurational disorder:<br />

Y 2 …t† ˆ 2 Rt R t<br />

d 1 d 2 f Y … 1 † f Y … 2 †‡ Rt R t<br />

d 1<br />

0<br />

0<br />

0<br />

0<br />

<br />

<br />

d 2 V‰X… 1 †Š V‰X… 2 †Š<br />

ˆ 2D 1 t ‡…W=2† Rt R t R 1<br />

d 1 d 2 dw exp ‰iw…X… 1 † X… 2 ††Š; …39†<br />

1<br />

0<br />

0<br />

where in the last line use was made of the representation Eqs. 25 <strong>and</strong> 26 with the flat spectrum<br />

of the delta-function. The remaining average on the rhs. of Eq. 39 is readily evaluated<br />

by remembering that it is the characteristic functional of the Brownian trajectory<br />

… 1 ; 2 ; w† exp‰iw…X… 1 † X… 2 ††Š ˆ exp w 2 D 1 j 1 2 j<br />

: …40†<br />

Inserting Eq. 40 into Eq. 39 one recovers the following result for the average squared<br />

displacement (ASD) in the direction of the flow field,<br />

<br />

Y 2 …t† ˆ 2D1 t ‡ 4W 3<br />

<br />

t 3 1=2<br />

: …41†<br />

p D 1<br />

One should remark that at times greater than t c =9pD 1 3 /(4W 2 ) the superlinear growth<br />

hY 2 …t†i t 3=2 in Eq. 41 dominates <strong>and</strong> the first, diffusive term can be neglected. One has<br />

then a superdiffusive behaviour with an exponent of g = 3/2 in Eq. 1 [9].<br />

We now turn to the analysis of the behaviour of polymers in MdM flow fields.<br />

3.4 Polymer chains in MdM flow fields<br />

The conformational properties <strong>and</strong> the dynamics of polymers in solutions under various<br />

types of flows have been a subject of considerable interest within the last decades. Much<br />

progress has been gained in the explanation of experimental data for systems in which the<br />

flow velocities are given functions in space <strong>and</strong> time, see Refs. [16–19]. On the other h<strong>and</strong>,<br />

the behaviour of polymers in r<strong>and</strong>om flows is less understood. In recent works [11] we<br />

(Oshanin <strong>and</strong> Blumen) succeeded in establishing analytically the behaviour of Rouse polymers<br />

[20] in MdM flow fields. The presentation here follows closely Ref. [11].<br />

38


3.4 Polymer chains in MdM flow fields<br />

In the Rouse model N monomers (beads) are coupled to each other via harmonic<br />

springs [16, 17, 20]. As is well-known, the forces are of entropic origin. It is customary to<br />

revert to a continuous picture in which n, the bead’s running number, takes real values. For a<br />

detailed discussion see Doi <strong>and</strong> Edwards [17]. The Langevin equations of motion for such a<br />

polymer in the MdM flow field are<br />

@X n…t†<br />

@t<br />

@Z n…t†<br />

@t<br />

@Y n…t†<br />

@t<br />

ˆ K @2 X n …t†<br />

@n 2 ‡ f x …n; t† ; …42†<br />

ˆ K @2 Z n …t†<br />

@n 2 ‡ f Z …n; t† ; …43†<br />

ˆ K @2 Y n …t†<br />

@n 2 ‡ V‰X N …t†Š ‡ f Y …n; t† ; …44†<br />

see Ref. [11]. Equations 42–44 are the generalization of Eqs. 27–29 to polymers. They are to<br />

be solved subject to the Rouse boundary conditions at the chain’s ends, n = 0 <strong>and</strong> n = N [17]:<br />

@X n …t†<br />

@n<br />

ˆ @Y n…t†<br />

@n<br />

ˆ @Z n…t†<br />

@n<br />

ˆ 0 :<br />

…45†<br />

As before, the fluctuating forces on the rhs. of Eqs. 42–44 are Gaussian <strong>and</strong> also<br />

delta-correlated with respect to the running index [11, 17]. For the X <strong>and</strong> Z components,<br />

which are not subject to the flow, the procedure is st<strong>and</strong>ard [17]. Say, for the averaged X<br />

component of the end-to-end vector one has<br />

P 2 X …t† ˆ…X 0…t† X N …t†† 2 ˆ b2 N<br />

; …46†<br />

3<br />

where b is the so-called persistence length.<br />

Furthermore, the X component of the radius of gyration is<br />

X 2 g ˆ 1<br />

2N 2 Z N<br />

0<br />

Z N<br />

0<br />

dndm …X n …t†<br />

X m …t†† 2 ˆ b2 N<br />

18 : …47†<br />

For isotropic situations (in the absence of flow fields) the end-to-end vector P R <strong>and</strong><br />

the radius of gyration R g of the polymer are related to P X <strong>and</strong> X g through P R 2 =3P X 2 <strong>and</strong><br />

R g 2 =3X g 2 . Because of the anisotropy one has to consider in the MdM model the different<br />

components separately.<br />

The dynamics of a flexible polymer chain is richer than that of a single bead. In the<br />

Rouse model the dynamics of X n (t) depends essentially on the time of observation t <strong>and</strong> on<br />

the Rouse time t R ,<br />

R ˆ b2 N 2<br />

3p 2 T : …48†<br />

39


3 Theoretical Aspects of Anomalous Diffusion in Complex Systems<br />

Now t R is the largest internal relaxation time of the chain [17, 20]. Exemplarily for<br />

t P t R one finds for the mean-squared displacement of a bead of the chain, say the zeroth one,<br />

<br />

…X 0 …t† X 0 …0†† 2 ˆ 2b D 1=2<br />

1t<br />

: …49†<br />

3p<br />

Equation 49 is subdiffusive with g = 1/2 in Eq. 1. This is due to the fact that the trajectory<br />

of a bead in a chain is spatially confined by its neighbours. In the limit t p t R the<br />

chain diffuses as one entity <strong>and</strong> the bead’s trajectory follows mainly the motion of the<br />

chain’s center of mass. The chain’s center of mass obeys<br />

X 2 …t† ˆ2D R …N† t<br />

…50†<br />

with D R (N) =D 1 /N { T/(zN).<br />

Let us now turn to the Y component of the center of mass for which Eq. 44 leads readily<br />

to<br />

Y…t† 1 N<br />

so that we obtain<br />

Z N<br />

0<br />

dnY n …t† ˆ 1 Z t<br />

d<br />

N<br />

0<br />

Z N<br />

0<br />

<br />

dn VX ‰ n …† Š‡ 1 f Y …n; † ; …51†<br />

<br />

Y 2 …t† ˆ 2DR …N† t ‡ W Z t Z t Z 1<br />

d 1 d 2<br />

2N<br />

0 0 1<br />

dwg…w; 1 ; 2 †;<br />

…52†<br />

where g(w;t 1 ,t 2 ) denotes the dynamic structure factor of the chain,<br />

g…w; 1 ; 2 †ˆ 1<br />

N<br />

Z N Z N<br />

0<br />

0<br />

dn dm exp‰iw…X n … 1 † X m … 2 ††Š : …53†<br />

It turns out [11], that all these integrations can be performed analytically. In the longtime<br />

limit t p t R one finds<br />

<br />

Y 2 …t† ˆ 2DR …N† t ‡ 4W 3<br />

whereas for t P t R the short-time behaviour is given by<br />

40<br />

<br />

N 1=2<br />

t 3=2 <br />

1 O…t 1=2 <br />

† ; …54†<br />

pT<br />

p<br />

<br />

Y 2 …t† 2DR …N† t ‡ 3 W <br />

bN 1=2 t2 1 O…t 1=4 <br />

† : …55†


3.4 Polymer chains in MdM flow fields<br />

The interpretation of Eq. 54 is that at long times the t 3/2 dynamics dominates the picture;<br />

the Rouse chain behaves like a compact bead. At short times the term t 2 may become<br />

important. This g = 2 case in Eq. 1 is called ballistic; at very short times the center of mass<br />

of the chain hardly moves <strong>and</strong> it practically does not change the flow pattern to which it is<br />

subjected.<br />

We remark that at short times the motion of the center of mass of the chain <strong>and</strong><br />

the motion of a tagged bead are characterized by different dependences on time. The<br />

mean-squared displacement along the Y-axis of a tagged bead, say the zeroth one, grows<br />

in time as<br />

<br />

Y0 2…t†<br />

<br />

e2DR …N† t ‡ W 1=4<br />

D<br />

b1=2 1<br />

t 7=4 ; …56†<br />

see Ref. [11]. This g = 7/4 dependence is, of course, related to the fact that the segmental<br />

motion at short times is confined, the number of distinct flow layers visited by the bead<br />

growing as t 1/4 . Results of such fractal-type behaviour may be formulated exactly [21, 22].<br />

We now turn to the question of the elongation of the Rouse chain in the MdM flow<br />

<strong>and</strong> sketch the evaluation of the end-to-end distance along the Y-axis [11].<br />

The solution of Eq. 44 under the boundary conditions Eq. 45 has the form of a Fourier<br />

series,<br />

Y n …t† ˆY…0;t†‡2 X1<br />

pˆ1<br />

<br />

cos ppn <br />

Y…p; t† ;<br />

N<br />

where the Y(p,t), p = 0, 1, …, denote the normal coordinates [17]. At t = 0 the chain is assumed<br />

to be in thermal equilibrium, i. e. to have a Gaussian conformation. This can be accounted for<br />

automatically by stipulating it to be subject to the thermal fluctuations since t =–?. Furthermore,<br />

the MdM flow fields are switched on at t = 0. This leads to<br />

…57†<br />

Y…p; t† ˆ 1 Rt 1<br />

d exp… p 2 …t †= R † ~ f Y …p; †<br />

‡ 1 N<br />

Z t<br />

0<br />

d<br />

Z N<br />

0<br />

<br />

dn cos ppn <br />

V‰X n …t †Š exp… p 2 = R † ; …58†<br />

N<br />

where the functions ~ f Y denote the Fourier components of the thermal fluctuations [17]. The<br />

Y component of the end-to-end vector follows now from:<br />

P Y …t† ˆY 0 …t†<br />

Y N …t† ˆ2 P1<br />

… 1 … 1† p †Y…p; t† : …59†<br />

pˆ1<br />

After performing the averaging over both, the thermal fluctuations <strong>and</strong> the realizations<br />

of the flow fields, one obtains from Eqs. 58 <strong>and</strong> 59 in the long-time limit the equilibrium<br />

value of the end-to-end vector for delta-correlated flows [11],<br />

41


3 Theoretical Aspects of Anomalous Diffusion in Complex Systems<br />

<br />

P 2 Y …1† b 2 <br />

N ˆ 1 ‡ CW2 bN 5=2 <br />

3<br />

T 2 ; …60†<br />

where C is a constant. Therefore, a Rouse polymer stretches along the Y-axis under MdM<br />

flows, taking the form of a prolate ellipsoid. The correction term to the Gaussian behaviour<br />

grows as N 7/2 .<br />

It is now interesting to confront this finding with the situation in simple shear flows,<br />

for which the result is [18, 19]:<br />

P 2 Y …t† ˆb2 N<br />

3<br />

<br />

1 ‡ 3<br />

6480<br />

_ 2 2 b 4 N 4<br />

T 2<br />

<br />

: …61†<br />

In Eq. 61 _ st<strong>and</strong>s for the (constant) shear rate. Because all flow lines point now in<br />

the same direction, the correction term shows a stronger N-dependence <strong>and</strong> obeys a N 5 law.<br />

3.5 Conclusions<br />

In this review several situations were presented, which lead naturally to the appearance of<br />

anomalous diffusion. Besides the already well-discussed subdiffusive behaviour, which is often<br />

seen in disordered media, we also considered superdiffusive dynamics, such as encountered<br />

in layered r<strong>and</strong>om flows. Anomalous diffusion was analysed through several models<br />

<strong>and</strong> we focussed on the behaviour of polymeric materials under such conditions. The basic<br />

aspect underlying these phenomena is dynamical scaling, which is often encountered experimentally<br />

<strong>and</strong> theoretically.<br />

On the other h<strong>and</strong> not all systems scale with time, <strong>and</strong> care is required in applying the<br />

models presented here; one has to be aware of intrinsic limitations of the scaling range (e. g.<br />

size limitations, other temporal scales involved). Because of this, close cooperations between<br />

experimentalists <strong>and</strong> theoreticians are much needed for the analysis of systems similar to the<br />

ones described here. Furthermore, despite the success of the now-closing Collaborative Research<br />

Center 213, much work still remains to be done.<br />

Acknowledgements<br />

The research collaboration with Prof. D. Haarer, Prof. J. Klafter, Dr. G. Oshanin, Dr. H.<br />

Schnörer, <strong>and</strong> Dr. G. Zumofen in our joint work reviewed here was always very helpful <strong>and</strong><br />

pleasant. The support of the Deutsche Forschungsgemeinschaft through the Sonder-<br />

42


References<br />

forschungsbereich 213 <strong>and</strong> Sonderforschungsbereich 60 was fundamental for the whole project.<br />

Additional help was provided by the Fonds der Chemischen Industrie <strong>and</strong> – in the later<br />

stages – by the PROCOPE-Program of the DAAD.<br />

References<br />

1. A. Blumen, J. Klafter, G. Zumofen: in: I. Zschokke (ed.): Optical Spectroscopy of Glasses, Reidel,<br />

Dordrecht, p. 199 (1986)<br />

2. D. Haarer, A. Blumen: Angew. Chem. Int. Ed. Engl., 27, 1210 (1988)<br />

3. A. Blumen, H. Schnörer: Angew. Chem. Int. Ed. Engl., 29, 113 (1990)<br />

4. G. Zumofen, J. Klafter, A. Blumen: in: R. Richert, A. Blumen. (eds.): Disorder Effects on Relaxational<br />

Processes: Glasses, Polymers, Proteins, Springer, Berlin, p. 251 (1994)<br />

5. J. Klafter, G. Zumofen, A. Blumen: J. Phys., A24, 4835 (1991)<br />

6. E.W. Montroll, G.H. Weiss: J. Math. Phys., 6, 167 (1965)<br />

7. H. Scher, M. Lax: Phys. Rev. B, 7, 4491; 4502 (1973)<br />

8. M.F. Shlesinger: J.Stat. Phys., 36, 639 (1984)<br />

9. G. Matheron, G. de Marsily: Water Resour. Res., 16, 901 (1980)<br />

10. G. Zumofen, J. Klafter, A. Blumen: Phys. Rev. A, 42, 4601 (1990)<br />

11. G. Oshanin, A. Blumen: Macromol. Theory Simul. 4, 87 (1995); G. Oshanin, A. Blumen: Phys.<br />

Rev. E, 49, 4185 (1994)<br />

12. H. Scher, E.W. Montroll: Phys. Rev. B, 12, 2455 (1975)<br />

13. H. Schnörer, H. Domes, A. Blumen, D. Haarer: Philos. Mag. Lett., 58, 101 (1988)<br />

14. D. Haarer, H. Schnörer, A. Blumen: Dynamical Processes in Condensed Molecular Systems, in: J.<br />

Klafter, J. Jortner, A. Blumen (eds.): World Scientific, Singapore, p. 107 (1989)<br />

15. M. Fixman: J. Chem. Phys., 42, 3831 (1965)<br />

16. P.G. de Gennes: in: Scaling Concepts in Polymer Physics, Cornell Univ. Press, Ithaca, N.Y., (1979)<br />

17. M. Doi, S.F. Edwards: in: The Theory of Polymer Dynamics, Oxford Univ. Press, Oxford, (1986)<br />

18. R.B. Bird, C.F. Curtiss, R.C. Armstrong, O. Hassager: in: Dynamics of Polymeric Liquids, Vol.2,<br />

2nd Ed. Wiley, New York, (1987)<br />

19. W. Carl, W. Bruns: Macromol. Theory Simul., 3, 295 (1994)<br />

20. P.E. Rouse: J. Chem. Phys., 21, 1273 (1953)<br />

21. J.-U. Sommer, A. Blumen: Croat. Chem. Acta, 69, 793 (1996)<br />

22. G. Zumofen, J. Klafter, A. Blumen: J.Stat. Phys., 65, 991 (1991)<br />

43


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong><br />

Attenuation, Specific Heat <strong>and</strong> Thermal Conductivity<br />

in Polymers<br />

Andreas Nittke, Michael Scherl, Pablo Esquinazi, Wolfgang Lorenz, Junyun Li,<br />

<strong>and</strong> Frank Pobell<br />

We have measured the long time (t =5 h to 200 h) heat release of polymethylmethacrylate<br />

(PMMA) <strong>and</strong> polystyrene (PS) at 0.070 K ^ T ^ 0.300 K. After cooling from a temperature<br />

(the charging temperature) of 80 K the heat release in PMMA shows a t –1 -dependence<br />

in the measured time <strong>and</strong> temperature ranges in agreement with the tunneling model. In contrast,<br />

for PS we observe strong deviations from a t –1 -dependence <strong>and</strong> a heat release smaller<br />

than in PMMA in by a factor of ten, in apparent contradiction to specific heat <strong>and</strong> thermal<br />

conductivity data for PS.<br />

To compare the heat release with other low-temperature properties <strong>and</strong> to verify the consistency<br />

of the tunneling model we have measured also the acoustical properties (sound velocity<br />

<strong>and</strong> attenuation), the specific heat <strong>and</strong> the thermal conductivity of PMMA <strong>and</strong> PS in the<br />

temperature ranges 0.070 K ^ T ^ 100 K, 0.070 K ^ T ^ 0.200 K <strong>and</strong> 0.3 K ^ T ^ 4K,<br />

respectively. We show that the anomalous time dependence of the heat release of PS is due to<br />

the thermally activated relaxation of energy states with excitation energies above 15 K.<br />

4.1 Introduction<br />

A disordered material releases heat after cooling it from an equilibrium or charging temperature<br />

T 1 to a measuring temperature T 0 [1]. This heat release _Q (T 1 ,T 0 , t) depends on the<br />

charging temperature T 1 as well as on the temperature T 0 , at which the measurement is performed,<br />

<strong>and</strong> the elapsed time during <strong>and</strong> after cooling [2, 3]. The time-dependent heat release,<br />

observed in several disordered systems, is a consequence of the long-time relaxation<br />

of the low-energy excitations, identified as two-level tunneling systems (TS) [4–6]. As a<br />

consequence of the finite relaxation time of the TS through their interaction with thermal<br />

phonons [7] the specific heat depends also on the time scale of the experiment [5], as first<br />

measured by Zimmermann <strong>and</strong> Weber [1].<br />

In recent publications [5, 6] it was shown that within the tunneling model we can<br />

quantitatively underst<strong>and</strong> the observed temperature <strong>and</strong> time dependence of the specific heat<br />

44 Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


4.1 Introduction<br />

<strong>and</strong> heat release of vitreous silica (SiO 2 ) over ten orders of magnitude in time. It has also<br />

been pointed out that the measurements of the heat release at workable long times is probably<br />

the only feasible method to obtain at least approximately the low-energy limit of the<br />

tunnel splitting D 0 of the distribution function of TS. This low-energy limit is considered in<br />

the literature by the cut-off parameter u min =(D 0 /E) min of the distribution function (E is the<br />

energy splitting of the TS) <strong>and</strong> is introduced to keep the number of TS finite, avoiding divergences<br />

in the calculated properties.<br />

The interpretation of the heat release data in terms of the tunneling model is a difficult<br />

task due to the not well-known:<br />

a) temperature <strong>and</strong> time dependence of the specific heat due to TS at T >3K;<br />

b) influence of relaxation processes of the TS at T > 3 K other than one-phonon tunneling<br />

relaxation, i. e. high-order phonon tunneling <strong>and</strong> thermally activated relaxation;<br />

c) influence of the cooling procedure.<br />

In a recently published paper Parshin <strong>and</strong> Sahling [8] showed the complexity of the interpretation<br />

of the heat release data when thermally activated relaxation is taken into account<br />

within the framework of the soft potential model. Further theoretical work on the residual<br />

properties of two-level systems <strong>and</strong> its dependence on the cooling procedure has been published<br />

by Brey <strong>and</strong> Prados [9].<br />

In this paper we present a further example of the complexity in the interpretation of<br />

the heat release data <strong>and</strong> an experimental proof of the influence of the relaxation, probably<br />

by thermally activated processes, of excited states that contribute to the heat release of the<br />

TS at low temperatures. We have studied the long-time heat release of two amorphous polymers<br />

with similar low-temperature specific heats <strong>and</strong> thermal conductivities, cooled under<br />

similar conditions. In spite of those similarities we have found a large difference in the absolute<br />

value <strong>and</strong> in the time dependence of the heat release between the two polymers when<br />

cooled from temperatures above 15 K. The similarities <strong>and</strong> differences in the low-temperature<br />

properties, their interpretation within the tunneling model, <strong>and</strong> the influence of thermally<br />

activated relaxation are the main scope of this work. Preliminary results were published<br />

in Ref. [10].<br />

The paper is organized in five sections. In Section 4.2 we describe the phenomenological<br />

theory for the heat release, based on the tunneling model <strong>and</strong> the influence of different<br />

relaxation rates, <strong>and</strong> the cooling process. In Section 4.3 we briefly describe the experimental<br />

procedures <strong>and</strong> samples. In Section 4.4 we show <strong>and</strong> discuss the experimental results. Conclusions<br />

are drawn in Section 4.5.<br />

45


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

4.2 Phenomenological theory for the heat release<br />

4.2.1 Generalities<br />

New theoretical work for the calculation of the heat release within the soft potential model<br />

has been published recently [8]. In order to simplify the calculations, to minimize the number<br />

of free parameters, <strong>and</strong> to assure a more transparent interpretation of the results we<br />

decided, however, to interpret our results in terms of the st<strong>and</strong>ard tunneling model. According<br />

to Ref. [8] <strong>and</strong> to our numerical calculations the differences between the soft potential<br />

<strong>and</strong> st<strong>and</strong>ard tunneling model are not significant at the low temperatures of our measurements.<br />

The st<strong>and</strong>ard tunneling model, including a thermally activated relaxation of the twolevel<br />

systems, was used successfully to interpret the acoustical properties of vitreous silica<br />

in eight orders of magnitude in phonon frequency from approximately 0.1 K up to room<br />

temperature [11].<br />

For a system of N two-level systems with energy difference E, the difference in the population<br />

of the two levels at a given temperature T <strong>and</strong> at thermal equilibrium is given by<br />

n 0 ˆ N tanh…E=2k B T† :<br />

…1†<br />

If the thermodynamic equilibrium of the system is slightly perturbed, e. g. the system<br />

is rapidly cooled to a temperature T, the dynamical behaviour of the population difference<br />

can be calculated according to the relaxation time approximation formula:<br />

d…n…t†<br />

n 0 …T††<br />

dt<br />

ˆ n…t† n 0 …T†<br />

…E;T†<br />

; …2†<br />

where t(E,T) is the relaxation time of a tunneling system with energy E at a temperature T.<br />

The heat released by the N two-level systems after the temperature change is given by<br />

_Q ˆ _nE=2 :<br />

…3†<br />

The problem of calculating _Q simplifies to calculating n from Eq. 2. However, the<br />

cooling process has to be taken into account. In this case Eq. 2 must be rewritten as<br />

_n ˆ @n 0<br />

@T<br />

dT<br />

dt<br />

n n 0<br />

<br />

: …4†<br />

Equation 4 describes the response of two-level systems during a temperature change<br />

given by the function T(t). In the general case Eq. 4 has to be solved numerically since n 0<br />

<strong>and</strong> t are temperature-dependent variables.<br />

46


4.2 Phenomenological theory for the heat release<br />

4.2.2 The st<strong>and</strong>ard tunneling model with infinite cooling rate<br />

If N two-level systems are in thermal equilibrium at a temperature T 1 <strong>and</strong> they are cooled to T 0<br />

with infinite cooling rate, i. e. the time to cool the sample from T 1 to T 0 is zero, from Eq. 2 we<br />

obtain<br />

n…t† ˆ…n 0 …T 1 † n 0 …T 0 †† exp… t= …E;T 0 †† ‡ n 0 …T 0 † : …5†<br />

The st<strong>and</strong>ard tunneling model assumes that at low temperatures (T < 2 K) the<br />

one-phonon process is the dominating mechanism. The one-phonon relaxation rate is given<br />

by<br />

1<br />

p ˆ 1<br />

pm u2<br />

…6†<br />

with<br />

1<br />

pm ˆ AE3 coth…E=2K B T† ;<br />

…7†<br />

where A =(g 2 l /n 5 l +2g 2 t n 5 t )/2 pr –4 . The indices l, t refer to the longitudinal <strong>and</strong> transversal<br />

phonon branches, r is the mass density, g l <strong>and</strong> g t are the coupling constants between phonons<br />

<strong>and</strong> TS, <strong>and</strong> u=D 0 /E (D 0 is the tunneling splitting). For symmetrical TS, i. e. D 0 = E (u = 1),<br />

t p reaches its minimum value t pm .<br />

Furthermore, the st<strong>and</strong>ard tunneling model assumes that the distribution function of<br />

TS is constant in terms of two independent variables, namely the asymmetry D <strong>and</strong> the tunneling<br />

parameter l, i.e.<br />

P…;†dd ˆ Pdd:<br />

…8†<br />

According Ref. [7] the distribution function can be written in terms of the variables E<br />

<strong>and</strong> u as<br />

P…E;u† ˆ P=u…1 u 2 † 1=2 : …9†<br />

It is also convenient to have the distribution function in terms of the asymmetry <strong>and</strong><br />

the barrier height V between the potential wells. Following the work of Tielbürger et al. [11]<br />

<strong>and</strong> assuming two well-defined harmonic potentials, it can be shown that in a first approximation<br />

l = V/E 0 where E 0 represents the zero-point energy. In this case the distribution function<br />

is<br />

Pdd ˆ<br />

P<br />

E 0<br />

ddV :<br />

…10†<br />

Replacing the total number N of TS with the integrals in E <strong>and</strong> u, the heat release is<br />

given by<br />

47


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

_Q ˆ PV Z 1 Z<br />

E<br />

E<br />

1<br />

dEE tanh tanh<br />

E 0 2k B T 0 2k B T 1<br />

0<br />

u min<br />

du<br />

p<br />

u 1 u 2<br />

1 …T 0 † exp… t=…T 0 †† ;<br />

(11)<br />

where V is the sample volume <strong>and</strong> u min is the cut-off in the distribution function at u ?0.<br />

Figure 4.1 shows the heat release at T 0 = 0.1 K as a function of time for different<br />

charging temperatures T 1 following Eq. 11 <strong>and</strong> taking into account one-phonon relaxation<br />

rate, Eqs. 6 <strong>and</strong> 7. The calculation was performed with parameters appropriated for vitreous<br />

silica, Ak 3 B =4610 6 s –1 K –3 , P = 1.6610 38 J –1 g –1 <strong>and</strong> u min =5610 –8 [5]. We observe that<br />

at short times <strong>and</strong> small T 1 the heat release _Q ! t –1 T 2 1. This dependence follows from the<br />

logarithmic time dependence of the specific heat <strong>and</strong> holds for tPt m /u 2 min. In this limit the<br />

heat release follows the often used approximation from Eq. 11 [1]:<br />

_Q<br />

p2<br />

24 k2 B PV…T 2 1 T 2 0 † 1 t ; …12†<br />

where _Q <strong>and</strong> P are measured in W <strong>and</strong> (Jg) –1 . We recognize in Fig. 4.1, however, that at<br />

longer times the theory deviates from the t –1 -dependence. This deviation comes from the<br />

exponential term in Eq. 11 that decreases strongly at long times <strong>and</strong> is determined by the<br />

product Au 2 min. The smaller this product the larger is the time-range where the approximation<br />

given by Eq. 12 holds. The influence of the cut-off u min can be recognized comparing the<br />

calculated heat release in Fig. 4.1 with that in Fig. 4.2 (dashed lines) calculated with a<br />

smaller u min .<br />

In Fig. 4.1 we note also that at large times the heat release becomes independent of<br />

T 1 . This feature was recognized experimentally in Refs. [2, 3] <strong>and</strong> was discussed in Ref. [6].<br />

This T 1 -independence within the assumptions described above is a direct consequence of the<br />

finite number of TS given by the cut-off u min . We should note, however, that in SiO 2 a saturation<br />

of the heat release for large charging temperatures (T 1 > 20 K) is observed <strong>and</strong> still<br />

a t –1 -dependence was measured [2]. This result cannot be explained within the st<strong>and</strong>ard tun-<br />

Figure 4.1: Heat release as a function of time t (in s) according to the st<strong>and</strong>ard tunneling model <strong>and</strong> infinite<br />

cooling rate with Ak 3 B =4610 6 s –1 K –3 , P = 1.6610 38 J –1 g –1 <strong>and</strong> u min =5610 –8 at a measuring<br />

temperature of 0.1 K <strong>and</strong> at different charging temperatures T 1 , bottom: 5 K, top: 80 K.<br />

48


4.2 Phenomenological theory for the heat release<br />

Figure 4.2: Heat release as a function of time t (in s) within the st<strong>and</strong>ard tunneling model <strong>and</strong> infinite<br />

cooling rate. The parameters are the same as in the previous figure but with u min =5610 –9 (dashed<br />

lines) <strong>and</strong> an additional constant (energy-independent) relaxation time t TA =10 3 s (continuous lines).<br />

neling model. The reason for this discrepancy is the relaxation rate, which is different from<br />

the one-phonon process at higher temperatures <strong>and</strong> the influence of a finite cooling rate (see<br />

below).<br />

4.2.3 Influence of higher-order tunneling processes <strong>and</strong> a finite cooling rate<br />

At temperatures above 3 K tunneling processes of higher-order are, in principle, possible. One<br />

particular high-order process was studied theoretically in Ref. [12] <strong>and</strong> is similar to the optical<br />

Raman process; the relaxation time of the TS is given by an interaction that involves two phonons.<br />

Following Ref. [12] we added this Raman relaxation rate to the one-phonon relaxation<br />

rate (Eq. 6). Following Ref. [12] a new free parameter, i. e. the coupling constant R for the Raman<br />

process, is assumed. Due to its strong temperature dependence (T 7 ), the Raman process<br />

can influence the population difference of the energy states of the TS mainly at high temperatures<br />

(T > 5 K) during the cooling process. We should note that for infinite cooling rate <strong>and</strong> for<br />

the temperature range of our measurements (T 0^1 K) the contribution of the Raman process<br />

to the time <strong>and</strong> temperature dependence of the heat release is negligible. This follows from Eq. 5<br />

where only the relaxation of the TS at the measurement temperature T 0 enters.<br />

In order to take into account the cooling process we used the algorithm explained below.<br />

The zero-time point is chosen as the time when the cooling process is started. The function<br />

n (E,u,t) represents the difference in population of the TS energy states at any time t<br />

<strong>and</strong> n 0 (E,T) means this difference in equilibrium (t ? ?) at a temperature T. Analog to<br />

Eq. 11 the heat release can be written as:<br />

_Q ˆ 1<br />

2 PV<br />

Z 1<br />

0<br />

dEE<br />

Z 1<br />

u min<br />

du<br />

p 1 …T 0 † e 1 …T 0 † t …n 0 …E;T 0 † n…E; u; t ˆ 0†† …13†<br />

u 1 u 2<br />

49


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

Note that the population difference n depends on E <strong>and</strong> u because the relaxation rates,<br />

that influence the relaxation of the TS, depend on these variables. As already described<br />

above, to obtain the function n we need to solve the differential equation given by Eq. 4. In<br />

general this is only possible by numerical methods.<br />

The time dependence of the temperature during the cooling process performed in this<br />

work can be well approximated by a linear function given by<br />

T…t† ˆT 1 ‡ T 0 T 1<br />

t A<br />

: …14†<br />

We assume that the sample is at t =0(t A )atT 1 (T 0 ); t A is the time needed to cool the<br />

sample from T 1 to T 0 . The calculations have been done splitting the function T(t) inN steps<br />

of time Dt (Fig. 4.3). During the time Dt the temperature remains constant <strong>and</strong> n (E,u,t)<br />

shows an exponential behaviour given by Eq. 5; this is qualitatively shown in Fig. 4.3. To<br />

obtain the population difference at the time Dt cooling from T 1 at t = 0 for a given E <strong>and</strong> u<br />

we calculate iteratively the value<br />

n…t ˆ 0 ‡ t† ˆn 0 …0 ‡ t† …n 0 …0 ‡ t† n…0† e 1 …T…0‡t††t : …15†<br />

The calculations given in this work have been made taking into account 10 to 200<br />

time-steps or iterations depending on the convergence of the numerical results.<br />

Figure 4.4 shows the heat release as a function of time taking into account one-phonon<br />

(dashed lines) <strong>and</strong> the Raman process (Rk 7<br />

B = 100 s –1 K –7 ) with a cooling time<br />

t A = 3000 s from T 1 to T 0 = 0.1 K. Taking into account our experimental time scale the<br />

calculations were performed for the time interval 10 4 s ^ t ^ 10 6 s only. For one-phonon<br />

Figure 4.3: Time dependence of the sample temperature during a cooling process <strong>and</strong> its splitting in N<br />

steps (dashed lines) assumed for numerical calculations. Bottom: qualitative behaviour of the population<br />

difference of the tunneling systems n 0 (T) as a function of time.<br />

50


4.2 Phenomenological theory for the heat release<br />

Figure 4.4: Heat release as a function of time for a finite cooling time T A = 3000 s <strong>and</strong> with (continuous<br />

lines) or without (dashed lines) Raman processes.<br />

process only, our calculations indicate that a finite cooling rate with t A = 3000 s to 5000 s<br />

has a negligible influence (< 3%) on the heat release in our experimental time scale (compare<br />

the dashed lines in Fig. 4.4 with the dashed lines in Fig. 4.2 obtained with an infinite<br />

cooling rate).<br />

The influence of the Raman process at T 0 = 0.1 K can be well observed if we take<br />

into account a finite cooling rate. In comparison with the results using the one-phonon process<br />

only <strong>and</strong> for T 1 >20K, _Q is smaller <strong>and</strong> shows a slightly different time dependence.<br />

Note that the results for T 1 6 20 K resemble those obtained taking into account the one-phonon<br />

process only but with a smaller charging temperature T 1 . We note also that at T 1^ 20 K<br />

the heat release still shows a T 1 2 -dependence but reaches its saturation at a smaller T 1 in<br />

comparison with the results with the one-phonon process only (Fig. 4.4).<br />

We conclude that taking into account Raman processes, that influence the relaxation<br />

rate of TS at temperatures larger than 2 K, it is possible to underst<strong>and</strong> qualitatively the results<br />

of, for example, SiO 2 where _Q reached a saturation at T & 10 K but still shows a t –1 -<br />

dependence at t ^ 3610 5 s [2]. This result is only valid if the product Au 2 min in Eq. 7 remains<br />

small enough.<br />

Because of the uncertainty in the coupling constant R between TS <strong>and</strong> phonons a<br />

quantitative comparison, however, of the heat release results with the theory taking into account<br />

the contribution of higher-order processes is not useful. Phenomenologically <strong>and</strong> in order<br />

to simplify the calculations, one can take into account the influence of higher-order processes<br />

by choosing an appropriate smaller charging temperature than the real one. In that<br />

sense charging temperatures T 1 6 10 K can be considered for comparison with theory as<br />

free parameters. From our calculation we learn that it is practically impossible to obtain reliable<br />

information on the density of states of TS from heat release experiments for charging<br />

temperatures in the region of the T 1 -saturation of _Q.<br />

51


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

4.2.4 The influence of a constant <strong>and</strong> thermally activated relaxation rate<br />

There are different attempts to link the high-temperature (T > 10 K) with the low-temperature<br />

properties of disordered solids [13, 14]. In particular, it has been proposed that the maximum<br />

in the attenuation of phonons observed in amorphous materials at T > 10 K can be interpreted<br />

assuming a thermally activated relaxation of the two-level systems. In this Section<br />

we discuss the influence of thermally activated relaxation rate on the heat release for a finite<br />

cooling rate.<br />

For a single activation barrier V 0 between the two potential wells <strong>and</strong> at a measuring<br />

temperature T 0 the thermally activated relaxation rate is given by<br />

1<br />

TA ˆ 1<br />

0 e V 0=k B T 0<br />

: …16†<br />

Taking into account this relaxation rate in addition to quantum tunneling (Eq. 6) <strong>and</strong><br />

assuming that both processes are independent, the total relaxation rate can be written as<br />

1 ˆ 1<br />

P ‡ 1<br />

TA :<br />

…17†<br />

Obviously, at a given temperature T 0 <strong>and</strong> for a single activation barrier we add a constant<br />

relaxation rate to the tunneling process. Figure 4.2 shows the time dependence of the<br />

heat release at different temperatures T 1 <strong>and</strong> using Eq. 17 with t TA =10 3 s. As expected, at<br />

t P t TA the introduction of an additional constant relaxation rate does not influence the heat<br />

release. At t & t TA the heat release is larger than taking only into account the one-phonon<br />

process (Fig. 4.2). At longer times _Q decreases exponentially with time. This decrease can<br />

be easily understood: the energy levels with long (tunneling) relaxation time have been depopulated<br />

at t & t TA at a larger rate. Since the total amount of heat released by the TS<br />

should be finite a decrease below the st<strong>and</strong>ard result (dashed lines in Fig. 4.2) is expected.<br />

To take into account the distribution of potential barriers V <strong>and</strong> a thermally activated<br />

rate we will follow the approach used by Tielbürger et al. [11] <strong>and</strong> transform Eq. 11 in a double<br />

integral in the variables D <strong>and</strong> V using the approximation l = V/E 0 <strong>and</strong> D 0 & 2E 0 e –l /p:<br />

_Q ˆ PV<br />

2<br />

Z 1<br />

0<br />

d<br />

ZV max<br />

0<br />

dV E <br />

E<br />

tanh<br />

E 0 2k B T 0<br />

<br />

E<br />

tanh 1 …T 0 † exp… t=…T 0 †† : …18†<br />

2k B T 1<br />

For the following discussion it is not relevant to introduce in Eq. 18 a specific distribution<br />

of potential barriers V [11]. This is done below for comparison with experimental results<br />

<strong>and</strong> for computing the acoustic properties. After finding the range of potential barriers<br />

relevant for the heat release we divided the calculations in two regions: T 1 ^ 1 K <strong>and</strong><br />

T 1 > 1 K as described below.<br />

In order to find the range of potential barriers relevant for the heat release in our measuring<br />

time <strong>and</strong> temperature ranges we have calculated it following Eq. 18, splitting the V-<br />

limits of the inner integral from V min = V max –10KtoV max , using only the one-phonon process,<br />

<strong>and</strong> measuring temperature T 0 = 0.200 K. We have recognized that for V < 130 K <strong>and</strong><br />

52


4.2 Phenomenological theory for the heat release<br />

at t&10 3 s the TS are mostly relaxed <strong>and</strong> do not contribute appreciably to the heat release<br />

in our measured time range (10 4 s ^ t ^ 10 6 s) any more. The TS that relax through the<br />

one-phonon process <strong>and</strong> are relevant to the heat release are found to be those with potential<br />

barriers 130 K < V < 200 K. The contribution of tunneling systems with V > 200 K to the<br />

heat release is negligible.<br />

At T 1 < 1 K the contribution of a thermally activated rate for TS with potential barriers<br />

130 K < V < 200 K is irrelevant, i. e. using t&10 –13 s (from acoustic measurements,<br />

see below) we obtain a t TA (V = 130 K k B ) of several years. The potential barriers which are<br />

relevant in our time <strong>and</strong> temperature range through a thermally activated rate are V/k B 1 K. Figure 4.5 shows the calculated heat release as a function of T 1 at a<br />

given time t = 1.5610 5 s taking into account the cooling process <strong>and</strong> one-phonon process<br />

as well as thermally activated relaxation rate. The numerical results (Fig. 4.5) show clearly a<br />

saturation of _Q for T 1 > 5 K, i. e. at lower temperatures in comparison with the results without<br />

thermally activated relaxation, in qualitative agreement with the results assuming Raman<br />

processes <strong>and</strong> experimental results [2]. It is also interesting to note that a thermally activated<br />

relaxation rate decreases slightly the exponent in the time dependence of the heat release,<br />

i. e. _Q ! t –a with a < 1. This result is shown in Section 4.4.<br />

Figure 4.5: Heat release at t = 1.5610 5 s as a function of charging temperature T 1 calculated with <strong>and</strong><br />

without thermally activated processes <strong>and</strong> a finite cooling rate.<br />

53


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

4.3 Experimental details<br />

The experimental setup for measuring the heat release <strong>and</strong> the specific heat consists of a calorimeter<br />

on which the sample is mounted. The samples were cooled from 80 K to 0.050 K<br />

in about (5–8)610 3 s. The calorimeter is attached to a holder through a thermal resistance<br />

or a superconducting heat switch made of an Al strip. The holder is screwed to the mixing<br />

chamber of a top loading dilution refrigerator.<br />

Two sample holders were used, one from plastic <strong>and</strong> a second from Teflon. The first<br />

one showed a heat release approximately proportional to t –1 whereas for the second one no<br />

heat release was measured in agreement with previous measurements [2].<br />

We used two different methods to measure the heat release. Firstly, after cooling to<br />

low temperatures (T&0.060 K) the heat switch was opened <strong>and</strong> the time dependence of the<br />

sample temperature (warmup rate) was measured till about 0.120 K. Then the sample was<br />

cooled again to about 0.050 K before starting another warmup run. The heat release is obtained<br />

from the slope of the temperature in the warmup run dT (t)/dt with the knowledge of<br />

the specific heat of the sample plus calorimeter C(T), i. e. _Q = C(T) dT (t)/dt. The measurements<br />

were completely automated with a personal computer. The background heat leak was<br />

about (0.03–0.50) nW at 0.090 K <strong>and</strong> at t & 10 5 s depending on the sample holder <strong>and</strong> the<br />

vibration of the connecting cables. To test the measuring setup <strong>and</strong> procedure we have produced<br />

well-known heat leaks with a electrical heater, fixed to the sample.<br />

The second method for measuring the heat release is based as before on the measurement<br />

of the time dependence of the sample temperature at constant bath (mixing chamber)<br />

temperature. The difference lies on the selection of a fixed thermal resistance R th between<br />

sample <strong>and</strong> holder that enables the semiadiabatic measurement of the heat release in a time<br />

scale larger than the intrinsic thermal relaxation time of the arrangement. Approximately<br />

30 minutes after reaching a constant bath temperature T b the decrease of the sample temperature<br />

T s with time provides directly the heat release, i. e. _Q =(T s – T b )/R th (T s ,T b ). The<br />

measurement of the thermal resistance together with the background contribution to the heat<br />

release is made in situ after reaching the time-independent minimum temperature. Within<br />

experimental error (&10 %) both methods show the same results.<br />

The specific heat of both polymers was measured with the heat pulse technique in semiadiabatic<br />

fashion. The thermal conductivity was measured with a top loading 3 He refrigerator<br />

using the st<strong>and</strong>ard procedure. The acoustic properties, sound velocity <strong>and</strong> attenuation, were<br />

measured with the vibrating reed technique [16] in the frequency range (0.2–3) kHz.<br />

Both polymers were prepared following st<strong>and</strong>ard procedures. The PMMA sample had<br />

additionally 10 –2 mol% of tetra-4-tert.butyl-phthalocyamin (dye molecule) because it was<br />

used in an early optical hole burning experiment [15]. For the specific heat <strong>and</strong> heat release<br />

measurements the mass <strong>and</strong> density of the PMMA sample were determined as 11.97 g <strong>and</strong><br />

1.15 g/cm 3 , respectively. We have measured two PS samples prepared from different<br />

batches. The densities of these samples were 1.05 g/cm 3 , <strong>and</strong> the masses 11.4 g (sample<br />

PS1) <strong>and</strong> 38.0 g (sample PS2).<br />

For the thermal conductivity measurements two slices were cut from the bulk samples<br />

with length l = 1 cm, width w = 2 mm, <strong>and</strong> thickness d = 300 mm. For the acoustic measurements<br />

the reeds had the geometry l = (0.7–1) cm, w = (0.1–0.3) cm <strong>and</strong> d = (100–300) mm.<br />

54


4.4 Experimental results <strong>and</strong> discussion<br />

4.4 Experimental results <strong>and</strong> discussion<br />

4.4.1 Specific heat <strong>and</strong> thermal conductivity<br />

Figure 4.6 shows for both polymers the specific heat devided by the temperature as a square<br />

function of temperature. These measurements were performed only when the heat release is<br />

negligible (t & 10 6 s). For t ?? the specific heat can be written in terms of two contributions:<br />

c (T) =c 1 T + c 3 T 3 with coefficients for the tunneling system c 1 <strong>and</strong> phonon contribution<br />

c 3 . From Fig. 4.6 we obtain the values c 1 = (3.0+0.3) mJ/gK 2 <strong>and</strong> c 3 = (93 + 18) mJ/gK 4<br />

for PMMA, <strong>and</strong> c 1 = (4.6+0.5) mJ/gK 2 <strong>and</strong> c 3 = (77 +23) mJ/gK 4 for PS. The values of the<br />

linear term are in fair agreement with those published by Stephens [17, 18] c 1 = 4.6 mJ/gK 2<br />

for PMMA <strong>and</strong> c 1 = 5.1 mJ/gK 2 for [PS], but we obtained larger phonon contribution<br />

(Stephens: c 3 =29mJ/gK 4 for PMMA <strong>and</strong> c 3 =45mJ/gK 4 for PS). In Figs. 4.7 <strong>and</strong> 4.8 we<br />

compare our specific heat results with data from Ref. [18]. Within the scatter of the data it<br />

is difficult to conclude whether the c values are really different.<br />

Figure 4.6: Specific heat divided by the temperature as function of the squared temperature for PMMA<br />

(a) <strong>and</strong> PS (b).<br />

Figure 4.7: Comparison of our specific heat for PMMA with data from Ref. [18].<br />

55


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

Figure 4.8: Comparison of our specific heat for PS with data from Ref. [18].<br />

According to the tunneling model <strong>and</strong> for t ?? the specific heat due to the TS is given<br />

by [5]:<br />

c…T† ˆp2<br />

6 k2 2<br />

BPV ln… †T ˆ c 1 T:<br />

u min<br />

…19†<br />

If we assume (from heat release data, see below) u min =10 –10 for PMMA <strong>and</strong><br />

u min =6610 –9 for PS <strong>and</strong> from the measured c 1 we obtain a density of states P = 4.0610 38<br />

(7.5610 38 ) 1/Jg for PMMA (PS) in reasonable agreement with the values obtained in earlier<br />

specific heat measurements [17] <strong>and</strong> also from acoustic measurements for PMMA [24].<br />

Recent optical hole burning experiments on PMMA <strong>and</strong> PS [15] indicate that the density of<br />

states of TS for PS is about two times larger than for PMMA in reasonable agreement with<br />

our specific heat measurements.<br />

Figure 4.9: Thermal conductivity as a function of temperature for PMMA <strong>and</strong> PS.<br />

56


4.4 Experimental results <strong>and</strong> discussion<br />

Figure 4.9 shows the thermal conductivity for PMMA <strong>and</strong> PS as a function of temperature.<br />

For PMMA <strong>and</strong> at T < 0.7 K we obtain k =28T 1.84 610 –3 W/mK 2.84 <strong>and</strong> for PS<br />

k =19T 1.93 610 –3 W/mK 2.93 , in agreement with earlier measurements [17, 19–22], (Figs. 4.10<br />

<strong>and</strong> 4.11), <strong>and</strong> with measurements in epoxies [23]. According to the tunneling model <strong>and</strong> for<br />

T < 1 K the thermal conductivity is given by:<br />

…T† ˆk3 B v<br />

6p 2 … P 2 † 1 T 2 ; …20†<br />

where v is the sound velocity <strong>and</strong> r the mass density. From Eq. 20 <strong>and</strong> with the sound velocity<br />

from Ref. [17] we obtain Pg 2 = 9.2610 5 J/m 3 (11.5610 5 J/m 3 ) for PMMA (PS). If we<br />

replace the density of states P from the specific heat results we obtain similar values for the<br />

coupling constant between TS <strong>and</strong> phonons in both materials g = 0.28 eV (0.27 eV) for<br />

PMMA (PS). The results for the thermal conductivity indicate that the density of states of<br />

TS for the PS sample is about a factor of two larger than for the PMMA sample.<br />

Figure 4.10: Comparison of our thermal conductivity data for PMMA with earlier publications. The<br />

dashed line has been calculated within the tunneling model with parameters taken from our internal friction<br />

data.<br />

Figure 4.11: Comparison of our thermal conductivity data for PS with earlier publications. The dashed<br />

line has been calculated within the tunneling model with parameters taken from our internal friction<br />

data.<br />

57


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

4.4.2 Internal friction <strong>and</strong> sound velocity<br />

Figures 4.12 <strong>and</strong> 4.13 show the internal friction of PMMA <strong>and</strong> PS. Below 1 K the internal<br />

friction of PMMA <strong>and</strong> PS is approximately temperature-independent. According to the tunneling<br />

model the sound absorption should be temperature-independent in the region ot p P1<br />

(fulfilled in our temperature range) where o =2pv is the phonon frequency <strong>and</strong> t p the relaxation<br />

time of the TS (Eq. 6). In this temperature range a simple relationship is found for<br />

the internal friction [11, 16]<br />

Q 1 ˆ p<br />

2 C …21†<br />

Figure 4.12: Squares: PMMA internal friction as a function of temperature at a frequency of 535 Hz.<br />

Full circles: indicate the attenuation data taken from the ultrasonic measurements at 15 MHz from<br />

Ref. [24]. Solid <strong>and</strong> dotted lines: the calculated values for v=535 Hz <strong>and</strong> v=15 MHz following the<br />

modified tunneling model considering a thermally activated relaxation rate. For more details see text.<br />

Figure 4.13: Squares <strong>and</strong> Circles: PS internal friction as a function of temperature at two different frequencies.<br />

Solid line: the calculated values for v=0.24 kHz following the modified tunneling model described<br />

in the text.<br />

58


4.4 Experimental results <strong>and</strong> discussion<br />

with the constant C = Pg 2 /ru 2 <strong>and</strong> the sound velocity u. Equation 21 should hold even if<br />

higher-order processes dominate the relaxation of the TS (if the relaxation rate is proportional<br />

to u 2 !D 0 2 , Eq. 6). From the temperature-independent value of the internal friction<br />

<strong>and</strong> assuming no background contribution (e. g. due to the clampling) we obtain for PMMA<br />

(PS) C % 2.6610 –4 (8.3610 –4 ). Taking into account the mass density, sound velocity, <strong>and</strong><br />

the coupling constant obtained from the thermal conductivity <strong>and</strong> specific heat results described<br />

above, the values for the parameters C from the internal friction indicate that the<br />

density of states of TS for PS is larger than for PMMA by a factor &2.5. This is slightly<br />

larger than that obtained from the specific heat (about 1.5). This might be attributed to the<br />

unknown clamping contribution to the measured attenuation (which is always present) or to<br />

different u min values (Eq. 19). In Figs. 4.10 <strong>and</strong> 4.11 we show the calculated thermal conductivity<br />

using Eq. 20 with Pg 2 values from our internal friction at the plateau. Excellent agreement<br />

is obtained for PMMA (Fig. 4.10) but for PS the thermal conductivity is by a factor<br />

of two smaller. This difference might be attributed partially to the background contribution<br />

which is not subtracted from the measurement before computing the parameter C.<br />

The behaviour of the internal friction above about 3 K is qualitatively different for<br />

both polymers. PMMA shows a decrease in the internal friction reaching a minimum at<br />

about 30 K <strong>and</strong> increasing monotonously to the highest temperature of our experiment,<br />

120 K (not shown in Fig. 4.12) in very good agreement with previous work at similar frequencies<br />

[25]. On the contrary, the internal friction of PS shows the typical temperature dependence<br />

measured for other amorphous materials, for example SiO 2 [16]. It increases,<br />

reaching a frequency-dependent maximum at T&37 K (;40 K), (Fig. 4.13), <strong>and</strong> increases<br />

again at T > 60 K (70 K) at the frequency 0.24 kHz (3.2 kHz). This internal friction increase<br />

at T > 60 K for PS together with the increase at T > 30 K for PMMA will not be discussed<br />

here. Instead, we will discuss the behaviour of the internal friction at lower temperatures in<br />

terms of an extension of the tunneling model following the procedure described in Ref. [11].<br />

We assume the simplified relation for a thermally activated relaxation rate given by<br />

Eq. 16 <strong>and</strong> add it to the tunneling rate (Eq. 17). For two well-defined harmonic potentials<br />

<strong>and</strong> within the approximations given by Eq. 10 <strong>and</strong> with l = V/E 0 , it can be shown that the<br />

internal friction increases linear with temperature just above the temperature-independent region<br />

(plateau) [11]:<br />

Q 1 ˆ pCk BT<br />

E 0<br />

: …22†<br />

This increase is observed for PS, (Fig. 4.13); applying Eq. 22 to the results below<br />

20 K we obtain a zero point energy E 0 =13+ 2 K. In the same temperature region <strong>and</strong> due<br />

to the influence of the thermally activated relaxation the relative change of the sound velocity<br />

can be written as [11]:<br />

uu<br />

u ˆ Ck BT<br />

E 0<br />

ln…! 0 † : …23†<br />

A nearly linear temperature dependence of uu/u is observed in both polymers at temperatures<br />

about below 50 K (Figs. 4.14 <strong>and</strong> 4.15). At T ~ 1 K a crossover, due to tunneling<br />

<strong>and</strong> due to thermally activated relaxation, from the linear T-dependence to the logarithmic<br />

59


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

Figure 4.14: Relative change of sound velocity at v=535 Hz for PMMA. The solid line represents the<br />

linear temperature dependence of the sound velocity at temperatures below 20 K used for the numerical<br />

calculations. Inset: the same data but below 10 K in a semilogarithmic scale.<br />

Figure 4.15: Relative change of sound velocity at two different frequencies for PS. The solid line was<br />

calculated according to the modified tunneling model described in the text. Inset: the same data but below<br />

10 K in a semilogarithmic scale.<br />

T-dependence can be observed for both polymers (insets in Figs. 4.14 <strong>and</strong> 4.15). Although<br />

the tunneling model with the assumption of thermally activated relaxation of the TS provides<br />

a reasonable fit for the linear temperature dependence, its origin is still controversial. We<br />

note that a linear temperature dependence of the sound velocity above a few Kelvin is a<br />

rather general behaviour observed in several amorphous [26, 27], disordered [28], <strong>and</strong> polycrystalline<br />

metals [29]. Nava [28] argued recently against an interpretation in terms of thermally<br />

activated relaxation of the TS for the linear T-dependence of the sound velocity. However,<br />

new acoustical results in polycrystalline materials indicate a linear T-dependence of the<br />

sound velocity comparable with those found in amorphous materials [29].<br />

Applying Eq. 23 to the measurements at the two phonon frequencies for PS <strong>and</strong> with<br />

the value of E 0 from the internal friction we obtain t 0 % (1 +0.5)610 –17 s. As discussed<br />

in Ref. [11] the meaning of the prefactor t 0 is still unknown. The very small value of t 0<br />

found for PS in comparison with the one for SiO 2 (t 0 %10 –13 s) should be taken as an ef-<br />

60


4.4 Experimental results <strong>and</strong> discussion<br />

fective value until a systematic comparison of the applied model to other experimental data<br />

is available.<br />

To obtain a maximum in the internal friction an upper limit of the potential height<br />

must be assumed. Instead of a cut-off in the distribution P (D, V) <strong>and</strong> following Ref. [11], a<br />

Gaussian distribution with a width s 0 will be assumed:<br />

P…;V†ˆ<br />

P<br />

E 0<br />

e … V2 =2 2 0 † : …24†<br />

Figures 4.13 <strong>and</strong> 4.15 show the result of numerical calculations with only the width of<br />

the distribution s as free parameter. The fits were obtained assuming s 0 = 1200 K <strong>and</strong> with<br />

the values C = 8.3610 –4 , v = 0.24 kHz, t 0 =10 –17 s <strong>and</strong> E 0 = 13 K using the equations<br />

given in Ref. [11]. A reasonable agreement with the experimental data is achieved for both<br />

acoustic properties in the expected temperature region. The minimum obtained numerically<br />

at T ~ 3 K (Fig. 4.13) has been discussed in Ref. [11] <strong>and</strong> is attributed to the difference in<br />

the number of TS contributing to the relaxation process in the tunneling or thermally activated<br />

regime.<br />

As pointed out above, the internal friction of PMMA behaves differently from PS<br />

above the plateau. It decreases at T > 3 K <strong>and</strong> reaches a minimum at T ~ 30 K. The predicted<br />

linear temperature dependence of the internal friction (Eq. 22), at the measured frequency<br />

(v = 535 Hz) is not observed for PMMA. It is tempting to interpret the decrease of<br />

the internal friction results above 3 K assuming an upper bound of the TS density of states<br />

around 15 K. This assumption has been indeed used to interpret the irreversible line broadening<br />

of the optical hole burning experiments [30]. However, ultrasonic attenuation measurements<br />

at 15 MHz for PMMA [24] show clearly a thermally activated maximum at 12 K.<br />

Therefore we have decided to search for a set of parameters that might explain the low <strong>and</strong><br />

high frequency results under the assumptions described above <strong>and</strong> as done for PS.<br />

From the sound velocity data below 20 K (Fig. 4.14) <strong>and</strong> assuming the validity of<br />

Eq. 23 (ot P1) we obtain (E 0 /k B )/ln (ot 0 ) %C/(u ln (u)/uT) ~ 0.9 +0.2. If we assume<br />

t 0 ~10 –17 s like PS, we obtain E 0 /k B ~ 30 K. With this value we are not able to obtain a reasonable<br />

set of parameters that explain the data. If we take the value for SiO 2 from Ref. [11],<br />

t 0 % 10 –13 s, we obtain E 0 /k B % (20 + 5) K, a value comparable to that for SiO 2 . After several<br />

trials, eventually we have found a set of parameters that fits reasonably the low <strong>and</strong><br />

high frequency measurements without the assumption of an energy cut-off.<br />

In Fig. 4.12 we compare the experimental results of the internal friction with the numerical<br />

calculations for frequencies at 535 Hz <strong>and</strong> at 15 MHz using E 0 /k B =10K,t 0 =10 –13 s,<br />

<strong>and</strong> a rather small distribution width s 0 = 150 K. The model reproduces fairly well the position<br />

of the maximum in the attenuation at 15 MHz as well as its relative value in comparison<br />

with the plateau [24]. It is worth pointing out, that for PMMA, since the changes produced by<br />

the free parameters E 0 , t 0 <strong>and</strong> s 0 compete with each other, it is not possible to find another set<br />

of values which can explain the results as well as we can. It is important to note that the internal<br />

friction data for both polymers indicate a much smaller thermally activated contribution to<br />

the phonon attenuation for PMMA than for PS, in spite of the fact that in the quantum tunneling<br />

regime both samples show similar results. This difference may influence the low-temperature<br />

heat release because the number of excited states that are available after the cooling process<br />

will be different.<br />

61


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

4.4.3 Heat release<br />

Figure 4.16 shows the heat release as a function of time for the 12 g PMMA sample at<br />

T = 0.090 K <strong>and</strong> after cooling from 80 K. The heat release follows very well the predicted<br />

t –1 law from the tunneling model (Eq. 12). The same result was obtained in different runs<br />

with slightly different cooling rates <strong>and</strong> measuring temperatures 0.070 K^T 0 ^0.110 K.<br />

For finite cooling rates <strong>and</strong> for charging temperatures T 1 p 1 K this independence cooling<br />

rate is to be expected (Section 4.2.3).<br />

As discussed in Section 4.2 the unknown higher-order relaxation processes are taken<br />

into account through an effective charging temperature T 1 . With the tunneling systems density<br />

of states from the specific heat <strong>and</strong> using Eq. 12 we obtain T 1 % 26 K. As stated in<br />

Section 4.2 for charging temperatures T > 10 K it is not possible to obtain reliable values of<br />

the density of states for TS from heat release measurements.<br />

From the observed t –1 -dependence <strong>and</strong> using the value for the coupling constant between<br />

TS <strong>and</strong> phonons from Ref. [24], Ak B 3 = 1.6610 9 K –3 s –1 , we obtain an upper limit for<br />

the parameter u min


4.4 Experimental results <strong>and</strong> discussion<br />

<strong>and</strong> (5). Even worse, the computed curve (5) with thermal activation shows much smaller<br />

values for the heat release by almost two orders of magnitude. This result indicates that too<br />

many energy states were depopulated during the cooling process through thermal activation.<br />

This disagreement might be ascribed to the assumed potential distribution function (Eq. 24)<br />

or the approximation l = V/E 0 . Curve (2), which is calculated without thermally activated relaxation,<br />

lies only a factor of two higher than the measured one (Fig. 4.16) <strong>and</strong> is in qualitative<br />

agreement with the acoustical data where no influence from thermal activation were taken<br />

into account.<br />

For PS, according to the specific heat <strong>and</strong> acoustical data (at T < 3 K) <strong>and</strong> due to a larger<br />

density of states of tunneling systems, we expected a larger heat release than for PMMA.<br />

Surprisingly, the opposite is observed. After cooling from 80 K with the same cooling rate,<br />

the heat release for PS measured at the same temperature as before was at least of a factor of<br />

ten smaller than for PMMA (Figure 4.16), <strong>and</strong> it does not follow the t –1 -dependence.<br />

In order to underst<strong>and</strong> the observed deviations we have measured the heat release of<br />

PS at lower charging temperatures. Figure 4.17 shows the heat release as a function of time<br />

at charging temperatures 0.5 K^T 1 ^1 K. The heat release follows very well the t –1 -dependence<br />

as well as an increase with T 1 2 (Eq. 12, Fig. 4.18). It is interesting to note that a<br />

fair quantitative agreement with the prediction of the tunneling model is obtained. If we calculate<br />

the density of states of tunneling systems P from these data with Eq. 12 (straight line<br />

in Fig. 4.18) we obtain P = 9.0610 –38 J –1 g –1 which is similar to the one obtained from the<br />

specific heat assuming u min =6610 –9 .<br />

Figure 4.17: Heat release as a function of time (in hours) for Polystyrene at a measuring temperature<br />

T 0 = 0.200 K cooling from different T 1 . The straight lines have a t –1 -dependence.<br />

At higher temperatures T 1 we observe the expected saturation of the heat release<br />

(Fig. 4.18), however, no deviation from the t –1 -dependence has been measured within experimental<br />

error. Deviations are observed if we charge the sample at temperatures T 1 >15K<br />

(Fig. 4.16). In this case we have cooled the sample from 80 K <strong>and</strong> measured the heat release<br />

at 0.090 K (sample PS1) <strong>and</strong> 0.300 K (sample PS2). According to theoretical estimates the<br />

observed difference in the heat release between the two PS samples is not attributed to the<br />

difference in measuring temperatures. The rather abrupt saturation of the heat release at<br />

T 1 6 7 K is attributed to the depopulation of the excited states through thermally activated<br />

relaxation in the cooling process (Figs. 4.5 <strong>and</strong> 4.18), as it will become clear below.<br />

63


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

Figure 4.18: Heat release multiplied by the time t as a function of the difference between the squares of<br />

charging temperature T 1 <strong>and</strong> measuring temperature T 0 for PS. The solid line follows the theoretical<br />

prediction with a density of states of tunneling systems 1.45 larger than the one obtained from the specific<br />

heat results.<br />

The following experiment provides further verification that the non-simple time dependence<br />

of the heat release for PS is due to energy states excited above 15 K <strong>and</strong> depopulation<br />

during the cooling process, very likely through thermally activated relaxation. We have<br />

charged the sample at 80 K for several hours <strong>and</strong> cooled it then to 1 K. We have left the sample<br />

for 22 h at 1 K <strong>and</strong> later continued to 0.3 K, the temperature at which the heat release<br />

was measured (Curve (1) in Fig. 4.19). We observe now a clear deviation from the t –1 law.<br />

That means that even after 22 h at 1 K the energy states excited above 15 K have not relaxed<br />

completely. After 60 h at 0.3 K we warmed the sample to 1 K for 17 h, cooled it to 0.3 K<br />

<strong>and</strong> measured the heat release (Curve (2) in Fig. 4.19). The heat release follows the theoretical<br />

t –1 -dependence very well.<br />

We have calculated the heat release taking one-phonon,thermally activated processes,<br />

<strong>and</strong> the experimental cooling process into account. We used the same parameters obtained<br />

from the fit to the acoustical data (Figs. 4.13 <strong>and</strong> 4.15). With only one free parameter, u min ,<br />

Figure 4.19: Heat release as a function of time for PS. (1): cooled from 80 K to 1 K <strong>and</strong> leaving the<br />

sample 17 h at this temperature, T 0 = 0.300 K. (2): the sample was warmed to a charging temperature<br />

T 1 = 1 K for 22 h <strong>and</strong> cooled again to T 0 = 0.300 K. The solid line in (1) is only a guide <strong>and</strong> the solid<br />

line (2) has a t –1 -dependence.<br />

64


4.5 Conclusions<br />

we can reproduce the measured heat release of PS reasonably well (Fig. 4.16). In this figure<br />

we present the two curves (3) <strong>and</strong> (4) using the parameter u min =10 –10 <strong>and</strong> 6610 –9 . This<br />

fit indicates that we can observe the influence of a finite u min through the steeper decrease<br />

of the heat release at long times. Curve (1) in Fig. 4.16 was calculated with parameters for<br />

PS but without thermally activated relaxation.<br />

4.5 Conclusions<br />

We conclude that thermal conductivity, specific heat <strong>and</strong> acoustical properties at low temperatures<br />

(T < 1 K) can be quantitatively interpreted within the tunneling model for PMMA<br />

<strong>and</strong> PS. At higher temperatures we have measured very different temperature dependence of<br />

the internal friction for the two polymers. At the used frequencies <strong>and</strong> at 0.1 K^T^120 K<br />

PMMA shows no thermally activated dissipation peak. PMMA shows a minimum in the internal<br />

friction at T ~ 30 K where PS shows a maximum. Nevertheless, within the tunneling<br />

model <strong>and</strong> introducing a thermally activated relaxation rate as well as a Gaussian distributon<br />

density of states of tunneling systems, it is possible to underst<strong>and</strong> the acoustical properties.<br />

From the fits we may conclude that the density of states of tunneling systems for PMMA is<br />

restricted to smaller energies, i. e. a smaller width s 0 , than for PS.<br />

Although the thermal conductivity, specific heat, <strong>and</strong> sound velocity temperature dependence<br />

for both samples are similar, the heat release shows different time dependences as<br />

well as different values in apparent contradiction with the other measured low-temperature<br />

properties. Taking into account the internal friction results, it is tempting to correlate the differences<br />

in the heat release between both samples to the difference in the contribution of<br />

thermally activated processes in the relaxation of the tunneling states during the cooling process.<br />

We have demonstrated that the deviations of the time dependence of the heat release<br />

from the t –1 -dependence predicted by the tunneling model is due to the relaxation of energy<br />

states excited above 15 K. The smaller values of the measured heat release of PS in comparison<br />

with the st<strong>and</strong>ard tunneling model can be explained by the depopulation of energy<br />

states relaxing mainly by thermally activated processes during the cooling process.<br />

Without detailed knowledge of the short relaxation times at T 1 > 1 K, a quantitative<br />

fit or even a qualitative underst<strong>and</strong>ing of the heat release _Q(T 1 ,T 0 ,t) data for high charging<br />

temperatures seems to be impossible. Therefore, the heat release with high charging temperatures<br />

can be hardly used to obtain information on the density of states of tunneling systems.<br />

Heat release experiments at high measuring temperatures T 0 > 1 K are necessary to<br />

obtain information on the dynamics of the tunneling systems with a relaxation rate not given<br />

by the tunneling one-phonon process. These experiments may clarify the influence of the<br />

two different relaxation mechanisms discussed nowadays in the literature to interpret the<br />

low-temperature properties of amorphous solids above 1 K, namely incoherent tunneling<br />

65


4 Low-Temperature Heat Release, Sound Velocity <strong>and</strong> Attenuation, …<br />

[31] <strong>and</strong> thermal activation within the soft potential model [8]. Both approaches have been<br />

used with impressive success in the last years. It is, however, not yet clear if incoherent tunneling,<br />

a mixture of the two processes (incoherent <strong>and</strong> thermal activation), or only thermal<br />

activation is the main relaxation mechanism of the TS above 1 K.<br />

Acknowledgements<br />

We wish to acknowledge W. Joy for the sample preparation, <strong>and</strong> S. Hunklinger, D. Haarer<br />

<strong>and</strong> J. Friedrich for valuable discussions. This work was supported by the Deutsche Forschungsgemeinschaft<br />

through the Sonderforschungsbereich 213. A. Nittke was supported by<br />

the “Graduiertenkolleg Po 88/13” of the Deutsche Forschungsgemeinschaft.<br />

References<br />

1. J. Zimmerman, G. Weber: Phys. Rev. Lett., 46, 661 (1981)<br />

2. M. Schwark, F. Pobell, M. Kubota, R.M. Mueller: J. Low Temp. Phys., 58, 171 (1985)<br />

3. S. Sahling, A. Sahling, M. Kolac: Solid State Commun., 65, 1031 (1988)<br />

4. P.W. Anderson, B.I. Halperin, C.M. Varma: Phil. Mag., 25, 1 (1972); W.A. Phillips, J. Low Temp.<br />

Physics 7, 351 (1972)<br />

5. M. Deye, P. Esquinazi: Z. Phys. B – Condensed Matter 76, 283 (1989)<br />

6. M. Deye, P. Esquinazi: in: S. Hunklinger, W. Ludwig, G. Weiss (eds.): Phonon 89, World Scientific,<br />

Singapore, p. 468, (1990)<br />

7. J. Jäckle: Z. Phys., 257, 212 (1972)<br />

8. D. Parshin, S. Sahling: Phys. Rev. B, 47, 5677 (1993)<br />

9. J.J. Brey, A. Prados: Phys. Rev. B, 43, 8350 (1991)<br />

10. P. Esquinazi, M. Scherl, Li Junyun, F. Pobell: in: M. Meissner, R. Pohl (eds): 4th International<br />

Conference on Phonon Scattering in Condensed Matter, Springer Series in Solid State Sciences,<br />

Vol. 112, p. 287 (1993)<br />

11. D. Tielbürger, R. Merz, R. Ehrenfels, S. Hunklinger: Phys. Rev. B, 45, 2750 (1992)<br />

12. P. Doussineau, C. Frenois, R. G. Leisure, A. Levelut, J. Y. Prieur: J. Phys. (Paris), 41, 1193 (1980)<br />

13. S. Hunklinger, W. Arnold: in: W. P. Mason, R. N. Thurston (eds.): Physical Acoustics, Vol XII,<br />

Academic Press, New York, 1976<br />

14. W.A. Phillips: in: S. Hunklinger, W. Ludwig, G. Weiss (eds.): Phonon 89, World Scientific, Singapore,<br />

p. 367 (1990)<br />

15. K.-P. Müller, D. Haarer: Phys. Rev. Lett., 66, 2344 (1991)<br />

16. A.K. Raychaudhuri, S. Hunklinger: Z. Phys. B – Condensed Matter, 57, 113 (1984)<br />

17. R.B. Stephens: Phys. Rev. B, 8, 2896 (1973)<br />

18. R.B. Stephens, G.S. Cieloszyk, G.L. Salinger: Physics Letters, 28A, 215 (1972)<br />

66


References<br />

19. D. Cahill, R. Pohel: Phys. Rev. B, 35, 4067 (1987)<br />

20. C. Choy, G. Salinger, Y. Chiang: J. of Applied Physics, 41, 597 (1970)<br />

21. J. Freeman, A. C. Anderson: Phys. Rev. B, 34, 5684 (1986)<br />

22. J. Mack, J. Freeman, A.C. Anderson: J. of Non Crystalline Solids, 91, 391 (1987)<br />

23. G. Hartwig: Progr. Colloid <strong>and</strong> Polymer Sci., 64, 56 (1978)<br />

24. G. Federle, S. Hunklinger: J. de Physique, C9(12), Tome 43, p. C9–505 (1982); G. Federle: PhD<br />

Thesis, M. Planck Stuttgart, 1983, unpublished; A.K. Raychaudhuri: in: T.V. Ramakrishnan, M.<br />

Raj Lakshmi (eds.): Non-Debye Relaxation in Condensed Matter, World Scientific, p. 193 (1987)<br />

25. J. Crissman, J. Sauer, E. Woodward: J. of Polymer Science, A2, 5075 (1964)<br />

26. G. Bellesa: Phys. Rev. Lett., 40, 1456 (1978)<br />

27. J.-Y. Duquesne, G. Bellesa: J. de Physique Lettres, 40, L-193 (1979)<br />

28. R. Nava, R. Oentrich: J. of Alloys <strong>and</strong> Compounds, in press; R. Nava, Phys. Rev. B, 49, 4295<br />

(1994)<br />

29. E. Gagemidze, P. Esquinazi, R. König: Europhys. Letters, 31, 13 (1995)<br />

30. W. Köhler, J. Zollfrank, J. Friedrich: Phys. Rev. B, 39, 5414 (1989)<br />

31. A. Würger: in: From coherent tunneling to relaxation, Springer Tracts in Modern Physics,Vol. 135,<br />

(1997)<br />

67


5 Spectral Diffusion due to Tunneling Processes<br />

at very low Temperatures<br />

Hans Maier, Karl-Peter Müller, Siegbert Jahn, <strong>and</strong> Dietrich Haarer<br />

5.1 Introduction<br />

Pioneered by the work of Zeller <strong>and</strong> Pohl [1] it was discovered in the early 1970s that amorphous<br />

solids show low-temperature thermal <strong>and</strong> acoustic properties which are very different<br />

from those observed in crystals. For reviews see for example Refs. [2, 3]. The most wellknown<br />

of these anomalous properties is the specific heat, which is in general considerably<br />

larger than would be expected from the Debye model <strong>and</strong> varies linear with temperature in<br />

contrast to the Debye T 3 -dependence. Other anomalies are the temperature dependence of<br />

the thermal conductivity, the properties of phonon echoes, <strong>and</strong> ultrasonic absorption. These<br />

features seem to be quite universal for all kinds of amorphous solids, irrespective of their<br />

chemical composition <strong>and</strong> structure, i. e. inorganic as well as organic or polymeric.<br />

Very soon after the first experimental evidence for these anomalous low-temperature<br />

excitations Anderson et al. [4] <strong>and</strong> independently Phillips [5] developed a theoretical description,<br />

called the tunneling model [4, 5]. It is based on the assumption that the low energy<br />

degrees of freedom in amorphous solids arise from tunneling motions of atoms or groups of<br />

atoms between local energetic minima which should exist in any disordered solid. If only<br />

the lowest energy level of each minimum is considered, this leads to two eigenstates. For<br />

this reason, these degrees of freedom are often referred to as two-level system (TLS). An<br />

important consequence of this model is the formal analogy to a system of particles with<br />

spin 1/2, which for example immediately explains the existence of phonon echoes. Another<br />

very striking feature of the tunneling model was the prediction of a time-dependent specific<br />

heat [4]. This was confirmed experimentally several years later [6]. Furthermore, the model<br />

can explain the heat release [7] of amorphous solids, which is a time-dependent non-equilibrium<br />

phenomenon. From these kinds of experiments, it could be concluded that TLS dynamics<br />

occur on time scales extending over many orders of magnitude.<br />

The sensitivity of optical experiments on amorphous solids was hindered, in many instances,<br />

by the large inhomogeneous broadening which arises from the distribution of local<br />

environments of the involved optical transitions. Therefore a very important step was the<br />

discovery of persistent spectral hole burning in 1974 [8, 9]. This method of high-resolution<br />

laser spectroscopy eliminates the inhomogeneous effects induced by the static disorder in<br />

68 Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


5.2 The optical cryostat<br />

amorphous hosts. For this purpose dye molecules are embedded in the solid which change<br />

their absorption spectra when irradiated with resonant narrow b<strong>and</strong> laser light. This produces<br />

a dip in the inhomogeneous absorption b<strong>and</strong>, called a spectral hole. A narrow spectral hole<br />

can be regarded as a highly sensitive spectral probe for any kind of distortion of the matrix,<br />

which induces small spectral changes [10]. In this way variations of strain fields caused by<br />

pressure changes of almost some 10 hPa can be detected [11]. Spectral holes are also sensitive<br />

to small changes of other external parameters including electric fields [12, 13].<br />

After observing quite a few anomalous properties of optical transitions in glasses <strong>and</strong><br />

attributing them to the dynamics of TLS [14], the tunneling model was adopted by Reinecke<br />

[15] to explain the low-temperature line widths of optical transitions in amorphous solids<br />

using the concept of spectral diffusion. This concept had originally been developed for the<br />

description of spin resonance experiments [16] <strong>and</strong> had already been applied to the theoretical<br />

treatment of the above mentioned ultrasonic properties of glasses [17]. Soon after this<br />

step, the possibility of a connection between thermal <strong>and</strong> optical properties of amorphous<br />

solids was supported by the observation of time dependence of spectral hole widths [18].<br />

The application of the tunneling model to the description of spectroscopic properties<br />

proposed an important link suggesting a correlation between the specific heat <strong>and</strong> the optical<br />

line width. This implies that, besides the mentioned calorimetric <strong>and</strong> acoustic methods,<br />

optical spectroscopy yields information about the dynamics of the same low energy excitations<br />

of amorphous solids, which dominates the calorimetric experiments. A crucial point in<br />

establishing this connection is the temperature dependence of these physical solid state parameters.<br />

The tunneling model in its original form predicts linear temperature dependence for<br />

the specific heat as well as for the optical line width. This prediction, however, is only valid<br />

at temperatures where the contribution of phonons is of minor importance. Earlier hole burning<br />

measurements on the temperature dependence of optical line widths [19] indicated the<br />

necessity of exp<strong>and</strong>ing the temperature range accessible to optical absorption spectroscopy<br />

to values far below 1 Kelvin in order to exclude any other mechanisms except TLS dynamics.<br />

For this reason we have constructed a 3 He/ 4 He dilution refrigerator with optical<br />

windows allowing transmission spectroscopy at temperatures down to 0.025 K, a temperature<br />

regime in which only very few data from optical spectroscopy existed before [20, 21].<br />

In addition, cooling with a 3 He/ 4 He mixture can be performed continuously for very long<br />

times (in principle unlimited) if the necessary care in cryostat design <strong>and</strong> operation is taken.<br />

In our experiments we have reached operation times up to three months.<br />

5.2 The optical cryostat<br />

In the design of a cryostat for optical spectroscopy two main problems have to be overcome.<br />

In an absorption experiment the sample has to dissipate the absorbed energy in order to<br />

avoid overheat during the measurement. Therefore the best possible thermal contact to the<br />

cold stage is of extreme importance. The second problem is how to guide the laser light to<br />

the sample without creating an intolerably large additional heat leak, for example by infrared<br />

69


5 Spectral Diffusion due to Tunneling Processes at very low Temperatures<br />

irradiation from the windows. Since various construction methods for optical cryostats in the<br />

temperature range below 0.500 K did not yield the desired results, special care had to be taken<br />

in the design of our dilution refrigerator. For example it had been reported [22] that for<br />

samples located outside the mixing chamber temperatures below 0.300 K could not be<br />

reached, even if the best thermal contact is realized by pressing the samples to the wall of<br />

the mixing chamber with indium metal as contact material. In another kind of experiment a<br />

glass fibre winding around the cold finger of a dilution refrigerator shows a significant temperature<br />

gradient with respect to the mixing chamber below 0.100 K [20]. For these reasons,<br />

we decided to place the sample directly into the dilute phase of the mixing chamber. This<br />

way optimal thermal contact <strong>and</strong> a constant temperature of the sample was achieved via the<br />

superfluid 4 He. For performing optical experiments the mixing chamber was designed as a<br />

glass cylinder.<br />

The optical path, used in our cryostat, is shown in Figure 5.1 [23]. Three sets of windows<br />

are mounted on cold copper shields to eliminate most of the room temperature radiation<br />

in three steps: at liquid nitrogen temperature (77 K), liquid helium temperature (4.2 K),<br />

<strong>and</strong> at approximately 1 K, the latter being achieved by cooling with the pumped 3 He of the<br />

distillation chamber. In contrast to other constructions, the vessel containing the dilution refrigerator<br />

part of the cryostat is in our design not surrounded by the helium tank which has<br />

the advantage that the laser beam is not scattered by boiling liquid helium. We estimate that<br />

the total heat leak caused by room temperature radiation is below 50 nW. The minimum<br />

temperature reached is 0.024 K. The cooling power is about 6 mW at 0.100 K. Although this<br />

is a very low value, it is sufficient since the power of the absorbed laser light was always far<br />

below 1 mW. The light powers used for hole burning in our experiments were on the order of<br />

several nanowatts to several tens of nanowatts depending on the temperature. For hole detection<br />

the power is reduced to the picowatt range. The cooling power of our cryostat <strong>and</strong> the<br />

thermal conductivity of the involved polymers are high enough to guarantee a uniform temperature<br />

distribution with no significant sample heating during the hole burning process.<br />

4<br />

He - shield mixing - chamber<br />

LN 2 - shield 1-K - shield<br />

sample<br />

Figure 5.1: Alignment of optical windows <strong>and</strong> sample in the cryostat. The sample is placed inside the<br />

mixing chamber of the dilution refrigerator. The shields are connected to the distillation chamber, the<br />

helium tank, <strong>and</strong> the nitrogen tank, respectively. The liquid N 2 windows possess an infrared reflective<br />

coating.<br />

70


5.3 Theoretical considerations<br />

The temperature is measured with a RuO 2 thick film resistor (TFR) supplied by Phillips<br />

(type RC-01). The heat capacity <strong>and</strong> thermal relaxation time of the TFR are equivalent<br />

to those of a carbon resistor but its thermal reproducibility during temperature cycles is better<br />

[24]. The resistor is calibrated against several primary <strong>and</strong> secondary thermometers including<br />

a NBS fixed point device, CMN <strong>and</strong> Ge, thermometers. Its accuracy is better than<br />

2% in the temperature range between 4.2 K <strong>and</strong> 0.025 K. The thermometer is placed into<br />

the dilute phase of the mixing chamber in direct thermal contact to the sample. The resistance<br />

is measured using a four wire ac resistance bridge (AVS-46 RV Elektroniikka, Finl<strong>and</strong>).<br />

At the minimum temperature, the energy dissipated in the resistor is less than 0.5 pW.<br />

5.3 Theoretical considerations<br />

In the tunneling model [4] the TLS was described by two parameters (Fig. 5.2), the asymmetry<br />

parameter D <strong>and</strong> the overlap parameter l, which contains the barrier height, the distance<br />

of the two minima, <strong>and</strong> the mass of the tunneling particle. They are related with the total energy<br />

splitting E of the two levels <strong>and</strong> the tunneling matrix element D 0 by the relations<br />

q<br />

E ˆ 2 ‡ 2 0<br />

<strong>and</strong> 0 ˆ k exp… † ; …1†<br />

where kO is a typical zero point energy. One of the most important ingredients of the TLS<br />

model is the distribution function characterizing the two parameters. The originally used assumption<br />

for this function, which has also been applied to explain spectral diffusion [15,<br />

17], is a flat distribution in both parameters, namely<br />

P…;†dd ˆ Pdd:<br />

…2†<br />

Figure 5.2: Double minimum potential as a model for a TLS. All relevant parameters are shown in the<br />

figure.<br />

71


5 Spectral Diffusion due to Tunneling Processes at very low Temperatures<br />

It is convenient, however, to use experimentally accessible quantities as variables, i. e.<br />

the energy E <strong>and</strong> a dimensionless relaxation rate R, which can be defined as: R = r/r max =<br />

D 2 0 /E 2 . The transformation to the new variables yields for the distribution function [25]:<br />

1<br />

P…E;R† ˆ P p : …3†<br />

2R 1 R<br />

A very important feature of this distribution is the fact that P (E, R) is independent of<br />

E. In order to keep the total number of TLS finite, some cut-off value for R ? 0 has to be<br />

introduced [3], which is usually denoted by R min .<br />

It was shown in Ref. [15] that for optical transitions in glasses the TLS dynamics results<br />

in spectral diffusion, which shows up in the experiment as a time <strong>and</strong> temperature-dependent<br />

Lorentzian line broadening. The width of this Lorentzian line must be calculated by<br />

averaging over the distribution of energies <strong>and</strong> relaxation rates P(E, R). It can be written as:<br />

…t; T† ˆ2p2 <br />

3k C ij<br />

<br />

<br />

Z Emax<br />

E min<br />

Z 1<br />

dEn…T†<br />

R min<br />

dR E P…E;R†…1 exp‰ tr max RŠ† : …4†<br />

Here AC ij S represents an average coupling constant between the TLS <strong>and</strong> the optical<br />

transition, n (T) is a thermal occupation factor of the TLS states. For the system which is investigated<br />

in this work, the parameter r max can be estimated from experimental data on ultrasonic<br />

attenuation [26] to be on the order of 10 7 s –1 at a temperature of 0.100 K <strong>and</strong> even larger<br />

at higher temperatures. Therefore, the condition t 7r max p 1 is well fulfilled on all time<br />

scales exceeding several microseconds. In this limit the rate integration in Eq. 4 can be performed<br />

analytically [27], which leads to the well-known logarithmic time dependence of<br />

spectral diffusion:<br />

!…t; T† ˆ p2 <br />

3k C <br />

ij P k B T ln…r max t† : …5†<br />

Equation 5 represents the theoretical prediction of the tunneling model for the time<br />

<strong>and</strong> temperature-dependent broadening of spectral holes. This result, however, is the result<br />

of a particular form of the density of tunneling states P (E, R), which is based on the a<br />

priori assumption of Eq. 2. A uniform density of states in D is a physically reasonable<br />

choice; the independence of l, however, is difficult to justify, since l consists of several<br />

parameters [28]. The temperature dependence of spectral diffusion is dominated by D the<br />

time evolution stems mainly from l. Therefore, these two predictions from Eq. 5 do not<br />

have the same validity. In our experiments we have investigated time <strong>and</strong> temperature dependence<br />

separately.<br />

72


5.4 Temperature dependence<br />

5.4 Temperature dependence<br />

In these experiments, we investigated polystyrene (PS) doped with phthalocyanine <strong>and</strong> polymethylmethacrylate<br />

(PMMA) doped with tetra-4-tert-butyl phthalocyanine. Both sample<br />

materials have also been investigated by heat release <strong>and</strong> specific heat measurements [29].<br />

The samples had optical densities of about 0.4 at a typical thickness of 3 mm. The samples<br />

were prepared by bulk polymerisation of the solution of the dye in the monomer.<br />

The observed hole shapes are Lorentzian with no detectable deviations in all cases. To<br />

eliminate saturation effects at least 10 holes with different energies were burned at each temperature<br />

<strong>and</strong> the homogeneous line width was determined by extrapolation to zero burning<br />

fluence [30]. Due to their low relative depths the extrapolation was done assuming a linear<br />

dependence of the hole area on energy [10]. A weak variation of the line width with light<br />

power was observed at our minimum temperature of 0.025 K for PMMA. We attribute this<br />

variation to sample heating.<br />

The temperature dependence of the hole width is shown in Fig. 5.3 [23]. The results are:<br />

a) both systems display a linear temperature dependence over the whole temperature range<br />

from 0.500 K down to 0.025 K <strong>and</strong> no crossover to a constant line width is seen. The linearity<br />

of the data plot shows that the density of states P (E, R) is indeed independent of the<br />

TLS energy;<br />

b) the relative magnitude of the line width in the two systems correlates with the respective<br />

specific heat [31];<br />

c) the extrapolated line width for T ? 0 is nearly the same for both systems, confirming<br />

that in an amorphous system the line width reaches the lifetime limited value G h = 1/2 pT 1<br />

(Heisenberg limit) of the electronic transition only when T ?0. This limit is reached in our<br />

experiment within 10%!<br />

Figure 5.3: Temperature dependence of the hole burning line width for PS <strong>and</strong> PMMA between<br />

0.025 K <strong>and</strong> 0.500 K.<br />

73


5 Spectral Diffusion due to Tunneling Processes at very low Temperatures<br />

In the case of PS, where unsubstituted phthalocyanine was used as dye, the result for<br />

the extrapolated line width G (T ? 0) is in very good agreement with the fluorescence lifetime<br />

measurements of this molecule [32]. We are not aware of measurements of the fluorescence<br />

lifetime of substituted phthalocyanine, but our results show that there is no major difference<br />

between both dyes as far as their excited state lifetimes are concerned.<br />

These results imply that in amorphous solids there are dynamical processes of twolevel<br />

systems with an energy spectrum extending down to values as low as k B 70.025 K.<br />

These low-energy excitations dominate the line widths of optical transitions even at low<br />

temperatures <strong>and</strong> the lifetime limited value can only be reached by extrapolating to zero<br />

temperature.<br />

5.5 Time dependence<br />

After the system has been cooled down spectral holes are burned immediately <strong>and</strong> after<br />

some delay times, while the sample was at constant temperature, we observe a hole broadening<br />

which depends on this delay time. This behaviour is demonstrated in Fig. 5.4. Dataset A<br />

is the broadening of a hole which is burned immediately after the sample (phthalocyanine in<br />

PMMA) has been cooled from liquid nitrogen temperature to 0.300 K. The curves show the<br />

evolution of holes burned about one day, three days, <strong>and</strong> one week after cooling down. The<br />

amount of spectral diffusion in a given time interval decreases continuously until a small residual<br />

effect becomes observable, which is independent of the time delay after reaching the<br />

final temperature. This is represented by dataset B in Fig. 5.4.<br />

A<br />

B<br />

Figure 5.4: Time evolution of spectral holes burned at different times after sample cooling. Datasets<br />

corresponding to A <strong>and</strong> B are investigated in the following.<br />

74


5.5 Time dependence<br />

Investigating the temperature dependence of the time evolution of the hole width [33],<br />

we performed experiments on a PMMA sample doped with phthalocyanine at different temperatures<br />

T, varying from 0.100 K to 1 K. For each of the temperatures T = 0.100, 0.300,<br />

0.500, <strong>and</strong> 0.700 K, the respective experiments were carried out in a separate run. It took<br />

about one hour to cool the system from 77 K to 4 K. The time for cooling down from 4 K<br />

to the final temperature T depends slightly on T, but is in the range of about (6+1) hours.<br />

Immediately after reaching T, holes are burned <strong>and</strong> their subsequent broadening is observed<br />

for about one week. These experiments yield data corresponding to set A in Fig. 5.4.<br />

The results of these measurements are shown in Fig. 5.5 [33]. In this plot the time origin for<br />

each run is identical, namely the time at which the cooling down procedure from 77 K was<br />

started. For better visibility, however, the data corresponding to T = 0.500, 0.300, <strong>and</strong><br />

0.100 K are shifted by 0.5, 1.0, <strong>and</strong> 1.5 orders of magnitude along the logarithmic time axis,<br />

respectively. Except for the data taken at 0.500 K the total observation time was 6–10 days.<br />

Figure 5.5: Broadening of holes burned immediately after reaching T. Data shifted for better visibility<br />

(see text).<br />

Keeping the samples at constant temperature new holes are burned after 1–2 weeks<br />

<strong>and</strong> their broadening is observed for several hours, corresponding to set B in Fig. 5.4. These<br />

data are shown in Fig. 5.6 [33]. Here the observation time is 2–16 hours. Note that the range<br />

of the DG-axis is about 10 times less as compared with Fig. 5.5.<br />

Comparing our data of Figs. 5.5 <strong>and</strong> 5.6, it is obvious that the measurements which are<br />

started immediately after cooling down show no variation of the observed broadening with the<br />

phonon bath temperature, while the residual behaviour observed many days later exhibits a<br />

rather pronounced dependence on the experimental temperature. It has to be emphasized that<br />

the data shown in Fig. 5.5 <strong>and</strong> the corresponding data in Fig. 5.6 for the same values of Twere<br />

measured without any change of temperature in between the two subsequent experiments.<br />

In spite of the variation of the final temperature over almost one order of magnitude,<br />

all four datasets in Fig. 5.5 show the same hole broadening behaviour.<br />

We attribute this to the existence of a non-equilibrium state of the TLS ensemble due<br />

to the fast cooling procedure: The large amount of spectral diffusion observed after fast<br />

75


5 Spectral Diffusion due to Tunneling Processes at very low Temperatures<br />

Figure 5.6: Broadening of holes burnt after 1–2 weeks at low temperature.<br />

cooling is due to the relaxation of TLS to their thermal equilibrium [34]. This relaxation<br />

process is known to produce the thermal phenomenon of heat release [7]. Thus, it can be<br />

concluded that we have established an important connection between optical <strong>and</strong> calorimetric<br />

phenomena.<br />

The data of Fig. 5.6 exhibit a pronounced dependence on temperature, as can be expected<br />

from theoretical considerations outlined in Section 5.3 <strong>and</strong> from the non-time resolved<br />

experiments of Section 5.4. The time evolution, however, shows a distinct non-logarithmic<br />

behaviour in contradiction to the prediction of the tunneling model. For short times<br />

the Fig. 5.6 shows clearly that the hole broadening starts fairly logarithmic. We have performed<br />

transient hole burning experiments on the millisecond time scale, which also yielded<br />

a logarithmic behaviour [35]. At longer times, however, the diffusional broadening increases<br />

faster than logarithmic. This increase occurs later at lower temperatures, which is to be expected<br />

from the temperature dependence of the relaxation rates. These new features are in<br />

qualitative agreement with theoretical work that invokes strong coupling of TLS with phonons<br />

[28, 36].<br />

We believe that our experimental work shows that the method of spectral hole burning<br />

spectroscopy is suitable for studying the low-temperature properties of amorphous solids.<br />

Especially for the investigation of the long-time behaviour of TLS equilibrium dynamics,<br />

where other methods cease to function, it is a very powerful experimental technique.<br />

References<br />

1. R.C. Zeller, R.O. Pohl: Phys. Rev. B, 4, 2029 (1971)<br />

2. Amorphous Solids. Low Temperature <strong>Properties</strong>, in:W.A. Phillips (ed.): Topics in current Physics,<br />

Vol. 24, Springer, Berlin, (1981)<br />

76


References<br />

3. S. Hunklinger, A. K. Raychaudhuri: in: D. F. Brewer (ed.), Progress in Low Temp. Physics,Vol. IX,<br />

Elsevier Science, Amsterdam, p. 265 (1986)<br />

4. P.W. Anderson, B.I. Halperin, C.M. Varma: Philos. Mag., 25, 1 (1972)<br />

5. W.A. Phillips: J. Low Temp. Phys., 7, 351 (1972)<br />

6. M.T. Loponen, R.C. Dynes,V. Narayanamurti, J.P. Garno: Phys. Rev. Lett., 45, 457 (1980)<br />

7. J. Zimmermann, G. Weber: Phys. Rev. Lett., 46, 661 (1981)<br />

8. B.M. Kharlamov, R.I. Personov, L.A. Bykovskaya: Opt. Commun., 12, 191 (1974)<br />

9. A.A. Gorokhovskii, R.K. Kaarli, L.A. Rebane: JETP Lett., 20, 216 (1974)<br />

10. J. Friedrich, D. Haarer: Angew. Chem., 96, 96 (1984); Angew. Chem. Int. Ed. Engl., 23, 113<br />

(1984)<br />

11. Th. Sesselmann, W. Richter, D. Haarer, H. Morawitz: Phys. Rev. B, 36(14), 7601 (1987)<br />

12. A.P. Marchetti, M. Scozzafara, R.H. Young: Chem. Phys. Lett., 51(3), 424 (1977)<br />

13. V.D. Samoilenko, N.V. Rasumova, R.I. Personov: Opt. Spectr., 52(4), 580 (in Russian) (1982)<br />

14. P.M. Selzer, D.L. Huber, D.S. Hamilton, W.M. Yen, M.J. Weber: Phys. Rev. Lett., 36, 813 (1976)<br />

15. T.L. Reinecke: Solid State Commun., 32, 1103 (1979)<br />

16. J.R. Klauder, P.W. Anderson: Phys. Rev., 125(3), 912 (1962)<br />

17. J.L. Black, B. I. Halperin: Phys. Rev. B, 16, 2879 (1977)<br />

18. W. Breinl, J. Friedrich, D. Haarer: J. Chem. Phys., 81(9), 3915 (1984)<br />

19. G. Schulte, W. Grond, D. Haarer, R. Silbey: J. Chem. Phys., 88(1), 679 (1988)<br />

20. M.M. Broer, B. Golding, W.H. Haemmerle, J.R. Simpson, D.L. Huber: Phys. Rev. B, 33, 4160<br />

(1986)<br />

21. A. Gorokhovskii,V. Korrovits, V. Palm, M. Trummal: Chem. Phys. Lett., 125, 355 (1986)<br />

22. Korrovits, M. Trummal: Proceedings of the Academy of Sciences of the Estonian SSR, 35(2), 198<br />

(1986)<br />

23. K.-P. Müller, D. Haarer: Phys. Rev. Lett., 66, 2344 (1991)<br />

24. W.A. Bosch, F. Mathu, H.C. Meijer, R.W. Willekers: Cryogenics, 26, 3 (1985)<br />

25. J. Jäckle: Z. Phys., 257, 212 (1972)<br />

26. G. Federle: PhD thesis, Max-Planck-Institut für Festkörperforschung Stuttgart (1983)<br />

27. S. Hunklinger, M. Schmidt: Z. Phys. B, 54, 93 (1984)<br />

28. K. Kassner: Z. Phys. B, 81, 245 (1990)<br />

29. A. Nittke, M. Scherl, P. Esquinazi, W. Lorenz, Junyun Li, F. Pobell: J. Low Temp. Phys., 98(5/6),<br />

517 (1995)<br />

30. L. Kador, G. Schulte, D. Haarer: J. Phys. Chem., 90, 1264 (1986)<br />

31. K.P. Müller: PhD thesis, Universität Bayreuth (1991)<br />

32. W.H. Chen, K.E. Rieckhoff, E.M. Voigt, L.W. Thewalt: Mol. Phys., 67(6), 1439 (1989)<br />

33. H. Maier, D. Haarer: J. Lum., 64, 87 (1995)<br />

34. S. Jahn, K.-P. Müller, D. Haarer: J. Opt. Soc. Am. B, 9, 925 (1992)<br />

35. S. Jahn, D. Haarer, B.M. Kharlamov: Chem. Phys. Lett., 181, 31 (1991)<br />

36. K. Kassner, R. Silbey: J. Phys. Condens. Matter, 1, 4599 (1989)<br />

77


6 Optically Induced Spectral Diffusion in Polymers<br />

Containing Water Molecules: A TLS Model System<br />

Klaus Barth, Dietrich Haarer, <strong>and</strong> Wolfgang Richter<br />

6.1 Introduction<br />

In the past decade much experimental <strong>and</strong> theoretical work was done on spectral diffusion in<br />

amorphous solids. It turned out that at low temperatures, optical dephasing phenomena [1]<br />

as well as the numerous caloric data [2] can be well described by assuming the presence of<br />

low-energy excitations, the so-called two-level systems (TLS). Photochemical hole burning<br />

[3] <strong>and</strong> optical echo experiments [4] have provided experimental evidence for different dynamics<br />

of optical transitions in glassy systems as compared to crystalline matrices. For<br />

glassy organic solids the TLS concept was first proposed by Small <strong>and</strong> co-workers [5, 6]<br />

<strong>and</strong> was later used by Reinecke [7] to explain low-temperature optical line widths. Especially<br />

the photochemical hole burning data over long observation periods (observation times<br />

longer than three months) at temperatures down to 0.050 K allowed a critical test of the theoretical<br />

approach within the tunneling model [8]. Theoretical <strong>and</strong> experimental results to<br />

this topic are also given in this book by H. Maier et al.<br />

At temperatures above 1 K tunneling transitions <strong>and</strong> localized vibrations influence the<br />

physical properties [9]. Both types of mechanisms have recently been incorporated in the<br />

soft potential model [10] which contains the well-known tunneling model as a special case.<br />

A microscopic interpretation of optical line broadening phenomena in terms of e. g.<br />

local excitations of the matrix or switching of optically addressed chemical groups suffered<br />

from the lack of experimental data. In the past, several experimental techniques have been<br />

developed to shed light on the microscopic mechanisms of the phenomena. These techniques<br />

are all based on the generation of phonons in the matrix material, especially in the neighbourhood<br />

of the spectral probe. Heat pulses from external heaters [11] produce a broad distribution<br />

of phonon frequencies inside the chromophore host system <strong>and</strong> permit the investigation<br />

of the dynamics of barrier crossings <strong>and</strong> the coupling between the TLS <strong>and</strong> the dye<br />

molecules [12]. Light-induced phonon generation in hole burning systems can in principle<br />

be achieved either by exciting non-radiative decay processes in appropriate chromophores<br />

[13] or by direct absorption of IR light in the matrix material [13, 14]. The non-radiative decay<br />

method, which has been used to study transient <strong>and</strong> irreversible spectral diffusion processes,<br />

yields also a broad distribution of phonon frequencies <strong>and</strong> gives rise to similar experimental<br />

processes as e. g. the heat pulse technique. On the other h<strong>and</strong>, the IR absorption<br />

78 Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


6.2 Experimental setup for burning <strong>and</strong> detecting spectral holes<br />

method where the sample is illuminated with narrow b<strong>and</strong> IR light allows us to generate<br />

phonons of high density <strong>and</strong> of a comparatively small energy distribution in the direct neighbourhood<br />

of the optical probe. The contribution of the broad b<strong>and</strong> black-body background<br />

radiation can be minimized by using an appropriate intensity of a narrow b<strong>and</strong> IR source.<br />

For the optical investigation of amorphous systems, photochemical hole burning (PHB)<br />

is a high-resolution technique <strong>and</strong> a powerful tool to examine local sites in polymers at low<br />

temperatures. Dye molecules embedded in polymers at low concentration exhibit strongly inhomogeneously<br />

broadened absorption b<strong>and</strong>s in a conventional spectroscopic experiment. Selecting<br />

an ensemble of these dye molecules with narrow b<strong>and</strong> laser light, which gives rise to a<br />

photoreaction, leads to a sharp spectral dip at the laser frequency, the so-called spectral hole.<br />

Due to the high resolution, this method is appropriate for detecting small perturbations in the<br />

matrix such as a transient change of the phonon distribution or permanent matrix rearrangements<br />

in the vicinity of the dye molecules. Such rearrangements can be due to a configurational<br />

change of parts of the polymer main chain or due to a reorientation of small molecules<br />

(water), which may be embedded in the matrix in addition to the dye molecules.<br />

In the following, two different experimental conditions are discussed which are based<br />

on the absorption of IR light. Both experimental situations lead to a change of the local interaction<br />

of the optical probe with its local matrix environment. In the first part the phonons<br />

are treated as a time-dependent temperature bath. In the second part experiments are discussed<br />

where local groups of the matrix are selectively addressed by IR radiation of very<br />

low intensity. The observed enhanced spectral diffusion shows a characteristic dependence<br />

on the IR frequency <strong>and</strong> coincides with only a few of the numerous IR absorption b<strong>and</strong>s. A<br />

quantitative description of this new process together with the identification of the resonant<br />

vibrations is given within a simple kinetic model.<br />

6.2 Experimental setup for burning <strong>and</strong> detecting spectral holes<br />

The experimental setup for burning <strong>and</strong> detecting spectral holes with a narrow b<strong>and</strong> laser is<br />

described elsewhere [16]. The experiments were performed at 1.8 K in a bath cryostat in<br />

which the sample was immersed in superfluid helium. As samples we used different polymeric<br />

matrices such as polymethylmethacrylate (PMMA), polyamide (PA), polyethylene, polystyrene,<br />

etc. doped with free base phthalocyanine (H 2 Pc) at low concentration (10 –2 mol%). Their<br />

preparation is described in Ref. [16]. The water content of the samples has been controlled by<br />

heating them under vacuum conditions. Because of its strong hydrophilic nature PMMA has<br />

a high capacity of water absorption at atmospheric conditions. In order to investigate the influence<br />

of the water molecules the experiments, described in Section 6.4, were carried out either<br />

with a nearly water-free PMMA sample or with a PMMA sample containing water molecules<br />

under equilibrium conditions. The optical windows of the cryostat consisted of crystalline<br />

BaF 2 which allowed the study of hole burning spectra in the lowest vibrational b<strong>and</strong> of the<br />

electronic S 0 ? S 1 transition of H 2 Pc at about 0.69 mm as well as under IR illumination between<br />

2 mm <strong>and</strong> 11 mm. In this setup, the broad b<strong>and</strong> background radiation with an emission<br />

79


6 Optically Induced Spectral Diffusion in Polymers Containing Water Molecules<br />

maximum at 10 mm has an integrated intensity of about 300 mW/cm 2 at the location of the<br />

sample. As IR radiation sources we used a Globar with a maximum emission intensity near<br />

3 mm <strong>and</strong> different CO 2 laser lines around 10 mm.<br />

An acousto-optic modulator was used for reducing the intensity of the CO 2 laser down<br />

to 1 W/cm 2 <strong>and</strong> for switching with times some microseconds. The radiation of the Globar<br />

was dispersed in a monochromator <strong>and</strong> with different dielectric filters to a b<strong>and</strong>width between<br />

0.05 mm <strong>and</strong> 0.2 mm with typical intensities around 100 mW/cm 2 .<br />

6.3 Reversible line broadening phenomena<br />

In this Section the mechanisms of reversible line broadening phenomena under continuous<br />

or pulsed irradiation conditions are discussed. Phonons are generated by illuminating the<br />

polymeric sample with the light of a single CO 2 laser line. The wavelength of the CO 2 laser<br />

was selected to generate high frequency phonons only within a small penetration depth of<br />

less than 10 mm. The dye molecules, which are homogeneously distributed throughout the<br />

sample thickness of some 100 mm, are a local probe for the increase in the phonon density<br />

during the IR irradiation. Due to the thermalisation processes, phonons show a broad energy<br />

distribution. This change of the phonon distribution gives rise to an increased effective temperature<br />

under cw irradiation conditions. Because of the high cooling rate of the surrounding<br />

superfluid helium, typical irradiation intensities of about 1 W/cm 2 were used to obtain a<br />

significant increase in the phonon density. The influence on the homogeneous line width of<br />

the dye molecules can be investigated in a zero burning fluence experiment, i. e. the limit of<br />

the hole width is given by extrapolating to zero burning fluence. The lower curve in Fig. 6.1,<br />

which was measured without any IR irradiation, yields an extrapolated value for the line<br />

Figure 6.1: Zero burning fluence experiment under different phonon density conditions in polyamide<br />

doped with H 2 Pc (see text).<br />

80


6.3 Reversible line broadening phenomena<br />

width of about 375 MHz. The two upper curves show increased hole widths due to an enhanced<br />

phonon generation either by IR irradiation during the burning process only(triangles)<br />

or during the detection process only(circles).<br />

The profile of a spectral hole can be written as<br />

Z 1 1<br />

A…! ! L †/<br />

nz…! 0 ! L †z…! ! 0 †d! 0 ; …1†<br />

where nz(o' – o L ) is the site distribution function of the molecular line contributing to the<br />

hole spectrum <strong>and</strong> z(o – o') is the absorption profile of a single molecule.<br />

On the time scale of this experiment (up to some minutes) the spectral hole profiles<br />

are mainly reversible (see also Fig. 6.2 <strong>and</strong> Fig. 6.3). Therefore, the homogeneous line profile<br />

function z(o) is much more affected by the altered phonon density than the site distribution<br />

function nz(o). Thus, it is obvious that the two upper curves in Fig. 6.1 are similar <strong>and</strong><br />

give for the hole width nearly the same zero burning fluence value of 440 MHz. Both curves<br />

are well separated from the lower curve which was obtained without any additional phonon<br />

generation. Applying an effective temperature model the difference between these two zero<br />

burning fluence values corresponds to a temperature change of about 0.3 K within the sample.<br />

Because of the strongly inhomogeneous experimental conditions during phonon generation<br />

<strong>and</strong> phonon diffusion through the sample, however, each different subensemble of dye<br />

molecules experiences a different distribution in phonon energies <strong>and</strong>, as a consequence, a<br />

bath description in terms of one single temperature is not appropriate. By using different op-<br />

Figure 6.2: Lower part: Spectral profile of a photochemical hole with <strong>and</strong> without phonon generation<br />

during the detection process. Upper part: Directly measured difference between these two profiles as<br />

obtained by modulating the IR light at a frequency of 150 Hz.<br />

81


6 Optically Induced Spectral Diffusion in Polymers Containing Water Molecules<br />

Figure 6.3: Decay of the transmission maximum during a laser pulse.<br />

tical energies for the phonon generation, it is in principle possible to study the contribution<br />

of different matrix vibrations on the dephasing of a single absorber function z(o).<br />

The time scale of the hole burning experiments discussed above is several minutes. It<br />

is about 3 orders of magnitude larger than the typical thermal relaxation time of the sample<br />

as given by<br />

ˆ d2 c<br />

<br />

; …2†<br />

where d is the dimension (thickness) of the sample, c the specific heat, r the mass density,<br />

<strong>and</strong> k the thermal conductivity. For the temperature of our experiment (1.8 K) <strong>and</strong> a typical<br />

sample dimension of 1 mm the thermal relaxation time t is calculated to 10 ms.<br />

A consequence of this short relaxation time is shown in Fig. 6.2 <strong>and</strong> Fig. 6.3. The two<br />

spectral holes in the lower part of Fig. 6.2 were obtained with <strong>and</strong> without phonon generation<br />

during the hole detection process. A modulation of the IR light during the detection<br />

process results in a continuous switching between these two hole profiles. In the upper part<br />

of Fig. 6.2, a modulation frequency of 150 Hz was used during the scan of the optical frequency.<br />

The switching time between the two hole profiles is of the same order of magnitude<br />

as the thermal relaxation time. The continuous switching between the two hole profiles is<br />

detected with lock-in technique. The result shows that the phonon generation <strong>and</strong> relaxation<br />

process is sufficiently fast to follow the 150 Hz modulation of the IR light.<br />

A time-resolved experiment is shown in Fig. 6.3. The optical transmission at the peak<br />

of a hole spectrum was monitored while the sample was being illuminated with a single<br />

CO 2 laser pulse of 600 ms duration. The decay of the transmission signal cannot be described<br />

with a single exponential function. It is well represented by the superposition of two<br />

exponentials with a fast contribution t 1 & 10 ms <strong>and</strong> a slow contribution t 2 & 150 ms. The<br />

fast contribution is close to the thermal relaxation time at the temperature of the helium<br />

bath. The slow contribution contains all the inhomogeneous experimental conditions as mentioned<br />

above, in particular a temperature gradient inside the sample that gives rise to a variation<br />

of the thermal relaxation time. Since the amplitude of the fast contribution is larger by<br />

a factor of 4, the switching time between the two curves in Fig. 6.2 is mainly determined by<br />

the fast contribution t 1 .<br />

82


6.4 Induced spectral diffusion<br />

The experimental results of this Section show that optically generated phonons can be<br />

used to study the transient broadening of the optical line shape of a single absorber. In principle,<br />

the dependence on the phonon frequency can be studied in such an experiment. In the<br />

same experiment the time-resolved dynamics of phonon diffusion <strong>and</strong> phonon decay in polymeric<br />

systems can be investigated via the dephasing mechanism of the optical probe.<br />

6.4 Induced spectral diffusion<br />

In this Section we will demonstrate that the resonant absorption of IR light in a polymeric<br />

system can also lead to irreversible line broadening phenomena. Even at a very low irradiation<br />

intensity (a typical value is 100 mW/cm 2 ) the investigation of specific types of relaxation<br />

processes is possible. Such low irradiation intensities do not increase the temperature of<br />

the matrix by more than 0.01 K <strong>and</strong> hence yield no measurable dephasing effect via a<br />

change of the phonon distribution. Because of the irreversible nature of the optical line<br />

broadening, such phenomena are usually ascribed to spectral diffusion processes. The IR induced<br />

spectral diffusion is caused by addressing specific moieties within the matrix via the<br />

excitation of local vibrations. In the following weakly bound water molecules are studied.<br />

When a change in the spatial configuration takes place during the relaxation process of a<br />

water molecule the optical transition of a neighbouring dye molecule will be affected. Optical<br />

absorbers in glasses are therefore very sensitive probes for local rearrangements <strong>and</strong> are<br />

suited for a microscopic study of this topic.<br />

Due to the hydrophilic nature of the matrix the investigated guest host system, PMMA<br />

doped with H 2 Pc, contains a comparatively large number of water molecules (up to one volume<br />

percent). The experimental procedure is the following: A spectral hole is burned <strong>and</strong> its<br />

spectrum is repeatedly recorded while the sample is being exposed to IR radiation of a certain<br />

wavelength at a constant irradiation intensity. Figure 6.4 shows the typical increase of the hole<br />

Figure 6.4: Time evolution of the hole broadening induced by different IR wavelengths. From top to<br />

bottom: 2.80 mm, 2.86 mm, 2.90 mm, 2.96 mm, <strong>and</strong> 3.04 mm.<br />

83


6 Optically Induced Spectral Diffusion in Polymers Containing Water Molecules<br />

Figure 6.5: Transmission spectra of PMMA without (curve A) <strong>and</strong> with (curve B) natural water content.<br />

The quotient C = B/A shows the H 2 O absorption lines. Dependence of the hole broadening on the<br />

IR wavelength after t 0 = 35 min (curve D) due to induced spectral diffusion.<br />

width <strong>and</strong> the strong dependence of this behaviour on the IR wavelength. To illustrate the pronounced<br />

spectral selectivity, one value of each hole broadening curve at t 0 = 35 min is plotted<br />

in Fig. 6.5 (curve D) versus the corresponding IR wavelength. The data points do not simply<br />

reflect the absorption behaviour of the sample. Although the PMMA matrix has many absorption<br />

b<strong>and</strong>s between 2 mm <strong>and</strong> 11 mm (curve B in Fig. 6.5), there are only two distinct wavelengths<br />

which give rise to induced spectral diffusion processes.<br />

The small concentration of water molecules in the matrix gives rise to non-saturated<br />

absorption b<strong>and</strong>s near 2.8 mm <strong>and</strong> 6.1 mm. They correspond to the fundamental stretching<br />

<strong>and</strong> bending vibrations of H 2 O, respectively. The difference spectrum (curve C in Fig. 6.5)<br />

of water free <strong>and</strong> water saturated PMMA yields the exact position of the H 2 O absorption<br />

b<strong>and</strong>s, which are in good agreement with the resonances found in our experiment. On an exp<strong>and</strong>ed<br />

scale in Fig. 6.5, even the asymmetric shape of the inhomogeneous H 2 O b<strong>and</strong> is reflected<br />

by the experimental data. Replacing protonated with deuterated water yields a red<br />

shift of the IR resonances which also agrees very well with the corresponding absorption<br />

spectra of heavy water.<br />

In analogy to the tunneling model, which is based on the assumption of two potential<br />

minima, we assume two stable sites for each water molecule (Fig. 6.6). Since the experimental<br />

results indicate that a permanent change in the matrix occurs transitions between the two<br />

sites are only allowed via the first excited vibrational states. Using such a simple four-level<br />

model for the underlying kinetics, we are able to explain the temporal behaviour of the induced<br />

spectral diffusion in a quantitative fashion.<br />

84


6.4 Induced spectral diffusion<br />

Figure 6.6: Energy level scheme of a H 2 O molecule in a two-site model.<br />

The two ground states are labelled 1 <strong>and</strong> 2, the first vibrationally excited states 3 <strong>and</strong><br />

4, respectively. The IR induced <strong>and</strong> the spontaneous transition rate for each site are denoted<br />

by the Einstein coefficients B <strong>and</strong> A, respectively. In our notation, B is proportional to the irradiated<br />

IR intensity. Transitions between the two sites are denoted by the conversion rate k.<br />

The transitions shown in Fig. 6.6 lead to a system of 4 coupled linear rate equations.<br />

Since we used in our experiments very low intensities of the IR light, we consider only the<br />

limit B P A where the population of the excited levels 3 <strong>and</strong> 4 is negligible at all times. The<br />

result is a time-dependent number of flips between the two ground states n f = n 1?2 + n 2?1 .<br />

Only those flips are taken into account which contribute to a change in the total configuration<br />

of all water molecules with respect to the time t = 0 of the hole burning process.<br />

According to Reinecke [7] the width of the spectral diffusion kernel Do is proportional<br />

to the number of flips n f . Though using the approximation that each flip yields the<br />

same contribution to Do, when taking into account the decrease of the IR intensity across<br />

the sample thickness as well as a r<strong>and</strong>om orientation of the H 2 O molecules, the result for<br />

Do can only be calculated numerically. Using a Taylor expansion, however, a very good analytical<br />

approximation of the width of the induced spectral diffusion is given by<br />

!…t† ˆG N <br />

0<br />

4pk 1 1 ‡ 2Bp !<br />

f t b<br />

: …3†<br />

b<br />

The difference between this formula <strong>and</strong> the numerical result is less than 5%. p f is the<br />

flip probability of a single H 2 O molecule as defined by p f = k /(A + k), b is a dimensionless<br />

geometry parameter of about 2, N 0 is the density of the H 2 O molecules <strong>and</strong> G is the coupling<br />

constant between dye <strong>and</strong> water molecules. Equation 3 describes very well the dependence<br />

of the broadening of a spectral hole on the number of absorbed IR photons (Fig. 6.4).<br />

It is possible to perform a test of the model. Equation 3 contains only two fitting parameters,<br />

the flip probability p f , <strong>and</strong> the coupling constant G. These two parameters can be obtained<br />

from a single experiment.<br />

An example is given in Fig. 6.7. One set of the experimental data (filled circles) is obtained<br />

at an IR intensity of 80 µW/cm 2 . The solid line is a fit according to Eq. 3 <strong>and</strong> yields<br />

a flip probability of 18% <strong>and</strong> a coupling constant of G = 3.6 7 10 –44 Jcm 3 .(Note: the slight<br />

difference with respect to the data given in reference [17] is due to the extension of the<br />

model by including the r<strong>and</strong>om orientation of the water molecules <strong>and</strong> the decrease of the<br />

IR intensity according to Beer’s law.) With these two values it is possible to calculate the<br />

85


6 Optically Induced Spectral Diffusion in Polymers Containing Water Molecules<br />

Time (s)<br />

Figure 6.7: Test of the model. Hole broadening versus IR irradiation time for three different IR intensities.<br />

The solid line is a fit, the dashed lines are calculated according to Equation 3.<br />

time dependence of the induced spectral diffusion for any other irradiation intensity. The<br />

dotted lines in Fig. 6.7 are calculated with Eq. 3 for the obtained values <strong>and</strong> the intensities<br />

230 µW/cm 2 <strong>and</strong> 20 µW/cm 2 . The experimental results under these conditions, represented<br />

by the squares <strong>and</strong> triangles, are in very good agreement with the theoretical predictions.<br />

The high quantum yield of 18% for the flip of a single water molecule becomes obvious<br />

by a comparison of the number of absorbed IR photons (typically 10 14 s –1 ) with the<br />

total number of water molecules (about 10 18 ) in the sample volume. Since a strong induced<br />

spectral diffusion is observed within several minutes, a local reorientation process with a<br />

high quantum yield must be involved.<br />

The coupling strength between water <strong>and</strong> dye molecules can be estimated by taking<br />

into account the static dipole moment of the H 2 O molecule. A reorientation of the H 2 O molecules<br />

affects the dye molecules by the concomitant change in the local electric field at the<br />

location of the chromophores. Assuming a mean distance of 15 Å <strong>and</strong> inserting data from<br />

Stark effect experiments [18], one obtains a value of approximately 2710 –44 Jcm 3 for the<br />

coupling constant which is very close to our fitted value. Together with the resonant phenomena<br />

discussed above, this yields a detailed microscopic underst<strong>and</strong>ing of this special<br />

kind of interaction.<br />

We can conclude that spectral diffusion induced by the absorption of IR photons in<br />

the frequency range 2–11 µm is a resonant process. Its microscopic origin is a spatial rearrangement<br />

of weakly bound water molecules after vibrational excitation. Assuming a configurational<br />

model with only two possible sites for each water molecule <strong>and</strong> applying the theoretical<br />

model of spectral diffusion by Reinecke [7], an analytical description of the resonant<br />

hole broadening behaviour can be obtained. A variation of the IR irradiation intensity shows<br />

that the model description yields excellent agreement with the experimental data.<br />

86


References<br />

References<br />

1. S. Jahn, K. P. Müller, D. Haarer: J. Opt. Soc. Am. B, 9, 925 (1992)<br />

2. M. Deye, P. Esquinazi: Zeitschr. für Phys. B, 39, 283 (1989)<br />

3. J. Friedrich, D. Haarer: Angew. Chem., Int. Ed. Engl,. 23, 113 (1984)<br />

4. C. A. Walsh, M. Berg, L. R. Narashiman, M. D. Fayer: Chem. Phys. Lett., 130, 6 (1986)<br />

5. J. M. Hayes G. J. Small: Chem. Phys., 27, 151 (1978)<br />

6. G. J. Small: Persistent nonphotochemical hole burning <strong>and</strong> the dephasing of impurity electronic<br />

transitions in organic glasses, in: V. M. Agranovich R. M. Hochstrasser (eds.): Spectroscopy <strong>and</strong><br />

Excitation Dynamics of Condensed Molecular Systems, North-Holl<strong>and</strong>, Amsterdam, p. 515 (1983)<br />

7. T. L. Reinecke: Sol. Stat. Com., 32, 1103 (1979)<br />

8. H. Maier, D. Haarer: J. Lumin., 64, 87 (1995)<br />

9. R. Greenfield, Y. S. Bai, M. D. Fayer: Chem. Phys. Lett., 170, 133 (1990)<br />

10. D. A. Parshin: Phys. Sol. State, 36, 991 (1994)<br />

11. U. Bogner: Phys. Rev. Lett., 37, 909 (1976)<br />

12. T. Attenberger, K. Beck, U. Bogner: in: S. Hunklinger, W. Ludwig, G. Weiss (eds.): Proc. of the<br />

3rd Int. Conf. on Phonon Physics, p. 555<br />

13. A. A. Gorokhovskii, G. S. Zavt, V. V. Palm: JETP Lett., 48, 369 (1988)<br />

14. W. Richter, Th. Sesselmann, D. Haarer: Chem. Phys. Lett., 159, 235 (1989)<br />

15. W. Richter, M. Lieberth, D. Haarer: JOSA B, 9, 715 (1992)<br />

16. G. Schulte, W. Grond, D. Haarer, R. Silbey: J. Chem. Phys., 88, 679 (1988)<br />

17. K. Barth W. Richter: J. of Lumin., 64, 63 (1995)<br />

18. R. B. Altmann, I. Renge, L. Kador, D. Haarer: J. Chem. Phys., 97, 5316 (1992)<br />

87


7 Slave-Boson Approach to Strongly Correlated Electron<br />

Systems<br />

Holger Fehske, Martin Deeg, <strong>and</strong> Helmut Büttner<br />

7.1 Introduction<br />

The problem of underst<strong>and</strong>ing high-temperature superconductivity has been a challenge to<br />

theoreticians from a wide variety of fields. Many theoretical investigations have been carried<br />

out in order to identify the mechanism of this fascinating phenomenon as well as to establish<br />

the canonical model itself [1]. Although excellent progress is being made in deducing<br />

a consistent description of the high-temperature superconductivity systems from a<br />

first-principles theory, at present no microscopic theory can account for their unconventional<br />

normal-phase data in its entirety. The interplay of charge <strong>and</strong> spin dynamics in the<br />

normal state seems to hold the key to the underst<strong>and</strong>ing of the physical mechanism behind<br />

high-temperature superconductivity in the cuprates. Both macroscopic measurements of<br />

transport <strong>and</strong> magnetic properties as well as microscopic measurements probing the charge<br />

<strong>and</strong> spin excitation spectra are fundamental in establishing the anomalous normal-state<br />

properties of these materials [1].<br />

The nature of spin excitations of the high-temperature superconductivity cuprates has<br />

been experimentally studied by means of nuclear magnetic/quadrupole resonance (NMR/<br />

NQR) <strong>and</strong> inelastic neutron scattering (INS) techniques clarifying the persistence of strong<br />

antiferromagnetic (AFM) correlations in the normal <strong>and</strong> superconducting states [2]. Detailed<br />

investigations of the wave-vector dependence of the low-frequency spin fluctuation spectrum<br />

have revealed remarkable differences between the YBa 2 Cu 3 O 6+x <strong>and</strong> La 2–x Sr x CuO 4 families<br />

at low doping level x; the dynamic structure factor S (~q, o) keeps its maximum at (p,p) in<br />

the YBCO system, while in LSCO, the peaks are displaced from the commensurate position<br />

to the four incommensurate wave-vectors p(1 +q 0 , 1), p (1,1+q 0 ), where q 0 F2x. On the<br />

other h<strong>and</strong>, the analysis of the NMR data shows that the relaxation rates on the planar Cu<br />

sites are similar in all materials.<br />

Two striking features are associated with 63 T 1 –1 :<br />

a) as a result of strong local spin fluctuations on 63 Cu sites it is enhanced by one order of<br />

magnitude over the oxygen rate 17 T 1 –1 ; <strong>and</strong><br />

b) in sharp contrast to the Korringa-like behaviour at the planar 17 O sites, its temperature<br />

dependence does not follow the Korringa law [2].<br />

88 Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


7.1 Introduction<br />

The physical origin of the contrasting ~q-dependence of the spin fluctuation spectrum<br />

is still under discussion. Millis <strong>and</strong> Monien [3] have argued that the spin dynamics <strong>and</strong>, in<br />

particular, the temperature dependence of the spin susceptibility w s (T) in LSCO are caused<br />

by a spin density wave instability, whereas in the YBCO family they are due to in-plane<br />

AFM fluctuations <strong>and</strong> a novel non-Fermi liquid spin singlet pairing of electrons in adjacent<br />

planes. On the other h<strong>and</strong>, the magnetic properties are intimately related to the energy b<strong>and</strong><br />

dispersion of the non-interacting system within a Fermi liquid based framework, i. e., in this<br />

way the observed spin dynamics can be attributed to different Fermi surface (FS) geometry<br />

of LSCO-type <strong>and</strong> YBCO-type, respectively. Along this line, details of the spin fluctuation<br />

spectrum are studied using a nearly antiferromagnetic Fermi liquid approach by Monthoux<br />

<strong>and</strong> Pines [4]. From a more microscopic point of view, the important effects of FS shape on<br />

the magnetic properties were confirmed by Si et al. [5] within a large Coulomb-U auxiliary<br />

boson scheme <strong>and</strong> by Fukuyama <strong>and</strong> co-workers [6] on the basis of a resonating valence<br />

bond (RVB) slave-boson mean-field approach to the extended t-J model. Furthermore, Ito et<br />

al. [7] have recently reported that the charge transport in the CuO 2 plane is determined by<br />

dominant spin scattering, i. e., the spin dynamics are manifest in the extraordinary transport<br />

properties of the high-temperature superconductivity cuprates as well.<br />

Encouraged by these findings it is the aim of this report to study magnetic <strong>and</strong> transport<br />

phenomena of high-temperature superconductors probably in terms of the most simple<br />

effective one-b<strong>and</strong> model describing both correlation <strong>and</strong> b<strong>and</strong> structure effects, the socalled<br />

t-t'-J model:<br />

H t t 0 J ˆ t X hi;ji;<br />

~c y i ~c j t 0 X<br />

h<br />

hi;jii;<br />

~c y i ~c j ‡ J X ij h i<br />

<br />

~n i ~n j<br />

~S i<br />

~S j<br />

4<br />

: …1†<br />

H t–t'–J acts in a projected Hilbert space without double occupancy, where ~c …y†<br />

c …y†<br />

i …1 ~n i † is the electron annihilation (creation) operator, ~S i ˆ 1<br />

P<br />

0 ~cy i ~ 0 ~c i<br />

i =<br />

0;<br />

2<br />

<strong>and</strong><br />

~n i ˆ P ~cy i ~c i. J measures the AFM exchange interaction, t <strong>and</strong> t' denotes hopping processes<br />

between nearest-neighbour (NN; Ai,jS) <strong>and</strong> next nearest-neighbour (NNN; AAi,jSS) sites<br />

on a square lattice. Compared to the original t-J model the t'-term incorporates several important<br />

effects near half-filling. Starting from a rather complex three-b<strong>and</strong> Hubbard or Emery<br />

model [8] for the CuO 2 planes, quantum cluster calculations [9] have revealed that the<br />

relative large direct transfer between NN oxygen sites (t pp < t pd /2) leads to a sizeable NNN<br />

hopping t' in the context of an effective one-b<strong>and</strong> description. More recently the t'-term has<br />

been introduced to reproduce the FS geometry observed in ARPES experiments [6, 10–12].<br />

Fitting the quasi-particle dispersion relation<br />

" ~k ˆ 2t…cos k x ‡ cos k y † 4t 0 cos k x cos k y ; …2†<br />

involved in Eq. 1 to experimental <strong>and</strong> b<strong>and</strong> theory results yields t in the order of 0.3 eV <strong>and</strong>,<br />

e. g., for the case of YBCO, t'&–0.4 t [4, 5, 13]. Moreover, Tohyama <strong>and</strong> Maekawa [12]<br />

have emphasized that a t-t'-J model with t' > 0 can be used to describe the electron-doped<br />

systems, e. g. Nd 2–x Ce x CuO 4 (NCCO). In this case one has to shift the momentum ~ k ? ~ k +<br />

(p,p) [12], i. e., within a b<strong>and</strong>-filling scenario one obtains a hole pocket-like FS centred at<br />

(p,p)-point which shrinks with increasing doping [14]. <strong>Final</strong>ly, as pointed out by Lee [15],<br />

89


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

in a locally AFM environment doped holes can propagate coherently only on the same sublattice<br />

without disturbing spins. Therefore, the t'-term coupling the same sublattice becomes<br />

crucial for the low-lying magnetic excitations. This clearly is a correlation effect related to<br />

the NNN hopping processes.<br />

7.2 Slave-boson theory for the t-t'-J model<br />

Apart from numerical techniques, the slave-particle methods have been employed extensively<br />

in studying the effects of electron-electron correlation in the (extended) t-J model [6, 16–21].<br />

The two commonly used approaches, the NZA slave-boson [16, 17] <strong>and</strong> slave-fermion [18]<br />

schemes, however, yield quite different results concerning the (mean-field) ground-state phase<br />

diagram <strong>and</strong> spin/charge excitations for this model [22]. As yet, the relationship between both<br />

types of slave-field theories is not well understood. Within the NZA formulation, for example,<br />

the Hamiltonian can be solved by a mean-field approximation with the (uniform) RVB order<br />

parameter. A serious difficulty of this approach is the absence of AFM correlations [18, 22],<br />

e. g., in the half-filled case, the lowest energy state, the energy of which is considerable higher<br />

than numerical estimates indicate, does not satisfy the Marshall sign rule <strong>and</strong> fails to show the<br />

expected long-range Néel order [23]. On the other h<strong>and</strong>, the mean-field slave-fermion schemes<br />

[18] are known to give reasonable results for the spin susceptibility as well as for the spin correlation<br />

length [21, 22], but also suffer from neglecting important correlation effects, especially<br />

the fermion charge degrees of freedom are not described sufficiently well. In contrast, the fourfield<br />

slave-boson (SB) technique, introduced by Kotliar <strong>and</strong> Ruckenstein (KR) [24] in the context<br />

of the Hubbard model, has the advantage of treating spin <strong>and</strong> charge degrees of freedom on<br />

an equal footing. Starting from the scalar KR SB representation of the Hubbard model, one may<br />

generate an intersite exchange interaction via a loop expansion in the coherent state functional<br />

integral [19]. However, the effective t-J Lagrangian, derived in this way, contains only an Ising<br />

interaction term, i. e., important spin-flip exchange processes are neglected. As we shall see below,<br />

to bosonize the complete exchange interaction term one has to use the spin-rotation-invariant<br />

(SRI) extension of the KR SB theory from the beginning [25, 26].<br />

7.2.1 SU(2)-invariant slave-particle representation<br />

For the sake of definiteness, we return to the extended t-J Hamiltonian (Eq. 1), that may be<br />

cast into the form<br />

H t t 0 J ˆ X<br />

t ij<br />

~ y i<br />

i;j<br />

~ j<br />

J<br />

4<br />

X <br />

hiji<br />

~ y i <br />

<br />

~ i ~ y<br />

j ~ <br />

j : …3†<br />

90


7.2 Slave-boson theory for the t-t'J model<br />

In order to bring out the SU(2) symmetry of the system, we have used in Eq. 3 a spinor<br />

representation, where the one-row [one-column] matrices ~ y i ˆ…~ y i ; ~ i †‰ ~ i Š are built up<br />

by the projected fermion creation [annihilation] operators ~ y i ~cy i ‰ ~ i ~c i Š: ‰ Š denotes<br />

the contravariant [covariant] four-component vector (m =0,x, y, z) of Pauli’s matrices.<br />

To preserve SRI, we apply to H t-J the manifest [SU(2) 6 U(1)] invariant SB scheme [26,<br />

27] based on the SRI SB approach developed for the Hubbard model by Li et al. [28]. Accordingly,<br />

we define scalar boson fields e ({) i <strong>and</strong> bosonic matrix operators p ({) i (representing<br />

empty <strong>and</strong> singly occupied sites, respectively) <strong>and</strong> pseudofermion spinor fields C ({) i in the<br />

following way:<br />

j0 i i ˆ e y i jvaci;<br />

j i i ˆ X y<br />

<br />

i py i vac<br />

j i: …4†<br />

The unphysical states in the extended Fock space of pseudofermionic <strong>and</strong> bosonic<br />

states are eliminated by imposing two sets of local constraints:<br />

C …1†<br />

i<br />

ˆ e y i e i ‡ 2Trp y i p i<br />

1 ˆ 0 ; …5†<br />

expressing the completeness of the bosonic projectors, <strong>and</strong><br />

C …2†<br />

i<br />

ˆ i <br />

y<br />

i ‡ 2p y i p i<br />

o ˆ 0 ;<br />

…6†<br />

relating the pseudofermion number to the number of p-type slave-bosons (hereafter underbars<br />

denote a 262 matrix in the spin variables). Obviously, double occupancy has been projected<br />

out. In the transformation of the fermionic spinor fields analogous to [26, 28],<br />

~ i ! z i i ; …7†<br />

the non-linear bosonic hopping operators<br />

z i ˆ‰ o 2p y i p i Š 1=2 e y i ‰1 ‡ ey i e i ‡ 2Trpy i p i Š1=2 p i<br />

‰…1 e y i e i † o 2 ~p y i ~p i Š 1=2 …8†<br />

yield a correlation-induced b<strong>and</strong> renormalization where ~p ({)<br />

irr' = rr'p ({)<br />

i,–r',r. Exploiting the<br />

SU(2),O(3) homomorphism, the matrix operators p i<br />

may be decomposed into scalar (singlet)<br />

p o ({) <strong>and</strong> vector (triplet) ~p i =(p ix , p iy , p iz ) components as<br />

p …y†<br />

i<br />

ˆ 1<br />

2<br />

X<br />

<br />

p …y†<br />

i ;<br />

…9†<br />

where the p im obey the usual Bose commutation rules ‰p i; p y j 0Šˆ ij 0. Consequently, the pseudofermions<br />

i ˆ… i; i † T satisfy anticommutation relations of the form f i ; y<br />

j 0gˆ ij 0.<br />

91


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

This way the interaction term is converted due to<br />

~ y i ~ i ! 2Trp y i p i<br />

; …10†<br />

where, more explicitly, the locally defined particle number <strong>and</strong> spin operators are just<br />

~n i ! ~n i …p i †ˆ2Trp y p ˆ P p y p ;<br />

i i i i<br />

~S i ! ~S i …p i †ˆTrp y i<br />

<br />

~p i ˆ 1<br />

2 …py io ~p i ‡ ~p y i p io i~p y i ~p i † : …11†<br />

Actually, the components of the bosonized SB spin operator act as generators of rotations<br />

in spin space, i. e., ~S i …p i † satisfies the spin algebra. Since C (1) i <strong>and</strong> C …2†<br />

i commute with<br />

the SB t-J Hamiltonian, the constraints in Eq. 5 <strong>and</strong> Eq. 6 can be ensured by introducing the<br />

time-independent Lagrange multipliers …1†<br />

i <strong>and</strong> …2†<br />

i<br />

ˆ P<br />

…2†<br />

i .1 As a consequence, the<br />

Hamiltonian of the t-J model (Eq. 3) in terms of the slave-boson <strong>and</strong> pseudofermion operators<br />

has to be replaced by<br />

H SB<br />

t t 0 J ˆ P<br />

t ij<br />

i;j<br />

‡ P i<br />

y<br />

i zy i z j j<br />

… …1†<br />

i C …1†<br />

i<br />

J P <br />

<br />

Trp y i p i<br />

Trp y j p j<br />

hiji<br />

‡ Tr …2†<br />

i C …2†<br />

i † : …12†<br />

In the physical subspace, H SB<br />

t–J possesses the same matrix elements for the basis states<br />

(Eq. 4) as the original t-J Hamiltonian (Eq. 3) for the purely fermionic states. To verify this<br />

directly, special attention has to be paid to the bosonization of AFM exchange interaction<br />

term<br />

<br />

~ y <br />

i ~ i<br />

~ y j ~ <br />

j ˆ 2 P …~c y i ~cy j ~c j ~c i <br />

~c y i ~cy j ~c j ~c i † ; …13†<br />

<br />

<br />

which includes besides the Ising exchange contributions # i " j $ <br />

#i " j the spin-flip processes<br />

# i " j $ "i # j . Let us emphasize that the matrix elements of the spin-flip terms<br />

are not reproduced in the scalar KR SB approach [19]. By contrast, within the SRI SB approach<br />

it is a straightforward exercise to show that these contributions can be expressed in<br />

terms of the p ({) irr' :<br />

X<br />

p y i py j 0 p j p <br />

1<br />

0 i i j ˆ<br />

4 <br />

<br />

i j :<br />

0<br />

…14†<br />

Thus our SRI SB scheme (Eqs. 4–12) provides a consistent bosonization of the extended<br />

t-J model.<br />

1 Note that additional constraints do not exist. Especially, if the constraint ~p i ~p i ˆ 0 is added [25] the<br />

spin algebra is not satisfied.<br />

92


7.2 Slave-boson theory for the t-t'J model<br />

7.2.2 Functional integral formulation<br />

To proceed further, it is convenient to represent the gr<strong>and</strong> canonical partition function for<br />

the redefined SB Hamiltonian (Eq. 12) in terms of a coherent-state functional integral [29]<br />

over Grassmann fermionic <strong>and</strong> complex bosonic fields as<br />

Z<br />

R <br />

Z ˆ D‰ ; ŠD‰e ;eŠD‰p ; p Š d‰ …1† Š d‰ …2†<br />

dL…†<br />

Š e 0 ; …15†<br />

<br />

L…† ˆP<br />

e i …@ ‡ …1†<br />

i †e i ‡ 2Trp T ‰…@ ‡ …1†<br />

i † o …2†T<br />

i<br />

‡ P <br />

<br />

Trp ~p i i Trp~p Trp p j j i i Trpp j j<br />

hi;ji<br />

‡ P <br />

<br />

i …@ † o ‡ …2†<br />

i i ‡ P t ij<br />

i z i z j j :<br />

i<br />

i;j<br />

i<br />

i<br />

Šp T i<br />

…1†<br />

i<br />

<br />

…16†<br />

Apart from the above symmetry considerations the SB functional-integral formalism<br />

reveals additional global <strong>and</strong> local gauge invariances [25, 30–32].<br />

The action S ˆ R <br />

0<br />

dL…† is invariant under the following site <strong>and</strong> time-dependent<br />

phase transformations (y i (t)w i<br />

(t)), i. e., under the local symmetry group SU(2) 6U(1),<br />

i ! e ii e i i i ;<br />

e i ! e i e i i<br />

; …17†<br />

p i<br />

! p i<br />

e i i ;<br />

provided that the (five) Lagrange parameters act as time-dependent gauge fields, l (1) i (t) <strong>and</strong><br />

…t†, absorbing the time derivatives of the phase factors:<br />

l …2†<br />

i<br />

…1†<br />

i<br />

! …1†<br />

i ‡ i _ i ;<br />

…2†<br />

i<br />

! e i i …2†<br />

i e i i i _ i<br />

‡ i _ i o : …18†<br />

Next, in the continuum limit, the radial gauge is introduced by representing the Bose<br />

fields by modulus <strong>and</strong> phase,<br />

e i ˆ je i je i'…ei† ;<br />

p i ˆ 1<br />

2<br />

X<br />

<br />

p i<br />

e<br />

i …p † i ;<br />

<br />

…19†<br />

93


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

(for a time-discretized version of this gauge fixing, see [31]). Exploiting the gauge freedom<br />

of the action, we can now fix the five real-valued coefficients y i (t), w im (t) to remove five<br />

phases (j i (t), f im (t)) of the Bose fields e i , p im in the radial gauge. As a consequence, all<br />

the Bose fields {e i , p im } become real, in contrast to the Hubbard model, where one SB field<br />

remains complex [25, 32]. At this point one should notice that the particle number <strong>and</strong> spin<br />

operators are changed into ~n i …p i †ˆP p2 i ; ~S i …p i †ˆp io ~p i (the previous notations e i , p io<br />

<strong>and</strong> ~p i now denote the radial parts of the corresponding Bose fields). Then, using the familiar<br />

identity for Gaussian integrals over Grassmann fields [29],<br />

R<br />

D‰ ; Š e ‰ G 1 Š 0 <br />

ˆ e Tr ln‰ G 1Š ; …20†<br />

the fermionic degrees of freedom can be integrated out <strong>and</strong> we obtain the following exact representation<br />

of the gr<strong>and</strong> canonical partition function<br />

Z ˆ R D‰Š e S ef f<br />

…21†<br />

in terms of the real-valued bosonic fields f ia ,<br />

i …† ˆ… i …†† ˆ …e i ;p io ; …2†<br />

io ;…1† i<br />

; p ix ; …2†<br />

ix ; p iy; …2†<br />

iy ; p iz; …2†<br />

iz † :<br />

…22†<br />

In Eq. 21 the effective bosonic action S eff takes the form<br />

(<br />

S ef f ˆ R d P<br />

0 i<br />

<br />

…1†<br />

i<br />

e 2 i ‡ P <br />

<br />

…1†<br />

i<br />

…2†<br />

io<br />

<br />

p 2 i<br />

2p io ~p i ~ …2†<br />

i<br />

…1†<br />

i<br />

<br />

‡J P )<br />

1<br />

p io ~p i ~p j p jo<br />

hiji<br />

4 …p2 io ‡ ~p2 i †…p2 jo ‡ ~p2 j †<br />

h<br />

i<br />

Tr ij; 0 ; 0 ln G ij; 1 † ; …23†<br />

where the (inverse) SB Green propagator is given by<br />

h<br />

<br />

Gij; 1 0…; 0 †ˆ @ ‡ …2†<br />

io<br />

0<br />

i<br />

~ …2†<br />

i ~ 0 ij … 0 †<br />

t ij …z y i z j† 0 ; 0…1 ij† : …24†<br />

It may be remarked that since the bosons are taken to be real, their kinetic terms,<br />

being proportional to the time derivatives in Eq. 16, drop out due to the periodic boundary<br />

conditions imposed on Bose fields (f ia (b) =f ia (0)). Strictly speaking it follows from this<br />

property that all the Bose fields do no longer have dynamics of their own [25].<br />

94


7.2 Slave-boson theory for the t-t'J model<br />

7.2.3 Saddle-point approximation<br />

The evaluation of Eq. 23 proceeds via the saddle-point expansion 2 , where at the first level<br />

of approximation we look for an extremum S … i † of the bosonized action S eff with respect<br />

to the Bose <strong>and</strong> Lagrange multiplier fields f ia :<br />

@S ef f<br />

@ i<br />

ˆ!<br />

0 ) S ˆ S ef f<br />

<br />

iˆ i<br />

: …25†<br />

The physically relevant saddle-point F i is determined to give the lowest free energy<br />

(per site)<br />

f t t 0 J ˆ =N ‡ n ; …26†<br />

where at given mean electron density n =1–d, the chemical potential m is fixed by the requirement<br />

n ˆ 1 @<br />

N @<br />

…27†<br />

( O ˆ S= denotes the gr<strong>and</strong> canonical potential). Obviously, an unrestricted minimization of<br />

the free energy functional is impossible for an infinite system. To keep the problem tractable,<br />

we use the ansatz<br />

~m i ˆ m~u i ; ~u i ˆ…cos ~Q ~R i ; sin ~Q ~R i ; 0† …28†<br />

for the local magnetization ~m i ˆ 2 ~S i ˆ 2p io ~p i . Following earlier analyses of spiral states<br />

for the Hubbard model [34] the unit vector ~u i is chosen as a local spin quantization axis<br />

pointing in opposite directions on different sublattices. Thereby, the order parameter wavevector<br />

~Q is introduced as a new variational parameter to describe several magnetic ordered<br />

states: PM, FM ( ~Q = 0), AFM ( ~Q =(p,p)), <strong>and</strong> incommensurate (1,1)-spiral ( ~Q =(Q, Q)),<br />

(1,p)-spiral ( ~Q =(Q,p)) <strong>and</strong> (0,1)-spiral ( ~Q = (0,Q)) states. Note that since fluctuations of<br />

the charge density are not incorporated the scalar Bose fields are homogeneous: e i = e, p io =<br />

p o , l (1) i = l (1) , <strong>and</strong> l (2) io = l (2) o . The vector fields exhibit the same spatial variation as the magnetization:<br />

~p i ˆ p~u i <strong>and</strong> l ~ …2†<br />

i ˆ l …2† ~u i . Then transforming Eq. 23 in the ( k; ~ o)-representation<br />

the trace in S eff can be easily performed to give the free energy functional<br />

2 Mean-field approximations to the functional integral are achieved by replacing the bosonic fields by<br />

their time averaged values. For the Hubbard model, a comparison of the paramagnetic <strong>and</strong> antiferromagnetic<br />

SB solutions with quantum Monte Carlo results shows that such a mean-field like approach<br />

yields an excellent quantitative agreement for local observables [33] <strong>and</strong> therefore can give a qualitative<br />

correct picture with relatively little effort.<br />

95


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

where<br />

f t t 0 J ˆ …1† …e 2 ‡ p 2 o ‡ p2 1† o …2† …p2 o ‡ p2 † 2 …2† p o p ‡ n<br />

<br />

<br />

‡J p 2 1<br />

o p2 …cos Q x ‡ cos Q y †<br />

2 …p2 o ‡ p2 † 2<br />

‡ 1<br />

N<br />

X<br />

ln‰1 n ~k Š ; …29†<br />

~k<br />

n ~k ˆ‰expf…E ~k<br />

†g ‡ 1Š<br />

1<br />

…30†<br />

holds. The renormalized single-particle energies E ~k … ˆ†are obtained by diagonalizing<br />

the kinetic part of Eq. 23 [26]. Requiring that f t-J be stationary with respect to the variation of<br />

the magnetic order vector, the wave-vector ~Q can be obtained from the extremal condition<br />

sin Q x;y ˆ 1<br />

Jp 2 o p2 1<br />

N<br />

X<br />

~k<br />

n ~k<br />

@E ~k<br />

@Q x;y<br />

:<br />

…31†<br />

If one substitutes Eq. 31 together with f a (obtained from the solution of the coupled<br />

self-consistency equations (Eq. 25)) into Eq. 29, the free energy of the t-t'-J model is obtained<br />

as<br />

f t t 0 J ˆ J<br />

4 ‰m2 …cos Q x ‡ cos Q y † 2n 2 Š‡ …2† m ‡… o …2† †n<br />

1 X <br />

ln 1 ‡ e …E k ~ ; …32†<br />

N<br />

~k<br />

where the quasi-particle energy takes on the form<br />

E ~k ˆ …1 ‡ †…" ~k ‡ " ~k<br />

†=y ‡ …2†<br />

~Q o<br />

h i 1=2<br />

2 …" ~k " ~k ~Q †2 =y ‡‰m…" ~k ‡ " ~k ~Q †=y ‡ …2† Š 2<br />

…33†<br />

with m=2 p o p <strong>and</strong> y=(1+d) 2 – m 2 .<br />

In particular, at the spatially uniform paramagnetic saddle-point, F …PM† =(e, p o , l o (2) ,<br />

l (1) ; 0, 0; 0,0; 0,0) the remaining bosonic fields<br />

96<br />

e 2 ˆ ;<br />

p 2 o ˆ 1 ;<br />

…1† ˆ<br />

…2†<br />

o<br />

ˆ<br />

2 ‡ 3 2<br />

1 2 ~"…0† ;<br />

2<br />

~"…0† …1 †J; …34†<br />

2<br />

…1 ‡ †


7.2 Slave-boson theory for the t-t'J model<br />

are explicitly given in terms of J, d <strong>and</strong> a single energy parameter ~" (0) defined by<br />

~"…~q† ˆ 2<br />

N<br />

X<br />

" ~k ~q<br />

… E ~k † …35†<br />

~k<br />

(at T=0). Here, the quasi-particle energy E ~k is<br />

E ~k ˆ 2" ~k =…1 ‡ †‡ …2†<br />

0 ; …36†<br />

<strong>and</strong> m can be determined from d =1– 2 N<br />

free energy becomes simply<br />

P<br />

~k … E ~ k<br />

†. <strong>Final</strong>ly, in this approximation, the<br />

f …PM†<br />

t t 0 J ˆ z2 ~"…0†<br />

1<br />

2 J …1 †2 ‡ …1 † : …37†<br />

7.2.4 Magnetic phase diagram of the t-t'-J model<br />

In the numerical evaluation of the self-consistency loop we proceed as follows: at given<br />

model parameters J <strong>and</strong> d, we solve the remaining saddle-point equations for m, l (1) , l o<br />

(2)<br />

<strong>and</strong> l (2) together with the integral equation for m using an iteration technique. Then, in an<br />

outer loop, the order parameter wave-vector ~Q is obtained from the extremal Eq. 31 by<br />

means of a secant method. Convergence is achieved if all quantities are determined with relative<br />

error less than 10 –6 . Note that our numerical procedure allows for the investigation of<br />

different metastable symmetry-broken states corresponding to local minima of the variational<br />

free energy functional (Eq. 32).<br />

At first, let us consider the case t' = 0, i. e., the pure t-J model. The resulting groundstate<br />

phase diagram in the J-d plane is shown in Fig. 7.1. For the case d = 0 (Heisenberg<br />

model), we obtain an AFM ground state. At J=0, the FM is lowest in energy up to a hole concentration<br />

of 0.327, where a first-order transition to the (1, p)-spiral takes place; above<br />

d = 0.39, we find a degenerate ground state with wave-vector (0,p). At d = 0.63 the PM becomes<br />

the lowest in energy state (second-order transition). This coincides with the U ?? SB<br />

results of the Hubbard model (J=4t 2 /U) [35]. The PM-FM instability occurs at d PM-FM = 0.33<br />

[32]. In contrast, the slave-fermion phase diagram [20] exhibits a much larger FM region.<br />

This can be taken as an indication that correlation effects are treated less accurate within the<br />

slave-fermion mean-field scheme [20, 36]. For finite exchange interaction J the (1,1)-spiral<br />

is the ground state at small doping. With increasing d a transition to the FM takes place,<br />

which becomes unstable against the (1,p)-spiral at larger doping concentrations. For<br />

J/t > 0.08, we find a transition from (1,1)-spirals to (1,p)-spirals at about d & 0.2. In<br />

Figure 7.1, the dotted line separates the (1, p)-spiral state from the region, where the (1, p)<br />

<strong>and</strong> (0, p) states are degenerate. If we admit only homogeneous phases, the phase boundaries<br />

(1,1)-spiral ⇔ (1,p)-spiral at d & 0.2 <strong>and</strong> (1,p)/(0, p)-phase ⇔ PM at d &0.63 remain<br />

nearly unchanged at larger values of J. However, analyzing the thermodynamic stability of<br />

97


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

Figure 7.1: SB ground-state phase diagram of the t-J model.<br />

the saddle-point solutions (i. e. the curvature of f t J …d), one observes a tendency towards<br />

phase separation into hole-rich <strong>and</strong> AFM regions above the ‘diagonal’ solid curve in the left<br />

upper part of the J-d phase diagram (see below).<br />

Figure 7.2 displays the variation of the extremal spiral wave-vector ~Q as function of<br />

doping. At d =0, ~Q =(p, p) indicates the AFM order. At low doping the (1,1)-spiral order vector<br />

decreases approximately linear. With decreasing J the AFM exchange is weakened, consequently<br />

the deviation of the order vector from (p, p) increases. The discontinuities reflect the<br />

first-order transition from (1,1)-spirals to (1, p)-spirals. For J/t =0.05 the transition to the FM<br />

state takes place, with Q x jumping down to zero. The (1,p)-spiral wave-vector shows a monotonous<br />

decrease of Q x until at d = 0.63 the transition to the PM (Q x = 0) occurs. Comparing<br />

the magnitude of the theoretical order vector of the spiral solutions to results from inelastic<br />

neutron scattering experiments on La 2–x Sr x CuO 4 [37], we find good agreement for an exchange<br />

interaction strength of J/t =0.4 (which seems to be a reasonable value with respect to<br />

the strong electron correlations observed in the high-temperature superconductors).<br />

Figure 7.2: The x-component of the spiral wave-vector ~Q as a function of doping d (compared with experiments<br />

[_] on LSCO [37]).<br />

98


7.2 Slave-boson theory for the t-t'J model<br />

Next, we investigate the ground-state properties of the t-t'-J model, where we fix<br />

J/t =0.4. Comparing the free energies of several (homogeneous) symmetry-broken states,<br />

we obtain the SB phase diagram shown in the t'/t-d plane in Fig. 7.3. Obviously, we can distinguish<br />

two regions. In the parameter region |t'/t| ^ 0.2, Figure 7.3 resembles the groundstate<br />

phase diagram of the pure t-J model (Fig. 7.1), i. e., we found large regions with incommensurate<br />

spiral order. However, compared to the pure t-J model, the t'-term stabilizes Néel<br />

order in a finite d region near half-filling. The increasing stability of AFM configurations<br />

can be intuitively understood because the t'-term moves electrons without disturbing the<br />

Néel-like background [12]. For larger ratios |t'/t| we have a completely different situation. In<br />

this parameter regime only commensurate states (AFM, FM, PM) occur, where for<br />

t' < t' c = –1.4 t we obtain the AFM state for all d. Note the rather large differences to the value<br />

of t' c obtained within a semiclassical (1/N)-expansion [38]. By varying the exchange coupling<br />

J, the phase boundaries in the t'/t-d plane are not much affected, e. g. for J/t =1 <strong>and</strong><br />

t'/t =–0.4, the transitions AFM ⇔ (1,1)-spiral <strong>and</strong> (1,1)-spiral ⇔ FM take place at d = 0.17<br />

<strong>and</strong> d = 0.6, respectively. We would like to point out that the main qualitative features of our<br />

SB phase diagram do confirm recent studies of magnetic long-range order in the t-t'-J<br />

model [38, 39].<br />

Figure 7.3: Restricted SB ground-state phase diagram of the 2D t-t'-J model at J/t =0.4.<br />

In Fig. 7.4 we plot the order parameter wave-vector as a function of doping at t'/t<br />

= +0.16 <strong>and</strong> t'/t =–0.4. The behaviour of Q x reflects a series of transitions AFM ⇔ (1,1)-<br />

spiral ⇔ (1, p)-spiral ⇔ PM. The corresponding (sublattice) magnetization abruptly changes<br />

at the (1,1)-spiral ⇔ (1, p)-spiral first-order transition. From Fig. 7.4 the asymmetry between<br />

hole (t' < 0) <strong>and</strong> electron doping (t' > 0) becomes evident. In contrast to recent Hartree-<br />

Fock results for the Hubbard model [40] we found the AFM phase near half-filling for both<br />

electron-doped <strong>and</strong> hole-doped cases (provided t'( 0). In the absence of t ' hopping, for arbitrarily<br />

small doping the AFM is found to be unstable against the (1,1)-spiral phase (cf.<br />

Fig. 7.3). Obviously, the AFM correlations are strongly enhanced by a positive t'-term,<br />

which is also in qualitative agreement with exact diagonalization studies of the t-t'-J model<br />

[12] <strong>and</strong> confirms the experimental findings for the electron-doped system NCCO [41, 42].<br />

Note that the stability region of the AFM phase agrees surprisingly well with the combined<br />

99


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

Figure 7.4: The x-component of the SB spiral wave-vector ~Q away from half-filling for the negative values<br />

t'/t =–0.16 (solid) <strong>and</strong> t'/t =–0.4 (long-dashed), i. e., hole doping (d > 0), <strong>and</strong> for the positive one<br />

t'/t =0.16 (dashed), i. e., electron doping.<br />

phase diagram for La 2–x Sr x CuO 4 <strong>and</strong> Nd 2–x Sr x CuO 4 obtained from neutron scattering [43]<br />

<strong>and</strong> muon spin relaxation measurements [44], respectively. For the YBCO parameter t'/t =<br />

–0.4, the AFM disappears around d = 0.1 whereas in the phase diagram of YBCO, determined<br />

by neutron diffraction [45], this transition takes place at about x = 0.4 oxygen content.<br />

However, there exits strong evidence that at least up to x = 0.2 no holes are transferred<br />

from CuO chains to CO 2 planes.<br />

As we have already noted, the phase diagrams in Fig. 7.1 <strong>and</strong> Fig. 7.3 result from the<br />

relative stability of various homogeneous states. On the other h<strong>and</strong>, there are arguments for<br />

the existence of inhomogeneous, e. g., phase-separated states in the t-J <strong>and</strong> related models<br />

[46]. Using very different methods, it was realized by several groups [47–51], that at large<br />

J/t the ground state of the t-J model separates into hole-poor (AFM) <strong>and</strong> hole-rich regions.<br />

Unfortunately, in the physically interesting regime of small exchange coupling (J/t ~ 0.2–<br />

0.4) <strong>and</strong> low doping level this point is still controversial. To gain more insight into the phenomenon<br />

of phase separation in t-J-type models of strongly correlated electrons it seems to<br />

be important to investigate the effect of an additional NNN hopping term t' as well. Therefore<br />

we study the free energy as a function of hole density d, where a (concave) convex curvature<br />

indicates local thermodynamic (in)stability implying a (negative) positive inverse isothermal<br />

compressibility k –1 = n 2 @2 f<br />

If k –1 < 0, the domain of the two-phase regime is determined<br />

performing a Maxwell construction for the anomalous increase of the chemical poten-<br />

@n 2 :<br />

tial m by doping.<br />

The results of our analysis of thermodynamic stability are depicted in Figs. 7.5a <strong>and</strong><br />

7.5 b for the t-J <strong>and</strong> t-t'-J model, respectively. The boundary of phase separation for the t-J<br />

model is given by the solid curve in the J-d plane shown in Fig. 7.5 a. As can be seen from<br />

Fig. 7.1, the different phase separated domains are built up by the (AFM) states at half-filling<br />

(d = 0) <strong>and</strong> the corresponding hole-rich state on the right boundary of the respective region.<br />

At J=0, where we recover the U ?? result of the Hubbard model [32], the free energy<br />

is a convex function Vd, i. e., our SRI SB theory does not support recent arguments<br />

[47] for phase separation in this limit. In the opposite limit of large J, complete charge se-<br />

100


7.2 Slave-boson theory for the t-t'J model<br />

Figure 7.5: Phase diagram of the t-(t')-J model including phase separated states. The phase separation<br />

boundary (solid curve) for the t-J model is shown in the J-d plane (a). We include the transition lines of<br />

Refs. [49] (dashed), [50] (dotted), <strong>and</strong> [51] (chain dashed). The triangles are the Lanczos results of<br />

Ref. [47]. The phase diagram of the t-t'-J is calculated in the t'/t-d plane at J/t =0.4 (b). Here the twophase<br />

region consists of AFM <strong>and</strong> (1,1)-spiral states. For further explanation see text.<br />

paration takes place for J/t > J PS /t=4.0, which seems to be an essentially classical result.<br />

From exact diagonalization (ED) we obtain J PS /t=4.1 +0.1 [52] compared with 3.8 derived<br />

by means of a high-temperature expansion [51]. We note that the homogeneous magnetic<br />

phases are always unstable close to half-filling (provided J/t > 0). This is in qualitative<br />

agreement with results obtained from ED studies [47] as well as from semiclassical [49] or<br />

renormalization-group calculations [50]. Also plotted in Fig. 7.5 a are the results of the hightemperature<br />

expansion method [51], where phase separation may occur only above a critical<br />

exchange J/t =1.2 as d?0, contrary to all the other approaches. The line separating the<br />

two-phase region from the stable states was determined by Marder et al. [49] within a semiclassical<br />

theory to vary as J/t =4 d 2 whereas our theory yields an approximately linear dependence<br />

at small d. We believe that, due to an improved treatment of spin correlations in<br />

our approach, the region of incomplete phase separation is reduced. The instability towards<br />

phase separation at small J can be taken as an indication that charge correlations may play<br />

an important role as well.<br />

The dotted lines of zero inverse compressibility in Fig. 7.5b show that also for |t'/t| >0<br />

there is a finite range of d over which the (1,1)-spiral is locally unstable. Similar results were<br />

recently obtained by Psaltakis <strong>and</strong> Papanicolaou [38]. But it is important to stress that in our<br />

theory the AFM state is locally stable for both signs of t'. In addition, based on the Maxwell<br />

construction, we can show that near half-filling the AFM state remains also globally stable<br />

against phase separation for the t-t'-J model (cf. Fig. 7.5 b, where the two-phase region is<br />

bounded by the solid lines). At larger values of |t'/t| 6 0.5, the phase-separated region is due<br />

to the first-order nature of the transition AFM ⇔ (1,1)-spiral (dashed curve) [38].<br />

<strong>Final</strong>ly to demonstrate the quality of our SRI SB approach, for the t-J model the expectation<br />

value of the kinetic energy is compared with the results from ED for a finite 16-<br />

site [48] (36-site [52]) lattice in Fig. 7.6. We find an excellent agreement between SB results<br />

<strong>and</strong> ED data. Obviously, this result does not depend on the interaction strength J, i. e. the SB<br />

101


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

Figure 7.6: Expectation value of the kinetic energy as a function of doping at several interaction<br />

strengths J in comparison to ED results for the 16 (36) site lattice [48, 52].<br />

theory well describes important correlation effects. Note that AH t S t–J /t is directly related to<br />

the effective transfer amplitude of the renormalized quasi-particle b<strong>and</strong>, which is taken as input<br />

for the calculation of transport coefficients in the following section.<br />

7.3 Comparison with experiments<br />

7.3.1 Normal-state transport properties<br />

As one of the main normal-state puzzles of the CuO 2 based high-temperature superconductors,<br />

the anomalous transport properties, in particular the temperature <strong>and</strong> doping dependence<br />

of the Hall resistivity R H (T,d) [53–56], has been under extensive experimental <strong>and</strong><br />

theoretical study. Quasi-particle transport measurements suggest a small density of charge<br />

carriers <strong>and</strong> hence a small pocket-like Fermi surface (FS) [57]. On the other h<strong>and</strong>, direct angle-resolved<br />

photoemission (ARPES) probes of the FS [14, 58] yield a large FS which satisfies<br />

Luttinger’s theorem <strong>and</strong> might be well described by (LDA) b<strong>and</strong> structure calculations<br />

[13, 59]. In principle, this contrasting behaviour is found for hole-doped (La 2–x Sr x CuO 4 ,<br />

YBa 2 Cu 3 O 6+x ) <strong>and</strong> electron-doped (Nd 2–x Ce x CuO 4 ) copper oxides as well.<br />

Adopting the hypothesis of Trugman [60], most of the normal-state properties of the<br />

cuprates may be explained by the dressing of quasi-particles due to magnetic interactions<br />

<strong>and</strong> the subsequent modification of their dispersion relation. Then, once the quasi-particle<br />

b<strong>and</strong> E ~k has been obtained, the Hall resistivity R H = s xyz /s xx s yy can be calculated in the relaxation<br />

time approximation, using st<strong>and</strong>ard formulas for the transport coefficients:<br />

102


7.3 Comparison with experiments<br />

ˆ<br />

ˆ<br />

e 2 X @n ~k<br />

k 2 u<br />

V k ~ u k ~ ;<br />

@E ~k<br />

~k<br />

e 3 2 X<br />

k 4 cV<br />

~k<br />

u ~ k<br />

" u ~ k<br />

@u ~ k<br />

@k <br />

@n ~k<br />

@E ~k<br />

:<br />

(38)<br />

Here n ~k is given by Eq. 30,V denotes the volume of the unit cell, e lkg is the completely<br />

antisymmetric tensor, <strong>and</strong> u ~ k ˆ @E ~k =@k . Note that R H does not depend on the relaxation<br />

time t.<br />

To make the discussion more quantitative, let us now consider the doping dependence of<br />

R H (d,T) in terms of the t-t'-J model using the saddle-point <strong>and</strong> relaxation time approximations,<br />

where FS <strong>and</strong> correlation effects are involved via the renormalized SB b<strong>and</strong> E ~k (Eq. 33). As we<br />

have pointed out above, in our approach the SB quasi-particle b<strong>and</strong> dispersion E ~k has to be determined<br />

in a self-consistent way at each doping level d. This should be in contrast to the NZA<br />

SB mean-field approach to the t-t'-J model of Chi <strong>and</strong> Nagi [61] where, in the J ? 0 limit, the<br />

calculation of transport properties is based on the simple replacement " ~k ! ~" ~k =<br />

–2td [(cos k x + cos k y )+2(t'/t)cos k x cos k y ] of the non-interacting b<strong>and</strong> dispersion (Eq. 2).<br />

Figure 7.7 shows the theoretical Hall resistivity as a function of carrier density in<br />

comparison to experiments on LSCO [53], YBCO [54] <strong>and</strong> NCCO [55, 56]. In the LSCO<br />

<strong>and</strong> NCCO systems, the concentration of chemically doped charge carriers in the CuO 2<br />

planes (d) definitely agrees with the composition (x) of the substitutes Sr <strong>and</strong> Ce. This simple<br />

relation, however, no longer holds for YBCO, i. e., the number of holes transferred into<br />

the planes does not increase linearly with the oxygen content. Indeed, the magnetic properties<br />

indicate d & 0uptox=0.2 [45]. In order to compare our theoretical model with the<br />

R H (x) data found on oxygen-doped YBCO, we use the relation d =(x – 0.2)/2 [62].<br />

Figure 7.7: Doping dependence of the Hall resistivity for hole-doped (left panel) <strong>and</strong> electron-doped<br />

(right panel) systems. The slave-boson results for J/t =0.4 <strong>and</strong> different ratios t'/t = 0 (solid), –0.16<br />

(dashed), –0.4 (chain dotted) <strong>and</strong> t'/t =0.16 (dotted) are compared with experiments on LSCO (_) [53]<br />

(at 80 K), YBCO (O) [54] (at 100 K) <strong>and</strong> NCCO (Z) [55, 56] (at 80 K), respectively. The inset shows<br />

the temperature dependence of R H for t'/t =0atd = 0.1 (short dashed) <strong>and</strong> d = 0.15 (dotted).<br />

103


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

Figure 7.7 clearly demonstrates the importance of the NNN transfer term t' for a consistent<br />

theoretical description of the experimental Hall data. For J/t =0.4, an excellent agreement<br />

with experiments on LSCO <strong>and</strong> NCCO, including the sign change of R H (d) at a very similar<br />

value, can be achieved using the parameter values t'/t =0 <strong>and</strong> t'/t = 0.16, respectively. It<br />

should be noted that in the case of LSCO we obtain t'/t =0 from R H (d) while LDA calculations<br />

yield a ratio t'/t =–0.16 [13, 59]. For YBa 2 Cu 3 O 6+x , where the experiments [54] give<br />

R H > 0 up to x=1, a negative t'-term suffices to give the correct tendency to R H (d). Using<br />

t'/t =–0.4, our theory yields a sign change of R H at d & 0.7. The strong increase (decrease)<br />

of the positive (negative) Hall coefficient as d?0 can be attributed to the formation of<br />

small hole (electron) pockets in the FS, which is a correlation effect. A recent analysis of resistivity<br />

saturation in LSCO [57], based on Boltzmann transport, has been taken as an indication<br />

of a small FS as well, however, the existence of a pocket-like FS is still a subtle <strong>and</strong><br />

unresolved issue. We found that the temperature dependence of R H (d,T) is at least in qualitative<br />

agreement with experiments on LSCO (see inset Fig. 7.7 (left panel)). Note that the<br />

quasi-particle dispersion E ~k exhibits extremely flat minima implying the presence of a new<br />

small energy scale D. Therefore, when the temperature becomes comparable to D the hole<br />

pockets are washed out <strong>and</strong> a sign change of R H occurs (cf. Fig. 7.7 at fixed d (inset)). 3<br />

The FS of the interacting system are shown in Fig. 7.8 for typical ratios t'/t at J/t =0.4<br />

<strong>and</strong> d = +0.1, where the diagonal (1,1)-spiral phase is lowest in energy. As observed for t-J<br />

<strong>and</strong> Hubbard models as well [64], we obtain small hole (or electron) pockets with a volume<br />

! |d|. The calculated FS are very anisotropic. As |d| increases, the pockets grow, until the<br />

FS topology changes completely at a critical doping value d c (R H (d c ) = 0 (cf. Fig 7.7), reflecting<br />

the transition from hole to electron carriers for t'/t < 0 <strong>and</strong> vice versa for t'/t >0.<br />

Figure 7.8: Quasi-particle Fermi surface in the (1,1)-spiral phase at J/t =0.4 for d = 0.1 (hole-doped<br />

system) <strong>and</strong> d = –0.1 (electron-doped system).<br />

3 We recently learned of a related exact diagonalization study of Dagotto et al. [63], where, similar in<br />

conclusion, the doping <strong>and</strong> temperature dependence of R H was calculated using a strongly renormalized<br />

flat quasi-particle dispersion.<br />

104


7.3 Comparison with experiments<br />

We want to point out that the renormalization of the quasi-particle b<strong>and</strong> E ~k strongly depends<br />

on both interaction strength J <strong>and</strong> doping level d [62] which, in fact, calls into question<br />

the frequently used rigid b<strong>and</strong> approximation. Due to the strong coupling of spin <strong>and</strong><br />

charge dynamics the characteristic energy scale for the coherent motion of the charge carriers<br />

is J <strong>and</strong> not t (provided t > J).<br />

7.3.2 Magnetic correlations <strong>and</strong> spin dynamics<br />

In this Section we try to underst<strong>and</strong> the INS <strong>and</strong> NMR experiments on the basis of the SRI<br />

SB mean-field theory. Using a generalized RPA expression for the spin susceptibility <strong>and</strong> assuming<br />

that the AFM correlations are spatially filtered by various ~q-dependent hyperfine<br />

form factors, we focus, in particular, on the spin dynamical properties in the paraphase of<br />

the t-t'-J model. Our starting point is an RPA-like form for the exchange-enhanced spin susceptibility<br />

[5, 6, 65]<br />

o …~q; !†<br />

s …~q; !† ˆ<br />

1 ‡ J 2 …cos q x ‡ cos q y † o …~q; !†<br />

: …39†<br />

The irreducible part 4<br />

o …~q; i! m †ˆ<br />

2 X<br />

N<br />

~k;n<br />

G… ~ k; i! n † G… ~ k ‡ ~q; i! n‡m †<br />

…40†<br />

contains the (dressed) SB Green propagators G… k; ~ i! n †ˆ‰i! n E ~k ‡ ~Š 1 describing noninteracting<br />

electrons with the renormalized b<strong>and</strong> structure E ~k …d† (Eq. 26).<br />

Once the dynamic spin susceptibility has been obtained both, INS measurements <strong>and</strong><br />

NMR experiments, can be explored. Probed by INS from the fluctuation-dissipation theorem<br />

the q-dependent <strong>and</strong> o-dependent spin structure factor is related to the dynamical susceptibility<br />

by<br />

S …~q; !† ˆ 1 1<br />

p 1 exp… k!† Im s…~q; !† …41†<br />

On the other h<strong>and</strong>, the nuclear spin-lattice relaxation rate a T –1 1a (a =k,k), e. g. for a<br />

field H a applied parallel to the c-axis, given by [66]<br />

a T 1k T 1 k B<br />

1ˆ<br />

2 2 B k lim 1<br />

!!0 N<br />

X<br />

~q<br />

a F 2 ? …~q† Im s…~q; !†<br />

k!<br />

…42†<br />

4 o n = 2np/b [o m = (2m +1)p/b] denote the fermionic [bosonic] Matsubara frequencies.<br />

105


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

<strong>and</strong> the transverse spin-spin relaxation rate, T –1 2G , for the RKKY coupling of the nuclear Cu<br />

spins, given by [67]<br />

T 2<br />

2G ˆ<br />

2<br />

0<br />

123<br />

c 1 X 2<br />

8k 2 …2 B † 4 Fk 2 N<br />

…~q† 1 X<br />

4<br />

s…~q† @ Fk 2 N<br />

…~q† s…~q† A 5 ; …43†<br />

~q<br />

provide local, atomic site (a = {63, 17} specific information. Here m B denotes the Bohr<br />

magneton <strong>and</strong> the constant c=0.69 is the natural-abundance fraction of the 63 Cu isotope.<br />

The form factors are given for 63 Cu <strong>and</strong> the planar 17 O nucleus as [68]<br />

63 F …~q† ˆA ‡ 2B…cos q x ‡ cos q y † ; …44†<br />

~q<br />

17 F …~q† ˆ2C cos q x<br />

2 : …45†<br />

Together with the anisotropy of the Cu relaxation rates the measurements of the<br />

Knight shift, a K ˆ 2<br />

a B n k lim ~q!0 a F …~q† s …~q; ! ˆ 0†, have been used to determine the hyperfine<br />

coupling constants A a , B <strong>and</strong> C on the basis of the Mila-Rice Hamiltonian [69]. Following<br />

Ref. [66] we take A || & –4B, A k & 0.84B, C & 0.87B <strong>and</strong> B & 3.3610 –7 eV, where<br />

63 gk = 7.5610 –24 erg/G <strong>and</strong> 17 gk = 3.8610 –24 erg/G.<br />

As we know, the RPA susceptibility contains an unphysical instability of the paramagnetic<br />

phase at some particular wave-vector ~q below a critical doping d c as signaled by the<br />

zero in the denominator of Eq. 39 at o = 0 (Stoner condition). Therefore the use of Eq. 39<br />

only makes sense if the system is far from the magnetic instability, i. e., d > d c , where for<br />

J=0.4 we have d c (t'/t =–0.16) H 0.27 <strong>and</strong> d c (–0.4) H d c (0) H 0.17.<br />

7.3.3 Inelastic neutron scattering measurements<br />

We begin with a discussion of the RPA dynamical spin structure factor S (~q; o†. The ~qdependence<br />

of S (~q; o† along the main symmetry axis of the Brillouin zone is shown in<br />

Fig. 7.9 at J=0.4 <strong>and</strong> hole density d = 0.3 for different ratios t'/t. Here the temperature<br />

is T=35 K <strong>and</strong> the frequency ko = 0.010 eV. For LSCO-type parameters (t'/t =0,<br />

–0.16) we found four pronounced incommensurate peaks located at the points<br />

p (1+q o , 1), p (1,1+q o ). The incommensurate modulation wave-vectors move with increasing<br />

doping level d away from the corner of the Brillouin zone along the directions<br />

(1, p) or(p, 1) (square lattice notation). Note, that the incommensurate peak position obtained<br />

from a three-b<strong>and</strong> RPA calculation of S (~q; o† [5] can be parametrized consistent<br />

with the experimental observation that q 0 H 2 x [37], while all the effective one-b<strong>and</strong><br />

RPA approaches [6, 10] yield an incommensurability scaling rather as q 0 H d. A more<br />

detailed investigation of the LSCO-type ~q-scans show that the ~q-variation of S(~q; o† is<br />

mainly governed by that of w o …~q; o† <strong>and</strong>, in accordance with experiments [70], the in-<br />

106


7.3 Comparison with experiments<br />

commensurate peaks considerably broaden when temperature or energy transfer are increased<br />

[71]. By contrast, the same plot for YBCO-type (t'/t =–0.4) parameters shows a<br />

broad <strong>and</strong> nearly T-independent [71] maximum around the (p, p)-point [72] (Fig. 7.9)<br />

which, due to the flat topology of w o , mainly reflects the ~q-dependence<br />

J (~q) =J (cosq x + cosq y ) (cf. Eq. 39). In this way, our calculations confirm recent arguments<br />

[5, 6, 73] for the importance of b<strong>and</strong> structure (Fermi surface geometry) effects<br />

in explaining the difference between observed LSCO <strong>and</strong> YBCO spin dynamics. Nevertheless,<br />

whether the incommensurate signals arise from an intrinsic magnetic structure or<br />

whether they result from the formation of domains (charge superstructures) in the LSCO<br />

system remains unanswered by INS [2, 68].<br />

Figure 7.9: Dynamic magnetic structure factor S(~q; o† is plotted along the (1,1) <strong>and</strong> (1, p)-directions of<br />

the Brillouin zone for different ratios t'/t.<br />

In a next step, we calculate the longitudinal or spin-lattice relaxation rate, T –1 1 , using the<br />

hyperfine form factors (Eq. 45). In Figure 7.10 the temperature dependences of 63 T –1 1k <strong>and</strong><br />

17 T –1 1k (inset) are shown for d = 0.35 <strong>and</strong> t'/t =–0.4 in comparison to experiments [74] on<br />

fully oxygenated YBCO materials (x =1). Although our theory does not succeed in giving<br />

the correct amplitude of a T –1 1k the qualitative features of the NMR data are described surprisingly<br />

well. Obviously, the broad magnetic peak in S (~q; o† at the AFM wave-vector<br />

~Q AFM =(p,p) strongly enhances the relaxation rate on Cu sites while, due to a geometrical<br />

cancellation ( 17 F a (~q ~Q AFM † 0†, the corresponding oxygen rate is rather insensitive to<br />

nearly commensurate AFM fluctuations <strong>and</strong> therefore is governed by the long wavelength<br />

part ~q 0 of the spin susceptibility [2]. For 63 Cu the nominal Korringa ratio S : (1/T 1 TK 2 S )<br />

(K S denotes the spin part of the Knight shift) is at least one order of magnitude larger. As<br />

can be seen from Fig. 7.10, for YBa 2 Cu 3 O 7 -type parameters, a Korringa (1/T 1 ! T) dependence<br />

(dotted line) holds at both Cu <strong>and</strong> O sites below T* ~ 120K, demonstrating the existence<br />

of a characteristic temperature T* as well as in all other near-optimum T c compounds<br />

[2, 75]. T* is in good agreement with the coherence energy scale suggested experimentally<br />

[75]. Above T* the 17 O NMR relaxation remains linear whereas the 63 Cu relaxation time<br />

does not follow the Korringa law.<br />

107


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

Figure 7.10: Spin-lattice relaxation rates 63 T –1 1k <strong>and</strong> 17 T –1 1k (inset) as a function of temperature. SB results<br />

((+); t'/t =–0.4, d = 0.35) are compared with experiments (D) on YBa 2 Cu 3 O 7 [74].<br />

In the oxygen-deficient compound YBa 2 Cu 3 O 6.6 , 1/( 63 T 1k T) shows a broad maximum<br />

at about 150 K (Fig. 7.11), which reflects a strong deviation from the canonical Korringa behaviour.<br />

In the normal-state regime the theoretical results agree even quantitatively with the<br />

experimental 63 Cu NMR data [76]. At this point it is important to stress that the present theory<br />

incorporates considerable b<strong>and</strong> renormalization effects already via w 0 …~q; o†, especially<br />

at low doping level. Thus a rather moderate strength J = 0.4 of the AFM exchange interaction<br />

yields the experimentally observed enhancement of 63 T 1k . In striking contrast to the optimally<br />

doped YBCO system, the Korringa relation is no longer satisfied for the planar 17 O<br />

nucleus sites in the underdoped material (cf. inset Fig. 7.11). Instead, a different behaviour<br />

17 T 1k T 17 K S = const was suggested to hold down to T c [76]. The unconventional T-scaling of<br />

17 T 1k has been taken as a signature for another important feature of the normal-state spin dynamics,<br />

the so-called spin-gap behaviour [2].<br />

Figure 7.11: 63 Cu <strong>and</strong> planar 17 O relaxation data (^) for underdoped YBa 2 Cu 3 O 6.6 [76] are plotted vs<br />

temperature. Theoretical results (6) are given at t'/t =–0.4, <strong>and</strong> d = 0.2.<br />

108


7.4 Summary<br />

Complementary measurements of the transverse spin-spin relaxation rate, T –1 2G ,have<br />

provided further insights into the drastic change in the magnetic properties when passing<br />

from the overdoped to the underdoped regime [67, 77]. As experimentally observed, we<br />

found that T –1 2G increases (decreases) with increasing (decreasing) hole doping (temperature)<br />

[78]. In order to detect the opening of a spin-pseudogap as a function of T, a powerful technique<br />

is to measure the ratio T 2G /T 1 T [79] which is nearly constant above ~ 200 K for deoxygenated<br />

YBa 2 Cu 3 O 6.6 [67]. The calculated temperature dependence of this quantity is<br />

shown in Fig. 7.12 together with recent experimental results [67]. Most notably, the opening<br />

of a spin-pseudo gap at T* ~ 135 K [79], i. e. well above T c , is clearly seen as a decrease of<br />

below T*. Note that for YBa 2 Cu 3 O 7 , as predicted by Fermi liquid theories, the ratio T 2 2G/T 1 T<br />

is approximately constant above 150 K [2].<br />

Figure 7.12: T 2G /( 63 T 1k T) in YBa 2 Cu 3 O 6.6 (^) as measured by Takigawa [67] compared with the SB<br />

data (6) att'/t =–0.4, <strong>and</strong> d = 0.2.<br />

7.4 Summary<br />

In this work we have used a spin-rotation-invariant SB approach to investigate magnetic <strong>and</strong><br />

transport properties of the 2D t-t'-J model. Our main results are the following (see also [81]):<br />

a) We present a detailed magnetic ground-state phase diagram of the 2D t-(t')-Jmodel, including<br />

incommensurate magnetic structures <strong>and</strong> phase separated states. At finite t', a main<br />

feature of the phase diagram, we would like to emphasize, is the existence of an AFM state<br />

away from half-filling, which is locally <strong>and</strong> also globally stable against phase separation.<br />

This result agrees with the experimentally observed AFM long-range order in the weakly<br />

doped LSCO <strong>and</strong> YBCO compounds. In contrast, for the simple t-J model we observe no<br />

AFM long-range order at any finite doping due to phase separation.<br />

109


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

b) The next nearest-neighbour hopping process (t') incorporates important correlation <strong>and</strong><br />

b<strong>and</strong> structure effects near half-filling. In particular, the t'-term can be used to reproduce<br />

the FS geometry of LSCO, YBCO, <strong>and</strong> NCCO. Also the NNN hopping provides a possible<br />

origin for the experimentally observed asymmetry in the persistence of AFM order of holedoped<br />

<strong>and</strong> electron-doped systems.<br />

c) The quality of the SRI SB approach was demonstrated in comparison with exact diagonalization<br />

results available for the t-J model on finite square lattices with up to 36 sites. In<br />

this case, the SB method yields an excellent estimate for the quasi-particle b<strong>and</strong> renormalization.<br />

d) Within the saddle-point <strong>and</strong> relaxation time approximation, our SB calculation of the<br />

Hall resistivity in the t-t'-J model provides a reasonable explanation of the experimentally<br />

observed doping dependence of R H on both hole-doped (La 2–x Sr x CuO 4 , YBa 2 Cu 3 O 6+x ) <strong>and</strong><br />

electron-doped (Nd 2–x Ce x CuO 4 ) copper oxides.<br />

e) Using a generalized RPA expression for the spin susceptibility <strong>and</strong> assuming that the<br />

AFM correlations are spatially filtered by the hyperfine form factors, we have calculated the<br />

temperature dependences of spin-lattice <strong>and</strong> spin-spin relaxation rates for planar copper <strong>and</strong><br />

oxygen sites. The results agree qualitatively well with various NMR experiments on YBa 2-<br />

Cu 3 O 6+x . In addition, we can attribute the contrasting ~q-dependence of the magnetic structure<br />

factor S (~q; o† seen in INS experiments for LSCO-type <strong>and</strong> YBCO-type systems to differences<br />

in their fermiology.<br />

Recently, our theory was improved to include (Gaussian) fluctuations beyond the paramagnetic<br />

saddle-point approximation [27, 80]. We derived a concise expression for the spin<br />

susceptibility w s (~q; o† of the t-t '-J model which does not have the st<strong>and</strong>ard RPA form. Then<br />

we were able to show that the instability line obtained from a divergence of w s (~q; 0† is in<br />

agreement with the PM ⇔ spiral state phase boundary in the saddle-point phase diagram,<br />

which in fact proves the consistency of both approaches [27].<br />

References<br />

1. For a recent overview of the experimental <strong>and</strong> theoretical situation see: Proceedings of the Int.<br />

Conf. M 2 S HTSC IV, Grenoble 1994, in: Physica C, 235–240 (1994)<br />

Proceedings of the Int. Euroconf. on Magnetic Correlations, Metal-Insulator-Transitions, <strong>and</strong><br />

Superconductivity in Novel Materials, Würzburg 1994, in: J. Low. Temp. Phys., 3/4 (1995)<br />

2. A.P. Kampf: Physics <strong>Report</strong>s, 249, 219 (1994)<br />

3. A.J. Millis, H. Monien: Phys., Rev. Lett. 70, 2810 (1993)<br />

4. P. Monthoux, D. Pines: Phys. Rev. B, 49, 4261 (1941)<br />

5. Q. Si, Y. Zha, K. Levin, J.P. Lu: Phys. Rev. B, 47, 9055 (1993)<br />

6. T. Tanamoto, K. Kuboki, H. Fukuyama: J. Phys. Soc. Jpn., 60, 3072 (1991)<br />

T. Tanamoto, H. Kohno, H. Fukuyama: J. Phys. Soc. Jpn., 62, 717 (1993)<br />

110


References<br />

T. Tanamoto, H. Kohno, H. Fukuyama: J. Phys. Soc. Jpn., 62, 1455 (1993)<br />

T. Tanamoto, H. Kohno, H. Fukuyama: J. Phys. Soc. Jpn., 63, 2739 (1994)<br />

7. T. Ito, K. Takenaka, S. Uchida: Phys. Rev. Lett., 70, 3995 (1993)<br />

8. V.J. Emery: Phys. Rev. Lett., 58, 2794 (1987)<br />

9. M.S. Hybertsen, M. Schlüter, N. E. Christensen: Phys. Rev. B, 39, 9028 (1989)<br />

10. P. Bénard, L. Chen, A.-M. S. Tremblay: Phys. Rev. B, 47, 15217 (1993)<br />

11. M. Lavagna, G. Stemmann: Phys. Rev. B, 49, 4235 (1994)<br />

12. T. Tohyama, S. Maekawa: Phys. Rev. B, 49, 3596 (1994)<br />

13. J. Yu, S. Massida, A.J. Freeman: Phys. Lett. A, 122, 198 (1987)<br />

14. D.M. King et al: Phys. Rev. Lett., 70, 3159 (1993)<br />

15. P.A. Lee: Phys. Rev. Lett., 63, 680 (1989)<br />

16. D.M. News, M. Rasolt, P. Pattnik: Phys. Rev. B, 38, 6513 (1988)<br />

17. Z. Zou, P.W. Anderson: Phys. Rev. B, 37, 627 (1988)<br />

18. D. Yoshioka: J. Phys. Soc. Jpn., 58, 1516 (1989)<br />

19. H. Kaga: Phys. Rev. B, 46, 1979 (1992)<br />

20. S.K. Sarker: Phys. Rev. B, 47, 2940 (1993)<br />

21. J. Igarashi, P. Fulde: Phys. Rev. B, 48, 998 (1993)<br />

22. Y.M. Li, D.N. Sheng, Z. B. Su, L. Yu: Phys. Rev. B, 45, 5428 (1992)<br />

23. S.D. Liang, B. Doucot, P.W. Anderson: Phys. Rev. Lett., 61, 365 (1988)<br />

24. G. Kotliar, A.E. Ruckenstein: Phys. Rev. Lett., 57, 1362 (1986)<br />

25. R. Frésard, P. Wölfle: Int. J. Mod. Phys. B, 5&6, 685; Erratum ibid. 3087 (1992)<br />

26. M. Deeg, H. Fehske, H. Büttner: Europhys. Lett., 26, 109 (1994)<br />

27. M. Deeg, H. Fehske: Phys. Rev. B, 50, 17 874 (1994)<br />

28. T. Li, P. Wölfle, P.J. Hirschfeld: Phys. Rev. B, 40, 6817 (1989)<br />

29. J.W. Negele, H. Orl<strong>and</strong>: Quantum Many-Particle Systems, Addison-Wesley, Reading, MA (1988)<br />

30. T. Jolicoeur, J.C.L. Guillou: Phys. Rev. B, 44, 2403 (1991)<br />

31. E. Arrigoni et al.: Physics <strong>Report</strong>s, 241, 291 (1994)<br />

32. M. Deeg, H. Fehske, H. Büttner: Z. Phys. B, 88, 283 (1992)<br />

33. L. Lilly, A. Muramatsu, W. Hanke: Phys. Rev. Lett., 65, 1379 (1990)<br />

34. R. Frésard, M. Dzierzawa, P. Wölfle: Europhys. Lett., 15, 325 (1991)<br />

E. Arrigoni, G.C. Strinati: Phys. Rev. B, 44, 7455 (1991)<br />

35. B. Möller, K. Doll, R. Frésard: J. Phys. Condens. Matter, 5, 4847 (1993)<br />

36. D. Boies, F.A. Jackson, A.-M.S. Tremblay: Int. J. Mod. Phys. B, 9, 1001 (1995)<br />

37. S.-W. Cheong et al.: Phys. Rev. Lett., 67, 1791 (1991)<br />

38. G.C. Psaltakis, N. Papanicolaou: Phys. Rev. B, 48, 456 (1993)<br />

39. H. Mori, M. Hamada: Phys. Rev. B, 48, 6242 (1993)<br />

40. A.P. Kampf, W. Brenig: J. Low Temp. Phys,. 95, 335 (1994)<br />

41. V. Chechersky et al.: Phys. Rev. Lett., 70, 3355 (1993)<br />

42. M. Matsuda et al.: Phys. Rev. B, 45, 12548 (1992)<br />

43. G.M. Luke et al.: Phys. Rev. B, 42, 7981 (1990)<br />

44. T.R. Thurston et al.: Phys. Rev. B, 40, 4585 (1990)<br />

45. J. Rossat-Mignod et al.: Physica B, 180 & 181, 383 (1992)<br />

46. V.J. Emery, S. A. Kivelson: Physica C, 209, 597 (1993)<br />

47. V.J. Emery, S. A. Kivelson, H. Q. Lin: Phys. Rev. Lett., 64, 475 (1990)<br />

48. H. Fehske,V. Waas, H. Röder, H. Büttner: Phys. Rev., B 44, 8473 (1991)<br />

49. M. Marder, N. Papanicolaou, G. C. Psaltakis: Phys. Rev. B, 41, 6920 (1990)<br />

50. E.V.L. de Mello: J. Phys. Condens. Matter, 6, 117 (1994)<br />

51. W.O. Puttika, M. Luchini, T.M. Rice: Phys. Rev. Lett., 68, 538 (1992)<br />

52. V. Waas, H. Fehske, H. Büttner: Phys. Rev. B, 48, 9106 (1993)<br />

53. H. Takagi et al.: Phys. Rev. B, 40, 2254 (1989)<br />

54. E.C. Jones et al.: Phys. Rev. B, 47, 8986 (1993)<br />

55. H. Takagi, S. Uchida, Y. Tokura: Phys. Rev. Lett., 62, 1197 (1989)<br />

56. S.J. Hagen, J.L. Peng, Z.Y. Li, R.L. Greene: Phys. Rev. B, 43, 13606 (1991)<br />

111


7 Slave-Boson Approach to Strongly Correlated Electron Systems<br />

57. H. Takagi et al.: Phys. Rev. Lett., 69, 2975 (1992)<br />

58. J.C. Campuzano et al.: Phys. Rev., Lett. 64, 2308 (1990)<br />

59. W.E. Pickett: Rev. Mod. Phys., 61, 433 (1989)<br />

60. S.A. Trugman: Phys. Rev. Lett., 65, 500 (1990)<br />

61. H. Chi, A.D.S. Nagi: Phys. Rev. B, 46, 421 (1992)<br />

62. H. Fehske, M. Deeg: Solid State Commun., 93, 41 (1995)<br />

63. E. Dagotto, A. Nazarenko, M. Bonsinsegni, Phys. Rev. Lett., 73, 728 (1994)<br />

64. S. Sarker, C. Jayaprakash, H.R. Krishnamurthy, W. Wenzel: Phys. Rev. B, 43, 8775 (1991)<br />

65. R. Grempel, M. Lavagna: Solid State Commun., 83, 595 (1992)<br />

66. H. Monien, P. Monthoux, D. Pines: Phys. Rev. B, 42, 167 (1990)<br />

67. M. Takigawa: Phys. Rev. B, 49, 4158 (1994)<br />

68. V. Barzykin, D. Pines, D. Thelen: Phys. Rev. B, 50, 16052 (1994)<br />

69. F. Mila, T.M. Rice: Physica C, 157, 561 (1989)<br />

70. T.E. Mason, G. Aeppli, H.A. Mook: Phys. Rev. Lett., 68, 1414 (1992)<br />

71. H. Fehske, M. Deeg: J. Low. Temp. Phys., 3/4, 425 (1995)<br />

72. J. Rossat-Mignod et al.: Physica C, 185–189, 86 (1991)<br />

73. P. Littlewood, J. Zaanen, G. Aeppli, H. Monien: Phys. Rev. B, 48, 487 (1993)<br />

74. M. Takigawa et al.: Physica C, 162, 853 (1989)<br />

75. P.C. Hammel et al.: Phys. Rev. Lett., 63, 1992 (1989)<br />

76. M. Takigawa et al.: Phys. Rev. B, 43, 247 (1991)<br />

77. Y. Itoh et al.: J. Phys. Soc. Jpn., 61, 1287 (1992)<br />

78. M. Deeg: PhD thesis, Universität Bayreuth, (1995)<br />

79. C. Berthier et al.: Physica C, 235–240, 67 (1994)<br />

80. M. Deeg et al., Z. Phys. B, 95, 87 (1994)<br />

81. H. Fehske, Spin Dynamics, Charge Transport, <strong>and</strong> Electron-Phonon Coupling Effects in Strongly<br />

Correlated Electron Systems, Habilitationsschrift, Universität Bayreuth (1995)<br />

112


8 Non-Linear Excitations <strong>and</strong> the Electronic Structure<br />

of Conjugated Polymers<br />

Klaus Fesser<br />

8.1 Introduction<br />

Conjugated polymers have attracted quite substantial research activities during the last 15<br />

years [1]. On one h<strong>and</strong> these materials promise interesting technological applications most<br />

of which are related to the possibility of a reversible charging <strong>and</strong> decharging of these systems.<br />

These include various battery designs as well as storage of (charged) pharmaceuticals<br />

which in turn can be released in a controlled way by application of an electrical current. For<br />

the design of non-linear optical components they play an important role as organic materials<br />

due to their processibility <strong>and</strong> fine tuning of their physical properties via suitable side<br />

groups. Recently light-emitting diodes made from these systems have made conjugated polymers<br />

to possible c<strong>and</strong>idates for the construction of thin displays [2].<br />

On the other h<strong>and</strong> a theoretical description of the whole class of these materials poses<br />

interesting questions which are worth to study on their own right. As essentially quasi onedimensional<br />

systems they give rise to the hope that many of these questions might be answered<br />

analytically. The main problems are the nature of the insulator-metal transition observed<br />

during doping <strong>and</strong> the origin of the intragap states, which are mainly responsible for<br />

the relaxation processes relevant for the light-emitting properties. We have addressed both<br />

aspects within this project <strong>and</strong> this article is organized accordingly. In Section 8.2 we present<br />

the theoretical model stressing the relevant physics <strong>and</strong> discuss the related fundamental symmetries.<br />

Then in Section 8.3 we adopt the view that the doping process mainly introduces<br />

disorder into these systems <strong>and</strong> calculate various properties (density of states, optical absorption)<br />

from this assumption. In Section 8.4 we investigate the non-linear excitations responsible<br />

for the intragap states in more detail <strong>and</strong> close in Section 8.5 with an outlook on<br />

still open problems.<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

113


8 Non-Linear Excitations <strong>and</strong> the Electronic Structure of Conjugated Polymers<br />

8.2 Models<br />

Common to all conjugated polymers is the existence of a carbon backbone with an alternating<br />

sequence of (short) double <strong>and</strong> (long) single bonds. The various systems, as polyacetylene,<br />

polyparaphenylene, polypyrrole, polythiophene, polyparaphenylenevinylene, etc.,<br />

differ only – from a physcist’s point of view – in the side groups <strong>and</strong> other chemical structures<br />

which are energetically far away from the Fermi energy. Emphasizing the general aspects<br />

of these systems, these chemical details are safely neglected if one restricts to properties<br />

which are mainly due to the electrons around the Fermi energy. Thus the parameters in<br />

the simple model presented below should be regarded as effective parameters including<br />

some aspects of the interactions which are otherwise neglected.<br />

Beside some biological systems such as b-retinol where a finite chain of a conjugated<br />

polymer is present (<strong>and</strong> thus the properties of this polymer might be of relevance for the<br />

biological functions of this material), there may be other, related systems where similar theoretical<br />

concepts are or can be applied. These include the metal-halogen (MX) chains where<br />

instead of one, as in the conjugated polymers, now two electron b<strong>and</strong>s govern the essential<br />

physics. On the same level the polyanilines can be modelled where the phonons of the conjugated<br />

polymers have to be replaced by the librons, i. e. oscillations of the quinoid/benzoid<br />

rings. <strong>Final</strong>ly, a true one-dimensional modification of carbon, carbene with alternating single<br />

<strong>and</strong> triple bonds, can also be understood along these lines.<br />

So far we have only mentioned the construction of adequate models for the investigation<br />

of physical properties. However, there are competing approaches to address the same<br />

set of questions. One of these approaches uses sophisticated quantum-chemical codes [1c,d]<br />

to calculate structure <strong>and</strong> electronic states of these materials. Although such a procedure<br />

may be able to reproduce the observed properties of a specific material quite well it is very<br />

difficult to obtain information about general trends <strong>and</strong> physical mechanisms. Therefore we<br />

did not follow this route. Another well-established method for calculating ground-state properties,<br />

namely the local density functional, has been used to some extent [3]. The drawback<br />

of this method, however, is the finite size of the system which can be calculated within a<br />

reasonable amount of computer resources. Therefore only a few questions have been addressed<br />

via this approach.<br />

Therefore, we shall argue in favor of a simple model which is capable to include the<br />

most essential physics. Assuming the sp 2 -hybridisation of the carbon atom leaves one p-electron<br />

per atom the others being incorporated into the bonds as s-electrons. It is the physical<br />

behaviour of this single p-electron which is responsible for all the interesting effects.<br />

Since there is only this one electron per site the polymer would be metal-like in its<br />

ground state, in contrast to nature where one finds an energy gap of the order of 1 eV. In<br />

one dimension, however, these electrons are unstable when they are coupled to the lattice.<br />

Due to the Peierls effect, scattering off 2 k F phonons, a gap opens right at the Fermi energy<br />

in accordance with the observations. In addition, electron-electron interaction contribute to<br />

the size of this gap [4] stabilizing the semiconductor ground state even further. A single-particle<br />

model along these lines has been put forward by Su et al. [5] in the early stages of conjugate<br />

polymer research. Its parameters although, derived from a simple picture, should<br />

114


8.2 Models<br />

nevertheless be interpreted as effective parameters including parts of the electron-electron<br />

interaction which is otherwise neglected. It has been shown [6] that for some ground-state<br />

properties such redefinition can indeed be performed.<br />

It turns out that there is a typical length scale in this model v F /D (with the Fermi velocity<br />

v F <strong>and</strong> the electronic gap 2 D) which is considerably larger than the interatomic spacing<br />

a. Thus for most physical properties of interest a continuum approximation is valid.<br />

This stresses the generic aspects of the whole class of these materials even more.<br />

In rescaled variables this model now reads<br />

H ˆ X Z<br />

s<br />

dx ‡ s …x† f i 3@ x ‡ 1 …x†g s …x†‡ 1 Z<br />

2<br />

dx 2 …x†<br />

…1†<br />

Here c (x) is a two-component spinor describing electrons moving to the left (right)<br />

along the one-dimensional polymer, s is the spin index which, except for external fields, is<br />

not relevant here. The s 3 term is the kinetic energy originating from the hopping of electrons<br />

between neighbouring sites. D (x) is the lattice order parameter where a constant<br />

D (x) =D 0 describes a uniform dimerization of the lattice. The s 1 term is the electron-phonon<br />

coupling responsible for the Peierls distortion. <strong>Final</strong>ly we have an elastic energy for the<br />

lattice. The kinetic energy is neglected within an adiabatic approximation. All quantities<br />

have been scaled in order to have a single coupling constant (l & 0.2).<br />

At this stage we postpone the non-linear excitations of this model to a subsequent<br />

Chapter, we only discuss some of the symmetries of this model which are of relevance also<br />

for these excitations. First we note that the form of the electronic part of the Hamiltonian<br />

(Eq. 1) has exactly the form of a (relativistic invariant) Dirac operator in one dimension.<br />

This correspondence has been exploited [7] successfully in obtaining solutions of this model<br />

from results known in models of elementary particles. We remark here that similar analogies<br />

can be made for specific forms of the Fermi surface also in higher dimensions. Thus the<br />

connection between solid state physics <strong>and</strong> quantum field theory models can be used for a<br />

better underst<strong>and</strong>ing of both.<br />

In addition there are two symmetries which can be directly related to observable quantities.<br />

The charge-conjugation symmetry, which can be expressed as H being invariant under<br />

c?s 2 Kc (K complex conjugate operator), relates particle <strong>and</strong> hole states <strong>and</strong> thus is responsible<br />

for the single particle spectrum being symmetric with respect to the Fermi energy,<br />

which has been set to zero in H (Eq. 1). A more hidden property is the supersymmetry of<br />

the electronic part H el (Eq. 1),<br />

<br />

0 ˆ H el ;H ‰ el ; 3 Š : …2†<br />

We have shown [8] that this more formal property is responsible for the asymmetry of<br />

the optical absorption peaks of transitions involving intragap states. Both symmetries are absent<br />

in the real systems under consideration, but for most questions this breaking of symmetries<br />

is merely a question of quantity rather than of importance for the existence of these localized<br />

states.<br />

<strong>Final</strong>ly we mention that Eq. 1 has been derived as a model for polyacetylene where<br />

the ground-state order parameters D o <strong>and</strong> –D o yield the same energy. This is not the case for<br />

115


8 Non-Linear Excitations <strong>and</strong> the Electronic Structure of Conjugated Polymers<br />

the other polymers of this class. A simple extension of this model, introduced by Brazovski<br />

<strong>and</strong> Kirova [9], corrects this. As a result all the solutions of Eq. 1 can be carried over [10]<br />

to this more general model where only the location of the intragap states are now parameters<br />

which can be adjusted to a specific material of interest. In consequence the model (Eq. 1)<br />

can be considered as the most simple generic model of conjugated polymers.<br />

8.3 Disorder<br />

A major focus of theoretical investigations is the experimental observation that the electronic<br />

structure <strong>and</strong> consequently the physical properties change drastically upon doping. Here<br />

we restrict ourselves to the question how these properties are affected through a r<strong>and</strong>om<br />

force introduced by these impurities. For simplicity we only consider isoelectronic doping,<br />

i. e. the total number of carriers remains unchanged <strong>and</strong> the dopants introduce only scattering<br />

centers for the electrons.<br />

In consequence two different types (site or bond impurities) of scatterers can be identified.<br />

The first gives rise to a r<strong>and</strong>om contribution to the on-site energy of an electron<br />

whereas the latter modifies the hopping integral between neighbouring sites. In the spirit of<br />

the continuum approximation we are thus lead to consider only backward (Eq. 3) or forward<br />

(Eq. 4) scattering.<br />

P R<br />

H imp ˆ U b dx ‡ …x† 1 …x†…x x j † ; …3†<br />

j<br />

P R<br />

H imp ˆ U s dx ‡ …x†1 …x†…x x j † ; …4†<br />

j<br />

x j denotes the (r<strong>and</strong>om) position of an impurity. Since this problem cannot be solved exactly<br />

we have to redraw to approximate methods.<br />

We have used three main routes.<br />

In the first method we employ the first Born approximation for the impurity self-energy<br />

[11]. This enables us to formulate equations of motion for the full space-dependent<br />

Green functions <strong>and</strong> thus consider the influence of disorder on the non-linear excitations as<br />

well [12] (see next Chapter). Furthermore, the replacement of Eq. 3 <strong>and</strong> Eq. 4 by r<strong>and</strong>om<br />

(Gaussian) fields<br />

Z n<br />

H imp ˆ dx ‡ …x†<br />

1<br />

1<br />

o<br />

V b=s …x† …x† ;<br />

which is correct within the Born approximation, allows the determination of the Green function<br />

<strong>and</strong> higher correlation functions via a functional integral technique [13]. The averaging<br />

procedure is formulated through the introduction of additional Grassmann variables in a<br />

116<br />

…5†


8.4 Non-linear excitations<br />

supersymmetric way. We note that this supersymmetry is not related to the one discussed<br />

earlier. Using the algebraic properties of these variables the average can be performed <strong>and</strong><br />

the resulting functional integral can be calculated via a transfer method in one dimension.<br />

We note that this procedure still works when two coupled chains are considered but cannot<br />

be done exactly in any higher dimension. As result we obtain the density of states <strong>and</strong> also<br />

information about the extension of the (in principle) localized wave functions via the Thouless<br />

formula. It turns out [14] that for moderate disorder a typical realistic chain length of<br />

approximately 200 units is smaller than the localization length of the b<strong>and</strong> states not too<br />

close to the b<strong>and</strong> edgesunverständlich. Thus for such systems these states can indeed be considered<br />

as extended states even in the presence of disorder putting various models which<br />

treat the propagation of electrons along one chain as metal-like on a firmer basis. This<br />

method can easily be extended to higher dimensional systems which means that also the<br />

coupling of polymer chains can be taken into account. Thus we are able to calculate, within<br />

a saddle-point approximation, the optical absorption coefficient for a two-dimensional film.<br />

For an orientation along the chains we find [15] a typical disorder induced broadening of the<br />

absorption edge together with a shoulder on the low-energy side due to neighbouring chains<br />

coupling. Both features agree quite well with experimental findings. Concomitantly, the absorption<br />

perpendicular to the chain direction is featureless.<br />

In the second method, going beyond Born, we examined the density of states within<br />

the coherent potential approximation (CPA) which takes into account multiple scattering processes.<br />

One might think that on this level impurity states are introduced in the gap. However,<br />

we find [16] that the existence of such localized impurity states strongly depends on the relative<br />

strength of site vs. bond impurity. Only states in the gap due to disorder can be found<br />

if the site amplitude |U s | is stronger than the bond amplitude |U b |. Since CPA is an effective<br />

medium theory this result might be questionable in one dimension.<br />

In the third method, in order to check the CPA result, we performed a numerical simulation<br />

[17] where for a given r<strong>and</strong>om distribution the electronic eigenvalues have been calculated<br />

numerically <strong>and</strong> the results been averaged over a large number of realizations. It turned<br />

out that the CPA results could be reproduced quite satisfactorily thus establishing the different<br />

role of both types of impurities.<br />

8.4 Non-linear excitations<br />

Given the single particle states according to Eq. 1 with a uniform order parameter D o one<br />

might expect that the lowest excited state corresponds to the lowest level in the conduction<br />

b<strong>and</strong> being occupied thus requiring the amount 2D o in energy, since the half-filled case considered<br />

here the valence b<strong>and</strong> with E |D o |is<br />

empty. The model (2.1), however, exhibits the property that an additional carrier modifies<br />

the order parameter locally, yielding a non-homogeneous D (x), <strong>and</strong> at the same time creates<br />

through a rearrangement of the b<strong>and</strong> statesadditional state(s) deep in the gap. Mathematically<br />

this behaviour is due to the fact that Eq. 1 is a non-linear model, the non-linearity re-<br />

117


8 Non-Linear Excitations <strong>and</strong> the Electronic Structure of Conjugated Polymers<br />

sulting from the requirement that the total energy functional AH{D (x)}S has to be stationary<br />

with respect to a variation of the order parameter D (x). This gives rise to a self-consistency<br />

equation where this order parameter is governed by the occupied electronic states,<br />

…x† ˆ<br />

P ‡ …x† 1 …x† : …6†<br />

occ<br />

Various exact solutions to this problem are known. The most prominent one is the<br />

kink (soliton) for the case of polyacetylene being,<br />

p<br />

…x† ˆ o thx= 2 :<br />

…7†<br />

As already mentioned, this kink does not exist for the other polymers of this class because<br />

the ground state is non-degenerate in D o . Therefore the simplest non-linear excitation<br />

in a more general sense is the bound kink-antikink pair (polaron) which is characterized by<br />

two localized electronic states in the gap at +o o (e. g. o o /D o & 0.5 for polythiophene). In<br />

addition there exist periodic solutions, e. g. the kink lattice with a periodicity determined by<br />

the concentration of excess charges. Physically these states lower the total energy of the system<br />

through an inhomogeneous order parameter D (x), which raises the energy (cf. Eq. 1)<br />

<strong>and</strong> a much larger compensation through the intragap states, which altogether give a smaller<br />

value of the total energy than the simple single-particle picture. For the polaron this gain in<br />

energy amounts to E p /2D o = 0.98 < 1 for polythiophene. For trans-polyacetylene this number<br />

is 0.90.<br />

One can now envisage the processes which are involved in generating visible light. A<br />

sufficiently strong electric field can promote a single electron into the lowest unoccupied<br />

conduction b<strong>and</strong> state. This state is unstable <strong>and</strong> relaxes on a fast time scale (femtosecond)<br />

into a polaron-like state which then can recombine to the ground state under the emission of<br />

radiation. A full microscopic underst<strong>and</strong>ing of all the processes involved is only possible if<br />

for the dynamics of the lattice the degrees of freedom are fully taken into account as well as<br />

the residual Coulomb interactions.<br />

One step in this direction has been made within this project by Bronold [18]. In his<br />

doctoral thesis he treats on an equal footing electron-electron (exciton) <strong>and</strong> electron-phonon<br />

(polaron) interactions. The coupling of this system to short laser pulses gives rise to characteristic<br />

changes of position <strong>and</strong> shapes of absorption/emission lines in optically stimulated<br />

emission <strong>and</strong> inverse Raman scattering experiments. As these effects have a very short time<br />

scale (femtoseconds), experiments are difficult to perform <strong>and</strong> a comparison with existing<br />

theoretical predictions is not convincing. Nevertheless, one expects that this line of approach<br />

will finally give a detailed underst<strong>and</strong>ing of the functioning of organic light-emitting diodes.<br />

Having gained some insight into the non-linear mechanisms giving rise to localized<br />

electronic (intragap) states, how dopants, mentioned in the previous Chapter, might influence<br />

these states. Two alternatives are feasible:<br />

a) the non-linear aspect dominates, i. e. the picture developed so far is still valid but only<br />

some details are modified due to the disorder. On the basis of the Born approximation we<br />

have indeed calculated [12] the electronic structure of the kink solution in the presence of<br />

impurities <strong>and</strong> found that the spatial extension of this structure is enlarged when the doping<br />

118


8.5 Perspective<br />

concentration is raised. In accordance with the closure of the gap at a critical concentration<br />

we find that the width of the kink tends to infinity at this value. The more interesting case<br />

of the polaron, however, could not be solved satisfactorily due to numerical instabilities,<br />

which could not be avoided. For details see Ref. [12].<br />

b) the more subtle case of the competition between this non-linear mechanism <strong>and</strong> the<br />

multiple scattering processes off the impurities, which is treated in terms of the T-matrix or<br />

the CPA gives also rise to localized states. The experimental observation that the formation<br />

of these states is independent of microscopic details leads to the conclusion that the non-linear<br />

aspect is the dominant one. A fully self-consistent treatment of this problem with an impurity<br />

located at x o <strong>and</strong> (for simplicity) a kink at x 1 gives complicated coupled integral equations<br />

[19] which have not been solved. A recent investigation for the case of a kink lattice<br />

shows both mechanisms working quite independently, however, the method employedwelche<br />

Methode ? does not give a full self-consistent solution..<br />

Summing up this Chapter we note that the non-linear excitations play a dominant role<br />

in various physical applications of conjugated polymers. But a full underst<strong>and</strong>ing of the interplay<br />

of various mechanisms giving rise to these localized states has not been reached yet.<br />

8.5 Perspective<br />

The potential technical applications have stimulated a myriad of experimental <strong>and</strong> theoretical<br />

studies. It is obvious that similar investigations have also been performed, mostly along<br />

different lines, by various groups. The disorder aspect, responsible for the observed metalinsulator<br />

(or semiconductor) transition, in conjunction with a kink (soliton) or polaron lattice<br />

has been treated by many authors [20] in all kinds of approximate approaches. All these studies<br />

resulted in the same prediction that such an M-I transition would occur at the experimentally<br />

observed dopant concentration level. It is now clear from the foregoing Chapter<br />

that only a fully self-consistent treatment will give a satisfactory answer to this question.<br />

But since the applications for conjugated polymers envisaged at present focus on optical<br />

properties rather than electronic transport this question has been lost out of sight. Still, the<br />

nature <strong>and</strong> dynamics of localized electronic states in these materials must be fully understood.<br />

In addition, from an application point of view, these polymers appear to be similar to<br />

conventional semiconductors. The recent proposal [21] that the non-linear excitations actually<br />

modify the conventional picture, e. g. at the interface between a polymer <strong>and</strong> a metal (or<br />

conventional semiconductor) space charge regions (depletion layers) differ from an inorganic<br />

semiconductor, finds renewed interest because transistors made from conjugated polymers<br />

are feasible <strong>and</strong> of interest for the integration of optical <strong>and</strong> electronic components.<br />

From a theoretical point of view the functional integral techniques promise interesting<br />

opportunities to make contacts to other areas of research. Universal properties have been dis-<br />

119


8 Non-Linear Excitations <strong>and</strong> the Electronic Structure of Conjugated Polymers<br />

covered [22] in the absorption spectrum, discussed earlier, as well as in the distribution of<br />

energy levels in certain disordered systems [23]. We propose that there is a close connection<br />

between both aspects. A careful treatment of the underlying correlation functions, including<br />

a more general type of disorder than discussed here, has to be performed. We expect that<br />

the result will lead to new universality classes which technically spoken will show up in different<br />

supersymmetric non-linear s models. Work along this line is in progress.<br />

In summary, conjugated polymers pose interesting problems, both for applications <strong>and</strong><br />

pure theoretical studies. Both aspects have matured during the past 15 years but still questions<br />

of a more general nature are left unanswered.<br />

Acknowledgements<br />

The author is indebted to all his collaborators for enlightening <strong>and</strong> stimulating discussions<br />

as well as fruitful collaborations. Thanks to A.R. Bishop, F. Bronold, H. Büttner, D.K.<br />

Campbell, K. Harigaya, U. Sum, Y. Wada, <strong>and</strong> M. Wolf.<br />

References<br />

1. a) A.J. Heeger, S. Kivelson, J.R. Schrieffer, W.P. Su: Rev. Mod. Phys. 60, 781 (1988)<br />

b) T.A. Skotheim (Ed.): H<strong>and</strong>book of Conducting Polymers. Dekker, New York (1986)<br />

c) J.L. Brédas, R. Silbey (Eds.): Conjugated Polymers. Kluwer, Dordrecht (1991)<br />

d) W.R. Salaneck, I. Lundström, B. Ranby (Eds.): Conjugated Polymers <strong>and</strong> Related Materials. Oxford<br />

University Press, Oxford (1993)<br />

e) Proc. Int. Conf. Science Technology of Synthetic Metals: ICSM ’90, Synth. Met. 41–43 (1991);<br />

ICSM ’92, Synth. Met. 55–57 (1993); ICSM ’94, Synth. Met. 69–71 (1995)<br />

2. J.H. Burroughes et al.: Nature 347, 539 (1990)<br />

3. P. Vogl <strong>and</strong> D.K. Campbell: Phys. Rev. Lett. 62, 2012 (1989); Phys. Rev. B 41, 12 797 (1990)<br />

4. For a review see: D. Baeriswyl, D.K. Campbell, S. Mazumdar, in: Conjugated Conducting Polymers,<br />

H. Spiess (Ed.), Springer Series in Solid State Sciences 102, 7 (1992)<br />

5. W.P. Su, J.R. Schrieffer, A.J. Heeger: Phys. Rev. B 22, 2099 (1980)<br />

6. D. Baeriswyl, E. Jeckelmann, in: Electronic Propteries of Polymers, H. Kuzmany, M. Mehring, S.<br />

Roth (Eds.), Springer Series in Solid State Sciences 107, 16 (1992)<br />

7. D.K. Campbell <strong>and</strong> A.R. Bishop: Nucl. Phys. B 200, 297 (1982)<br />

8. U. Sum, K. Fesser, H. Büttner: Ber. Bunsenges. Phys. Chem. 91, 957 (1987)<br />

9. S. Brazovskii <strong>and</strong> N. Kirova: Pis’ma Zh. Eksp. Teor. Fiz 33, 6 (1981) [JETP Lett. 33, 4 (1981)]<br />

10. K. Fesser, A.R. Bishop, D.K. Campbell: Phys. Rev. B 27, 4804 (1983)<br />

11. a) K. Fesser: J. Phys. C 21, 5361 (1988)<br />

b) K. Iwano <strong>and</strong> Y. Wada: J. Phys. Soc. Jpn. 58, 602 (1989)<br />

120


References<br />

12. F. Bronold <strong>and</strong> K. Fesser, in: Nonlinear Coherent Structures in Physics <strong>and</strong> Biology, M. Remoissenet<br />

<strong>and</strong> M. Peyrard (Eds.), Springer Lecture Notes in Physics 393, 118 (1991)<br />

13. K.B. Efetov: Adv. in Phys. 32, 53 (1983)<br />

14. M. Wolf <strong>and</strong> K. Fesser: Ann. Physik 1, 288 (1992)<br />

15. M. Wolf <strong>and</strong> K. Fesser: J. Phys. Cond. Matter 5, 7577 (1993)<br />

16. K. Harigaya, Y. Wada, K. Fesser: Phys. Rev. Lett. 63. 2401 (1989);<br />

Phys. Rev. B 42, 1268 <strong>and</strong> 1276 (1990)<br />

17. K. Harigaya, Y. Wada, K. Fesser: Phys. Rev. B 43, 4141 (1991)<br />

18. F. Bronold: Doct. Thesis, Univ. Bayreuth (1995)<br />

19. K. Fesser: Prog. Theor. Phys. Suppl. 113, 39 (1993)<br />

20. a) E.J. Mele <strong>and</strong> M.J. Rice: Phys. Rev. B 23, 5397 (1981)<br />

b) G.W. Ryant <strong>and</strong> A.J. Glick: Phys. Rev. B 26, 5855 (1982)<br />

c) S.R. Philpott et al.: Phys. Rev. B 35, 7533 (1987)<br />

d) E.M. Conwell, S. Jeyadev: Phys. Rev. Lett. 61, 361 (1988)<br />

21. a) S.A. Brazovskii, N. Kirova: Synth. Met. 55–57, 4385 (1993)<br />

b) G. Paasch <strong>and</strong> T.P.H. Nguyen: unpublished (1995)<br />

22. K. Kim, R.H. Mckenzie, J.W. Wilkins: Phys. Rev. Lett. 71, 4015 (1993)<br />

23. B.D. Simons <strong>and</strong> B.L. Altshuler: Phys. Rev. B 48, 5422 (1993)<br />

121


9 Diacetylene Single Crystals<br />

Markus Schwoerer, Elmar Dormann, Thomas Vogtmann, <strong>and</strong> Andreas Feldner<br />

9.1 Introduction<br />

Polydiacetylenes can be grown as macroscopic polymer single crystals [1–3]. This property<br />

is unique. They comprise one linear polymer axis <strong>and</strong> can have spatial extensions of up to<br />

several millimetres or more in all three spatial directions (Fig. 9.1). The covalent chemical<br />

bonds along the polymer axis make them mechanically strong along the corresponding crystallographic<br />

axis. Their Young’s modulus is about a quarter of the Young’s modulus of steel<br />

[4, 5] <strong>and</strong> their tensile strength along the polymer axis has been reported to exceed that of<br />

steel [6]. Weak bonds of van der Waals-type perpendicular to the polymer axis are responsible<br />

for extremely low dimensional – generally one-dimensional – macroscopic electronic<br />

properties of these polydiacetylene single crystals. They are insulators, they can become<br />

pyro- or ferroelectric, <strong>and</strong> they show large optical non-linearities. These electronic <strong>and</strong> optical<br />

properties are primarily determined by the p-electron system along the polymer axis.<br />

Most polydiacetylenes differ only in the substituents R <strong>and</strong> R' (Scheme 9.1). But the electro-<br />

Figure 9.1: Photo of macroscopic paratoluylsulfonyloximethylene-diacetylene (TS6) single crystals under<br />

polarized light. Monomer (top), polymer (bottom).<br />

122 Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


9.1 Introduction<br />

Scheme 9.1: Polydiacetylene in its isomeric structures: acetylene-type (a) <strong>and</strong> butatriene-type (b). R<br />

<strong>and</strong> R' are the substituents for different diacetylenes (see also Tab. 9.1).<br />

nic structure of their all-trans planar carbon chain is at first approximation of acetylene-type<br />

(a) rather than of butatriene-type (b). Both isomeric structures contain a non-interrupted<br />

p-electron system along the carbon chain.<br />

The term polydiacetylene is somewhat puzzling, at least for a physicist. However, it<br />

becomes quite clear if one takes into account the structure of the diacetylene monomer<br />

(Fig. 9.2). A typical substituent R is paratoluylsulfonyloximethylene (Scheme 9.2).<br />

The diacetylene with R = R' = paratoluylsulfonyloximethylene is termed TS6 (sometimes<br />

TS). It was shown by G. Wegner <strong>and</strong> his co-workers in a series of works, published in<br />

the early 1970s [1, 7], that large molecular crystals can be grown from a solution of TS6,<br />

e. g. in acetone, <strong>and</strong> that these monomer diacetylene crystals can be converted by a topochemical<br />

(or solid state) 1,4-addition reaction to the polydiacetylene single crystals<br />

(Fig. 9.2). The crystal structures of TS before <strong>and</strong> after the reaction have been investigated<br />

in detail by Kobelt <strong>and</strong> Paulus [8], Bloor et al. [8], <strong>and</strong> Enkelmann [9] <strong>and</strong> are sketched in<br />

Fig. 9.3 <strong>and</strong> in Tab. 9.1.<br />

Figure 9.2: The monomer diacetylene crystal is converted by a topochemical or solid state reaction to<br />

the polydiacetylene single crystal.<br />

123


9 Diacetylene Single Crystals<br />

Scheme 9.2: Diacetylene monomer with the substituents R = R' = paratoluylsulfonyloximethylene (TS6).<br />

Figure 9.3: Monomer <strong>and</strong> polymer crystal structure of TS6 deduced from X-ray data [9]. Note the reactive<br />

carbene at the chain end.<br />

Both crystal structures are monoclinic with two monomers or monomer units of different<br />

orientation per unit cell (Fig. 9.3 sketches only one of these two.). The chemical bond<br />

between two carbons of nearest neighbour diacetylenes (1,4-addition) results in the linear<br />

polymer chain <strong>and</strong> the small change in the lattice parameter along the b-axis prevents the<br />

destruction of the macroscopic single crystal during the topochemical reaction. Several surfaces<br />

of diacetylene crystals have been studied by atomic force microscopy (AFM) in order<br />

to investigate both, the single crystal surface structure <strong>and</strong> solid state reactions at the surface<br />

[129, 131].<br />

124


9.1 Introduction<br />

Table 9.1: TS monomer <strong>and</strong> polymer crystals (monoclinic, space group P2 1 /c) [9].<br />

T/K a/Å b/Å c/Å b/ 8 D x<br />

g/cm 3<br />

TS monomer 120 14.61(1) 5.11(1) 25.56(5) 92.0(5) 1.46<br />

TS monomer 295 14.60 5.15 15.02 118.4 1.40<br />

TS monomer 295 14.65(1) 5.178(2) 14.94(1) 118.81(3) 1.40<br />

TS polymer 295 14.993(8) 4.910(3) 14.936(10) 118.14(4) 1.483<br />

TS polymer 120 14.77(1) 4.91(1) 25.34(2) 92.0(5) 1.51<br />

The topochemical reaction can be induced thermally <strong>and</strong>/or photochemically <strong>and</strong>/or<br />

by electron-beam irradiation. For TS the thermal conversion versus time (Fig. 9.4) is<br />

strongly temperature dependent <strong>and</strong> highly non-linear. The thermodynamics of the integral<br />

reaction has been investigated extensively by Bloor et al. [10], Eckhardt et al. [11], Chance<br />

et al. [11], <strong>and</strong> others. The reaction diagram (Fig. 9.5) for TS shows that the dark reaction is<br />

thermally activated <strong>and</strong> has an activation energy of 1 eV per monomer. It is exothermic with<br />

a polymerization enthalpy of 1.6 eV per addition of one monomer. The entire reaction is irreversible<br />

<strong>and</strong> the TS6-polydiacetylene (PTS) crystals are not solvable in ordinary solvents.<br />

During the solid state reaction almost all properties of the diacetylene crystals change<br />

drastically, e. g. the transparent monomer crystals are converted to polymer crystals with<br />

highly dichroic, strongly reflecting surfaces which contain the b-axis. In transmission the<br />

polydiacetylene crystals can only be investigated as thin films (Fig. 9.6). Their absorption<br />

spectrum, e. g. for TS, clearly shows the vibronic spectrum due to the single, double <strong>and</strong> triple<br />

bonds of the polymer chain [12]. Batchelder et al. [13] investigated extensively the optical<br />

absorption, reflection, <strong>and</strong> Raman spectra of TS single crystals.<br />

While these experiments were directed towards the study of the electronic excitations<br />

of the bulk <strong>and</strong> their coupling to the vibrations, Sebastian <strong>and</strong> Weiser [14] investigated the<br />

Figure 9.4: Time conversion curves for the thermal polymerization of PTS at 60 8C (.), 70 8C (#), <strong>and</strong><br />

80 8C (d) [9].<br />

125


9 Diacetylene Single Crystals<br />

Figure 9.5: Reaction diagram for the thermal polymerization of TS (solid curve) <strong>and</strong> photopolymerization<br />

of 4BCMU (dashed curve) [10, 11].<br />

Figure 9.6: Absorption spectra of a thin TS diacetylene single crystal for light polarized parallel <strong>and</strong><br />

perpendicular to the polymer axis b (T = 300 K). Monomer (M) <strong>and</strong> polymer (P) absorption.<br />

defects by electroabsorption. As an example Fig. 9.7 shows the absorption <strong>and</strong> the electroabsorption<br />

spectra, which they have analyzed in great detail.<br />

Since about 1980 <strong>and</strong> especially during our work for the Sonderforschungsbereich 213<br />

we have synthesized several new diacetylenes, the substituents of which are shown in<br />

Tab. 9.2 [15]. For selected diacetylene single crystals we have investigated:<br />

126


9.1 Introduction<br />

Figure 9.7: Absorption a <strong>and</strong> electroabsorption Da of photoproducts in PTS-monomer as a function of<br />

ko. Full curve: experimental spectra, dashed curves: fit by Lorentzian <strong>and</strong> a charge transfer model for<br />

Da [14].<br />

a) in Section 9.2 the elementary steps <strong>and</strong> structures during the solid state photopolymerization<br />

by transient optical spectroscopy, by electron spin resonance (ESR), <strong>and</strong> by electron<br />

nuclear double resonance (ENDOR);<br />

b) in Section 9.3 the application of the photopolymerization of thick diacetylene single<br />

crystals as a very effective holographic storage process;<br />

c) in Section 9.4 the tailoring of diacetylenes as ferro or pyroelectric crystals, which do<br />

not dem<strong>and</strong> considerable efforts for the poling processes <strong>and</strong> which show good thermal stability;<br />

d) in Section 9.5 the optical non-linearity of second <strong>and</strong> of third order <strong>and</strong> their application<br />

for an optical device with femtosecond time resolution.<br />

The present paper is a review of our work with diacetylene single crystals.<br />

127


9 Diacetylene Single Crystals<br />

Table 9.2: A survey of the investigated diacetylenes [15].<br />

Survey of the investigated diacetylenes<br />

R 1 C C C C R 2<br />

DNP<br />

TS<br />

PD-TS<br />

CD -TS<br />

2<br />

FBS<br />

O 2 N<br />

CH 2 O NO<br />

CH 2 O SO 2<br />

D<br />

CD 2 O SO2<br />

D<br />

H<br />

CD2 O SO2<br />

H<br />

CH<br />

2<br />

O SO<br />

2<br />

2<br />

CH 3<br />

D<br />

CD 3<br />

D<br />

H<br />

CH 3<br />

H<br />

F<br />

R 2<br />

ability to polymerize<br />

therm. γ<br />

= R 1<br />

= R 1<br />

= R 1<br />

= R 1<br />

= R 1<br />

+++<br />

+++<br />

+++<br />

+++<br />

+++<br />

-<br />

+++<br />

+++<br />

+++<br />

+++<br />

IPUDO<br />

Name R 1<br />

CH O<br />

O CH 3<br />

(CH 2 ) 4 O C NH CH<br />

= R 1<br />

-<br />

+++<br />

NP/PU<br />

CH 2<br />

O NO2<br />

CH 3<br />

O<br />

2 C NH<br />

-<br />

+<br />

O<br />

NP/4-MPU CH2 O NO<br />

2 CH2 O C NH<br />

CH 3 - +<br />

NP/MBU<br />

DNP/MNP<br />

O 2<br />

O H<br />

CH2 O NO2 CH2 O C NH C + +<br />

CH 3<br />

(-)(S) or (+)(R)<br />

N<br />

O 2 N<br />

CH<br />

NO CH O CH<br />

+++<br />

2 O 2<br />

2<br />

3<br />

DNP/PU<br />

CH<br />

2<br />

N<br />

O<br />

O 2<br />

O<br />

NO2 CH2 O C NH<br />

- -<br />

DNP/4-MPU<br />

O2<br />

N<br />

CH<br />

2 O<br />

O<br />

NO2 CH2 O C NH CH 3<br />

- -<br />

O 2 N<br />

DNP/DMPU CH2 O<br />

NO2<br />

CH<br />

O CH3<br />

O NH<br />

2 C<br />

CH 3<br />

- -<br />

DNP/MPU<br />

O 2 N<br />

CH O<br />

2<br />

O H<br />

NO2 CH2 O C NH C<br />

CH<br />

(-)(S) or (+)(R)<br />

3<br />

- -<br />

TS/FBS CH2 O SO2<br />

CH 3 CH2<br />

O SO 2 F +++<br />

FBS/TFMBS CH2 O SO 2 F CH2 O SO2 CF 3 +++<br />

128


9.2 Photopolymerization<br />

9.2 Photopolymerization<br />

9.2.1 Carbenes<br />

The aim of this Chapter is to review our spectroscopic work towards the analysis of both,<br />

the electronic structures <strong>and</strong> the dynamics of the intermediate reaction products (Fig. 9.2),<br />

during the photopolymerization, i. e. after the excitation of the solid state reaction by light.<br />

For the entirely thermally activated solid state polymerization detailed spectroscopy of the<br />

intermediate states turned out to be difficult or not very efficient. One exception was the<br />

identification of carbenes as reactive species during the thermal solid state polymerization<br />

of TS6 (Fig. 9.8).<br />

In contrast to radicals with one non-bonded electron carbenes have two non-bonded<br />

electrons. It has been shown for the first time by Wassermann et al. [16] that the electronic<br />

ground state of pure methylene (:CH 2 ) is a triplet state, where the total spin quantum number<br />

S is 1. Because of their non-centrosymmetry most molecular triplet states show a splitting<br />

into three components even in the absence of an external field. This splitting is due to the<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

2hν<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R C R<br />

C<br />

C<br />

R<br />

C<br />

C<br />

R<br />

C<br />

C<br />

C<br />

R<br />

kT<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R R<br />

C<br />

C<br />

C<br />

C<br />

R C R<br />

C<br />

R<br />

C<br />

C<br />

C<br />

kT<br />

R<br />

C<br />

C<br />

C<br />

R C<br />

R<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R C R<br />

C<br />

C<br />

C<br />

R C R<br />

C<br />

C<br />

R C<br />

C R<br />

C<br />

C<br />

R<br />

C<br />

C R<br />

C<br />

C<br />

C<br />

R<br />

kT<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R C R<br />

R C<br />

C<br />

C<br />

C<br />

R C<br />

R<br />

C<br />

C<br />

C<br />

R<br />

Figure 9.8: UV photopolymerization of TS6: The monomer crystal is irradiated with an UV-flash. The<br />

dimer is formed from the monomer by a photoreaction, then a series of thermally activated monomer<br />

addition reactions leads via the diradicals (DR) <strong>and</strong> dicarbenes (DC) to the polymer.<br />

R<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

R C C R<br />

kT<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R C R<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

C<br />

R C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

R<br />

R C C R<br />

kT<br />

C<br />

R C<br />

R<br />

C<br />

C<br />

C<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

R C R<br />

R<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

R C R<br />

C<br />

C<br />

R C C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

R C<br />

C<br />

R<br />

R<br />

R<br />

R<br />

R<br />

C<br />

C<br />

C<br />

R<br />

kT<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

C<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

C<br />

R C<br />

R<br />

C<br />

C<br />

C<br />

R<br />

129


9 Diacetylene Single Crystals<br />

magnetic dipole-dipole coupling of the two unpaired electrons <strong>and</strong> is called zero-field splitting.<br />

In an external magnetic field the resulting fine structure of the ESR spectra is highly anisotropic,<br />

i. e. it strongly depends on the direction of the external magnetic field with respect<br />

to the orientation of the tensor describing the magnetic dipole-dipole interaction. This interaction<br />

is of course an intramolecular property <strong>and</strong> therefore single crystals are ideal c<strong>and</strong>idates<br />

for the measurement <strong>and</strong> the quantitative analysis of molecular triplet states by ESR. As originally<br />

shown by the work of Hutchison <strong>and</strong> Mangum [17], <strong>and</strong> van der Waals [18] the ESR<br />

spectrum of a molecular triplet state is described by the spin Hamiltonian T H s [19]<br />

T H s ˆ B B o g S^<br />

‡D S^ 2<br />

z ‡ E…S^ 2<br />

x S^ 2<br />

y † ; …1†<br />

where m B is Bohr’s magneton, B 0 the external magnetic field, g the spectroscopic splitting<br />

factor, ^S the spin operator for a spin with total spin quantum number S = 1, ^S u with u = x, y, z<br />

are the components along the principal axes of the magnetic dipole-dipole interaction tensor.<br />

D <strong>and</strong> E represent the two independent values of this tensor, the trace of which is zero. They<br />

usually are called zero-field splitting parameters:<br />

D ˆ o<br />

4p<br />

E ˆ o<br />

4p<br />

3<br />

4 g2 2 z 2<br />

B hr2 r 5 i ; …2†<br />

3<br />

4 g2 2 x 2<br />

B hy2 r 5 i ; …3†<br />

r is the distance of the two unpaired electrons <strong>and</strong> x, y, <strong>and</strong> z are the components of r along<br />

the principal axes of the magnetic dipole-dipole interaction tensor of these two electrons.<br />

Only for systems of cylindrical symmetry (around z) the zero-field splitting parameter E<br />

vanishes (E = 0). And only for spherical symmetry (Ax 2 S = Ay 2 S = Az 2 S = 1/3 Ar 2 S) the entire<br />

fine structure is zero. The values of D <strong>and</strong> E for the triplet carbenes as detected during the<br />

purely thermal solid state polymerization [20] are shown in Tab. 9.3 in comparison with typical<br />

values of different molecular triplet states. By these values <strong>and</strong> especially by the strong<br />

anisotropy of the fine structure in the ESR spectrum these carbenes (as sketched in Fig. 9.8)<br />

are clearly identified as reactive species during the pure thermal solid state polymerization<br />

of TS. They will play a major role in the electronic structures of the intermediate products<br />

during the photopolymerization as described in the following paragraphs.<br />

Table 9.3: Zero-field splitting parameters for the reactive species during the thermal solid state polymerization<br />

of TS [20], pure methylene [16], diphenylmethylene [21], <strong>and</strong> benzene [22] in their first excited triplet<br />

state.<br />

D<br />

hc =cm–1<br />

E<br />

hc =cm–1<br />

TS 0.2731 –0.0048<br />

:CH 2 0.6636 0.0003<br />

:C(C 6 H 5 ) 2 0.39644 –0.01516<br />

C 6 H 6 0.1581 –0.0046<br />

130


9.2 Photopolymerization<br />

9.2.2 Intermediate photoproducts<br />

Further experiments for analyzing the electronic structure of intermediate states were not<br />

very successful, until Sixl et al. [23] <strong>and</strong> Bubeck et al. [24] published their first low-temperature<br />

spectroscopic experiments on partially photopolymerized TS crystals. Thereby, the<br />

monomer crystal is cooled to 4.2 K in the dark. Then it is irradiated with UV light<br />

(l ^ 310 nm) for a short period. After this procedure a large number of different species<br />

show up in both, the optical absorption <strong>and</strong> the ESR spectrum. They persist at helium temperature.<br />

Subsequent annealing in the dark produces further intermediate reaction products<br />

which are also identified by their optical or ESR spectra. <strong>Final</strong>ly, further irradiation with<br />

visible light, which is absorbed only by the intermediate reaction products, produces still<br />

more <strong>and</strong> different reaction products. In a comprehensive series of investigations [25–47]<br />

most of these intermediate products have been identified <strong>and</strong> classified. Moreover, the mechanisms<br />

of their production <strong>and</strong> their reaction kinetics have been analyzed. According to<br />

Sixl, there exist three different series of intermediate products:<br />

1. The diradical-dicarbene series: DR 2 , DR 3 , DR 4 , DR 5 , DR 6 , DC 7 , DC 8 , DC 9 , DC 10 ,<br />

DC 11 … polymer;<br />

2. The asymmetric carbene series (AC);<br />

3. The stable oligomer series (SO).<br />

The DR-DC series leads directly from the monomer to the polymer. It is initiated <strong>and</strong><br />

processed in the following simple <strong>and</strong> clear way. The monomer crystal is irradiated with an<br />

UV flash <strong>and</strong> subsequently rests in the dark. The flash excites the monomers <strong>and</strong> produces<br />

dimers (DR 2 ). These react by a thermally activated step by step addition of monomers, as illustrated<br />

in Fig. 9.8. In the following paragraphs we will show for a few selected examples<br />

how these results have been achieved <strong>and</strong> we will present details of both, the electronic<br />

structures <strong>and</strong> the dynamics.<br />

The AC <strong>and</strong> the SO series are produced by additional irradiation with light, i. e. these<br />

are photoproducts which do not necessarily arise during the solid state polymerization of<br />

TS6. Although they are an important part of the entire variety of structures in partially polymerized<br />

TS6 crystals, we will not treat them in this review. They have been described extensively<br />

by Sixl in his papers cited above <strong>and</strong> in Ref. [48].<br />

9.2.3 Electronic structure of dicarbenes<br />

9.2.3.1 Electron spin resonance of quintet states ( 5 DC n )<br />

As an example for the electronic structure analysis of the intermediate states with ESR, we<br />

will review below the ESR spectra of the dicarbenes DC 7 …DC 13 [28, 32, 42]. They are<br />

characterized by their fine structure <strong>and</strong> temperature dependence. Figure 9.9 shows the ESR<br />

131


9 Diacetylene Single Crystals<br />

ESR-Signal<br />

T<br />

Q<br />

T<br />

T<br />

Q<br />

Q Q T<br />

Q<br />

Q<br />

Q<br />

T<br />

Q<br />

T=10 K<br />

ν =9,46 GHz<br />

T<br />

0 200 400 600<br />

Magnetic Field B 0/mT<br />

Figure 9.9: The ESR spectrum of perdeuterated TS after irradiation for 1000 s. The signals marked<br />

with T arise from triplet states <strong>and</strong> those with Q from quintets. The T-lines are microwave saturated. The<br />

magnetic field B 0 is oriented parallel to the z-axis of the quintet fine structure tensor [28].<br />

spectrum of a perdeuterated TS crystal which has been irradiated with UV light (313 nm) of<br />

a mercury high-pressure arc lamp (HBO 200) at 4.2 K for about 1000 s. All ESR lines in<br />

Fig. 9.9 labelled with Q are due to dicarbenes in their quintet state (S = 2). Prior to the UV<br />

irradiation no ESR signal is observed.<br />

The temperature dependences of the ESR signals (Fig. 9.10) show one common feature:<br />

the ESR intensities vanish for T ? 0, i. e. the ESR signals are thermally activated <strong>and</strong><br />

the ground state is spinless. But the temperatures for the maximum intensities are different<br />

for each ESR signal, indicating different activation energies.<br />

1<br />

10<br />

5<br />

T/K<br />

2<br />

Signal Intensity<br />

0,5<br />

0<br />

0<br />

0,2 0,4 0,6 0,8<br />

1/T/K -1<br />

Figure 9.10: Temperature dependence of the ESR intensities of dicarbenes DC 9 ,DC 10 , <strong>and</strong> DC 11 . The<br />

calculated lines have been fitted by Eq. 9 with the activation energies De* SQ for the quintet states [33, 47].<br />

132


9.2 Photopolymerization<br />

The resonance fields for fixed microwave frequencies are strongly anisotropic.<br />

Figure 9.11 shows as an example this anisotropy for the dicarbenes DC 9 ,DC 10 ,DC 11 , <strong>and</strong><br />

DC 12 , respectively. The external field has been rotated to the polymer backbone plane. Each<br />

anisotropy belongs to one distinct temperature dependence.<br />

The following model (Fig. 9.12 <strong>and</strong> inset Fig. 9.13) describes quantitatively the anisotropic<br />

fine structure <strong>and</strong> the temperature dependence. The ground state (|S>) of each dicarbene<br />

in first order is a singlet state with the total spin quantum number S = 0. The excited<br />

state (|Q>) in first order is a quintet state with S = 2. The excitation energy is De SQ . The<br />

quintet state is built up by the electronic coupling of two triplet carbenes at both ends of the<br />

oligomer via exchange interaction. R 12 is the distance between the two triplet carbenes. If<br />

the singlet-quintet splitting De SQ is large as compared to the magnetic dipole-dipole coupling<br />

(zero field splitting) within the triplet carbenes, the total spin quantum numbers S = 0<br />

<strong>and</strong> S = 2, respectively, are good quantum numbers, i. e. the quintet state is pure. The spin<br />

Hamiltonian for this case has been analyzed by Schwoerer et al. [34].<br />

Figure 9.11: Angular dependence of the resonance fields B 0 of the 4 dicarbene structures DC 9 ,DC 10 ,<br />

DC 11 , <strong>and</strong> DC 12 in perdeuterated TS-diacetylene crystals. The crystal is rotated so that the external<br />

magnetic field B 0 is in the plane of the polymer backbone. The b-axis is the direction of the polymer<br />

chain, y <strong>and</strong> z are the principal axes of the fine structure tensor. The curves are fitted to the experimental<br />

points by computer calculations. A <strong>and</strong> B indicate the two magnetically equivalent directions of the<br />

molecular orientation within the monoclinic unit cell. Dots: experimental values; lines: calculated by<br />

Q H S 0 (Eqs. 4–7) [32, 42].<br />

Figure 9.12: Dicarbene configuration. The dicarbene molecule consists of n diacetylene units with two<br />

identical triplet carbene chain ends. D t <strong>and</strong> E t are the triplet fine structure parameters of the S = 1 carbene<br />

species. j gives the orientation of the triplet fine structure z-axis with respects to the crystal b-<br />

axis.<br />

133


9 Diacetylene Single Crystals<br />

Figure 9.13: Experimental values for the singlet-quintet splitting De SQ for seven different dicarbenes<br />

DC 7 …DC 13 [42].<br />

If De SQ is smaller than or somewhere in the order of the magnetic dipole-dipole coupling<br />

then the singlet <strong>and</strong> quintet states are mixed. The spin Hamiltonian Q H S for this general<br />

case has been derived in the elegant work of Benk <strong>and</strong> Sixl [35]:<br />

Q H S ˆ g B B 0 …^S 1 ‡ ^S 2 †‡ 1 6 " SQ …^S 1 ‡ ^S 2 † 2 ‡H DD<br />

S<br />

…4†<br />

The first term of Eq. 4 represents the electronic Zeeman term, ^S 1 <strong>and</strong> ^S 2 the spin operators<br />

for two triplet carbenes, <strong>and</strong> the second term represents the electronic exchange interaction.<br />

If ^S 1 <strong>and</strong> ^S 2 couple to a quintet state (S = 2), (^S 1 + ^S 2 ) 2 = S 2 = S(S + 1) = 6. If<br />

they couple to a singlet, S = 0. Therefore, this term directly results in the energy level<br />

scheme, indicated in the inset of Fig. 9.13. The pure singlet <strong>and</strong> the pure quintet states are<br />

split by De SQ which turns out to be the characteristic property of each dicarbene. The third<br />

term of Eq. 4 represents the magnetic dipole-dipole coupling of the two triplet carbenes:<br />

H DD<br />

S<br />

^<br />

ˆ D …S 2 1z<br />

^<br />

1<br />

3 S2 1 †‡E…S2 1x<br />

X S^<br />

1x S^<br />

2x ‡ XAS^<br />

1y S^<br />

2y<br />

^<br />

S 2 ^<br />

1y†‡D…S 2 2z<br />

1<br />

^<br />

3 S2 2<br />

^<br />

†‡E…S 2 2x<br />

^<br />

S 2 2y†‡<br />

X…1 ‡ A† S^<br />

1z S^<br />

2z ‡ Xa…S^<br />

1z S^<br />

2y ‡ S^<br />

1y S^<br />

2z† …5†<br />

D <strong>and</strong> E are the fine structure parameters of the identical triplet carbene chain ends;<br />

x, y, <strong>and</strong> z are the principal axes of the corresponding fine structure tensors. The intercarbene<br />

magnetic dipolar interaction is represented by the parameter<br />

134


9.2 Photopolymerization<br />

X ˆ g 2 2 B … 0=4p†R 3<br />

12 : …6†<br />

The geometrical factors A <strong>and</strong> a are only dependent on the orientation j of the fine<br />

structure tensor z-axis with respect to the b-axis of the crystal.<br />

A ˆ 1 3 sin 2 '; ˆ 3 sin 2': …7†<br />

2<br />

The diagonalization Q H S of yields the allowed ESR transitions, their resonance fields,<br />

<strong>and</strong> their dependence on the orientation of the external magnetic field B 0 . At a first sight,<br />

the number of fit parameters seems to be high: D, E, g, the orientation of the triplet fine<br />

structure tensor, R 12 , <strong>and</strong> De SQ . But besides R 12 <strong>and</strong> De SQ these parameters are well-known<br />

from earlier ESR experiments on triplet carbenes [20, 49, 50]. Therefore, R 12 <strong>and</strong> De SQ are<br />

the only free to fit parameters. Figure 9.11 shows the result of the fit. The ESR anisotropy<br />

was calculated by exact diagonalization of Q H S with the fitting parameters R 12 <strong>and</strong> De SQ .In<br />

all cases the fit is almost perfect. Furthermore, it turns out that the intertriplet magnetic dipole-dipole<br />

interaction does not influence the results if R 12 exceeds 12 Å. Therefore, only<br />

the parameter De SQ remains to fit for each dicarbene. The obvious differences between the<br />

anisotropies of the different dicarbenes (Fig. 9.11) are due to De SQ only! The result is shown<br />

in Fig. 9.13 where De SQ is decreasing exponentially with increasing number n of monomer<br />

units:<br />

" SQ ˆ " 0 SQ e R=R 0<br />

; …8†<br />

with R = n1 m (1 m is the length of the monomer unit within the oligomer). The slope in<br />

Fig. 9.13 yields R 0 = 5.4 Å&10 a 0 (a 0 = Bohr radius). The origin of the abscissa in Fig. 9.13<br />

was deduced by Neumann [47]. But even without the knowledge of this origin the value of R 0<br />

as compared to a 0 shows that the triplet carbene is highly delocalized <strong>and</strong> not at all restricted<br />

to the end of the oligomer, as one might think because of Fig. 9.12. The model also gives the<br />

temperature dependencies of the dicarbene signals. Because in first order the ground state is<br />

spinless, the ESR intensities I of the quintet states (beyond Curie’s law) must be thermally activated:<br />

I / 1 T ‰5 ‡ exp…" SQ =kT†Š :<br />

…9†<br />

The lines in Fig. 9.10 are calculated with Eq. 9 by fitting De* SQ . For the dicarbenes<br />

DC 8 ,DC 9 ,DC 10 , <strong>and</strong> DC 11 , both values, De* SQ <strong>and</strong> De SQ , are determined respectively. Within<br />

the experimental uncertainty they are identical [32, 42]. This latter result is an excellent<br />

proof for the dicarbene model because De* SQ <strong>and</strong> De SQ were determined from two completely<br />

different properties: the anisotropy of the fine structure <strong>and</strong> the intensity of the ESR spectra.<br />

We therefore have no doubt that we really did observe dicarbenes (as sketched in Figs. 9.8<br />

<strong>and</strong> 9.12). They are an ideal modelling substance for short linear oligomers of diacetylenes<br />

which are perfectly oriented in the single-crystal lattice. The longest dicarbene (DC 13 ) observed<br />

has according to our model a length of (R =1364,9 Å) 64 Å!<br />

135


9 Diacetylene Single Crystals<br />

9.2.3.2 ENDOR of quintet states<br />

The most important result of the preceding Chapter is the delocalization of the spins S = 1<br />

of the two triplet carbenes which is necessary for their coupling over a distance of up to<br />

64 Å (!) to the well defined quintet states. It was therefore attractive to measure <strong>and</strong> analyze<br />

the ENDOR spectrum of at least one quintet dicarbene. ENDOR should detect at<br />

least the protons of the CH 2 groups of the substituents if the nuclear spins of these protons<br />

are hyperfine coupled with the electron spin of the triplet carbene. As compared to<br />

ENDOR with an electron spin of S = 1/2, the complication is the high anisotropy of the<br />

five electronic Zeeman levels Q u (u = 1 … 5) of the quintet state (Fig. 9.14). Not only<br />

their energy separation, i. e. the ESR transition fields (Fig. 9.11), are strongly dependent<br />

on the direction of the external field. Also their effective spin S eff , i. e. the expectation<br />

value of the spin, is dependent on the direction <strong>and</strong> on the strength of the external field<br />

B 0 . Therefore in this case the orientation quantum number m s is an unsuitable quantum<br />

number not only because of the large zero field splitting, as expressed by D, but also because<br />

of the singlet-triplet mixing, as expressed by De SQ . As described in the preceding<br />

paragraph this problem has been solved with high accuracy, Hartl et al. [36] measured<br />

the ENDOR spectra <strong>and</strong> their anisotropies for one quintet dicarbene ( 5 DC 10 ), which is accessible<br />

most comfortably at T = 4.2 K (Fig. 9.10) <strong>and</strong> for which the total spin quantum<br />

number (S = 2) is a good quantum number. The aim of these experiments was to determine<br />

the hyperfine coupling constants with the above-mentioned protons <strong>and</strong> subsequently<br />

to extract the electron spin density from these values, i. e. the delocalization of<br />

the triplet carbene quantitatively.<br />

Figure 9.14: Quintet state with an external field B o interacting with one proton (I = 1/2). All allowed<br />

ESR transitions (a–d) <strong>and</strong> NMR transitions (1–5) are shown. By ENDOR, in first order, only the two<br />

NMR transitions directly connected to the observed ESR line are detectable [36].<br />

136


9.2 Photopolymerization<br />

The spin Hamiltonian Q H S;i of a quintet dicarbene coupled to one individual proton,<br />

numbered i, is<br />

Q H s;i ˆ QH 0 s ‡ g I K B 0 ^I i ‡ ^SA i I i :<br />

…10†<br />

The first term in Eq. 10 is given by Eq. 4 <strong>and</strong> the second is the nuclear Zeeman energy.<br />

^I i is the nuclear spin operator for nuclear spin 1/2. The third term is the hyperfine interaction<br />

of the individual proton i, as defined by the hyperfine tensor A i . ^S is the total electron<br />

spin operator, ^S ˆ ^S 1 ‡ ^S 2 . As nuclear dipole-dipole interaction can be neglected we<br />

will omit the index i in the following.<br />

The nuclear terms in Eq. 10 are small as compared to the electronic terms. Therefore,<br />

we treat them in first order perturbation theory, taking the solutions of Q H 0 S as basis. Q H 0 S<br />

has five quintet eigenstates |Q u S , u = 1, 2, 3, 4, 5. For very high fields they become the<br />

high field states |Q m S, for which ^S z |Q m s<br />

S = km s |Q m s<br />

S,m s = +2, +1, 0, –1, –2. But for the<br />

fields used in ordinary ESR spectrometers the electronic Zeeman energy is in comparison to<br />

the fine structure not large <strong>and</strong> therefore the electron spin is not quantized along the external<br />

field B 0 [39]. This results in a strong |Q u S-dependence on the direction of B 0 with respect to<br />

the crystal axes.<br />

From Q H 0 S the effective spin<br />

S u eff ˆ<br />

D E<br />

Qu j ^S jQ u<br />

…11†<br />

can be calculated exactly [34]. It is this electron spin which interacts via A with the proton<br />

spin.<br />

The first order perturbation theory of the nuclear terms calculates the shift Dn u of the<br />

individual proton Larmor frequency with respect to the free proton Larmor frequency n F<br />

(ENDOR shift):<br />

hDn u = hn u – g I m K |B 0 | , (12)<br />

(g I m K |B 0 |/h = n F ). The result of the calculation of the Larmor frequencies n u of the hyperfine<br />

coupled protons is [36, 43]:<br />

h u ˆjS u eff A g I K B 0 j : …13†<br />

For a quintet state one should observe 5 different ENDOR lines per proton. This is illustrated<br />

in Fig. 9.14 where the observed NMR transitions in the ENDOR experiment are indicated<br />

by the ciphers (n) = (1), (2), (3), (4), <strong>and</strong> (5); the first order ESR transitions are indicated<br />

by the lower case letters (a), (b), (c), <strong>and</strong> (d).<br />

Fig. 9.15 shows four ENDOR spectra as detected via the ESR transitions (a), (b), (c),<br />

<strong>and</strong> (d), respectively. Four protons i = 1, 2, 3, <strong>and</strong> 4, respectively, are clearly separated from<br />

the free proton frequency n F . In the vicinity of n F a large number of weakly coupled protons<br />

are visible. They also have been resolved by expansion of the NMR frequency scale. For a<br />

few strongly coupled protons we are able to detect all five NMR transitions (1) to (5). This<br />

is a further <strong>and</strong> definite proof that we do observe quintet states.<br />

137


9 Diacetylene Single Crystals<br />

3(2)<br />

2(2)<br />

1(2)<br />

(a)<br />

Bo II X<br />

= 9.570 MHz<br />

ν F<br />

ν F<br />

(c)<br />

Bo II Y<br />

ν F = 14.193 MHz<br />

2(1)<br />

1(4) 2(4) 3(4)<br />

4(4)<br />

0<br />

ν F<br />

3(2) 2(2)<br />

10 20 30<br />

40<br />

0 10 20 30<br />

(b)<br />

Bo II Y<br />

= 10.911 MHz<br />

ν F<br />

2(4)<br />

ν F<br />

ν F<br />

(d)<br />

Bo II X<br />

= 15.553 MHz<br />

ν F<br />

3(4)<br />

4(4)<br />

4(2)<br />

1(2)<br />

0 10 20 30 0 8 12 16 20<br />

NMR Frequency ν/MHz<br />

NMR Frequency ν/MHz<br />

NMR FREQUENCY ν / MHz<br />

Figure 9.15: ENDOR spectra as detected by the ESR transitions (a–d); n F is the free proton frequency.<br />

(1) to (5) label the NMR transitions illustrated in Fig. 9.14, <strong>and</strong> 1, 2, 3 … number the individual protons.<br />

The external field B o is oriented along the y or x-axis of the fine structure tensor [36].<br />

The ENDOR shift anisotropy is shown in Fig. 9.16 for the strongly coupled protons<br />

i = 1, 2, 3, <strong>and</strong> 4. This anisotropy is mainly due to the anisotropy of the effective spin S u eff.<br />

The lines were calculated (via Eqs. 12 <strong>and</strong> 13) by fitting A i . In total we have analyzed<br />

22 protons, the hyperfine tensors A i of which are presented in Tab. 9.4. Two features can be<br />

Figure 9.16: ENDOR shift anisotropies for the four strongest coupled protons i = 1, 2, 3 <strong>and</strong> 4, detected<br />

via the ESR transition (b). x, y <strong>and</strong> z are the principal axes of the fine structure tensor. The experimental<br />

values were taken for the rotation of B 0 in the yz-plane <strong>and</strong> in the zx-plane, respectively. The ENDOR<br />

shifts Dn are calculated for (b)i(2) transitions (drawn out) <strong>and</strong> the (b)i(3) transitions (dashed), respectively<br />

[36].<br />

138


9.2 Photopolymerization<br />

Table 9.4: Complete hyperfine tensors for 22 protons, calculated by fitting the experimental anisotropy<br />

in a least-squares method. The A ij are diagonalized principal values of the fitted tensor, a is the isotropic<br />

coupling constant, B jj the dipolar anisotropic tensor, F, Y, <strong>and</strong> c are Euler angles of the hyperfine tensor<br />

axes relative to the fine structure axes [36].<br />

i A xx A yy A zz a B xx B yy B zz F Y C Assign-<br />

MHz MHz MHz MHz MHz MHz MHz degr. degr. degr. ment<br />

1 16.298 18.953 16.747 17.333 –1.035 1.620 –0.586 13.6 –51.8 2.8 CH2<br />

2 11.870 14.290 10.683 12.281 –0.411 2.009 –1.598 –39.2 76.4 –12.8 CH2<br />

3 2.348 3.737 5.667 3.917 –1.569 –0.180 1.749 28.0 80.5 –6.9 CH2<br />

4 1.905 1.797 1.253 1.652 0.253 0.146 –0.399 74.9 101.2 4.6 CH2<br />

5 1.788 0.710 –0.391 0.702 1.086 0.007 –1.093 56.0 –16.4 –21.8 CH2<br />

6 0.407 0.804 0.703 0.638 –0.231 0.166 0.065 10.3 10.1 –36.9 CH2<br />

7 1.456 0.089 –0.119 0.475 –0.981 –0.387 –0.594 26.3 –25.4 –4.4 CH2<br />

8 0.700 0.620 0.076 0.465 0.234 0.155 –0.389 57.4 –15.3 8.0 CH2<br />

9 –0.791 0.981 0.520 0.237 –1.028 0.744 0.284 –23.4 –80.7 –15.4 CH2<br />

10 –0.992 1.424 0.148 0.194 –1.185 1.230 –0.045 –23.2 1.6 –67.2<br />

11 0.219 –0.079 0.400 0.180 0.039 –0.259 0.220 –67.5 111.6 41.8<br />

12 0.098 0.161 0.271 0.177 –0.079 –0.016 0.094 21.5 90.6 –25.7<br />

13 –0.819 1.321 –0.275 0.076 –0.895 1.245 –0.351 –6.7 –5.7 –6.0 CH2<br />

14 0.075 –0.120 0.239 0.065 0.011 –0.185 0.174 12.6 7.3 –38.9<br />

15 1.969 –1.045 –0.889 0.012 1.958 –1.057 –0.901 –68.2 –7.4 49.5 ARYL<br />

16 –0.066 –0.165 0.235 0.002 –0.067 –0.166 0.234 –24.1 –5.2 38.5<br />

17 –0.456 –0.051 –0.203 –0.203 –0.254 0.254 0.000 91.1 114.5 27.5<br />

18 2.350 –2.533 –0.473 –0.219 2.569 –2.314 –0.254 –20.4 9.5 –53.6 ARYL<br />

19 –0.311 –0.237 –0.442 –0.330 0.019 0.093 –0.112 –23.5 91.3 89.0<br />

20 –0.777 –0.289 –0.143 –0.403 –0.374 0.114 0.260 –36.0 –30.2 18.1 ARYL<br />

21 –2.070 1.303 –1.288 –0.685 –1.385 1.988 –0.603 6.0 –49.3 6.5 CH2<br />

22 –1.072 –0.735 –0.652 –0.820 –0.252 0.085 0.168 4.7 18.6 –27.0 CH2<br />

extracted from Tab. 9.4 immediately: first, the anisotropic part of the hyperfine tensor is<br />

small as compared to the isotropic part for almost all protons, <strong>and</strong> second, several protons<br />

(i = 17, 18, 19, 20, 21, 22) show a negative value of the isotropic coupling constant a.<br />

We are able to assign 6 protons unambiguously: i = 1, 2, 3, 13, 21, <strong>and</strong> 22 (Fig. 9.17).<br />

The analysis of the hyperfine data shows that the spin density at C 1 (Fig. 9.18) is only 11%,<br />

–2,2% at C 2 , 17% at C 3 , –6% at C 4 , 7,9% at C' 1 <strong>and</strong> –2% at C' 2 , i. e. the spin is highly delocalized<br />

within the oligomer (Fig. 9.18). This corresponds with the above-described exchange<br />

coupling of the two carbenes <strong>and</strong> is to our opinion a very impressive demonstration<br />

of the power of ENDOR.<br />

9.2.3.3 ESR <strong>and</strong> ENDOR of triplet dicarbenes 3 DC n<br />

For a long time it was unclear whether the triplet state (S = 1) of the dicarbenes ( 3 DC) does<br />

exist, <strong>and</strong> if it does whether its energy is higher or lower than the energy De SQ of the dicarbene<br />

quintet state ( 5 DC). Müller-Nawrath et al. [51] have shown theoretically <strong>and</strong> experimentally<br />

that the ESR transitions of the triplet states of carbenes ( 3 C) <strong>and</strong> of dicarbenes ( 3 DC),<br />

respectively, are mutually degenerated in diacetylene oligomers of diacetylene crystals if the<br />

139


9 Diacetylene Single Crystals<br />

Figure 9.17: Assignment of 6 hyperfine tensors to methylene protons at the polymerization head. The<br />

arrows indicate the directions of the strongest main value of the respective hyperfine tensors. These 6<br />

tensors have been fitted simultaneously yielding the spin density distribution shown in Tab. 9.4 [36].<br />

20<br />

17<br />

15<br />

10<br />

7,9<br />

11<br />

5<br />

0<br />

-5<br />

-2<br />

-2,2<br />

-6<br />

-10<br />

C2' C1' C4 C3 C2 C1<br />

Figure 9.18: Carbene spin density at the reactive polymer chain end (Fig. 9.17).<br />

oligomers are long, i. e. if the number of added monomers n is more than 7. Therefore 3 C<br />

<strong>and</strong> 3 DC cannot be discriminated by ESR. Their ENDOR spectrum, however, is completely<br />

different. The ENDOR shifts Dn are related by Dn DC = –1/2 Dn C . The existence of several<br />

ENDOR lines, which fulfills the characteristic factor –1/2 in the above-mentioned relation,<br />

unambiguously shows the existence of the triplet state of dicarbenes. Its excitation energy is<br />

lower than the excitation energy De SQ of the quintet states of the same dicarbene [51].<br />

140


9.2 Photopolymerization<br />

9.2.4 Flash photolysis <strong>and</strong> reaction dynamics of diradicals<br />

A single UV laser flash with wavelength l = 308 nm, pulsewidth 15 ns, <strong>and</strong> flash energy<br />

1 mJ initiates photopolymerization by the production of the diradical DR 2 [52]. At low temperature,<br />

i. e. 4.2 K, this photoproduct is stable. It is detected by its absorption spectrum. In<br />

the spectral range the 0-0 transition peaks at 422 nm between monomer <strong>and</strong> polymer absorption<br />

(Fig. 9.20a). Annealing the crystal containing photoproduct DR 2 , prepared as described<br />

above, produces the diradicals DR 3 ,DR 4 ,DR 5 , <strong>and</strong> DR 6 respectively by addition of one<br />

monomer per step (Fig. 9.8). All these diradicals can be detected by optical absorption spectroscopy,<br />

for example by cooling the crystals to low temperature in order to slow down the<br />

dark reaction (Fig. 9.20b). Sixl et al. [48] were the first who detected these optical absorption<br />

spectra [131]. The dark reaction at low annealing temperatures has been investigated extensively<br />

by Gross [38, 46].<br />

If the reaction is photoinitiated by a single UV laser flash at high temperatures<br />

(T > 180 K) the entire time-dependent reaction series, DR 2 ? DR 3 ? DR 4 ?DR 5 ?DR 6 ,<br />

can be observed by monitoring the transient optical absorption of each of the products DR 2<br />

to DR 6 . By this experiment we were able to analyze the reaction kinetics of these thermally<br />

activated steps separately [37, 44]. As an example Fig. 9.19 shows the transient absorption<br />

0.23<br />

T=270 K<br />

DR 422nm<br />

2<br />

0<br />

0.23<br />

DR 514nm<br />

3<br />

∆OD<br />

0.29 0<br />

DR 578nm<br />

4<br />

0.5 0<br />

DR 664nm<br />

5<br />

0<br />

0<br />

10<br />

t/ µ s<br />

20<br />

Figure 9.19: Time sequence of the intermediate products DR 2 ,DR 3 ,DR 4 , <strong>and</strong> DR 5 at T = 270 K after<br />

the UV flash which is indicated by an arrow. DOD is the change of the optical density after the UV<br />

flash [37, 44].<br />

141


9 Diacetylene Single Crystals<br />

Figure 9.20a: Difference of the optical absorption spectra (DOD) of TS6 after <strong>and</strong> before one single<br />

UV pulse (l = 308 nm, t = 15 ns, E = 0.1 mJ, T = 80 K). The peak at 422 nm is due to the absorption<br />

of the Dimer DR 2 .<br />

Figure 9.20b: Appearance of DR n reaction intermediates after a single pulse irradiation at 308 nm <strong>and</strong><br />

additional annealing. (a): 5 K, 0 min; (b): 100 K, 6 min; (c): 100 K, additional 30 min plus 120 K,<br />

36 min; (d): 130 K, 36 min plus 140 K, 16 min [131].<br />

at T = 270 K as detected by the change of the optical density (DOD) after the flash vs. time.<br />

Each intermediate product is detected at the maximum of its optical absorption: DR 2 at<br />

422 nm, DR 3 at 514 nm, DR 4 at 578 nm, <strong>and</strong> DR 5 at 664 nm. The delay in the production<br />

of a subsequent intermediate is clearly demonstrated. The whole reaction passes in a 10 ms<br />

time scale at room temperature.<br />

Assuming that DR 2 is produced by the flash promptly, DR 3 by a dark reaction from<br />

DR 2 ,DR 4 by a subsequent dark reaction from DR 3 , <strong>and</strong> so on, we used a simple kinetic<br />

142


9.2 Photopolymerization<br />

model for the quantitative analysis. This model is described by the following equations for<br />

the concentrations n i <strong>and</strong> the rate constants K i for the DR i , i =1,2,3,4,5:<br />

dn i<br />

dt ˆ K i 1 n i 1 K i n i ; …14 a†<br />

K i ˆ K 0 e …E i=kT† : …14 b†<br />

As an example, the fit of this model to the transient DR 4 at 200 K is shown in<br />

Fig. 9.21. The fit does not show any significant deviation from the experimental curve.<br />

Figure 9.21: Experimental transient <strong>and</strong> model curve according to Eq. 14a for DR 4 in perdeuterated TS<br />

at T = 200 K. The difference between experiment <strong>and</strong> model is plotted around the baseline [37, 44].<br />

One result of these experiments is shown in Fig. 9.22. The addition reactions are thermally<br />

activated, the activation energies being about 0.25 eV per monomer, almost identical<br />

for each step (Tab. 9.5). Figure 9.22 also includes values for a product labelled V which is<br />

presumably due to long polymer chains with reactive carbene chain ends.<br />

A second result of these experiments is an estimation of the polymer yield for the<br />

photoreaction as defined by the number of polymerized monomer molecules per absorbed<br />

UV photon. Q is the quantum yield for the initiation process. Thus, the polymer yield P is<br />

the product of Q <strong>and</strong> the kinetic chain length L, i. e. the number of monomer molecules<br />

which are added to the chain after one initiation process:<br />

P = Q7L = 0.07 ± 0.02 (15)<br />

The polymer yield was determined from the increase of the polymer absorption DOD<br />

due to UV irradiation [53].<br />

If we assume a kinetic chain length of 100, which is a reasonable value, we get a quantum<br />

yield for chain initiation of 7610 –4 . P increases with decreasing temperature from<br />

300 K to 180 K by nearly a factor of five. If TS is perdeuterated P also increases by a factor<br />

of 2.5 [44, 54].<br />

Both, the temperature effect <strong>and</strong> the isotope effect on the polymer, can be explained<br />

qualitatively by an increase of the kinetic chain length due to a longer carbene chain end<br />

lifetime [37, 44, 54].<br />

143


9 Diacetylene Single Crystals<br />

Figure 9.22: Temperature dependencies of the rate constants for the decays of DR 2 ,DR 5 , <strong>and</strong> V. They<br />

can be described by the Arrhenius law in a range of three orders of magnitude [37].<br />

Table 9.5: Activation energies DE i <strong>and</strong> frequency factors K o for the reaction rates of DR i , evaluated from<br />

their Arrhenius plots [37].<br />

K 2 K 3 K 4 K 5<br />

DE i / eV 0.25 ± 0.03 0.26 ± 0.03 0.30 ± 0.03 0.30 ± 0.03<br />

K o /s –1 10 10 ± 1 10 11 ± 1 10 11 ± 1 10 11 ± 1<br />

9.3 Holography<br />

Diacetylenes have been subject to intense work due to their unique ability to undergo topochemical<br />

solid state polymerization, resulting in macroscopic polymer single crystals [1–3,<br />

25, 28, 37]. Whether this reaction takes place depends on the monomer stacking distance<br />

<strong>and</strong> the tilt angle (Fig. 9.3). Both can be influenced by varying the rest groups R. The polymerization<br />

can be initiated by heat, UV radiation, X rays, or g rays <strong>and</strong> is irreversible. The<br />

optical properties, especially the absorption coefficient a <strong>and</strong> the refractive index n, are<br />

known to change dramatically during polymerization.<br />

By means of UV photopolymerization high-efficiency holographic grating on diacetylene<br />

crystals can be recorded, as was first shown by Richter et al. [55, 56]. Utilizing a fre-<br />

144


9.3 Holography<br />

quency-doubled argon laser (l w = 257 nm) Richter et al. obtained surface phase gratings,<br />

due to the low penetration depth of this UV wave (Fig. 9.6). At high exposures higher diffraction<br />

orders up to five has been observed. Because of the 5% stacking distance mismatch<br />

between monomers <strong>and</strong> polymers, Richter et al. often observed a destructive surface peel<br />

off. This problem has been shown to become less important by using longer UV wavelengths<br />

<strong>and</strong>, therefore, higher penetration depths [57].<br />

The first aim within the Collaborative Research Centre 213 was to find a suitable technique<br />

for recording such gratings in an effective <strong>and</strong> reproducible way. Using this technique<br />

we have investigated the most important diffraction characteristics of these gratings: efficiency,<br />

thickness, angular selectivity <strong>and</strong> their dependence on exposure, sample thickness,<br />

<strong>and</strong> prepolymerization. The second aim, however, was to explain these characteristics within<br />

appropriate theoretical approach. Using this knowledge, the chain length of the polymers during<br />

UV polymerization <strong>and</strong> subsequent thermal treatment can be estimated. In the final investigation<br />

we have shown that images <strong>and</strong> even a holographic trick film can be recorded in the<br />

diacetylene crystals at room temperature. The peculiarity of the method is the difference of<br />

recording (UV) <strong>and</strong> reading (VIS) wavelengths. This difference allows prompt readout without<br />

a developing process <strong>and</strong> without perturbation of the hologram by the readout laser.<br />

9.3.1 Theory<br />

Figure 9.23 shows the writing <strong>and</strong> reading beams for recording <strong>and</strong> replay of the simple holographic<br />

grating, respectively. Provided that they are of equal intensity two coherent UV<br />

Figure 9.23: Schematic geometry of writing <strong>and</strong> reading of a holographic grating. Two coherent UV<br />

waves (dashed) form the grating of a spatial periodicity L. A VIS wave (solid) will generally be diffracted<br />

into different orders j.<br />

145


9 Diacetylene Single Crystals<br />

waves, impinging symmetrically the photoactive medium, form an intensity pattern on the<br />

surface of the photoactive material,<br />

I…x† ˆI 0 cos 2 …px=L† :<br />

…16†<br />

L is the grating distance given by<br />

L ˆ w =2 sin w ;<br />

…17†<br />

where l w is the vacuum wavelength <strong>and</strong> y w the angle of incidence outside the medium. This<br />

intensity pattern results in a photoproduct distribution of the same periodicity, represented<br />

by a grating vector K,<br />

K ˆ 2p ^x=L ;<br />

K ˆ jKj ˆ 2p=L ; …18†<br />

<strong>and</strong> results in a UV photopolymerization pattern, the refractive index n, <strong>and</strong> the absorption<br />

coefficient a of which can be described as<br />

n…x† ˆn 0 ‡ P1<br />

n h cos …Khx† ;<br />

hˆ0<br />

a…x† ˆa 0 ‡ P1<br />

a h cos …Khx† :<br />

hˆ0<br />

…19†<br />

The importance of the Fourier coefficients besides h = 1 generally depends on reaction<br />

kinetics, exposure, <strong>and</strong> saturation effects. In general, n <strong>and</strong> a can also vary with the z coordinate.<br />

In principle a readout light beam (vacuum wavelength l r ) will be diffracted by the<br />

grating resulting in different orders j as indicated in Fig. 9.23. Each of them have an amplitude<br />

S j <strong>and</strong> a propagation vector r j . The total electric field inside the medium is a superposition<br />

of all theses waves:<br />

E…z† ˆ<br />

jˆ1 P<br />

jˆ 1<br />

S j …z† ^s j exp… ir j r† : …20†<br />

Here ^s j are the polarization vectors. The wave vectors r j are coupled to the grating<br />

vector K via<br />

r j ˆ r 0 ‡ jK :<br />

…21†<br />

This treatment was used first by Magnusson <strong>and</strong> Gaylord [58] <strong>and</strong> leads to a system<br />

of coupled differential equations for the complex amplitudes S j :<br />

146<br />

uS j<br />

uz ‡ cos …a 0 ‡ i# j †S j ‡<br />

1 X1 <br />

i cos S j h …2pn h = r ia h † ^s j h ^s j ‡ S j‡h …2pn h = r ia h † ^s j‡h ^s j ˆ 0 …22†<br />

2<br />

hˆ1


9.3 Holography<br />

where is given by<br />

# j ˆ jK cos… '† …jK† 2 r =4pn 0 …23†<br />

with a slant angle j between K <strong>and</strong> the x direction; j =908 for the symmetric case shown<br />

in Fig. 9.23. This so-called coupled wave approach is a generalization of the work of Kogelnik<br />

[59], who assumed pure sine gratings read under the Bragg condition where only one<br />

transmitted <strong>and</strong> one diffracted wave is present. Kogelnik [59] gives analytical solutions for<br />

the amplitudes of the reference <strong>and</strong> the signal wave (zeroth <strong>and</strong> first order respectively) –<br />

only for the first ascent period of a growth curve, extending earlier work for the transparency<br />

region. For the efficiency Z j of the transmission grating with the thickness d<br />

j …d† ˆS j …d†S j …d†<br />

…24†<br />

he gets for j = 1 the well-known formula<br />

<br />

…d† ˆ sin 2 …pn 1 d=cos 0 †‡sinh 2 <br />

…a 1 d=2 cos 0 † exp… 2ad=cos 0 † ; …25†<br />

with the replay wavelength l <strong>and</strong> the Bragg angle y 0 . It should be pointed out that this formula<br />

is only valid for the special case of thick (volume) gratings, which show pure Bragg behaviour.<br />

That is they possess neither higher Fourier coefficients nor a modulation amplitude n 1 nor large<br />

enough a 1 , to produce higher diffraction orders, assuming n 1 <strong>and</strong> a 1 do not vary with z.<br />

The thickness of a holographic grating is often described in terms of the Q factor, defined<br />

by<br />

Q ˆ 2pd=L 2 n 0 :<br />

…26†<br />

Gratings with Q^1 are regarded as thin those with Q 610 as thick.<br />

9.3.2 Experimental setup<br />

We used diacetylene single crystal platelets, approx. 20 mm to200mm thick <strong>and</strong> some mm 2<br />

in area, cleaved from a parent TS6 crystal parallel to the (100) surface.<br />

The experimental setup is shown in Fig. 9.24. For writing holographic gratings we utilized<br />

a cw helium-cadmium laser (l w = 325 nm) or a xenon chloride excimer laser (l w = 308 nm).<br />

We gave preference to the recording geometry suggested by Bor et al. [62] <strong>and</strong> not to the wellknown<br />

beamsplitter geometry. The first order diffracted beams of a reflection grating R are reflected<br />

by a pair of parallel mirrors M1 <strong>and</strong> M2 (or pass a biprism instead) <strong>and</strong> are superimposed<br />

on the sample S. This geometry yields four main advantages:<br />

a) Given the fringe distance of the reflection grating, the resulting fringe distance on the<br />

sample is L = D/2 <strong>and</strong> wavelength-independent;<br />

147


9 Diacetylene Single Crystals<br />

b) Both writing beams are superimposed correctly because they have passed the same<br />

number of reflections. This is especially important if non-Gaussian beams are used;<br />

c) The intensities of both beams are equal <strong>and</strong> wavelength-independent;<br />

d) The setup is realizable compactly <strong>and</strong> can easily be adjusted for the use of short coherence<br />

lengths.<br />

Control of exposure is possible using an UV enhanced photodiode (De3). To read the<br />

gratings we used a 3 mW helium-neon laser (l r = 633 nm). At this wavelength the increase<br />

in refractive index can be expected to be high, whereas the absorption should not become<br />

too strong during the induction period of the polymerisation reaction (Fig. 9.4). The sample<br />

holder was mounted on the axis of a stepping motor for varying the angle of incidence. For<br />

analyzing transmitted or diffracted light a preamplifed large-area photodiode (De1) was<br />

mounted on a radius level of a second stepping motor being coaxial to the first one. Both<br />

motors include a reduction gear, giving an angular resolution of 0.06 mrad. Signal improvement<br />

was achieved by chopping the readout beam <strong>and</strong> using a second photodiode (De2) as<br />

an intensity reference. Both photodiode outputs were led to lock-in amplifiers. Two polarizers,<br />

Pol1 <strong>and</strong> Pol2, were used to attain both UV <strong>and</strong> VIS laser polarizations parallel to the<br />

sample’s b-axis.<br />

Figure 9.24: Experimental setup. De1, De2, De3: large area photodiodes, R: reflection grating, S: sample<br />

on sample holder, Sh: beam shutter, BS: beam splitter, M: mirrors, Pol1, Pol2: polarizers (l/2<br />

plates), Ch: chopper.<br />

Two sample orientations relative to the grating fringes were investigated:<br />

a) UV polarization,VIS polarization, <strong>and</strong> polymer axis b lying in the plane of incidence (E<br />

mode), henceforth denoted as b k orientation;<br />

b) UV polarization, VIS polarization, <strong>and</strong> polymer axis b were perpendicular to the plane<br />

of incidence (H mode), called b || orientation.<br />

148


9.3 Holography<br />

The high dichroism of the crystals <strong>and</strong> their well-formed habit can be used to align<br />

their orientation under a polarizing microscope. Of course, the optical anisotropy of the crystals<br />

as well as anisotropic reaction kinetics give rise to birefringence effects.<br />

9.3.3 General characterization<br />

Figure 9.25 shows for typical samples the maximum efficiency Z(y 0 )=Z as a function of<br />

the exposure. These growth curves show that an optimum can be reached between 0.5 <strong>and</strong><br />

1.0 J/cm 2 for l w = 308 nm <strong>and</strong> between 5 <strong>and</strong> 10 J/cm 2 for l w = 325 nm. For TS6 the penetration<br />

depth at 308 nm is about 110 mm, whereas at 325 nm it is about 420 mm. This difference<br />

<strong>and</strong> in addition a presumably lower quantum yield at 325 nm results in much slower<br />

growth curves for this wavelength. The decrease in efficiency is not only an effect of coupling<br />

back intensity from first to zeroth order, but also produced by an increase of absorption.<br />

Sample quality <strong>and</strong> thickness can influence these curves. A destruction of the surface<br />

is only observed if the optimal exposure is exceeded by a factor 2 to 3. This is due to the<br />

lattice mismatch of monomer <strong>and</strong> polymer stacking distance.<br />

The energy needed to reach the maximum is approximately by a factor 5 larger than<br />

at 257 nm [55, 56]. From the growth curves, i. e. from the exposure E which was needed to<br />

produce a certain efficiency Z, the holographic sensitivity S of the material given by<br />

S ˆ<br />

p<br />

=E<br />

…27†<br />

can be estimated. For TS6 <strong>and</strong> depending on sample quality this value can range from 0.4 to<br />

1.4 cm 2 /J. This is in the order of typical sensitivities for photorefractive crystals but far less<br />

0 12.5 25<br />

0.5 0.13<br />

Efficiency η<br />

0.25<br />

0.065<br />

0 0<br />

0<br />

1.5<br />

3.0<br />

Exposure E / Jcm -2<br />

Figure 9.25: Holographic growth curves of TS6 (solid curve: l w = 308 nm, d = 130 mm; dotted curve:<br />

l w = 325 nm, d = 270 mm) <strong>and</strong> IPUDO (dashed curve: l w = 308 nm, d = 260 mm). Orientation b k ,<br />

L = 3.3 mm. Z(y 0 )=Z: efficiency for readout Bragg condition.<br />

149


9 Diacetylene Single Crystals<br />

than for common silver halide materials [63]. Nevertheless, the resolution of TS6 is comparable<br />

to that of common photographic materials. From the experiments of Richter et al. [55]<br />

it is known that a resolution of 2500 lines/mm is achievable. In our electron beam lithography<br />

experiments with a resolution of 5000 lines/mm has been achieved [64].<br />

Supposing the thickness of the sample is equal to the grating thickness, we can calculate<br />

the factor Q for our gratings. At a grating distance of L = 3.3 mm this factor<br />

ranges from 2.5 to 60 for samples of thickness 25–500 mm, whereas at L = 0.8 mm Q<br />

factors of 800 can be achieved. These values are typical for thick phase gratings. Nevertheless,<br />

it should be pointed out that for example a grating with Q = 40 already shows<br />

higher diffraction orders. This confirms the arguments brought by Moharam <strong>and</strong> Young<br />

[60] that the r factor (r = L 2 /L 2 n 1 n 0 ) should be preferred when discussing grating diffraction.<br />

9.3.4 Angular selectivity<br />

The angular selectivity is a measure for diffracted intensity when readout is performed under<br />

off-Bragg conditions or, quantitatively speaking, it is the half-width Dy of the function Z (y),<br />

where y is the angle of incidence. This property is of central interest, when discussing holographic<br />

storage media, because it limits the number of holograms/holographic gratings<br />

which can be stored simultaneously.<br />

Only as a rule of thumb, Kogelnik [59] gives Dy & Ln 0 /d, whereas Magnusson <strong>and</strong><br />

Gaylord [58] do not give any analytical expression for this quantity. Both approaches can easily<br />

be implemented on a computer to simulate diffraction properties. Assuming n h = 0 for<br />

h 62 <strong>and</strong> only for the first ascent period of a growth curve we find a Dy-dependence of the<br />

following simple form<br />

ˆ n 0 =d<br />

…28†<br />

valid only for the first ascent period of a growth curve, because the angular selectivity<br />

curves get split beyond the first maximum [61].<br />

The dependence of Dy on sample thickness d proved to be very well reproducible with<br />

respect to adjustment <strong>and</strong> sample quality. Figure 9.26 shows two typical measurements of<br />

Z (y) representing samples of different thickness. In contrast to the early experiments of<br />

Richter et al. [55] <strong>and</strong> of Niederwald et al. [56] these values represent an improvement of<br />

Dy by up to two orders of magnitude.<br />

Figure 9.27 shows Dy values versus sample thickness for TS6. No significant difference<br />

could be observed with respect to exposure or orientation for this grating distance.<br />

There is also no significant difference between TS6 <strong>and</strong> IPUDO with respect to angular selectivity.<br />

Measurements at L = 0.8 mm are summarized in Fig. 9.28. Within the series the minimum<br />

angular selectivity of 0.188 arose, using a 380 mm thick TS6 platelet. The hyperbolas<br />

plotted in Figs. 9.27 <strong>and</strong> 9.28 are fits of the function Dy = Ln 0 /d to the data points. Averaged<br />

over all exposures n 0 ranges from 1.3 to 1.7.<br />

150


9.3 Holography<br />

1<br />

Rel. Efficiency<br />

0.5<br />

0<br />

-30 0<br />

30<br />

Readout Angle<br />

/ Deg.<br />

Figure 9.26: Dependence of relative efficiency Z as a function of the incidence angle y for two TS6<br />

samples of different thickness d.<br />

Figure 9.27: Angular selectivity Dy as a function of sample thickness d for TS6 (open circles) <strong>and</strong><br />

IPUDO (full circles). L = 3.3 mm, l w = 308 nm [65].<br />

Angular Selectivity ∆Θ / Deg.<br />

2<br />

1<br />

0<br />

0 250<br />

500<br />

Sample Thickness d / µ m<br />

Figure 9.28: Angular selectivity Dy as a function of sample thickness d for TS6. L = 0.8 mm,<br />

l w = 325 nm for both writing geometries, b || (full circles) <strong>and</strong> b k (open circles) [65].<br />

151


9 Diacetylene Single Crystals<br />

9.3.5 Prepolymerized samples<br />

Until now we were assuming that sample thickness <strong>and</strong> hologram thickness are equal. In<br />

fact, for large thickness values d one can observe slight deviations from the hyperbola function<br />

due to the finite penetration depth of the UV light. These deviations should become<br />

more obvious for smaller penetration depths. To observe this some Dy measurements were<br />

carried out using thermally prepolymerized (up to 5 h at 70 8C before recording gratings)<br />

TS6 samples. Penetration depths for these samples varies between 18 to 80 mm at 308 nm<br />

<strong>and</strong> from 16 to 420 mm at 325 nm in the b || orientation.<br />

The effect of prepolymerization on angular selectivity is shown in Fig. 9.29. The values<br />

for samples prepolymerized for 2 h still show for thin samples a weak thickness dependence.<br />

For thick samples Dy does not tend to decrease further. Samples prepolymerized for<br />

5 h do not show any dependence on d, indicating a penetration depth distinctly smaller than<br />

the smallest sample thickness used. Thus, for an increasing polymer concentration the holographic<br />

gratings get thinner <strong>and</strong> thinner. The possibility of producing high efficiency volume<br />

phase gratings is restricted to fresh monomer crystals.<br />

Figure 9.29: Angular selectivity of prepolymerized TS6 samples: 5h (upper open circles), 2 h (full circles),<br />

<strong>and</strong> fresh crystals (lower open circles) [66].<br />

9.3.6 Chain length, polymer profile, <strong>and</strong> grating profiles<br />

A model [66, 68] describing the spatially inhomogeneous reaction kinetics of diacetylenes<br />

must take into account a kinetic chain length L, which depends on the polymer conversion P.<br />

The monomolecular reaction in a simple homogeneous situation then is given by<br />

dP=dt L…P†…1 P† : …27†<br />

For the experiments described so far, L can be assumed constant during the exposure<br />

time, because only very little conversion takes place. Furthermore, for all experiments de-<br />

152


9.3 Holography<br />

scribed, the factor (1 – P) can be dropped, because the experiments were carried out in the<br />

low conversion regime of the induction period <strong>and</strong> because saturation effects affect development<br />

experiments in both orientations. For the two orientations, different situations must be<br />

investigated corresponding to the different interaction of the two characteristic lengths L<br />

<strong>and</strong> L (Fig. 9.30),<br />

b jj : dP=dt L…P† cos 2 …px=L† : …28†<br />

Here a polymer molecule grows parallel to the fringes contributing its whole chain<br />

length L to the polymer growth at point x, where it has been initiated. In the other case, a chain<br />

initiated in x contributes to the polymer growth in the whole interval [x – L/2, x + L/2]:<br />

x‡L=2<br />

b ? : dP=dt R<br />

cos 2 …px 0 =L†dx 0 :<br />

x L=2<br />

…29†<br />

Integrating both Equations we see that in the second case the chain growth smears out<br />

the photoproduct distribution for a certain amount. This effect should be stronger if L/L approaches<br />

1. For the quotient of both polymer modulations we get the ratio d,<br />

ˆ P ?<br />

P jj<br />

ˆ sin…pL=L†<br />

pL=L : …30†<br />

The resulting d values range for fresh samples from 0.85 to 0.95 <strong>and</strong> tend to decrease<br />

with increasing prepolymer content. For samples prepolymerized (5–6 hours, 3–4% polymer)<br />

we find that d is between 0.65 <strong>and</strong> 0.80. These values correspond to chain lengths (L)<br />

of 0.15 to 0.4 mm or 300 to 800 repeat units [66].<br />

Figure 9.30: <strong>Microscopic</strong> model to underst<strong>and</strong> the interaction of the two characteristic lengths L <strong>and</strong> L:<br />

The two geometries investigated change the angle between chain growth direction <strong>and</strong> grating fringes.<br />

153


9 Diacetylene Single Crystals<br />

9.3.7 Multrecording<br />

In additional experiments, we succeeded in recording more than only one grating into one<br />

thick TS6 crystal. After each exposure interval, we rotated the sample for a certain amount<br />

to slant the gratings relatively to each other. Figure 9.31 shows the first order diffraction versus<br />

the angle of incidence for a 310 mm thick sample containing 42 gratings tilted stepwise<br />

by 1.08 8. We observed that gratings already present are almost not influenced by the subsequent<br />

recording processes except by the increase of total absorption, which causes a loss of<br />

efficiency less than 10 %. Thus, we did choose an exposure of only 0.07 J/cm 2 for each grating.<br />

As can be seen from Fig. 9.31, the signal-to-noise ratio is bad for the last written gratings.<br />

It will get worse by decreasing the angular distance of the individual gratings but will<br />

hardly improve by increasing it [67]. A serious diffraction of the UV light by the gratings already<br />

recorded in the sample does not take place. UV diffraction efficiencies are extremely<br />

small in diacetylene crystals.<br />

-3<br />

Efficiency η / 10<br />

8<br />

4<br />

1<br />

2<br />

4<br />

0<br />

-10 20<br />

50<br />

Readout Angle Θ/ Deg.<br />

Figure 9.31: Efficiency of the first order of a multihologramm consisting of 42 single gratings successively<br />

recorded in a 310 mm thick TS6 crystal. The gratings were recorded in the order of the numbers<br />

indicated [69].<br />

9.3.8 Holography<br />

For the reconstruction of a real hologram with different writing <strong>and</strong> reading wavelengths, it<br />

was necessary to use a divergent reference wave. With a test platelet as object a resolution<br />

of 300 lines/mm was achieved in excellent TS6 samples (Fig. 9.32). Holograms up to 32<br />

pictures were written (l w = 325 nm) in one crystal using the angular selectivity of the thick<br />

phase. When rotating the sample the pictures can be reconstructed (l r = 633 nm) successively<br />

(holographic trick film).<br />

The storage density of about 6.7610 9 cm –3 , estimated from the real resolution <strong>and</strong><br />

angular selectivity, is about one to two orders of magnitude lower than the storage density<br />

calculated from the grating distance [70].<br />

154


9.4 Di-, pyro-, <strong>and</strong> ferroelectricity<br />

Figure 9.32: Reconstructed image from one of the 32 pictures of a holographic trick film [70].<br />

9.4 Di-, pyro-, <strong>and</strong> ferroelectricity<br />

The dielectric properties of diacetylene monomer single crystals are not strikingly different<br />

from those of other organic materials. Typically, they have permittivity values (earlier called<br />

dielectric constant e r ) of about 4–6. Solid state polymerization, however, results in a pronounced<br />

anisotropy of the electric permittivity with maximum values parallel to the polymer<br />

chain direction, due to the large polarizability of the extended p-orbitals. During our TOPO-<br />

MAK activities the change of e r , accompanying solid state polymerization, was analyzed<br />

quantitatively (Section 9.4.1.1). The results of the e r analysis can be applied for the in situ<br />

monitoring of the polymer content (Section 9.4.1.2).<br />

The tailoring of pyro or ferroelectric properties of diacetylenes (Section 9.4.2) is less<br />

straightforward than might be surmised from the well-ordered arrangement of the R, R' substituents<br />

in Scheme 1, <strong>and</strong> from the fact that polar side groups can easily be introduced as<br />

substituents R <strong>and</strong> R' (Tab. 9.2). The difficulties originate from the packing of the individual<br />

polar diacetylene monomer molecules in the elementary cell of the solution-grown single<br />

crystal. This, generally, gives rise to a center of inversion symmetry for a pair of molecules<br />

thus compensating the individual molecular electric dipole moments in an antiferroelectric<br />

arrangement. During solid state polymerization spurious electric polarization has been observed<br />

[71] resulting from intramolecular distortion of originally centrosymmetric monomer<br />

units. Here we want to emphasize that our systematic TOPOMAK investigations have realized<br />

pyroelectric (Section 9.4.2) as well as ferroelectric properties (Section 9.4.3) for appropriately<br />

substituted diacetylenes.<br />

155


9 Diacetylene Single Crystals<br />

9.4.1 Dielectric properties of diacetylenes<br />

Typically, the electric permittivity of monomer single crystals of substituted diacetylenes<br />

ranges from 4 to 6 at room temperature. These e r values are therefore larger than values of<br />

simple non-polar organic polymers, like polytetrafluoroethylene, in agreement with the existence<br />

of polar side groups. As is exemplified in Fig. 9.33, monoclinic crystals of 2,4-hexadiynylene<br />

di-p-toluenesulfonate (TS, see Tab. 9.2) shows generally a weak but non-negligible<br />

anisotropy for e r [72, 75].<br />

ε r (h)<br />

8<br />

7<br />

6<br />

5<br />

TS<br />

T = 60°C<br />

(1)<br />

(2)<br />

4<br />

3<br />

0<br />

10 20 30<br />

polymerization time t(h)<br />

Figure 9.33: Electric permittivity e r (t) as a function of the polymerization time for TS crystals at<br />

60 8C. The electric field was applied parallel to the chain direction of the monoclinic crystals (1) <strong>and</strong> in<br />

the two orthogonal directions (2) <strong>and</strong> (3). The permittivity was measured at 1 kHz for three different<br />

thin parallel-plate single crystal capacitors [75].<br />

(3)<br />

40<br />

9.4.1.1 Correlation of polymer content <strong>and</strong> electric permittivity<br />

Generally, a sigmoid time-conversion curve is observed for thermal solid state polymerization<br />

of diacetylenes. This behaviour was shown for TS in Fig. 9.4 <strong>and</strong> is reported for the unsymmetrically<br />

substituted 6-( p-toluenesulfonyloxy)-2,4-hexadiynyl-p-fluorobenzenesulfonate<br />

(TS/FBS, Tab. 9.2). In Fig. 9.34a the slow conversion of the initial induction period of the<br />

solid state polymerization lasts until a polymer content of about 10 % is achieved. For higher<br />

up to complete conversion, this is followed by an autocatalytic reaction enhancement. The<br />

st<strong>and</strong>ard technique for the derivation of time-conversion curves is the gravimetrical analysis<br />

of a large number of crystals in a point-by-point procedure. After thermal polymerization<br />

for a well-defined period both, the soluble monomer <strong>and</strong> the insoluble polymer portions, are<br />

determined gravimetrically (Fig. 9.34 a). This technique consumes a considerable number of<br />

crystals <strong>and</strong> a substantial amount of time. Furthermore, only averaged time-conversion<br />

curves are obtained.<br />

Figure 9.34b shows that during solid state polymerization of TS/FBS the electric permittivity<br />

parallel to the chain direction, e r|| , increases by a factor of about 2. The permittivity<br />

e r|| is almost a linear function of the polymer content, as is exemplified in Fig. 9.34 c. We<br />

have found comparable behaviour for the substituted diacetylenes TS, FBS, FBS/TFMBS,<br />

<strong>and</strong> DNP [72–75]. Because the reorientation of the side groups during the polymerization is<br />

weak, it does not influence substantially the increase of e r parallel to the chain direction.<br />

156


9.4 Di-, pyro-, <strong>and</strong> ferroelectricity<br />

Figure 9.34: Thermal solid state polymerization of TS/FBS (PTS is another acronym for TS).<br />

(a): Time-conversion curve derived by gravimetrical analysis. (b): Time-permittivity curve derived at<br />

1 kHz for a thin parallel-plate single-crystal capacitor oriented with the polymer chain direction (b-axis)<br />

parallel to the electric field. (c): Correlation of conversion <strong>and</strong> electric permittivity (with time as implicit<br />

parameter) obtained by combination of (a) <strong>and</strong> (b) [73].<br />

This proves that only the extended p-electron system of the diacetylene backbone is responsible<br />

for the enhanced polarizability. This conclusion is supported by the experimental analysis<br />

for three orthogonal directions in TS single crystals shown in Fig. 9.33 [75]. The minor<br />

variations for the two orthogonal directions (2) <strong>and</strong> (3) can be explained by the changes of<br />

the lattice parameters. In contrast, the change of the respective lattice parameters with polymer<br />

content does not suffice to explain the change of De r|| .<br />

The linear relation between permittivity <strong>and</strong> polymer content is in principle surprising<br />

because there is a distribution of chain lengths of solid state polymerization, differing between<br />

induction period <strong>and</strong> autocatalytic range. The linearity, shown in Fig. 9.34 c, reveals<br />

that the polarizability of the p-electron system saturates already at chain lengths below the<br />

shortest ones occurring during thermal solid state polymerization.<br />

The experimental range for e r is 1.4 to 2.2 derived for different substituted diacetylenes<br />

by Gruner-Bauer [72–75], which agrees with the observations of other groups [76–<br />

78]. These values compare favourably with the estimate De r & 1.6 obtained for TS by simplified<br />

model calculations [75]. For these theoretical estimates the method of Genkin <strong>and</strong><br />

157


9 Diacetylene Single Crystals<br />

Mednis [79] has been modified by Gruner-Bauer, extending earlier work for the transparency<br />

region [80].<br />

9.4.1.2 Application to topospecifically modified diacetylenes<br />

The linear relation of electric permittivity parallel to the chain direction vs. polymer content<br />

thus established can be used for the control of the polymer content of prepolymerized samples<br />

as well as for the derivation of time-conversion curves of individual single crystals in<br />

situ. We have used this technique to study the solid state polymerization of topospecifically<br />

<strong>and</strong> fully deuterated TS.<br />

Striking differences of their reactivities were reported before by Ch. Kröhnke [54].<br />

The limited amount of samples available was sufficient for the permittivity analysis.<br />

Figure 9.35 shows the behaviour of different topospecifically deuterated derivatives of TS<br />

at T =608C [73]. It should be stressed that the induction periods of single crystals with<br />

nominally the same history did not differ by more than 10% at the same polymerization<br />

temperature. The toposelective modification of these substituted diacetylenes by the deuteration<br />

of the methylene groups close to the triple bonds of the diacetylene monomer<br />

(that are engaged in the crankshaft-type motion of the monomer molecule around its center<br />

of mass during solid state polymerization) evidently has a drastic influence on the solid<br />

state polymerization.<br />

8<br />

7<br />

PTS<br />

T = 60°C<br />

6<br />

ε (t)<br />

5<br />

8<br />

7<br />

6<br />

5<br />

8<br />

7<br />

CD - PTS<br />

T = 60°C<br />

CD - PTS<br />

T = 60°C<br />

2<br />

6<br />

5<br />

0<br />

10<br />

Figure 9.35: Electric permittivity as function of time for solid state polymerization at T =608C for TS,<br />

PD-TS (fully deuterated TS), <strong>and</strong> CD 2 -TS (where only the methylene groups close to the triple bond of<br />

the diacetylene monomer unit are deuterated, see Tab. 9.2 a) [73].<br />

158<br />

20<br />

30<br />

polymerization time t/h<br />

40


9.4 Di-, pyro-, <strong>and</strong> ferroelectricity<br />

9.4.1.3 Additional applications<br />

Structural phase transitions accompanying solid state polymerization influence the behaviour<br />

of e r (t). Thus they can not be observed but additional hints concerning the type of structural<br />

changes can be obtained. For example, a structural phase transition for DNP (Tab. 9.2) occurs<br />

at a polymer content above 95% resulting in the loss of order perpendicular to the polymer<br />

chains [81]. The accompanying increase of the side-group mobility results in a distinct<br />

increase of the electric permittivity [75]. Even more dramatic changes of e r were observed<br />

at the transition to a fibrillar structure in polar crystals of DNP/MNP (Tab. 9.2) [74, 75].<br />

Thinking of applications outside fundamental research, the correlation of permittivity<br />

<strong>and</strong> polymer content can also be used for different kinds of ageing control. Since solid state<br />

polymerization is an activated process, the temperature-weighted time at elevated temperature<br />

– like a low-temperature radiation dose – influences the capacity of a diacetylene crystal-plate<br />

capacitor according to a well-defined characteristic history-capacity. Evidently, this<br />

can be adapted to an appropriate electronic ageing control.<br />

9.4.2 Pyroelectric diacetylenes<br />

One fascinating goal of diacetylene materials research is the realization of polar polymer single<br />

crystals without complicated poling procedures. Substituted diacetylenes R 1 –C:–C:C–R 2<br />

incorporating polar side groups R 1 <strong>and</strong> R 2 with large but different electric dipole moments<br />

have to be synthesized. Since solid state polymerization depends on the ability of individual<br />

molecule side groups, which can perform intramolecular torsion <strong>and</strong> giving rise to an overall<br />

crankshaft-like motion, it is difficult – <strong>and</strong> was impossible for us – to predict the packing arrangement<br />

of individual monomer molecules in the solid. Frequently, the unit cell of substituted<br />

diacetylenes was observed to accommodate pairs of formula units in a centrosymmetric<br />

arrangement, thus compensating the net electric polarization. Polymorphism turned out to be<br />

an obstacle to systematic tailoring of dielectric properties of diacetylene single crystals, because<br />

different crystal structures of the same diacetylene derivative could be obtained from<br />

different, <strong>and</strong> occasionally even from the same solvent [82]. Nevertheless, Strohriegl synthesized<br />

during our TOPOMAK activities several non-centrosymmetric diacetylenes, which crystallized<br />

also in a polar phase. Typically, their permittivities were relatively large <strong>and</strong> anisotropic<br />

[83]. We restrict this report to three examples.<br />

9.4.2.1 IPUDO<br />

IPUDO (for the molecular structure see Tab. 9.2) is an example of a symmetrically substituted<br />

diacetylene. For the monomer as well as polymer crystals, non-centrosymmetrical<br />

orthorhombic crystal structures can be found already at room temperature [84]. The c-axis<br />

is the polar axis of the monomer crystal. IPUDO can only be polymerized by g radiation<br />

( 60 Co). For 85% polymerized crystals pyroelectric properties were observed only in b direction.<br />

The permittivity of IPUDO is highly anisotropic, with maximum values of 8.5 for the<br />

159


9 Diacetylene Single Crystals<br />

monomer (parallel to a), or 11.6 for the 85% polymer (parallel to b) <strong>and</strong> with minimum values<br />

(parallel to c) smaller by a factor of 2 (3.5) for the monomer (polymer) [83]. The distortion<br />

of the long side groups by the development of hydrogen bonds between neighbouring<br />

–CO–NH- groups is supposed to be responsible for the non-centrosymmetry [83, 85, 86].<br />

Thus it is not surprising that the variation of the electric polarization of about<br />

3610 –8 Ccm –2 between room temperature <strong>and</strong> 4 K amounts to an unbalancing of only<br />

about 1% of the compensation of oppositely oriented C=O … HN dipole moments.<br />

9.4.2.2 NP/4-MPU<br />

NP/4-MPU (for the molecular structure see Tab. 9.2) is an example of a non-centrosymmetric<br />

diacetylene that forms polar monomer single crystals only, if it is grown from appropriate<br />

solvents, here from 2-propanol [82]. The resulting modification I is orthorhombic<br />

with the polar space group Fdd2 (Z = 16) <strong>and</strong> c as the polar axis [82]. This modification is<br />

not reactive thermally or under X-ray irradiation, because the monomer packing is outside<br />

the favourable range for solid state polymerization. The diacetylene rods make an angle of<br />

678 with the stacking axis c , <strong>and</strong> the stacking distance d = 4.61 Å. The permittivity of this<br />

diacetylene is highly anisotropic <strong>and</strong> shows the largest value of about e r &23 for an electric<br />

field applied in the direction of the polar axis [82, 87].<br />

Figure 9.36 shows the variation of the spontaneous electric polarization DP of NP/4-<br />

MPU with temperatures between 10 K <strong>and</strong> the melting point. DP amounts to about 15% of<br />

the electric polarization, which can be estimated from the volume density of molecular electric<br />

dipole moments of about 3 Debye (10 –29 Cm).<br />

The pyroelectric coefficient p(T) =dP S /dT =8.8610 –10 Ccm –2 K –1 is of the same<br />

order of magnitude (smaller by 1/3) as that of the well-known <strong>and</strong> commercially used pyro<br />

<strong>and</strong> ferroelectric polyvinylidenefluoride. Therefore NP/4-MPU single crystals can be used<br />

for the detection of radiation [89], which we showed by using chopped low-power laser light<br />

as radiation source. The pyroelectric current <strong>and</strong> the total variation of the surface charge<br />

Figure 9.36: Temperature-dependent change of the spontaneous polarization for a NP/4-MPU single<br />

crystal (modification I) along the polar c-axis for temperatures below the melting point T m . The pyroelectric<br />

coefficient P at 300 K is also given [87].<br />

160


9.4 Di-, pyro-, <strong>and</strong> ferroelectricity<br />

have been used for the detection. Furthermore, a transversal piezoelectric coefficient of<br />

1018 fCN –1 for NP/4-MPU was derived at room temperature. It is comparable with the value<br />

of a-quartz [88].<br />

9.4.2.3 DNP/MNP<br />

DNP/MNP (for molecular structure see Tab. 9.2) was the most successful one of Strohriegl’s<br />

syntheses [74]. It has a polar crystal structure for monomer <strong>and</strong> polymer crystals (space<br />

group P2 1 ). DNP/MNP polymerizes – thermally or exposed to UV radiation – extremely<br />

fast, because during solid state polymerization the molecules are packed optimally.<br />

The solid obtained by thermal polymerization of monomer crystals exhibits a fibrous<br />

texture, probably due to the large changes in lateral packing of the side groups. The temperature<br />

influence on the spontaneous electric polarization perpendicular to the chain axis c<br />

is shown in Fig. 9.37 for the monomer as well as the polymer crystal. The polarization varies<br />

up to room temperature by 7610 –8 Ccm –2 for the monomer crystal, with a pyroelectric<br />

coefficient of about 3.2610 –10 Ccm –2 K –1 at room temperature.<br />

0.0<br />

-2<br />

∆P(T) / 10 C cm<br />

-7<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

-1.0<br />

0<br />

monomer<br />

DNP / MNP<br />

parallel to polar axis<br />

100 200<br />

T/K<br />

polymer<br />

Figure 9.37: Variation of the spontaneous electric polarization parallel to the polar b axis of DNP/MNP<br />

crystals with temperature. The data of DP (T) =P (T)–P (5 K) were derived by charge integration during<br />

a temperature cycle [74].<br />

300<br />

9.4.2.4 Spurious piezo <strong>and</strong> pyroelectricity of diacetylenes<br />

Sample defects can be another origin of piezo <strong>and</strong> pyroelectric phenomena in substituted<br />

diacetylenes, which generally can be identified via sample dependence <strong>and</strong> smaller size of<br />

these effects [88]. Bloor et al. discussed such spurious pyroelectric effects, which seemed to<br />

be correlated with the occurrence of macroscopic deformations, such as screw dislocations<br />

in TS single crystals [90]. Similarly the analysis of weak polarization, caused by molecular<br />

distortion during solid state polymerization of TS, was reported by Bertault et al. [91]. We<br />

have observed comparable weak pyroelectric phenomena for TS/FBS [87].<br />

161


9 Diacetylene Single Crystals<br />

9.4.3 The ferroelectric diacetylene DNP<br />

Whereas several pyroelectric diacetylenes were identified, uniform ferroelectric phases seem<br />

to be rather the exception, according to our experience with many new substitutions of diacetylenes<br />

[92]. The ferroelectric low-temperature phase of the symmetrically disubstituted<br />

diacetylene DNP, i. e. 1,6-bis(2,4-dinitrophenoxy)-2,4-hexadiyne (Tab. 9.2), turned out to be<br />

one of the rare <strong>and</strong> interesting exceptions [93–98].<br />

On both ends DNP carries polar dinitrophenoxy groups, with an electric dipole moment<br />

of 10 –29 Cm, whose mutual twisting gives rise to the spontaneous electric polarization<br />

of the non-centrosymmetric low-temperature phase (space group P2 1 ) (Fig. 9.38). The DNP<br />

monomer crystal has besides a pyroelectric [93] low-temperature phase a ferroelectric one<br />

for T < T c = 46 K. According to our investigations, the direction of the spontaneous polarization<br />

(Fig. 9.39) can be influenced by an external electric field [97]. Structural defects, occurring<br />

in all DNP crystals, give rise to a distribution of transition temperatures <strong>and</strong> to the<br />

existence of domains, whose polarization can not be inverted with the accessible external<br />

fields – thus behaving like pyroelectrics. Only the domains with the highest transition temperatures<br />

could be poled [97].<br />

b<br />

a<br />

b<br />

a<br />

b<br />

a<br />

a)<br />

b)<br />

c)<br />

Figure 9.38: Packing arrangement of DNP molecules in the monomer crystal viewed along the c-axis<br />

at three temperatures (a): T = 296 K, (b): T = 145 K, (c): T = 5 K [96].<br />

The temperature dependence of the spontaneous polarization (Fig. 9.39) of the monomer<br />

crystal can be described in the framework of L<strong>and</strong>au’s phenomenological theory assuming<br />

a tricritical phase transition [97]. The maximum experimental value of the electric polarization<br />

of 2.4610 –7 Ccm –2 compares favourably with the polarization that was calculated<br />

from the intramolecular twisting angle of the polar dinitrophenoxy groups of 5.18 determined<br />

by X-ray structural analysis at 5 K [96].<br />

The electric permittivity (Fig. 9.40) reaches values of about e r & 150 at the transition<br />

temperature. Its temperature dependence is strongly influenced by defects [97]. Thus, it is<br />

less appropriate for a comparison with theoretical predictions.<br />

In the early days of TOPOMAK, H. Schultes has already observed that the phase transition<br />

of DNP shifts to lower temperatures (Fig. 9.41–43) with increasing polymer content<br />

162


9.4 Di-, pyro-, <strong>and</strong> ferroelectricity<br />

3<br />

DNP<br />

-2<br />

P(T) / 10 -7 Ccm<br />

2<br />

1<br />

0<br />

0<br />

10 20 30 40 50 60<br />

T/K<br />

Figure 9.39: Temperature dependence of the spontaneous polarization parallel to the polar b-axis for<br />

different DNP monomer single crystals [97].<br />

75 0.3<br />

DNP<br />

ε r (T)<br />

50<br />

25<br />

0.2<br />

0.1<br />

-1<br />

(χ ferro<br />

(T))<br />

0<br />

0<br />

20<br />

40<br />

Figure 9.40: Temperature dependence of the low-frequency electric permittivity (left axis) <strong>and</strong> the inverse<br />

of the ferroelectric part of the susceptibility (right axis) for a DNP monomer crystal [97].<br />

T/K<br />

60<br />

80 0<br />

-7<br />

polarization / 10 C cm<br />

-2<br />

1,0<br />

0,5<br />

10 h<br />

5h<br />

2h<br />

0h<br />

0<br />

25<br />

Figure 9.41: Variation of the zero-field electric polarization in crystallographic b-direction of DNP single<br />

crystal with duration of thermal polymerization at 130 8C [94] (Fig. 9.34 for conversion curve).<br />

50<br />

T/K<br />

163


9 Diacetylene Single Crystals<br />

ε (T) / ε (293 K)<br />

r<br />

r<br />

8<br />

4<br />

10 h<br />

5h<br />

13 h<br />

14 h<br />

16 h<br />

0<br />

0 25 50 75 100<br />

T/K<br />

Figure 9.42: Influence of duration of thermal solid state polymerization at 130 8C on temperature dependence<br />

of electric permittivity of DNP [94].<br />

50<br />

T c / K<br />

40<br />

30<br />

1,6<br />

1,2<br />

-1<br />

∆S /Jmol K<br />

-1<br />

20<br />

10<br />

0<br />

0<br />

X/%<br />

80<br />

40<br />

0<br />

0<br />

8 16<br />

t/h<br />

4 8<br />

16 0,0<br />

Figure 9.43: Variation of the transition temperature T c (e r maximum) <strong>and</strong> the conversion entropy (transition<br />

entropy change) DS with the duration t p of the thermal solid state polymerization of DNP at<br />

129 8C [83, 87]. The inset shows the typical variation of the polymer content X with t p .<br />

12<br />

t p/h<br />

0,8<br />

0,4<br />

(thermal polymerization) <strong>and</strong> can not be observed in the polymer crystal [94, 98]. This was<br />

a puzzle for our early attempts to underst<strong>and</strong> the ferroelectric phase transition of DNP.<br />

In the microscopic picture of the phase transition, nuclear magnetic resonance spectroscopy<br />

<strong>and</strong> relaxation of the DNP protons gave important additional information [95, 96].<br />

The proton NMR spectrum reflects the orientation of the proton-proton axis of the methylene<br />

groups close to the central diacetylene unit via the nuclear-spin magnetic dipole interaction<br />

in a rather clear-cut way (Fig. 9.44). For fixed crystal orientation <strong>and</strong> varied temperature<br />

the spectrum of the methylene group protons of the DNP monomer proved that both DNP<br />

moieties are twisted around the central C–C single bond of the diacetylene below T c . Since<br />

this degree of freedom is lost during the solid state polymerization the ferroelectric phase<br />

transition is suppressed.<br />

164


9.4 Di-, pyro-, <strong>and</strong> ferroelectricity<br />

r<br />

H0IIb<br />

r<br />

H 0<br />

b<br />

-40<br />

0 40<br />

-40 0 40<br />

Rel. Frequency (kHz)<br />

Figure 9.44: Simulated (left) <strong>and</strong> experimental (right) H NMR spectra as a function of the crystal orientation<br />

with respect to the external magnetic field, recorded at room temperature (n p = 200 MHz). The<br />

crystal was rotated in steps of 58 around its long axis (a-axis), which was oriented perpendicular to the<br />

external field [96].<br />

Additional information on the phase transition was obtained from proton-spin-lattice<br />

relaxation measured as function of the Larmor frequency, temperature, <strong>and</strong> orientation of<br />

the single crystals (Fig. 9.45). The analysis indicated the slowing down of a molecular motion<br />

on approaching the ferroelectric phase transition with the activation energy of about<br />

0.020 eV, which is in the range of known librational <strong>and</strong> torsional modes of diacetylenes<br />

[96].<br />

The phase transition of DNP could further be characterized by specific heat measurements<br />

for monomer <strong>and</strong> thermally polymerized single crystals (Fig. 9.46) [98]. These data<br />

support the description of the phase transition of the monomer crystals as a tricritical transition.<br />

This means it is a borderline case between a first-order <strong>and</strong> second-order phase transition,<br />

with a distribution of transition temperatures. The transition enthalpy was much lower<br />

than the corresponding order-disorder transition, in agreement with results obtained by Bertault<br />

et al. via Raman spectroscopy, which proved the importance of displacive contributions<br />

to the DNP phase transition [99].<br />

165


9 Diacetylene Single Crystals<br />

31 MHz<br />

-3 -1<br />

Spin-lattice relaxation rate (10 s )<br />

90 MHz<br />

200 MHz<br />

Temperature (Kelvin)<br />

Figure 9.45: Spin-lattice relaxation rate of the protons in a DNP monomer single crystal for three Larmor<br />

frequencies. Two rate maxima can be discerned with temperature dependence explained by the<br />

model of Bloembergen, Purcell, <strong>and</strong> Pound (solid line) [96].<br />

30<br />

-1<br />

∆/C/J mol K -1<br />

20<br />

2h<br />

2 p =0h<br />

10<br />

6h<br />

0<br />

30<br />

Figure 9.46: The ferroelectric contribution to the molar heat capacity of solid state polymerized DNP<br />

single crystals for different annealing times t p at 129 8C for [98].<br />

40<br />

T/K<br />

50<br />

9.4.4 Summary<br />

The increase of the electric permittivity for electric fields parallel to the polymer chain direction<br />

during solid state polymerization of diacetylenes can be used for in situ monitoring<br />

the monomer to polymer conversion of individual single crystals. The large librational am-<br />

166


9.5 Non-linear optical properties<br />

plitudes of the diacetylene moiety, required for solid state polymerization, are also the basis<br />

for the occurrence of interesting dielectric phase transitions. The tailoring of diacetylenes as<br />

ferro or pyroelectric crystals, which show good thermal stability <strong>and</strong> do not dem<strong>and</strong> considerable<br />

efforts for the poling process, is a trial-<strong>and</strong>-error process, but has been realized during<br />

our TOPOMAK activities in a number of cases. Thus, material properties useful for applications<br />

of pyro- or piezoelectricity thus have been obtained.<br />

9.5 Non-linear optical properties<br />

9.5.1 Aims of investigation<br />

Polydiacetylenes are polymers showing a one-dimensional semiconducting behaviour. This<br />

one-dimensional structure causes exceptionally high third order non-linearities (w (3) ) [100],<br />

also in off-resonant wavelength regions [101], with extremely short sub-picosecond switching<br />

times [102]. After this discovery it was believed that an optical amplifying switch (optical<br />

transistor) or even an optical computer was close at h<strong>and</strong>.<br />

At the start of our SFB 213 project the initial optimism about the application of the<br />

huge non-linearity of polydiacetylenes in optical switching was already somewhat damped.<br />

It was becoming clear, that the polydiacetylenes’ non-linear optical coefficients, despite belonging<br />

to the largest non-resonant non-linearities, were not sufficient for cascadable, intensity<br />

amplifying switches [103]. There was not much known about the mechanisms leading to<br />

the large non-linearities <strong>and</strong> different, sometimes contradicting theoretical models were proposed<br />

to describe them. On the other side, there was a dem<strong>and</strong> for tailor-made materials<br />

with predictable non-linearities, absorption b<strong>and</strong>s, <strong>and</strong> good optical quality.<br />

In this scenario, polydiacetylenes were nevertheless interesting, as the mechanisms of<br />

their large non-linearity form a good basis to build on. Only a few modifications were well<br />

characterised <strong>and</strong> there was a general lack of measurements with single crystals due to problems<br />

with sample preparation. Therefore, the aim was to characterise systematically the influence<br />

of side groups on the optical properties, preferably in the macroscopically ordered,<br />

stable, <strong>and</strong> reproducible framework of good <strong>and</strong> – later on – thin crystals.<br />

9.5.2 Experimental setup<br />

Mainly two kinds of experiments will be described here, two-beam pump-probe <strong>and</strong> degenerate<br />

four wave mixing (DFWM) measurements.<br />

In a pump-probe experiment, an intense pulsed pump beam <strong>and</strong> a weaker equally<br />

pulsed probe beam of the same wavelength are focused into the sample, <strong>and</strong> the transmission<br />

167


9 Diacetylene Single Crystals<br />

of the probe beam is measured. The timing of the two pulses can be shifted, to make it possible<br />

to probe the decay of excitations generated by the pump beam. Two contrary effects<br />

may occur, bleaching <strong>and</strong> induced absorption.<br />

Near-resonant pump-probe lifetime measurements were performed by W. Schmid<br />

[105, 123], using the same picosecond dye laser system as for DFWM, described below.<br />

Th. Fehn further on investigated the subject in the off-resonant wavelength region between<br />

720 nm <strong>and</strong> 820 nm, using a commercial Titan-Sapphire laser system (Coherent) with pulse<br />

lengths of ca. 120 fs, pumped by an argon ion laser (Coherent Mira).<br />

DFWM measurements can be done in a variety of geometries. Here, the forward mixing<br />

geometry is used (Fig. 9.47), where three beams, forming a right angle, are focused into the<br />

sample. This setup allows time-resolved measurements <strong>and</strong>, in contrast to third harmonic generation<br />

(THG) measurements, yields the w (3) (o;–o,o,–o) tensor which is related to an intensity-dependent<br />

refractive index, the interesting quantity for optical switching applications.<br />

Figure 9.47: Beam geometry for DFWM measurements.<br />

The interference pattern of the pump beams 1 <strong>and</strong> 2 forms horizontal stripes in the<br />

medium. As w (3) is directly related to an intensity-dependent refractive index, this interference<br />

pattern generates a refractive index pattern. The third probe beam is partially reflected<br />

on these horizontal planes, generating the signal beam 4. The efficiency of diffraction is related<br />

to |w (3) |. The pulse timing is adjustable, so the decay of the refractive index pattern can<br />

be probed. When the delay between the beams 2 <strong>and</strong> 3 is in the range of the laser’s coherence<br />

length, these pulses generate a diffraction grating for beam 1, resulting in an artificial<br />

raise of the signal, called coherence peak.<br />

All DFWM measurements were done with a commercially available synchronously<br />

pumped dye laser with a cavity dumper (Spectra Physics model 3500). When using Pyridine-1<br />

as radiant dye, the wavelength can be tuned between 670 nm <strong>and</strong> 750 nm. The pulse<br />

width is about 1 ps; the repetition rate can be adjusted between single shot <strong>and</strong> 8 MHz. This<br />

is important to reduce the average heat load absorbed by the crystals.<br />

To suppress stray light signals a well-known modulation technique was used, the modulation<br />

of two pump beams with different frequencies using a chopper blade with a set of<br />

different divisions. As the DFWM signal depends on the product of the input intensities, the<br />

signal can be detected at the sum or difference of the two modulation frequencies [104],<br />

168


9.5 Non-linear optical properties<br />

I 3 sig / I 1I 2 I 3 ˆ I 1 …0†…cos ! 1 t ‡ 1†I 2 …0†…cos ! 1 t ‡ 1†I 3 ˆ<br />

1<br />

ˆ I 1 …0†I 2 …0†I 3 ‰<br />

2 cos…! 1 ‡ ! 2 †t ‡ cos…! 1 ! 2 †t ‡ 2 cos ! 1 t ‡ 2 cos ! 2 tŠ: …31†<br />

W. Schmid proposed the separated detection of w (5) effects by an extension of this<br />

modulation method. A w (5) signal without w (3) contributions can be detected at twice the sum<br />

or difference frequency [105], as the w (5) signal contains terms depending on the square of<br />

the product of the modulated beam intensities,<br />

I sig / I 1 I 2 I 3 s k 3 ‡ k 35 … I 1 ‡ I 2 ‡ I 3 †‡k 5 … I 1 ‡ I 2 ‡ I 3 † 2 ; …32†<br />

where k 3 is a function of w (3) , k 5 of w (5) , <strong>and</strong> k 35 depends on both non-linearities <strong>and</strong> their<br />

relative phase.<br />

Although this modulation method permits qualitative measurements of w (5) effects,<br />

especially determination of relaxation times, quantitative evaluation of w (5) is not possible,<br />

due to the infinite number of modulation harmonics with ill-defined relative amplitudes<br />

caused by the trapezoidal modulation form provided by a chopper blade [106].<br />

This problem has been solved by A. Feldner. He developed a new type of modulator,<br />

working with two independently rotating polarizing foils (Fig 9.48). This type of modulator<br />

generates two harmonics in the intensity modulation spectrum with a fixed ratio of 4 : 1<br />

(cos 4 ot) [106].<br />

Figure 9.48: Scheme of intensity modulation by rotating polarizing foils (a). The vertical polarizer (b)<br />

ensures that an anisotropy of the material has no effect on the signal or the modulation spectrum.<br />

For time-resolved pump-probe measurements of shorter relaxation times, a Titan-sapphire<br />

laser pumped by an argon ion laser is employed. In the current configuration, the wavelength<br />

can be tuned between 720 nm <strong>and</strong> 820 nm; the pulse width is about 120 fs.<br />

169


9 Diacetylene Single Crystals<br />

9.5.3 Theoretical approaches<br />

The simplest approach is the free electron model [107, 108]. The electrons are treated to<br />

move freely in a one-dimensional box, subject to a potential V 0 cos (px/d) by the ion cores<br />

(d denotes the bond length). The polarizability a <strong>and</strong> hyperpolarizability g is derived from the<br />

2nd <strong>and</strong> 4th order perturbation energies caused by an electrical field along the chain. This<br />

rough model does not account for local field effects nor for the alternating bond length found<br />

along the axis in PDAs. The model yields a static w (3) = 3.0610 –11 esu along the axis.<br />

Agrawal, Cojan et al. [109–111] treated explicitly the linear <strong>and</strong> non-linear optical<br />

properties of PDAs. They computed the electronic energy levels using a Hückel formalism.<br />

They computed a static w (3) = 0.7610 –10 esu for PTS <strong>and</strong> w (3) = 0.25610 –10 esu for<br />

TCDU. This model neglects electron-electron interaction <strong>and</strong> therefore excitonic effects.<br />

The model yields the correct order of magnitude for w (3) , but the wrong sign.<br />

The phase space filling (PSF) model was developed to describe non-resonant NLO<br />

properties in semiconductors [112, 113], especially in quantum well structures [114]. Greene<br />

et al. adapted this model to one-dimensional polymer chains [115, 116]. The model is only<br />

applicable to systems were the low energy absorption b<strong>and</strong> is excitonic, as is the case with<br />

PDAs [117]. Formation of excitons is limited by the number of available electron states that<br />

are necessary to form the exciton. With an increasing number of excitons, the dipole momentum<br />

for forming a new exciton is reduced. The exciton b<strong>and</strong> bleaches.<br />

As the following measurements will show, the PSF model seems to describe best the<br />

non-linear optical behaviour of polydiacetylenes. A very important prediction of this model<br />

is the proportionality of w (3) <strong>and</strong> a in the near-resonant frequency regime [115]. This behaviour<br />

was found in polydiacetylenes, strongly supporting the PSF theory [118].<br />

9.5.4 Sample preparation<br />

For measurements well off the resonance, i. e. with wavelengths larger than 720 nm, p-TS6<br />

crystals were prepared by thermal polymerisation of monomer crystals, grown out of a saturated<br />

solution <strong>and</strong> manually cut using a shaver blade. Thickness of these crystals varies between<br />

40–100 µm.<br />

Later, for measurements closer to resonance, a method has been developed to grow<br />

thin mono-crystalline layers of TS6 <strong>and</strong> 4BCMU between glass substrates. To avoid crystal<br />

strains the monomer crystals have to be removed from the substrate before polymerising.<br />

The resulting crystal thickness can be made as low as 300 nm.<br />

9.5.5 Value <strong>and</strong> phase of the third order susceptibility w (3)<br />

From DFWM measurements |w (3) | was determined for several polydiacetylenes in the nearresonant<br />

to off-resonant wavelength region, 680 nm to 750 nm (Tab. 9.6). Concurrently, the<br />

170


9.5 Non-linear optical properties<br />

Table 9.6: w (3) values for some different Polydiacetylenes [118, 119].<br />

Material Modification |w (3) |/esu at wavelength<br />

PTS crystal 2610 –10 720 nm<br />

FBS crystal 2610 –10 720 nm<br />

4BCMU amorphous film 4610 –11 720 nm<br />

4BCMU thin monocrystalline film 3610 –10 720 nm<br />

4BCMU thin monocrystalline film 2610 –9 670 nm<br />

imaginary part of w (3) can be computed from the non-linear absorption coefficient b, obtained<br />

from measurements of the intensity dependence of the sample transmission. It was<br />

found that the real part of w (3) was dominating the imaginary part by a factor of 3 [118].<br />

This is predicted by the PSF model, from the bleaching of the exciton b<strong>and</strong>.<br />

The value of w (3) is nearly identical for p-TS6 <strong>and</strong> p-4BCMU, <strong>and</strong> 4 times lower for<br />

p-IPUDO. These values were reproducible within 30%. For p-FBS, the reproducibility was<br />

only within an order of magnitude. No difference was found between thermally <strong>and</strong> x-ray<br />

polymerised crystals [118].<br />

9.5.6 Relaxation of the singlet exciton<br />

In p-TS6 energy relaxation times could be resolved at wavelengths below 700 nm. As the relaxation<br />

times are of the same order of magnitude as the laser pulse width (1 ps), a model<br />

function has to be fitted to the measured curve to obtain the relaxation (Fig. 9.49).<br />

Figure 9.49: w (3) relaxation in p-TS6 at a wavelength of 680 nm <strong>and</strong> temperature of 200 K. For comparison,<br />

the dashed line gives the 3 rd order autocorrelation of the laser pulse. The relaxation time of the<br />

singlet exciton can be seen as a significant broadening effect [120].<br />

171


9 Diacetylene Single Crystals<br />

The DFWM signal decays with half the time constant of the exciton density. For measurements<br />

of the relaxation time, the pump-probe experiments described below are to be preferred,<br />

because the intensity-dependent transmission directly follows the exciton lifetime<br />

[120].<br />

In the off-resonant regime at 760 nm, the lifetime is about 0.5 ps, rising to more than 2 ps<br />

at 680 nm (Fig. 9.50). As with DFWM measurements, the lifetime can not be seen as an exponential<br />

signal decay, but as a broadening of the laser pulse autocorrelation signal. The lifetime<br />

Figure 9.50: Wavelength dependence of the lifetime of the singlet exciton in p-TS6, derived from the<br />

broadening of the autocorrelation function of picosecond pulses. Different symbols denote different<br />

crystals [123].<br />

can be estimated by fitting a model function accounting for the laser pulse shape.<br />

Th. Fehn investigated the subject, using a Titan-sapphire laser system with pulse<br />

lengths of 120 fs between 720 nm <strong>and</strong> 820 nm. Here, the exponential signal decay can be directly<br />

resolved, because T 1 of the singlet exciton is much longer than the laser pulse<br />

(Fig. 9.51). Accordingly, the relaxation time can be determined more accurately. In comparison,<br />

W. Schmid’s fits tend to yield a slightly longer relaxation time, but within the error<br />

ranges the results are consistent.<br />

9.5.7 The w (3) tensor components<br />

The w (3) tensor component parallel to the polymer chain of about 2–3610 –10 esu (off resonant)<br />

is the well dominating source of the non-linear response. The DFWM signal was found<br />

to be polarized parallel to the polymer chain <strong>and</strong> the signal decreased below the detection<br />

threshold if any or all of the pump beams were polarized perpendicular to the chain. From<br />

these observations w (3)<br />

perp can be estimated to be lower than 1610 –12 esu, i. e. at least a factor<br />

172


9.5 Non-linear optical properties<br />

Figure 9.51: The energy relaxation time of the singlet exciton is clearly resolvable with Ti-sapphire laser<br />

driven pump-probe experiments.<br />

of 200 lower than w (3)<br />

par [121, 122]. The behaviour of the different PDAs under investigation<br />

(p-TS6, p-4BCMU, p-FBS) was very similar [118].<br />

In the near-resonant frequency regime, a constant ratio of w (3) <strong>and</strong> the linear absorption<br />

coefficient a has been found, as predicted by the PSF model [115]. This model provides<br />

the best description of the experimental results [120].<br />

9.5.8 Signal saturation<br />

With high pump intensities saturation of the w (3) signal was observed (Fig. 9.52). Although this<br />

general behaviour was found in all PDA samples, the finer details, e. g. the dependence on the<br />

mean intensity, varied between different measurements. It turned out that mainly three effects<br />

are responsible for the damping of the signal, whose relative influence depends on the wavelength,<br />

the crystal thickness, <strong>and</strong> the general quality of the crystals. A thermo-optical effect<br />

widens the focus diameter thus decreasing the intensity of the pump/probe beams. Induced absorption<br />

<strong>and</strong> two-photon absorption decrease this intensity too. <strong>Final</strong>ly, polydiacetylenes show<br />

a significant influence of w (5) on the non-linear susceptibility, which is contrary to w (3) .<br />

The induced absorption was measured concurrently to the DFWM signal by the transmission<br />

of one of the pump beams. Thus, one can easily account for this effect in the measurements.<br />

Our measurements on manually cleaved thick (50–150 µm) crystals [118] showed a<br />

stronger damping with higher repetition rates of the laser system, i. e. higher heat load but<br />

lower peak intensities. No damping was observed at off-resonant frequencies, e. g. at 750 nm.<br />

Therefore, the damping is clearly related to the heat load absorbed by the crystal, an indica-<br />

173


9 Diacetylene Single Crystals<br />

Figure 9.52: The flattening of the intensity dependence of the DFWM signal can successfully be described<br />

by a significant value of the fifth order susceptibility w (5) (inset solid curve: fit with assumed<br />

contributions of both w (3) <strong>and</strong> w (5) ) [123].<br />

tion of a thermo-optical effect. An amplification of the effect by inclusions of the solvent<br />

was proposed.<br />

In samples prepared later on with better optical quality W. Schmid observed a different<br />

behaviour [120]. The damping became stronger with lower repetition rates, i. e. higher<br />

peak intensity. No thermo-optic effect was observed. This can be well explained by the advances<br />

in crystal preparation: solvent inclusions were reduced <strong>and</strong> the crystal quality improved,<br />

resulting in a lower absorption coefficient at the same, near-resonant wavelength<br />

(e. g. 720 nm). So, the absorbed heat load <strong>and</strong> the thermo-optical coefficients were reduced<br />

concurrently.<br />

The damping effect observed by W. Schmid has to be addressed to a w (5) effect contrary<br />

to w (3) [123]. Later measurements on thin crystalline films, using an improved measurement<br />

technique, sustained this interpretation [106].<br />

9.5.9 Spectral dispersion, phase, <strong>and</strong> relaxation of w (5)<br />

The value of w (5) has first been evaluated by Schmid [123], by fitting a polynomial to the<br />

measured curve of signal intensity versus pump laser intensity. At 748 nm, this fit yields a<br />

value of w (5) = 2.8610 –34 (m/V) 4 , with a phase angle of 1658 between w (3) <strong>and</strong> w (5) .<br />

With the improved modulation method, a w (5) value of (1.69 ± 0.79)610 –33 (m/V) 4<br />

(1.1610 –16 esu) was evaluated at 729 nm [106]. This is in reasonable agreement with<br />

Schmid’s measurements, considering the different wavelengths, different measurement techniques,<br />

<strong>and</strong> different samples.<br />

After successful growth of thin (a few micrometers) p-TS6 crystals, it was possible to<br />

study w (5) signals at near-resonant wavelengths without an interfering thermo-optical effect<br />

174


9.5 Non-linear optical properties<br />

Figure 9.53: Wavelength dependence of w (3) <strong>and</strong> w (5) in p-TS6 single crystals. The isolated dots are experimental<br />

values obtained from a thin crystal (about 5 µm), all other dots are from a thick crystal<br />

(about 50 µm) [106].<br />

[106]. The spectral dispersion of w (3) <strong>and</strong> w (5) has been measured with the improved modulation<br />

technique in the near resonant regime (Fig. 9.53). The w (5) amplitude increases roughly<br />

proportional to w (3) , as expected from the PSF model. In this total regime, a phase between<br />

w (5) <strong>and</strong> w (3) of about 1608 was found.<br />

In the near resonant regime, the w (5) relaxation shows the same behaviour as w (3) .No<br />

energy relaxation can be resolved in the off-resonant regime around ca. 720 nm, but the relaxation<br />

time increases with shorter wavelengths (Fig. 9.54).<br />

However, with even larger wavelengths, the w (5) relaxation times rise again (e. g. 2.3 ps<br />

at 748 nm) [120]. This can be successfully described proposing a three level energy scheme<br />

Figure 9.54: Relaxation time of the w (5) response in p-TS6. At 720 nm, only an 5 th order autocorrelation<br />

can be seen. At 685 nm a finite lifetime is clearly resolvable [106].<br />

175


9 Diacetylene Single Crystals<br />

Figure 9.55: At wavelengths between 720 nm <strong>and</strong> 750 nm, the w (5) signal relaxes in a two-step process [120].<br />

with an even state at 3.2 eV [123]. In addition to the fast dominant process, a very slow weak<br />

relaxation has been found, with relaxation times between 20 ps <strong>and</strong> 100 ps (Fig. 9.55) [120].<br />

9.5.10 Conclusion<br />

The polydiacetylenes’ non-linear optical behaviour was found to be in accordance with the<br />

predictions of the PSF model that assigns the non-linearity to the bleaching of the exciton<br />

b<strong>and</strong>. Theoretical approaches neglecting excitionic effects clearly fail. The exciton dominates<br />

the non-linear optics of polydiacetylenes [118].<br />

With increasing laser intensity, the w (3) signal shows significant deviations from the<br />

ideal I 3 -dependence. In thick crystals at near resonant wavelengths, a thermo-optic effect<br />

can be observed, which is amplified by crystal impurities [118]. However, also thin crystals<br />

show a similar damping of the w (3) signal. In contrast to the thermo-optical effect, this does<br />

not depend on the average heat load absorbed by the crystal, but on the peak intensity of the<br />

pump pulses. This effect can be described by a w (5) effect with a phase shift of 1608 against<br />

w (3) , by analysis of the intensity dependence of the net signal of both effects [120] as well as<br />

by direct measurements using a new modulation technique [106].<br />

Polydiacetylenes with different symmetrically substituted side groups differ mainly in<br />

terms of crystal quality, stability, <strong>and</strong> solubility; the influence on the third order susceptibility<br />

|w (3) | is rather small [118]. Interesting effects where found, however, in the unsymmetrically<br />

substituted diacetylene NP/4-MPU. This diacetylene forms a non-centrosymetric crystal<br />

with a significant w (2) [124, 125]. The second harmonic generation with this crystal can<br />

176


References<br />

be phase matched, making the material very interesting for commercial applications, e. g. as<br />

a multiplying medium for a laser pulse autocorrelator [126].<br />

In terms of optical quality, stability, <strong>and</strong> the value of w (3) polydiacetylenes still are<br />

superior to other polymers with a linear p-electron system, e. g. the polyparaphenylenevinylene<br />

investigated at our institute [127, 128, 130]. This is mainly due to their unique ability<br />

to form macroscopically ordered single crystals with aligned polymer chains.<br />

References<br />

1. G. Wegner: Z. Naturforschung, 24b, 824 (1969)<br />

G. Wegner: Macromol. Chem., 145, 85 (1971)<br />

2. H.-J. Cantow (Editor): Polydiacetylenes, Springer, (1984)<br />

3. D. Bloor, R.R. Chance: Polydiacetylenes, NATO ASI Series, Martinus Nijhoff Publishers (1985)<br />

4. R.J. Young: in: D. Bloor <strong>and</strong> R.R. Chance (eds.): Polydiacetylenes, NATO ASI Series, Martinus<br />

Nijhoff Publishers, (1985)<br />

5. R.H. Baughman, H. Gleiter, N. Sendfeld: J. Polym. Sci., Polym. Phys. Ed., 13, 1871 (1975)<br />

6. D.N. Batchelder, D. Bloor: J. Polym. Sci., Polym. Phys. Ed., 17, 569 (1979)<br />

7. G. Wegner: Macromol. Chem., 145, 85 (1971)<br />

G. Wegner: Recent progress in chemistry <strong>and</strong> physics of poly(diacetylenes): in: W.E. Hatfield<br />

(ed.): Molecular metals, Plenum press N.Y., 209 (1979)<br />

8. D. Kobelt, E. Paulus: Acta Cryst., B30, 232 (1974)<br />

D. Bloor, L. Koski, G.C. Stevens, F.H. Preston, D.J. Ando: J. Mater. Sci., 10, 1678 (1975)<br />

9. V. Enkelmann: Acta Cryst. B33, 2842 (1977)<br />

V. Enkelmann: Structural aspects of the topochemical polymerization of diacetylenes, in: H.-J.<br />

Cantow (ed.): Polydiacetylenes, Springer, 91 (1984)<br />

10. D. Bloor: Diacetylene polymerization kinetics, in: D. Bloor <strong>and</strong> R.R. Chance (eds.): Polydiacetylenes,<br />

NATO ASI Series, Martinus Nijhoff Publishers, 1 (1985)<br />

11. H. Eckhardt, T. Prusik, R.R. Chance: Solid State photopolymerization of diacetylenes, in: D. Bloor<br />

<strong>and</strong> R.R. Chance (eds.): Polydiacetylenes, NATO ASI Series, Martinus Nijhoff Publishers, 25<br />

(1985)<br />

12. H. Eichele: Diplomarbeit, Universität Stuttgart (1975)<br />

13. D.N. Batchelder: The study of electronic excitations of polydiacetylenes <strong>and</strong> resonance raman<br />

spectroscopy, in: D. Bloor <strong>and</strong> R.R. Chance (eds.): Polydiacetylene, NATO ASI Series, Martinus<br />

Nijhoff Publishers, 187 (1985)<br />

14. G. Weiser, L. Sebastian: Electric field sensitive defect states in fully <strong>and</strong> partially polymerized<br />

PTS, in: D. Bloor <strong>and</strong> R.R. Chance (eds.): Polydiacetylenes, NATO ASI Series, Martinus Nijhoff<br />

Publishers, 213 (1985)<br />

15. E. Dormann: private communication<br />

16. E. Wassermann et al.: Chem Phys. Lett., 7, 409 (1970)<br />

E. Wassermann et al.: J. Chem Phys., 54, 4120 (1971)<br />

17. C.A. Hutchison Jr., B.W. Mangum: J. Chem. Phys., 34, 908 (1961)<br />

C.A. Hutchison Jr., B.W. Mangum: J. Chem. Phys., 29, 952 (1958)<br />

18. J.H. van der Waals, M.S. de Groot: Mol. Phys., 2, 333 (1959)<br />

J.H. van der Waals, M.S. de Groot: Magnetic interactions Related to phosphorescence: in: A.B.<br />

Zahlan (ed.): The Triplet State, Cambridge University Press (1967)<br />

177


9 Diacetylene Single Crystals<br />

19. S.P. Mc Glynn, T. Azumi, M. Kinoshita: Molecular spectroscopy of the triplet state, Prentice Hall,<br />

New Jersey, (1969)<br />

20. H. Eichele, M. Schwoerer, R. Huber, D. Bloor: Chem. Phys. Lett., 42, 342 (1976)<br />

21. C.A. Hutchinson Jr., B.E. Kohler: J. Chem. Phys., 51, 3327 (1969)<br />

22. G. Herzberg: Molecular spectra <strong>and</strong> molecular structure, Vol. III, Van Nostr<strong>and</strong> Reinhold Company,<br />

New York (1966)<br />

23. H. Sixl, W. Hersel, H.C. Wolf: Chem. Phys. Lett., 53, 39 (1978)<br />

24. C. Bubeck, H. Sixl, H.C. Wolf: Chem. Phys., 32, 231 (1978)<br />

25. M. Schwoerer, H. Niederwald: Macromol. Chem. Suppl., 12, 61 (1985)<br />

26. W. Hersel, H. Sixl, G. Wegner: Chem. Phys. Lett., 73, 288 (1980)<br />

27. C. Bubeck, W. Neumann, H. Sixl: Chem. Phys., 48, 269 (1980)<br />

28. R.A. Huber, M. Schwoerer: Chem. Phys. Lett., 72, 10 (1980)<br />

29. W. Neumann, H. Sixl: Chem. Phys., 50, 273 (1980)<br />

30. C. Bubeck, W. Hersel, H. Sixl, J. Waldmann: Chem. Phys., 51, 1 (1980)<br />

31. H. Niederwald, H. Eichele, M. Schwoerer: Chem. Phys. Lett., 72, 242 (1980)<br />

32. R.A. Huber, M.Schwoerer, H. Benk, H. Sixl: Chem. Phys. Lett., 78, 416 (1981)<br />

33. W. Neumann, H. Sixl: Chem. Phys., 58, 303 (1981)<br />

34. M. Schwoerer, R.A. Huber, W. Hartl: Chem. Phys., 55, 97 (1981)<br />

35. H. Benk, H. Sixl: Mol. Phys., 42, 779 (1981)<br />

36. W. Hartl, M. Schwoerer: Chem. Phys., 69, 443 (1982)<br />

37. H. Niederwald, M. Schwoerer: Z. Naturforsch., 38a, 749 (1983)<br />

38. H. Gross, W. Neumann, H. Sixl: Chem. Phys. Lett., 95, 584 (1983)<br />

39. H. Sixl, W. Neumann, R. Huber,V. Denner, E. Sigmund: Phys. Rev. B, 31, 142 (1985)<br />

40. C. Bubeck: PhD thesis, Universität Stuttgart (1981)<br />

41. W. Hersel: PhD thesis, Universität Stuttgart (1981)<br />

42. R.A. Huber: PhD thesis, Universität Bayreuth (1981)<br />

43. W. Hartl: PhD thesis, Universität Bayreuth (1981)<br />

44. H. Niederwald: PhD thesis, Universität Bayreuth (1982)<br />

45. R. Müller: Diplomarbeit, Universität Bayreuth (1982)<br />

46. H. Gross: PhD thesis, Universität Stuttgart (1983)<br />

47. W. Neumann: PhD thesis, Universität Stuttgart (1984)<br />

48. H. Sixl: Spectroscopy of the intermediate state of the solid state polymerization reaction in diacetylene<br />

crystals, in: H.-J. Cantow (ed.): Polydiacetylene, Springer, 49 (1984)<br />

49. R. Huber, M. Schwoerer, C. Bubeck, H. Sixl: Chem. Phys. Lett., 53, 35 (1978)<br />

50. C. Bubeck, H. Sixl, H.C. Wolf: Chem. Phys., 32, 231 (1980)<br />

51. R. Müller-Nawrath, R. Angstl, M. Schwoerer: Chem. Phys., 108, 121 (1986)<br />

52. H. Sixl, W. Hersel, H.C. Wolf: Chem. Phys. Lett., 53, 39 (1978)<br />

53. H. Niederwald, K.H. Richter, W. Güttler, M. Schwoerer: in Photochemistry <strong>and</strong> photobiology, Vol.<br />

2, Harwood Academic Publishers, London, 1155 (1984)<br />

54. Kröhnke: PhD thesis, Universität Freiburg (1979)<br />

55. K-H. Richter, W. Güttler, M. Schwoerer: Appl. Phys., A32, 1 (1983)<br />

56. H. Niederwald, K.-H. Richter, W. Güttler, M. Schwoerer: Mol. Cryst. Liq. Cryst., 93, 247 (1983)<br />

57. B.E. Kohler, H.-D. Bauer, Be.E. Kohler, W. Güttler, M. Schwoerer: Chem Phys. Lett., 125, 251<br />

(1986)<br />

58. R. Magnusson, T.K. Gaylord: J. Opt. Soc. Am., 67, 1165 (1977)<br />

59. H. Kogelnik, Bell Syst. Tech. J., 48, 2909 (1969)<br />

60. M.G. Moharam, L. Young: Appl. Opt., 17, 1757 (1978)<br />

61. D.L. Staebler, W. Burke, W. Phillips, J.J. Amodei: Appl. Phys. Lett., 26, 182 (1975)<br />

62. Z. Bor, N. Racz, G. Szabo, A. Müller, H.-P. Dorn, Helv. Phys. Acta, 56, 383 (1983)<br />

63. H.M. Smith (ed).: Topics in Applied Physics, Vol. 20, Holographic Storage Materials, Springer,<br />

(1977)<br />

64. H. Niederwald, H. Seidel, W. Güttler, M. Schwoerer: J. Phys. Lett., 88, 1933 (1984)<br />

65. H.-D. Bauer, Th. Vogtmann, I. Müller, M. Schwoerer: Chem. Phys., 133, 303 (1989)<br />

178


References<br />

66. H.-D: Bauer, Th. Vogtmann, M. Schwoerer: Mol. Cryst. Liq. Cryst., 183, 421 (1990)<br />

67. Th. Vogtmann, H.-D. Bauer, I. Müller, M. Schwoerer: in: J.L. Bredas <strong>and</strong> R.R. Chance (eds.):<br />

Conjugated Polymeric Materials, NATO ASI Series E, 182 (1990)<br />

68. H.-D. Bauer: PhD thesis, Universität Bayreuth (1990)<br />

69. Th. Vogtmann, Diplomarbeit, Universität Bayreuth (1988)<br />

70. W. Lorenz, Diplomarbeit, Universität Bayreuth (1990)<br />

71. R. Nowak, J. Sworakowski, B. Kuchta, M. Bertault, M. Schott, R. Jakubas, H.A. Kolodziej: Chem.<br />

Phys., 104, 467 (1986)<br />

72. P. Gruner-Bauer, E. Dormann: The dielectric properties of diacetylenes during thermal solid state<br />

polymerization, Second International Conference Electrical, Optical <strong>and</strong> Acoustic <strong>Properties</strong> of<br />

Polymers, Canterbury, UK, (The Plastics <strong>and</strong> Rubber Institute, London, UK), 15/1 (1990)<br />

73. P. Gruner-Bauer, I. Müller, E. Dormann, Ch. Kröhnke: Makromol. Chem., 192, 1541 (1991)<br />

74. P. Strohriegl, J. Gmeiner, I. Müller, P. Gruner-Bauer, E. Dormann, V. Enkelmann: Ber. Bunsenges.<br />

Phys. Chem., 95, 491 (1991)<br />

75. P. Gruner-Bauer, E. Dormann: Phys. Rev. B, 45, 13938 (1992)<br />

76. M. Hurst, R.W. Munn: Chem. Phys., 127, 1 (1988)<br />

77. M. Orczyk, J. Sworakowski, M. Bertault, R.M. Faria: Synth. Metals, 35, 77 (1990)<br />

78. M. Orczyk: Chem. Phys., 142, 485 (1990)<br />

79. V.N. Genkin, P.M. Mednis: Sov. Phys. JETP, 27, 609 (1968)<br />

80. C. Cojan, G.P. Agrawal, C. Flytzanis: Phys. Rev. B, 15, 909 (1977)<br />

81. P.-A. Albouy: Thèse 3ième cycle, Université de Paris Sud (1982)<br />

82. P. Strohriegl, H. Schultes, D. Heindl, P. Gruner-Bauer, V. Enkelmann, E. Dormann: Ber. Bunsenges.<br />

Phys. Chem., 91, 918 (1987)<br />

83. E. Dormann, P. Gruner-Bauer, P. Strohriegl: Makromol. Chem., Macromol. Symp., 26, 141<br />

(1989)<br />

84. H. Bässler, H. Sixl, V. Enkelmann: Polydiacetylenes, Adv. Polymer Sci., Vol 63, Springer, Berlin,<br />

(1984)<br />

85. P. Gruner-Bauer, P. Strohriegl, E. Dormann: Dielectric <strong>and</strong> pyroelectric properties of substituted<br />

diacetylenes, in: Electrical, Optical <strong>and</strong> Acoustic <strong>Properties</strong> of Polymers, (The Plastics <strong>and</strong> Rubber<br />

Institute, Conference Publication, Canterbury, UK), 11/1 (1988)<br />

86. Hans Schultes: PhD thesis, Universität Bayreuth (1987)<br />

87. P. Gruner-Bauer, P. Strohriegl, E. Dormann: Ferroelectrics, 92, 15 (1989)<br />

88. M. Feile, P. Gruner-Bauer, P. Strohriegl, E. Dormann: Ferroelectrics, 106, 345 (1990)<br />

89. P. Gruner-Bauer: PhD thesis, Universität Bayreuth (1991)<br />

90. D.Q. Xiao, D.J. Ando, D. Bloor: Chem. Phys. Lett., 90, 247 (1982)<br />

91. M. Bertault, M. Schott, J. Sworakowski: Chem. Phys. Lett., 104, 605 (1984)<br />

92. H. Schultes, P. Strohriegl, E. Dormann: Z. Naturforsch., 42 a, 413 (1987)<br />

93. G.F. Lipscomb, A.F. Garito, T.S. Wei: Ferroelectrics, 23, 161 (1980)<br />

94. H. Schultes, P. Strohriegl, E. Dormann: Ferroel. Bulletin, 1, 36 (1986); H. Schultes, P. Strohriegl,<br />

E. Dormann: Ferroelectrics, 70, 161 (1986)<br />

95. H. Winter, E. Dormann: The pyroelectric phase transition of „DNP“, a disubstituted diacetylene<br />

single crystal, in: A. Blumen et al. (eds.): Dynamical processes in condensed molecular systems,<br />

World Scientific Publishing Co., 214 (1990)<br />

96. H. Winter, E. Dormann, M. Bertault, L. Toupet: Phys. Rev. B, 46, 8057 (1992)<br />

97. P. Gruner-Bauer, E. Dormann: J. Phys.: Condens. Matter, 4, 5599 (1992)<br />

98. G. Nemec, E. Dormann: J. Phys.: Condens. Matter, 6, 1417 (1994)<br />

99. J. Even, M. Bertault, A. Girard, Y. Delugeard, J.-L. Fave: Chem. Phys., 188, 235 (1994)<br />

100. C. Sauteret, J.-P. Hermann, R. Frey, F. Pradère, J. Ducuing: Phys. Rev. Let., 36, 956 (1976)<br />

101. G.M. Carter, M.K. Thakur, Y.J. Chen, J.V. Hryniewicz: Appl. Phys. Lett., 47, 457 (1985)<br />

102. G.M. Carter, J.V. Hryniewicz, M.K. Thakur, Y.J. Chen, S.E. Meyler: Appl. Phys. Lett., 49, 998<br />

(1986)<br />

103. D.R. Ulrich: Mol. Cryst. Liq. Cryst., 189, 3 (1990)<br />

104. Ithaco Scientific Instruments IAN 37 (1985)<br />

179


9 Diacetylene Single Crystals<br />

105. W. Schmid, Th. Vogtmann, M. Schwoerer: Opt. Comm., 121, 55 (1995)<br />

106. A. Feldner: Diplomarbeit, Universität Bayreuth (1996)<br />

107. J. Ducuing: in: P.G. Harper, B.S. Wherett (eds.): Nonlinear Optics, Academic Press (1977)<br />

108. K.C. Rustagi, J. Ducuing: Opt. Comm., 10, 258 (1974)<br />

109. G.P. Agrawal, C. Flytzanis: Chem Phys. Lett., 44, 366 (1976)<br />

110. C. Cojan, G.P. Agrawal, C. Flytzanis: Phys. Rev., B15, 909 (1977)<br />

111. G.P. Agrawal, C. Cojan, C. Flytzanis: Phys. Rev., B17, 776 (1978)<br />

112. H. Haug, S. Schmitt-Rink: J. Opt. Soc. Am., B2, 1135 (1985)<br />

113. D.S. Chemla, D.A.B. Miller: J. Opt. Soc. Am., B2, 1155 (1985)<br />

114. S. Schmitt-Rink, D.S. Chemla, D.A.B. Miller: Phys. Rev. B, 32, 6601 (1985)<br />

115. B.I. Greene, J. Orenstein, R.R. Millard, L.R. Williams: Phys. Rev. Lett., 58, 2750 (1997)<br />

116. B.I. Greene, J. Orenstein, S. Schmitt-Rink: Science, 24, 679 (1990)<br />

117. L. Sebastian, G. Weiser: Chem. Phys., 62, 447 (1981)<br />

118. Th. Vogtmann: PhD thesis, Universität Bayreuth (1993)<br />

119. Th. Vogtmann, W. Schmid, Th. Fehn, F. Bauer, I. Bauer, M. Schwoerer: Synth. Met., 55–57, 4018<br />

(1993)<br />

120. W. Schmid: PhD thesis, Universität Bayreuth (1994)<br />

121. Th. Fehn: Diplomarbeit, Universität Bayreuth (1992)<br />

122. Th. Fehn, Th. Vogtmann, J. Hübner, M. Schwoerer: Appl. Phys. B, 59, 203 (1994)<br />

123. W. Schmid, Th. Vogtmann, M. Schwoerer: Chem. Phys., 204, 147 (1996)<br />

124. M. Braun: Diplomarbeit, Universität Bayreuth (1996)<br />

125. M. Braun, F. Bauer, Th. Vogtmann, M. Schwoerer: J. Opt. Soc. Am., B14, 1699 (1997)<br />

126. H. Oester: Dplomarbeit, Bayreuth (1997)<br />

127. F. Bauer: PhD thesis, Universität Bayreuth (1996)<br />

128. F. Bauer, M. Schwoerer: Mol. Cryst. Liq. Cryst., 283, 131 (1996)<br />

129. S.N. Maganov, G. Bar, H.-J. Cantow, H.-D. Bauer, I. Müller, M. Schwoerer: Polymer Bulletin, 26,<br />

223 (1991)<br />

130. Th. Schimmel, B. Winzer, R. Kemnitzer, Th. Koch, J. Küppers, M. Schwoerer: Adv. Mat., 6, 308<br />

(1994)<br />

131. H.-D. Bauer, A. Materny, I. Müller, M. Schwoerer: Mol. Cryst. Liq. Cryst., 200, 205 (1991)<br />

180


10 Matrix-Molecule Interaction in Dye-Doped<br />

Rare Gas Solids<br />

Thomas Giering, Peter Geissinger, Wolfgang Richter, <strong>and</strong> Dietrich Haarer<br />

10.1 Introduction<br />

The investigation of the electronic states of polyatomic organic dye molecules is enhanced<br />

considerably when they are doped into suitable solid host matrices [1] at low concentrations.<br />

If the interaction between host <strong>and</strong> guest molecules is weak, the guest molecule will exhibit<br />

its characteristic electronic <strong>and</strong> vibronic signature except for a possible shift of the total<br />

spectrum, starting with the zero-phonon origin. The hosts are chosen according to their absorption<br />

because they must not overlap with the guest states under investigation. This can be<br />

achieved for a variety of polymers, n-alkanes, <strong>and</strong> rare gases, so that the optical absorption<br />

spectra of these guest-host systems are dominated by the guest molecules.<br />

A natural extension of these investigations was the incorporation of guest molecules into<br />

disordered systems to serve as probes of the host properties. In contrast to n-alkane hosts, in disordered<br />

host materials the optical absorption spectra of these guest-host systems are often characterized<br />

by broad <strong>and</strong> featureless b<strong>and</strong>s, the so-called inhomogeneous broadening. When a<br />

dye molecule is incorporated into a host matrix its transition energy will experience a shift due<br />

to the dye-matrix interaction. In a (hypothetic) perfect crystal all guest molecules will experience<br />

exactly the same shift, whereas in disordered hosts a distribution of local environments<br />

<strong>and</strong> therefore a distribution of transition energies is observed, which can be as large as several<br />

100 cm –1 . The inhomogeneous broadening prevents access to the homogeneous absorption<br />

lines, whose widths are related to energy <strong>and</strong> phase relaxation processes in the system.<br />

The fact that the inhomogeneous broadening is a consequence of statistically distributed<br />

local environments of the guest molecules means that it can be described by a stochastic<br />

model which dates back to Markoff [2]. For a review see Stoneham [3]. In this model, for<br />

details see next Section, the host material is dissected into matrix units, which in the case of<br />

rare gas hosts are identical with the atoms. For polymeric hosts the matrix units correspond<br />

to the respective monomer units. The inhomogeneous distribution of absorption frequencies<br />

is then given by averaging over all possible arrangements of matrix units around a cavity<br />

containing the dye molecule, which are weighted by the line shift caused by each respective<br />

configuration. The overall inhomogeneous shift is calculated by adding up the contributions<br />

of each matrix unit to the total line shift. Their contributions depend on their respective positions<br />

to the dye molecule. A modification of the interatomic distances, for example<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

181


10 Matrix-Molecule Interaction in Dye-Doped Rare Gas Solids<br />

through the application of external pressure, leads to a change of the interaction <strong>and</strong> therefore<br />

to a shift of the line. This suggests that pressure effects can also be taken into account<br />

within the framework of the stochastic model [4].<br />

However, in order to generate measurable changes of the entire inhomogeneous b<strong>and</strong>,<br />

pressure changes in the range of several gigapascals are required. These high pressures alter<br />

the structure of the investigated sample significantly which also affects the homogeneous<br />

width [5]. To observe the effects of small pressure changes the spectral resolution has to be<br />

increased significantly. This can be accomplished by the experimental technique of hole burning<br />

spectroscopy, which was introduced in 1974 [6, 7]. The hole burning method, for a description<br />

see Ref. [8] <strong>and</strong> Chapter 5, allows access to homogeneous line widths which are<br />

masked by the inhomogeneous broadening. Furthermore, within the inhomogeneous b<strong>and</strong>,<br />

spectral holes can serve as persistent narrow frequency markers. The changes to these frequency<br />

markers are due to: structural relaxations of disordered host materials (Chapter 5), IR<br />

induced spectral diffusion (Chapter 6), external perturbations – like electric fields – [9, 10],<br />

<strong>and</strong> pressure (Section 10.6) <strong>and</strong> can be monitored accurately over extended time periods.<br />

In the ideal case the hole width is twice the homogeneous width. Therefore the increase<br />

in spectral resolution is roughly given by the ratio of inhomogeneous to homogeneous<br />

width. This ratio, also referred to as the multiplexing factor, which is due to the possible application<br />

of the hole burning technique in data storage devices, is typically in the range of<br />

10 3 to 10 5 [8]. This means that for producing detectable changes in a spectral hole due to<br />

external pressure the magnitude of the external perturbation can be reduced by approximately<br />

the multiplexing factor. This in turn allows sample investigations near equilibrium<br />

conditions, meaning that the displacements of the matrix units from their zero-pressure positions<br />

are small. The results of the first pressure tuning experiments reveal a red shift of the<br />

hole center, when the pressure is raised, <strong>and</strong> a blue shift, when the pressure is lowered after<br />

the burning of the hole [11, 12]. Furthermore, in both cases a hole broadening is observed.<br />

Laird <strong>and</strong> Skinner’s extension of the stochastic approach [4] for describing pressure effects<br />

was first applied to dye molecules in various polymer hosts like polyethylene (PE),<br />

polystyrene (PS), <strong>and</strong> polymethylmethacrylate (PMMA). Their prediction of a frequencydependent<br />

pressure shift was duly verified. Within 20% the predicted value agreed with the<br />

experimental results. The apparent success of the stochastic model was surprising, because<br />

polymers meet the basic requirements of the model quite poorly. Usually it is assumed that<br />

one type of interaction, e. g. dispersive forces, between dye <strong>and</strong> matrix molecules predominates.<br />

Also, the matrix units, which in this case are the monomer units, are considered to be<br />

spherical <strong>and</strong> independent of each other yielding additive contributions to the solvent shift<br />

of the doped dye molecules, which are also assumed to be spherical. Additionally, the matrix<br />

units are assumed to be able to arrange themselves independently around the dye molecule.<br />

In the case of polymer hosts, the matrix units clearly are unable to arrange themselves independently,<br />

because they are connected by strong directional bonds. Moreover, the monomer<br />

units are often slightly polar, like PMMA.<br />

Furthermore, the validity range of the results of the stochastic model remained unclear,<br />

since in the course of the calculation two conflicting approximations with regard to<br />

the number density r of the matrix units within the interaction range of the dye molecule<br />

were made: the Gaussian approximation (valid for r??) <strong>and</strong> the continuum approximation<br />

(valid for r?0). The latter is neglecting correlations between matrix units, which arise<br />

from their mutual steric exclusion.<br />

182


10.2 Stochastic theory<br />

To clarify these inconsistencies we set out to systematically investigate the stochastic<br />

model, using model systems that come closest to its basic assumptions. For this reason we<br />

chose solid rare gases as host matrices. For these systems additional information is readily<br />

available [13, 14]. Furthermore, the interaction potential <strong>and</strong> its parameters are well-known for<br />

pure rare gases. Matrix parameters can be varied systematically by using different rare gases,<br />

while still retaining a structural similarity. Some rare gas mixtures show a different structure<br />

than pure rare gases (Sections 10.3 <strong>and</strong> 10.6). Through a systematic variation of the mixing ratio<br />

the impact of the involved structural transition onto the optical data can be investigated.<br />

10.2 Stochastic theory<br />

The above-mentioned stochastic model description considers an amorphous system of N matrix<br />

units, containing a small concentration of dye molecules. Each matrix unit will shift the<br />

electronic absorption line of the dye molecule by some amount n( ~R i †, where ~R i is the position<br />

of the i th matrix unit with respect to the dye molecule. The contributions of all matrix<br />

units are assumed to be additive. The total inhomogeneous distribution of absorption lines<br />

can then be written as [2–4]:<br />

I…† ˆ 1<br />

V N Z<br />

d ~R 1 ...d ~R N P… ~R 1 ; ...; ~R N †<br />

<br />

X N<br />

iˆ1<br />

!<br />

… ~R i † : …1†<br />

V is the volume of the sample. In order to further evaluate this expression, the function<br />

P ( ~R 1 ; ...; ~R N ), which is the probability of finding N matrix units at positions ~R 1 ; ...; ~R N ,<br />

has to be specified. If correlations between the matrix units are neglected, the (N+1) particle<br />

solute-solvent distribution function can be factorized into a product of two-particle distribution<br />

functions g( ~R):<br />

P… ~R 1 ; ...; ~R N †ˆ QN<br />

nˆ1<br />

g… ~R n † :<br />

…2†<br />

This is the so-called continuum approximation, which is valid in the case of small<br />

number densities r of matrix units.<br />

Furthermore, considering only a non-polar solute <strong>and</strong> solvent, the perturbation function<br />

n( ~R) of the transition frequency is of the familiar Lennard-Jones type<br />

( " <br />

s 12 #<br />

s 6<br />

… ~R† ˆ 4"<br />

R R 0 R R 0<br />

if R R 0 ;<br />

(3)<br />

1 if R


10 Matrix-Molecule Interaction in Dye-Doped Rare Gas Solids<br />

with the origin shifted by R 0 to account for the large size difference between solute <strong>and</strong> solvent<br />

[4]. In this representation R 0 + s/2 is the solute radius, while<br />

6p<br />

2s can be taken as the<br />

solvent diameter. This means that a matrix unit located at the position of the potential minimum<br />

R 0 +<br />

6p<br />

2s will shift the electronic absorption line of the dye by the potential depth e,<br />

as given in Eq. 3. In the case of r??, the inhomogeneous distribution is Gaussian, which<br />

is a consequence of the Central Limit Theorem [15].<br />

For the spatial distribution of the matrix units a simple step function was chosen [4],<br />

meaning that a matrix unit can be found anywhere outside a spherical cavity of radius<br />

R 0 + R c , but not inside. The parameter R c determines how close a matrix unit can come to<br />

the dye molecule. With the above conventions about solvent <strong>and</strong> solute dimensions, we have<br />

R c = s/2 (1 +<br />

6p<br />

2 ) &1.061 s.<br />

Inserting Eqs. 2 <strong>and</strong> 3, the inhomogeneous distribution (Eq. 1) can now be evaluated,<br />

resulting in expressions relating the experimentally accessible parameters, namely the solvent<br />

shift n s (spectral position of the b<strong>and</strong> maximum with respect to its gas phase position)<br />

<strong>and</strong> the full width at half maximum G s to the local number density r <strong>and</strong> to the microscopic<br />

parameters e, R 0 , s, <strong>and</strong> R c [15].<br />

We can now turn to the question of matrix correlations, which arises because of the<br />

application of two conflicting approximations in the course of the calculation: the Gaussian<br />

approximation for large <strong>and</strong> the continuum approximation for small number densities<br />

of the matrix units. A reasonable way out of this dilemma is to retain the Gaussian approximation<br />

<strong>and</strong> to introduce matrix correlations. Here the only correlation effect considered<br />

is the principle that two matrix units cannot be located at the same position. This<br />

means that the factorization used in Eq. 2 cannot be applied. Matrix correlations can be<br />

accounted for, within the framework of the stochastic model, by introducing a three-particle<br />

distribution function g 3 ( ~R; ~R 0 † [16–18]. Applying the Kirkwood superposition approximation<br />

[19],<br />

<br />

<br />

g 3 … ~R; ~R 0 †g… ~R†g… ~R 0 †g S ~R ~R 0<br />

; …4†<br />

we introduce a solvent-solvent distribution g S (| ~R ~R 0 |). In analogy to the solute-solvent distribution,<br />

we insert a simple step function for g S (| ~R ~R 0 |) with the cut-off radius given by<br />

6p<br />

2s . As in the case of neglected correlations it is possible to derive expressions that allow<br />

the calculation of the potential depth e <strong>and</strong> the local number density of the matrix units r<br />

from spectroscopic data (Section 10.5).<br />

As mentioned in the introduction, the stochastic approach can also provide a description<br />

of the effects of external pressure on spectral holes. In our experiments, however, the<br />

pressure change is always accompanied by a simultaneous change in temperature, see<br />

Ref. [20]. Therefore, the observed hole shifts <strong>and</strong> broadenings will not only be due to pressure<br />

changes, but also to the thermal expansion of the matrix. There may also be dynamical<br />

effects such as phonon scattering <strong>and</strong> (fast) tunneling systems (TLS) relaxations, which we<br />

will not treat in this contribution.<br />

Pure pressure effects have been accounted for by the Laird <strong>and</strong> Skinner theory [4].<br />

This theory can be conveniently exp<strong>and</strong>ed to also include the thermal expansion of the matrix<br />

<strong>and</strong> its influence on the dye molecule, for details see Refs. [20, 21]. In analogy to the<br />

inhomogeneous distribution (Eq. 1), the temperature-pressure kernel, i. e. the probability that<br />

184


10.2 Stochastic theory<br />

a guest molecule with the original solvent shift n will have a new transition frequency n'<br />

after a pressure change Dp <strong>and</strong> a temperature change DT, can be written as<br />

f … 0 ; p; T† ˆ<br />

Z<br />

1<br />

I…†V N<br />

<br />

<br />

P N<br />

iˆ1<br />

d ~R 1 ...d ~R N P… ~R 1 ; ...;R * N†<br />

<br />

<br />

… ~R i † 0 P N<br />

0 … ~R i ;p;T† : …5†<br />

The observed temperature-pressure shift was found to be linear, while the concomitant<br />

hole broadening can be described by a power law. The latter is clearly dominated by dynamical<br />

processes which are affected by the change in temperature. It is now assumed that, in analogy<br />

to pure pressure effects, the hole shift <strong>and</strong> the hole broadening due to the thermal volume<br />

expansion are linear functions of the temperature change. The function n' ( ~R i ; Dp; DT)<br />

in Eq. 5 can then be linearized to<br />

0 … ~R;p;T† ˆ… ~R† k @… ~R†<br />

@R<br />

iˆ1<br />

!<br />

R<br />

p ‡ g<br />

3<br />

@… ~R†<br />

@R<br />

!<br />

R<br />

T :<br />

3<br />

k is the compressibility <strong>and</strong> g the volume thermal expansion coefficient of the matrix.<br />

It is important to note that the temperature <strong>and</strong> pressure effects have opposite signs, see<br />

Eq. 6 <strong>and</strong> Refs. [11, 22].<br />

The further evaluation follows the lines of the calculation given in Ref. [21]. Within<br />

the above-mentioned approximations, the pressure-temperature kernel is found to be Gaussian.<br />

At this point we restrict our considerations to the evaluation of the hole shift, which is<br />

calculated as<br />

" Z ( )( )#<br />

…; p; T† ˆ d 3 R @… ~R†<br />

Rg…R†<br />

1 ‡… S † … ~R†<br />

3 @R<br />

s 2 …gT kp† :<br />

S<br />

(7)<br />

s S is related to the full width at half maximum of the inhomogeneous line shape G S by<br />

G<br />

s S = p S<br />

, while n is the burning frequency of the spectral hole with respect to the gas-phase<br />

2 2 ln 2<br />

position of the transition frequency of the dye molecule.<br />

For the modified Lennard-Jones potential (Eq. 3) Eq. 7 has to be evaluated numerically.<br />

It can, however, be simplified by considering only a purely attractive van der Waals<br />

potential. This approximation seems reasonable, since only matrix units located in the attractive<br />

part of the intermolecular potential can cause the observed red shift in pure pressure<br />

tuning experiments [11]. We obtain for the temperature-pressure shift:<br />

…; p; T† ˆ2 ‰ kp gTŠ: …8†<br />

With this equation, we found a simple way to extract the pure pressure shift from our<br />

pressure-temperature data. Due to the opposite sign of pressure <strong>and</strong> temperature shifts<br />

(Eq. 8), the pure pressure shift will be even larger with the inclusion of the temperature ef-<br />

…6†<br />

185


10 Matrix-Molecule Interaction in Dye-Doped Rare Gas Solids<br />

fects, leading to a larger compressibility after including the correction. The evaluation of<br />

our experimental data leads to an additional complication. The volume thermal expansion<br />

coefficient depends, contrary to the compressibility, strongly on the temperature, even in the<br />

small temperature interval from 1.8 K to 5 K. Fortunately, for crystalline samples experimental<br />

data are available [23] which allow us to substitute gDT ? R TDT<br />

T<br />

dT'g(T') in Eq. 8.<br />

With the help of this equation it can be shown that the pressure effects should clearly dominate<br />

the experiment [20], which is confirmed by the observed red shift in pure pressure-tuning<br />

experiments [11].<br />

The stochastic description of pressure effects can also be modified to include correlations<br />

between matrix units [20, 21]. It is interesting to note that for the purely attractive van<br />

der Waals potential, Eq. 8 will not acquire any additional terms suggesting that for the determination<br />

of the compressibility k matrix correlations will also play a minor role when the<br />

Lennard-Jones potential Eq. 3 is valid (Section 10.7).<br />

10.3 Rare gases<br />

Due to their simplicity rare gases have served as model systems for a long time. This means<br />

that a large data basis [13, 14, 24] is available. Of special importance for our investigations<br />

is the knowledge of the microscopic interaction potential between rare gas atoms. For its attractive<br />

region it was already calculated by F. London [25], to depend as 1/R 6 upon the interatomic<br />

distance. In the well-known Lennard-Jones (6,12) potential the exponential R-dependence<br />

for the repulsive contribution is approximated by an algebraic 1/R 12 term:<br />

s<br />

12 <br />

s<br />

6<br />

…R† ˆ4~"<br />

: …9†<br />

R R<br />

The potential depths ~e, which are directly connected with the polarizability a <strong>and</strong> the<br />

parameters s, which determine the distance of the potential minimum, are listed in Tab. 10.1.<br />

Table 10.1: Lennard-Jones parameters ~e <strong>and</strong> s [26, 27] <strong>and</strong> polarizability a [28].<br />

Gas ~e [cm –1 ] s [Å] a [10 –25 cm 3 ]<br />

Argon 83.7 3.405 16.3<br />

Krypton 113 3.65 24.8<br />

Xenon 161 3.98 40.1<br />

Another aspect is the structure of solid rare gases, which can be produced either by<br />

slow freezing the liquid or by condensation onto a sufficiently cold substrate. While the for-<br />

186


10.4 Experimental<br />

mer method yields single crystals with macroscopic dimensions, condensed rare gases are<br />

polycrystalline. Electron <strong>and</strong> X-ray diffraction [29, 30] experiments show the crystals <strong>and</strong><br />

crystallites to have a face-centred cubic structure, although theoretical simulations predicted<br />

the hexagonal close packing to be energetically favoured by about 0.01%.<br />

The situation is different for certain quench condensed rare gas mixtures, where computer<br />

simulations [31] predicted the possibility of an amorphous structure for rare gases<br />

whose atomic radii differ by at least 10%. X-ray diffraction measurements [30] indeed revealed<br />

an amorphous structure for Ar 1–x Xe x matrices with mixing ratios 0.2 < x < 0.7. This<br />

opens up the possibility to study the transition between the polycrystalline <strong>and</strong> amorphous<br />

state by varying the mixing ratio in a single system.<br />

10.4 Experimental<br />

The dye-doped rare gas matrices had to be prepared in situ. For this purpose a gas h<strong>and</strong>ling<br />

system was installed which provided alternatively the pure rare gases (Linde, 99.998% purity<br />

for argon, 99.990 % purity for krypton <strong>and</strong> xenon) <strong>and</strong> rare gas mixtures. The concentrations<br />

of the mixture components were determined by monitoring the pressure in the mixing<br />

chamber during the composition process. Commercially available phthalocyanine (Aldrich,<br />

used without further purification) was sublimated <strong>and</strong> subsequently mixed with the desired<br />

matrix gas in a Knudsen effusion furnace. Following Bajema et al. it consisted of a sublimation<br />

chamber (T&600 K) <strong>and</strong> a super-heating chamber (T&700 K). The gaseous matrix<br />

passed through a nozzle onto a sapphire substrate being in thermal contact with the cold finger<br />

of a continuous flow cryostat.<br />

Two different cryo-systems were used, which opened a wide temperature range for investigations.<br />

In system I temperatures down to 4 K could be produced using a commercially<br />

available continuous flow cryostat (Oxford Instruments). A home-built special outer vacuum<br />

container enabled either sample deposition or optical access through the sample. The system II<br />

is our own design. It consisted of a continuous flow cryostat which was immersed into liquid<br />

helium after the sample preparation. Temperatures down to 1.5 K could be reached by pumping<br />

the helium. Furthermore, hydrostatic pressure could be exerted by sealing off the helium<br />

bath. A typical dye concentration was approximately 2610 –3 mol/l with a sample thickness<br />

of about 10 mm. To characterize the samples, absorption spectra were recorded with a monochromator<br />

(Jobin Yvon THR 1500, used resolution about 0.2 cm –1 ). Hole burning studies<br />

were performed using a single mode dye laser (Coherent 599, b<strong>and</strong>width about 3 MHz). The<br />

optical setup is described in detail in Refs. [11, 32].<br />

187


10 Matrix-Molecule Interaction in Dye-Doped Rare Gas Solids<br />

10.5 Inhomogeneous absorption lines<br />

The visible absorption spectra of phthalocyanines show two regions with strong <strong>and</strong> characteristic<br />

absorption b<strong>and</strong>s. The lowest electronic energy transition, called Q b<strong>and</strong>, is degenerated<br />

in metallophthalocyanines due to a molecular D 4h symmetry. In free base phthalocyanines<br />

(H 2 Pc), however, the reduction of molecular symmetry from D 4h to D 2h , caused by<br />

the two H atoms in the inner ring, lifts the degeneracy of the Q b<strong>and</strong>. It splits into two<br />

b<strong>and</strong>s, which are conventionally referred to as Q x (for the lower energy component) <strong>and</strong> Q y<br />

(for the higher energy component). These two absorption b<strong>and</strong>s are shown in Fig. 10.1 for a<br />

H 2 Pc-doped krypton matrix. At higher frequencies a number of vibronic lines appear.<br />

Figure 10.1: Absorption spectrum of H 2 Pc in a krypton matrix.<br />

Our spectra agree quite well with spectra recorded previously by Bajema et al. [33].<br />

Due to the inherent local disorder in vapour condensed matrices [30, 34] the absorption<br />

spectra are inhomogeneously broadened. For the investigated systems Gaussian line shapes<br />

are expected from theoretical considerations [2, 3]. However, this prediction is difficult to<br />

test due to a slight asymmetry of the absorption spectra towards lower frequencies. This observation<br />

indicates a second preferential position of H 2 Pc molecules in the rare gas matrix<br />

<strong>and</strong> confirms earlier measurements [35]. In samples with a sufficiently high dye concentration<br />

this second site leads to a weaker set of spectral transitions shifted about 70 cm –1 towards<br />

lower energies [35, 36].<br />

Table 10.2 compiles the spectral positions of the Q x <strong>and</strong> Q y b<strong>and</strong>s as well as the<br />

widths of the Q x b<strong>and</strong>s <strong>and</strong> their solvent shifts n S , i. e. the frequency shifts of the respective<br />

absorption maxima with respect to the absorption of free H 2 Pc molecules, as measured in a<br />

supersonic free jet [37]. From argon to xenon increasing solvent shifts <strong>and</strong> inhomogeneous<br />

widths can be observed.<br />

188


10.5 Inhomogeneous absorption lines<br />

Table 10.2: Spectroscopic data (positions n, solvent shift n S , width G S ) of the absorption b<strong>and</strong>s of H 2 Pcdoped<br />

rare gas matrices. Units are in cm –1 .<br />

Matrix n (Q x ) G S (Q x ) n S (Q x ) n (Q y )<br />

Argon 14764 44 364 15731<br />

Krypton 14664 58 464 15605<br />

Xenon 14540 89 593 15466<br />

The solvent shift measures the strength of the dye-matrix interaction which is purely<br />

dispersive for the investigated systems. When going from argon to xenon the increasing solvent<br />

shift can therefore be explained by the increasing matrix polarizability. The latter is<br />

even effective enough to compensate the reduction of the number of interacting matrix<br />

atoms. Due to their increasing size the solvent shift is produced by fewer interacting units<br />

which, according to basic statistics, enlarges the relative fluctuation of the line shifts <strong>and</strong><br />

manifests itself in the observed increasing inhomogeneous width.<br />

However a quantitative analysis of the absorption b<strong>and</strong>s requires the formalism of the<br />

stochastic theory, outlined in Section 10.2, which is able to connect the measured solvent<br />

shifts <strong>and</strong> inhomogeneous b<strong>and</strong>widths to two microscopic parameters of the system, namely<br />

the respective number densities r of matrix units around a dye molecule <strong>and</strong> the depths e of<br />

the dye-matrix interaction potentials. While for polymer matrices the stochastic theory was<br />

to be used to determine geometric parameters [4], they are already known for our rare gas<br />

model systems from independent investigations. This enabled us to reduce the number of fit<br />

factors <strong>and</strong> calculate the r <strong>and</strong> e values as listed in Tab. 10.3.<br />

Table 10.3: Potential parameters e <strong>and</strong> number densities r calculated with (^r;^e) <strong>and</strong> without (r, e) matrix<br />

correlations. r cryst : number density for pure rare gas crystals.<br />

^e [cm –1 ] e [cm –1 ] ^r [Å –3 ] r [Å –3 ] r cryst [Å –3 ]<br />

Argon 23.1 1.57 0.00975 0.145 0.0267<br />

Krypton 33.4 2.45 0.00787 0.104 0.0222<br />

Xenon 47.5 4.58 0.00594 0.062 0.0173<br />

The widely used continuum approximation turned out to lead to the unreasonable result<br />

that r exceeds the number densities measured in bulk rare gas crystals! The physical<br />

reason for this discrepancy is that the continuum approximation does not prevent matrix<br />

units from occupying identical positions, since their mutual steric exclusion has been neglected!<br />

Therefore the calculation is based upon too many matrix units – reflected in the unrealistically<br />

large number density – giving insufficient weight to the individual matrix unit<br />

[20]. Taking matrix correlations into account, as sketched in Section 10.2, the depth of the<br />

dye-matrix interaction potential ^e is increased <strong>and</strong> the number density ^r reduced to appropriate<br />

values (Table 10.3).<br />

189


10 Matrix-Molecule Interaction in Dye-Doped Rare Gas Solids<br />

10.6 Pressure effects<br />

In Fig. 10.2 the effect of hydrostatic pressure on a spectral hole, as measured with our system<br />

II, in an argon matrix is shown. A pressure increase causes a red shift <strong>and</strong> broadening<br />

of the initial hole. It has to be emphasized, however, that due to the construction of our cryostat,<br />

pressure changes were always connected with temperature changes following the helium<br />

vapour pressure curve. Therefore the measured broadening <strong>and</strong> shift of spectral holes do<br />

have thermal <strong>and</strong> dynamical contributions <strong>and</strong> are not only due to the pure pressure effect.<br />

These two (thermal <strong>and</strong> dynamical) contributions to the hole shift are separated in<br />

Fig. 10.3 for the case of argon. The measured hole shift is plotted versus the pressure increase<br />

Dp. The pure pressure contribution to the hole shift was extracted by taking into account<br />

the thermal expansion of the matrix (Section 10.2). Figure 10.3 also shows that the<br />

Figure 10.2: Broadening <strong>and</strong> pressure shift of a spectral hole (H 2 Pc in an argon matrix). Trace (1): Original<br />

hole profile; Trace (2): Spectral hole after a pressure increase of 88 kPa.<br />

Figure 10.3: Shift of hole minimum vs. pressure change: Raw data (squares) <strong>and</strong> data after temperature<br />

correction (bullets).<br />

190


10.7 Rare gas mixtures<br />

pure pressure shift depends linearly on the pressure increase <strong>and</strong> is larger than the total observed<br />

pressure shift due to the opposite signs of temperature <strong>and</strong> pressure effect (Eq. 8).<br />

Using Eq. 7 it is now possible to calculate the matrix compressibility k. The observed<br />

linear pressure shift towards lower energies suggests that the interacting matrix atoms are located<br />

in the attractive part of the Lennard-Jones potential. The van der Waals potential is<br />

therefore expected to be a good approximation, which was employed in the derivation of the<br />

analytical expression for the pressure dependence (Eq. 8). Indeed, Tab. 4 shows that the results<br />

obtained, using the Lennard-Jones <strong>and</strong> the van der Waals potential respectively, are<br />

identical within the accuracy of our experiment.<br />

Table 10.4: Matrix compressibilities in units of GPa –1 determined with (^k) <strong>and</strong> without (k) correlations.<br />

Calculation using Lennard-Jones (LJ) <strong>and</strong> van der Waals (vdW) potentials.<br />

^k LJ k LJ k vdW<br />

Argon 0.2500 0.2429 0.2758<br />

Krypton 0.2899 0.2653 0.2895<br />

Xenon 0.3768 0.3919 0.4191<br />

Of further interest is the question whether matrix correlations are of the same importance<br />

for the calculation of the matrix compressibility as they are for the potential parameter<br />

<strong>and</strong> the number density. A detailed analysis, however, reveals that for our model systems as<br />

well as for H 2 Pc-doped polymers PE <strong>and</strong> PS the corrections to the simple equations due to<br />

matrix correlations are smaller than the experimental error. This explains the good agreement<br />

between optically determined compressibilities, calculated conventionally without consideration<br />

of matrix correlations, <strong>and</strong> mechanically determined bulk values for polymeric systems.<br />

10.7 Rare gas mixtures<br />

So far experiments with H 2 Pc-doped pure rare gas matrices have been reported. As discussed<br />

above, the transition from a polycrystalline to an amorphous solid can be studied in a<br />

single system using condensed rare gas mixtures, on which we will focus now.<br />

The Q x absorption b<strong>and</strong>s of H 2 Pc-doped matrices with various Ar-Xe compositions are<br />

plotted in Fig. 10.4 together with the spectra for the pure argon <strong>and</strong> xenon matrices. In contrast<br />

to the slight asymmetry of the spectra in the pure rare gas matrices, which was already mentioned<br />

above, the absorption b<strong>and</strong>s in rare gas mixtures can be well-described by a Gaussian line<br />

shape. For all composition ratios the Q x b<strong>and</strong>s are located within the spectral range given by the<br />

absorption lines of the two pure rare gas hosts. However, it is not possible to describe the spectra<br />

of the mixed matrices as a linear combination of the pure constituents spectra. This demonstrates<br />

that the gases are homogeneously mixed before condensation as well as in the solid matrix.<br />

191


10 Matrix-Molecule Interaction in Dye-Doped Rare Gas Solids<br />

Figure 10.4: Normalized absorption spectra of the Q x b<strong>and</strong> of H 2 Pc in solid Ar 1–x Xe x matrices. From<br />

left to right: x = 1 (pure Xe), x = 0.6, 0.4, 0.3, 0.15, x = 0 (pure Ar).<br />

More insight into the dependence of the spectra upon the composition ratio can be<br />

gained if the frequency of the absorption maximum is plotted against the Xe concentration.<br />

This is done in Fig. 10.5, where the solvent shift is also given. The latter rises continuously<br />

with increasing Xe concentration.<br />

The solid line in Fig. 10.5 was calculated with Eq. 1 extended to mixed systems. The<br />

density of the matrix units is interpolated from the values of the pure components assuming<br />

a constant atomic volume, as described in [39]. Using this simple approach, the calculated<br />

Figure 10.5: Position of the maximum of the Q x b<strong>and</strong> (left-h<strong>and</strong> scale) <strong>and</strong> solvent shift n S (right-h<strong>and</strong><br />

scale) of H 2 Pc in Ar 1–x Xe x matrices versus Xe concentration. The solid line corresponds to the interpolation<br />

of the mass density on the basis of constant atomic volume.<br />

192


10.7 Rare gas mixtures<br />

curve agreed quite satisfactory with the experimental data. The slight deviation from the theoretical<br />

curve may have its origin in uncertainties of the respective matrix densities [40] <strong>and</strong><br />

compositions due to different sublimation temperatures of argon <strong>and</strong> xenon.<br />

The quasi-homogeneous line width is obtained by extrapolating the spectral hole width<br />

to zero hole area, assuming that the latter depends linearly on the burning fluence [8]. Thus,<br />

saturation effects may be excluded [41].The quasi-homogeneous line widths are determined<br />

for all mixed Ar-Xe matrices with a sufficient optical density at a temperature of 1.8 K. The<br />

results are plotted in Fig. 10.6 versus the Xe concentration. Since the line width depends on<br />

the spectral position in the absorption b<strong>and</strong> [40] the values are interpolated to the frequency<br />

of the absorption maximum of each sample. The quasi-homogeneous line width in pure argon<br />

matrices is significantly lower than in pure xenon matrices. From argon to xenon concentrations<br />

up to 50 % only minor changes for H 2 Pc-doped rare gas mixtures are observed<br />

with no recognizable discontinuity at the expected transition from a polycrystalline to an<br />

amorphous matrix.<br />

Figure 10.6: Quasi-homogeneous width, as measured at the maximum of the Q x b<strong>and</strong>, of H 2 Pc in<br />

Ar 1–x Xe x matrices versus Xe concentration.<br />

It is well-known that at liquid helium temperatures the fluorescence lifetime plays a<br />

minor role only for the quasi-homogeneous width whereas the dominating contributions<br />

come from fast relaxations of TLS <strong>and</strong> coupling to local modes. The experimentally observed<br />

smaller widths in argon matrices as compared to xenon matrices can therefore be attributed<br />

to the smaller TLS density, which was determined by specific heat investigations<br />

[30], <strong>and</strong> to a higher local mode frequency, measured via hole burning [20]. These two relaxation<br />

mechanisms are thought to be characteristic for disordered systems. Thus, TLS <strong>and</strong><br />

local modes are not only present in amorphous solids but also in polycrystalline samples, in<br />

which these excitations can exist at grain boundaries as well as in the neighbourhood of im-<br />

193


10 Matrix-Molecule Interaction in Dye-Doped Rare Gas Solids<br />

purities inside the crystallites. The line width behaviour at the transition to the amorphous<br />

state can therefore be attributed to the high disorder <strong>and</strong> to the porous structure that is already<br />

present in the pure matrices.<br />

10.8 Summary<br />

The present investigation centred on a stochastic model, which was developed to describe<br />

the inhomogeneous broadening of absorption lines of dye molecules doped into a disordered<br />

host. Furthermore, the model seems to account well for the effects of external pressure on<br />

spectral holes burned into these b<strong>and</strong>s, making it possible to determine the host compressibility<br />

by performing a purely optical experiment. This model has also been widely used for<br />

dye-doped polymer hosts although they hardly satisfy the basic assumptions, making its application<br />

questionable. The aim of the present work was therefore to put the model to the<br />

test by investigating dye-doped rare gas solid systems that come closest to satisfying these<br />

requirements.<br />

In its original form the theory treats the matrix units in a continuum approximation by<br />

neglecting correlations between them. This leads to unreasonable results for microscopic<br />

parameters, which could be demonstrated for the first time in our model systems. Therefore,<br />

the theory was extended to take into account steric exclusion of matrix units. For the depth<br />

of the dye-matrix interaction potential <strong>and</strong> the local matrix density the modified theory produces<br />

physically realistic results.<br />

The behaviour is different with respect to pressure effects on spectral holes. Our investigations<br />

verified that the local matrix compressibility, measured in our experiments, is<br />

mainly sensitive to the dispersive part of the dye-matrix potential. Therefore, the details of<br />

the repulsive region of the dye-matrix potential as well as the consideration of matrix-matrix<br />

correlations cause only minor changes to the matrix compressibility, which are beyond the<br />

experimental accuracy. This results justifies previous determinations of the matrix compressibility<br />

using the original model.<br />

In order to investigate the influence of the matrix structure our experiments were extended<br />

to dye-doped rare gas mixtures of argon <strong>and</strong> xenon. In these matrices the transition,<br />

from polycrystalline pure rare gas matrices to amorphous mixed rare gas solids, can be monitored<br />

by varying the composition ratio. At the transition to the amorphous state the quasihomogeneous<br />

line width revealed no increase, which can be attributed to the high degree of<br />

structural disorder already present in the condensed pure rare gas matrices.<br />

Structural transitions can also be studied in annealing experiments. Therefore the next<br />

step in our investigations will focus on the changes of the inhomogeneous <strong>and</strong> quasi-homogeneous<br />

widths, occurring during thermal cycling experiments in rare gases <strong>and</strong> – turning to<br />

more complex systems – in dye-doped amorphous water [42]. The investigation of water<br />

matrices promises, in addition to the structural aspects, to shed some light on the importance<br />

of electrostatic interactions due to the polarity of the water molecules.<br />

194


References<br />

Acknowledgements<br />

The authors gratefully acknowledge many elucidating discussions with Dr. L. Kador.<br />

References<br />

1. E. Shpol’skǐ: Sov. Phys. Usp., 3, 372 (1960)<br />

2. A. Markoff: Wahrscheinlichkeitsrechnung, Teubner, Leipzig, (1912)<br />

3. A. Stoneham: Rev. Mod. Phys., 41, 82 (1969)<br />

4. B. Laird, J. Skinner: J. Chem. Phys., 90, 3274 (1989)<br />

5. A. Ellervee, R. Jaaniso, J. Kikas, A. Suisalu,V. Shcherbalov: Chem. Phys. Lett., 176, 472 (1991)<br />

6. B. Kharlamov, R. Personov, L. Bykovskaja: Opt. Commun., 12, 191 (1974)<br />

7. A. Gorokhovskií, R. Kaarli, L. Rebane: JETP Lett., 20, 216 (1974)<br />

8. J. Friedrich, D. Haarer: Angew. Chem. Int. Ed. Engl., 23, 113 (1984)<br />

9. L. Kador: PhD thesis, Universität Bayreuth, (1988)<br />

10. R. Altmann: PhD thesis, Universität Bayreuth, (1992)<br />

11. T. Sesselmann, W. Richter, D. Haarer, H. Morawitz: Phys. Rev. B, 36, 7601 (1987)<br />

12. S. Reul, W. Richter, D. Haarer: J. Non-Cryst. Sol., 145, 149 (1992)<br />

13. G. Pollak: Rev. Mod. Phys., 36, 748 (1964)<br />

14. M. Klein, J. Venables (eds): Rare Gas Solids, Academic Press, London, (1977)<br />

15. L. Kador: J. Chem. Phys., 95, 5574 (1991)<br />

16. H. Sevian, J. Skinner: Theor. Chim. Acta, 82, 29 (1992)<br />

17. L. Kador: J. Chem. Phys., 99, 7 (1993)<br />

18. L. Kador, P. Geissinger: Mol. Cryst. Liq. Cryst., 252, 213 (1994)<br />

19. S. Simon,V. Dobrosavljevic, R. Stratt: J. Chem. Phys., 93, 2640 (1990)<br />

20. P. Geissinger: PhD thesis, Universität Bayreuth, (1994)<br />

21. P. Geissinger, L. Kador, D. Haarer: Phys. Rev. B, 53(8), 4356 (1996)<br />

22. M. Sapozhnikov: J. Chem. Phys., 68, 2352 (1978)<br />

23. C. Tilford, C. Swenson: Phys. Rev. B, 5, 719 (1972)<br />

24. G. Horton: Am. J. Phys., 36, 93 (1968)<br />

25. F. London: Z. Phys., 63, 245 (1930)<br />

26. N. Ashcroft, N. Mermin: Solid State Physics, Saunders College, (1976)<br />

27. N. Bernardes: Phys. Rev., 112, 1534 (1958)<br />

28. L<strong>and</strong>olt-Börnstein: Zahlenwerte und Funktionen, in: A. Eucken (ed.): I. B<strong>and</strong>: Atom- und Molekülphysik,<br />

p. 399, Springer, (1950)<br />

29. S. I. Kovalenko, E. I. Indan, A. A. Khudoteplaya: Phys. Stat. Sol. (a), 13, 235 (1972)<br />

30. H. Menges, H. Löhneysen: J. Low Temp. Phys., 84, 237 (1991)<br />

31. M. Schneider, A. Rahman, I. Schuller: Phys. Rev. B, 34, 1802 (1986)<br />

32. P. Geissinger, D. Haarer: Chem. Phys. Lett., 197, 175 (1992)<br />

33. L. Bajema, M. Gouterman, B. Meyer: J. Mol. Spectry., 27, 225 (1968)<br />

34. K. Takeda, Y. Osamu, H. Suga: J. Phys. Chem., 99, 1602 (1995)<br />

35. V. Bondybey, J. English: J. Am. Chem. Soc., 101, 3446 (1979)<br />

36. P. Geissinger, W. Richter, D. Haarer: J. Lumin., 56, 109 (1993)<br />

195


10 Matrix-Molecule Interaction in Dye-Doped Rare Gas Solids<br />

37. P. Fitch, A. Haynam, D. Levy: J. Chem. Phys., 73(3), 1064 (1980)<br />

38. M. Anderson, C. Swenson: J. Phys. Chem. Solids, 36, 145 (1975)<br />

39. M. Loistl, F. Baumann: Z. Phys. B – Condensed Matter, 82, 199 (1991)<br />

40. T. Giering, P. Geissinger, L. Kador, D. Haarer: J. Lumin., 64, 245 (1995)<br />

41. L. Kador, G. Schulte, D. Haarer: J. Phys. Chem., 90, 1264 (1986)<br />

42. T. Giering, D. Haarer: J. Lumin., 66 & 67, 299 (1996)<br />

196


II<br />

Mainly Micelles, Polymers, <strong>and</strong> Liquid Crystals<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong><br />

<strong>Properties</strong> of Surfactant Solutions<br />

Heinz Hoffmann<br />

11.1 General behaviour of surfactants<br />

Surfactants consists of molecules with a hydrophobic <strong>and</strong> a hydrophilic part. Due to this amphiphilic<br />

nature these molecules adsorb from aqueous solutions onto interfaces, as is expressed<br />

in their name [1]. The molecules are densely packed in these adsorbed monolayers.<br />

In the aqueous bulk phase the surfactant molecules assemble above a characteristic concentration,<br />

called cmc, into micellar structures, which can be understood as interfaces in the<br />

bulk solution. The driving force for the adsorption <strong>and</strong> the aggregation is the same for both<br />

processes <strong>and</strong> is given by the hydrophobic interaction [2].<br />

The molecular packing of the surfactant molecules in films <strong>and</strong> micelles is mainly determined<br />

by the area a which a surfactant molecule requires at the interface. In both, films<br />

<strong>and</strong> micelles, the molecules will occupy about the same area [3]. If the area a is larger than<br />

the cross section a 0 of the hydrocarbon chain in its equilibrium conformation the interface<br />

of the micelle will be curved towards the hydrocarbon core. If it is the same the interface<br />

will be flat on a local scale <strong>and</strong> if it is smaller the interface will be curved the other way<br />

around. The importance of the area a for the structures of micelles was recognized by Tanford<br />

[2] <strong>and</strong> later by Ninham <strong>and</strong> co-workers [4].<br />

From simple geometrical considerations it follows that the shape of a micelle can be<br />

expressed by the packing parameter P = ar/v, where r is the length of the hydrocarbon chain<br />

<strong>and</strong> v its volume. The packing parameter varies from 3 for globules to 1 for bilayers. As a<br />

result of their packing parameter single-chain surfactants often form globular micelles <strong>and</strong><br />

double-chain surfactants tend to form bilayer structures.<br />

In a film at a macroscopic interface the area a of the molecule controls the thickness of<br />

the film <strong>and</strong> the order parameter of the hydrocarbon chains inside the film. If a>a 0 the film<br />

is thinner than the length of the surfactant chain <strong>and</strong> water borders directly on the hydrocarbon<br />

chains of the surfactant resulting in a large interfacial tension. If a &a 0 the film looks like<br />

one half of a bilayer of a real membrane. In this case the water is not in contact with the hydrocarbon<br />

chains <strong>and</strong> the interface has a low interfacial tension. In this qualitative model the<br />

parameter a controls the curvature of a micelle <strong>and</strong> the interfacial tension at an interface.<br />

Note that the differently shaped micelles are present only in surfactant solutions.<br />

When the micellar solutions are in contact with hydrocarbon the micelles will solubilize hy-<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

199


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

drocarbon <strong>and</strong> will be transformed into microemulsion droplets. Then the curvature of the<br />

droplets correlates with the interfacial tension [5]. The minimum of this interfacial tension<br />

corresponds to the maximum of the solubilization capacity of the surfactants [6, 7]. This relation<br />

has been used extensively to optimize surfactant systems for tertiary oil recovery. On<br />

the basis of theoretical work the values of interfacial tension at the minimum seems to depend<br />

on the bending constants of the films [8]. A high (low) minimum of interfacial tension<br />

would mean a high (low) bending constant.<br />

11.2 From globular micelles towards bilayers<br />

Amphiphilic substances with an extremely small hydrophilic group, for example aliphatic alcohols<br />

with intermediate chain lengths, are called cosurfactants. Normally cosurfactants do<br />

not form micelles in aqueous solutions but they are surface active <strong>and</strong> make up mixed micelles<br />

with normal surfactants. In cosurfactant/surfactant mixtures the mean area a per head<br />

group must vary smoothly with the mole fraction of the cosurfactant, between the value for<br />

the pure surfactant a s <strong>and</strong> the value for the cosurfactant a c . For most uncharged surfactants<br />

the value a s corresponds to a situation where normal globular micelles are formed while a c<br />

corresponds to a situation where inverse micellar structures of the L 2 phase exist. In principle,<br />

by varying the mole fraction X c of the cosurfactant in aqueous solutions, we should encounter<br />

all the types of micellar structures <strong>and</strong> mesophases that can possibly exist in surfactant<br />

solutions. While the area per head group can vary smoothly with X c , it is evident that<br />

the curvature of the micellar interface cannot vary continuously from strongly convex to<br />

strongly concave for micelles of the same type. The system has to switch to differently<br />

shaped micelles <strong>and</strong> mesophases if X c is varied.<br />

There is evidence that the systems use defects in mesophases to adjust the mean curvature<br />

that is set by X c . Strain in a mixed system can also be avoided by concentration fluctuations.<br />

The system can form structures with two different X c values that are in equilibrium,<br />

one with a lower X c <strong>and</strong> one with a higher X c than the average value. For all these reasons<br />

we find a rich variety of micellar structures <strong>and</strong> phases if X c is varied <strong>and</strong> the total concentration<br />

is kept constant. The diversity which is encountered in such situations is shown in<br />

Fig. 11.1 which represents the phase diagram of the ternary system tetradecyldimethylaminoxide/hexanol/water<br />

(C 14 DMAO/C 6 OH/H 2 O) at the water-rich corner. With increasing X c<br />

we observe the six visibly different single-phase regions L 1 (micellar solution), L 1<br />

*,L al ,L 3<br />

*<br />

(vesicle phases), L ah (lamellar phase with flat lamellae), <strong>and</strong> L 3 (bicontinuous sponge phase)<br />

[9]. All these phases have distinct properties. Furthermore there is a change in shape of the<br />

micelles in the L 1 phase from globules to rods. This transition is, however, not visible to the<br />

unaided eye. The phases L 1 ,L 1<br />

*,L 3<br />

*, <strong>and</strong> L 3 are optically isotropic while the phases L al <strong>and</strong><br />

L ah are birefringent (the L al phase is only weekly birefringent in most cases).<br />

The general features of this diagram are typical for such ternary systems <strong>and</strong> are practically<br />

independent of the chain lengths of the surfactant <strong>and</strong> the cosurfactant. Figure 11.1<br />

200


11.2 From globular micelles towards bilayers<br />

c C6 OH mM<br />

400<br />

2Φ<br />

300<br />

L 3<br />

2Φ<br />

200<br />

L αh<br />

L αl-h<br />

L<br />

100<br />

*<br />

αl<br />

L 3<br />

2Φ<br />

* L<br />

L 1<br />

1<br />

0<br />

0 50 100 150 200<br />

Figure 11.1: Section of the phase diagram of the ternary system C 14 DMAO/C 6 OH/H 2 O in the waterrich<br />

corner at 25 8C. For details see Section 11.4.1 <strong>and</strong> Fig. 11.11.<br />

c C 14 DMAO /mM 201<br />

shows that the phases with lamellar structures are already formed at total surfactant concentrations<br />

around 1 wt%. These phases, except the L al <strong>and</strong> L ah phase, which probably do not<br />

develop a phase boundary, are separated from each other by two-phase regions. Three phase<br />

regions can also exist in the phase diagram although all phases consist of about 99 wt%<br />

water. Therefore <strong>and</strong> because of the very similar densities of the different phases the phase<br />

separation takes a long time <strong>and</strong> hence the determination of such phase diagrams is dificult.<br />

Furthermore the refractive indices of the phases are almost the same, which makes it difficult<br />

to distinguish one <strong>and</strong> two-phase regions unambiguously as the single phases are often<br />

slightly turbid. Inspite of that most two-phase systems separate macroscopically into two<br />

phases after a sufficiently long time.<br />

The phase diagram of Fig. 11.1 serves as an example of what happens if the radius of<br />

curvature varies continuously with the mixing ratio X c at a constant surfactant concentration.<br />

For the classic non-ionic alkyl polyglycolethers (C 12 E 5 ) this can be done by varying the temperature<br />

[10]. For ionic surfactants this can be done by increasing the salinity [11], by mixing<br />

a cationic with an anionic surfactant, or vice versa [12]. In all these different situations<br />

we can expect to observe the same sequence of phases. The phase diagram in Fig. 11.1 was<br />

established on the basis of visual observation of samples. To check for birefringence the<br />

samples were viewed between crossed polarizers.<br />

Before we come to the microstructures there are several facts that are worth emphasizing.<br />

Note that the phase boundaries between the different phases are given by more or less<br />

straight lines, which means that the mixing ratio for the micelles is constant at the phase<br />

boundaries <strong>and</strong> does not depend on the concentration. Notice also that the equimolar mixtures<br />

of surfactant <strong>and</strong> cosurfactant are in the middle of the wide L a region. The combination<br />

therefore acts like a real double-chain surfactant. The phases from L 1<br />

* to L ah are actually<br />

subphases of L a , as has been demonstrated by freeze fracture electron micrographs. They<br />

consist of bilayer-type structures [13] but the topology of the bilayers is quite different.


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

11.3 Viscoelastic solutions with entangled rods<br />

11.3.1 General behaviour<br />

Surfactant solutions with globular micelles are generally Newtonian liquids with a low viscosity<br />

which increases linearly with the volume fraction F of the particles according to Einstein’s<br />

law<br />

ˆ s …1 ‡ 2:5F† :<br />

…1†<br />

Here Z s is the viscosity of the pure solvent. F is an effective volume fraction which<br />

also takes into account the hydration of the molecules. It can be up to three times the true<br />

volume fraction. But also in this case the viscosity of a 10 wt% surfactant solution is only<br />

about twice as high as the solvent viscosity. The same is true if anisometric micelles are present<br />

in the solutions as long as their rotational volumes do not overlap.<br />

On the other h<strong>and</strong>, many surfactant solutions are highly viscous even at low concentrations<br />

in the range of 1 wt%. From this observation it can be concluded that the micellar<br />

aggregates in these solutions must organize themselves into some kind of a supermolecular<br />

network. The viscosity of such systems strongly depends on parameters like surfactant concentration,<br />

ionic strength, temperature, or concentration of additives. The solutions usually<br />

also have elastic properties because the zero shear viscosity Z 0 is the result of a transient<br />

network of entangled rods that is characterized by a shear modulus G 0 <strong>and</strong> a structural relaxation<br />

time t according to<br />

0 ˆ G 0 t :<br />

…2†<br />

In this case the shear modulus is determined by the particle density n of entanglement<br />

points<br />

G 0 ˆ n kT :<br />

…3†<br />

The networks of entangled cylindrical micelles could be made visible by cryo-electron<br />

microscopy by Talmon et al. [14]. These pictures clearly show the shape <strong>and</strong> the persistence<br />

length of the rods but they do not reveal their dynamic behaviour. According to Eq. 2 the<br />

viscosity is the result of structure <strong>and</strong> dynamic behaviour, i. e. the structural relaxation time<br />

constant t which strongly depends on many parameters <strong>and</strong> can vary by many orders of<br />

magnitude for the same surfactant, if for instance the counterion concentration is changed<br />

[15]. This is shown in Fig. 11.2 where Z 0 for several cetylpyridiniumchloride (CPyCl) concentrations<br />

is plotted vs. the sodiumsalicylate (NaSal) concentrations. With increasing<br />

amounts of NaSal the viscosity passes a maximum, then a minimum, <strong>and</strong> finally a second<br />

maximum. This behaviour is due to a corresponding dependence of t on the NaSal concentration<br />

while G 0 is independent of this parameter for a constant CPyCl concentration. Simi-<br />

202


11.3 Viscoelastic solutions with entangled rods<br />

10 6 30mM CPyCl<br />

10 5<br />

60mM CPyCl<br />

100mM CPyCl<br />

η o /mPas<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 1 10 2 10 3<br />

c NaSal /mM<br />

Figure 11.2: Double logarithmic plot of the zero shear viscosity Z 0 for three different CPyCl concentrations<br />

vs. the concentration of added NaSal at 25 8C.<br />

lar results have been observed for many systems by various groups. The viscosity of a 1 wt%<br />

surfactant solution varies from the solvent viscosity <strong>and</strong> up to a 10 6 times higher value.<br />

Figure 11.3 shows a schematic sketch of a network of rod-like micelles. It is generally<br />

assumed that the crosslinks of the network that cause the elastic behaviour are entanglements<br />

[16]. However, this is not always true. Adhesive contacts between the micelles or a<br />

transient branching point, like a many armed disc-like micelle, can act as crosslinks [17,<br />

18]. Some experimental evidence for both possibilities have recently been observed. The entangled<br />

thread or worm-like micelles have typical persistence lengths between some 100 to<br />

1000 Å <strong>and</strong> they may, or may not, be fused together at the entanglement points.<br />

Figure 11.3: Schematic drawing of an entanglement network of long cylindrical micelles. Note the different<br />

length scales: k is the mesh size, l the mean distance between two knots, <strong>and</strong> m the contour length<br />

between two knots.<br />

The cylindrical micelles have an equilibrium network conformation. They constantly<br />

undergo translational <strong>and</strong> rotational diffusion processes. They also break <strong>and</strong> reform. If the<br />

network is deformed or the equilibrium conditions are suddenly changed it will take some<br />

time until the system reaches equilibrium again. If a shear stress p 21 is applied to a network<br />

203


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

solution in a much shorter time than the equilibration time, the solution behaves like a soft<br />

material that obeys Hooke’s law,<br />

p 21 ˆ G 0 g ;<br />

…4†<br />

with the spring constant G 0 <strong>and</strong> the deformation g. On the other h<strong>and</strong>, if the stress is applied<br />

for a longer time the system flows like a Newtonian liquid,<br />

p 21 ˆ 0 _g ;<br />

…5†<br />

with the zero shear viscosity Z 0 <strong>and</strong> the shear rate _g. A mechanical model for a viscoelastic<br />

fluid is the so-called Maxwell model which consists of a spring with the constant G 0 <strong>and</strong> a<br />

dashpot with the viscosity Z 0 . The zero shear viscosity of such a system can be expressed by<br />

the product of G 0 <strong>and</strong> t according to Eq. 2.<br />

Both quantities can be determined by oscillating rheological measurements [19].<br />

Many viscoelastic surfactant solutions can be described in a large frequency range by the<br />

Maxwell model with a single shear modulus G 0 <strong>and</strong> a single structural relaxation time constant<br />

t [20]. This is shown in Fig. 11.4 a. However, there are surfactant solutions which behave<br />

in a completely different manner as can be seen from Fig. 11.4b. The systems do not<br />

show a frequency-independent plateau value of the modulus <strong>and</strong> the viscosity cannot be expressed<br />

by a single G 0 or a single t value. In such situations the shear stress after a rapid deformation<br />

relaxes exponentially [21],<br />

<br />

t<br />

a<br />

p 21 ˆ ^p 21 exp : …6†<br />

t<br />

Figure 11.4a shows that the loss modulus G@ increases again with the frequency f.<br />

This increase can be related to Rouse modes of the cylindrical micelles. On the basis of a<br />

10 3<br />

10 1<br />

10 2<br />

G´,G´´ / Pa<br />

10 2<br />

10 0<br />

G´<br />

10 -1<br />

G´´<br />

|η * |<br />

10 -2<br />

10 -3 10 -2 10 -1 10 0 10 1<br />

f/Hz<br />

Figure 11.4a: Double logarithmic plot of storage modulus G', loss modulus G@, <strong>and</strong> complex viscosity<br />

|Z*| vs. frequency f for a solution with 100 mM CPyCl <strong>and</strong> 60 mM NaSal at 25 8C. The solution behaves<br />

like a Maxwell fluid with a single shear modulus G 0 <strong>and</strong> a single structural relaxation time t.<br />

204<br />

10 1<br />

10 0<br />

10 -1<br />

|η * |/Pas


11.3 Viscoelastic solutions with entangled rods<br />

G´, G´´ / Pa<br />

10 2<br />

10 1<br />

10 0<br />

G´<br />

G´´<br />

|η * |<br />

10 -1<br />

10 -3 10 -2 10 -1 10 0 10 1<br />

f/Hz<br />

Figure 11.4b: The same plot for a solution with 80 mM C 14 DMAO, 20mMSDS , <strong>and</strong> 55 mM C 6 OH.<br />

The solution does not behave like a Maxwell fluid. Note the differences to Fig. 11.4a: G@ does not pass<br />

over a maximum, G' does not show a plateau value but increases with f after the intersection with G@<br />

with a constant slope of 0.25.<br />

10 2<br />

10 1<br />

10 0<br />

10 -1<br />

|η * |/Pas<br />

theoretical model [22] the minimum value of G@ can be expressed by the storage modulus<br />

G', the entanglement length l e <strong>and</strong> the contour length l k of the cylindrical micelles according<br />

to<br />

G 00 min ˆ G0 l e<br />

l k<br />

:<br />

…7†<br />

11.3.2 Viscoelastic systems<br />

Rod-like micelles from ionic surfactants are usually formed at high ionic strengths with<br />

strongly binding, or hydrophobic counterions, or with large hydrophobic groups, like double-chain<br />

or perfluoro surfactants [4]. Solutions of such surfactants become highly viscous<br />

<strong>and</strong> viscoelastic with increasing concentration. In Fig. 11.5 some results are shown as double<br />

logarithmic plots of viscosity vs. concentration [23–25] in spite of the different chemistry<br />

of the surfactants. All these surfactants show a concentration region where the slopes<br />

are the same <strong>and</strong> in the range of 8.5, which is very high. The viscosity starts to rise<br />

abruptly at an overlap concentration c* <strong>and</strong> follows the scaling law within a small transition<br />

concentration<br />

0 /<br />

c x<br />

; …8†<br />

c <br />

where x is about 8.5 ± 0.5. The exponent must therefore be controlled by the electrostatic interaction<br />

in the solutions <strong>and</strong> is much bigger than the value of 4.5 ± 0.5 expected for large<br />

polymer molecules, which do not change their size with increasing concentration. From the<br />

large exponent it can be concluded that the rod-like micelles continue to grow with concentration<br />

above c*. This was postulated by McKintosh et al. [26].<br />

205


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

η ο /mPas<br />

10 6 Lec / C 14 DMAO / SDS<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

C 8 F 17 SO 3 NEt 4<br />

C 16 TMASal<br />

C 16 PyCl + NaSal<br />

C 16 C 8 DMABr<br />

10 0<br />

10 0 10 1 10 2<br />

c/mM<br />

Figure 11.5: Double logarithmic plot of zero shear viscosity Z 0 vs. concentration c for several solutions<br />

of charged surfactants. Note that all different systems show the same power law exponent within a limited<br />

concentration range above c*.<br />

Figure 11.6 shows logarithmic plots of Z 0 vs. c for the system CPyCl + NaSal [27]. It<br />

has been shown that this system can be described by the Maxwell model with one G 0 <strong>and</strong><br />

one t value above the first viscosity maximum. In this range the complex behaviour is due<br />

to the concentration dependence of t, while G 0 steadily increases with concentration.<br />

Furthermore, it was found that t is controlled by the kinetics of breaking <strong>and</strong> reforming of<br />

the micelles [28]. These processes are faster than the reptation of the rods under the given<br />

experimental conditions.<br />

As already mentioned, the same is true for the system with a constant CPyCl concentration<br />

with increasing amounts of NaSal, which is shown in Fig. 11.2. The dependence of<br />

the viscosity on the counterion concentration is controlled by a corresponding behaviour of<br />

t, while G 0 is independent of the NaSal concentration. The viscosity is therefore a result of<br />

the dynamics of the system <strong>and</strong> not of its structure. This can be proved by cryo-electron microscopy<br />

[14]. The electron micrographs show no differences between the structure of the<br />

10 5 CPyCl + NaSal<br />

CPyCl + 0.3MNaSal<br />

10 4<br />

uncharged<br />

η ο /mPas<br />

10 3<br />

10 2<br />

10 1<br />

10 0 10 1 10 2 10 3<br />

c/mM<br />

Figure 11.6: Double logarithmic plot of zero shear viscosity Z 0 vs. concentration c of CPyCl + NaSal<br />

for solutions at the first viscosity maximum (o), at the second maximum (p), <strong>and</strong> at the minimum (_)<br />

at 25 8C (Fig. 11.2).<br />

206


11.3 Viscoelastic solutions with entangled rods<br />

micelles for all four concentration regions. This is very remarkable because the micelles are<br />

differently charged in the concentration regions. Below the first maximum of the viscosity<br />

they are highly <strong>and</strong> positively charged, at the minimum they are completely neutral, <strong>and</strong> at<br />

the second maximum they carry a negative charge.<br />

The power law behaviour in the different concentration regions is also completely different,<br />

as can be seen from Fig. 11.6. The exponent at the first maximum is 8, at the minimum<br />

1.3 <strong>and</strong> at the second maximum 2.5. No theoretical explanation is available for the extremely<br />

low exponent of 1.3 which has also been observed for systems which completely<br />

differ in chemistry <strong>and</strong> conditions. Therefore the exponent seems to represent a general behaviour<br />

determined by fundamental physics.<br />

Figure 11.7 shows logarithmic plots of Z 0 vs. c for zwitterionic alkyldimethylaminoxide<br />

surfactants (C x DMAO) [17]. The data again show a power law behaviour over extended<br />

concentration regions. Some curves show a break indicating that also uncharged systems can<br />

switch the relaxation mechanism. At the lowest concentration region, where a power law behaviour<br />

is observed, the slope is the highest <strong>and</strong> almost the same as for the observed polymers.<br />

This is somewhat surprising because it could be expected that the length of the micellar<br />

rods increases with increasing concentration, which should lead to a higher exponent of<br />

the power law. It is therefore likely that the dynamics of the systems is already influenced<br />

by kinetic processes under these conditions.<br />

η o /mPas<br />

10 6 C 14 DMAO<br />

10 5 C 16 DMAO<br />

OleylDMAO<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 0 10 1 10 2 10 3<br />

c/mM<br />

Figure 11.7: Double logarithmic plots of zero shear viscosity Z 0 vs. concentration c for solutions of alkyldimethylaminoxide<br />

surfactants of various chain lengths at 25 8C.<br />

For higher concentrations a smaller exponent is observed. Obviously a new mechanism<br />

is operating under these conditions which is more effective in reducing a stress than the mechanism<br />

in the low concentration region. Generally a mechanism can only become dominant<br />

with increasing concentrations if it is faster than the one at low concentrations. The moduli<br />

increase with the same exponent in the various concentration regions. Furthermore, Fig. 11.7<br />

shows that the absolute value of Z 0 for C 16 DMAO <strong>and</strong> ODMAO (Oleyl) differ by an order<br />

of magnitude even though the slope is the same in the high concentration region. This is due<br />

to the fact that in the kinetically controlled region the breaking of the micelles depends very<br />

much on the chain length of the surfactant.<br />

207


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

According to the theory of micelle formation, addition of cosurfactants to surfactant<br />

solutions leads to a transition of spherical micelles to rods or to a growth of rod-like micelles<br />

[29]. As a consequence the viscosity of a surfactant solution increases with increasing<br />

cosurfactant concentration. This is shown in Fig. 11.8 for the system of 100 mM C 14 DMAO<br />

with various cosurfactants. The viscosity first increases <strong>and</strong> then passes a maximum. The situation<br />

is similar as shown in Fig. 11.2. It is likely that the micelles still grow steadily with<br />

increasing cosurfactant concentration but the system switches from one mechanism on the<br />

left side of the maximum to a faster mechanism on the right side. The reason for the switch<br />

is probably that the rods become more flexible with increasing cosurfactant concentration.<br />

The different mechanisms become obvious in Fig. 11.9 where log(Z 0 ) is plotted vs. log (c)<br />

for C 14 DMAO with different amounts of decanol (C 10 OH), which is completely solubilized<br />

in the micelles due to its poor solubility in water. The plot shows that the slopes of mixtures<br />

at the left side of the maximum are the same while the mixture with the highest cosurfactant/surfactant<br />

ratio has the lowest slope of 1.3. This value is equal to the one for CPySal at<br />

the viscosity minimum. Systems with similar low slopes from the literature [30] are shown<br />

η o /mPas<br />

10 3 hexanol<br />

octanol<br />

decanol<br />

10 2<br />

lecithin<br />

10 1<br />

10 0<br />

0 10 20 30 40 50 60<br />

c/mM<br />

Figure 11.8: Semilogarithmic plot of zero shear viscosity Z 0 of a 100 mM solution of C 14 DMAO vs.<br />

the concentration c of added cosurfactants at 25 8C. Note that all curves pass a maximum.<br />

10 4 TDMAO/C 10 OH=5:1<br />

TDMAO/C 10 OH=6.6:1<br />

10 3 TDMAO/C 10 OH=10:1<br />

TDMAO/C 10 OH=20:1<br />

TDMAO<br />

η o /mPas<br />

10 2<br />

10 1<br />

10 0<br />

10 1 10 2 10 3<br />

C 14 DMAO / mM<br />

Figure 11.9: Double logarithmic plot of zero shear viscosity Z 0 of a mixture of C 14 DMAO <strong>and</strong> C 10 OH<br />

with different molar ratios of cosurfactant/surfactant vs. the concentration of C 14 DMAO at 25 8C.<br />

208


11.3 Viscoelastic solutions with entangled rods<br />

η o /mPas<br />

10 3<br />

10 2<br />

CPyCl+NaSal (T=20 o C)<br />

C 14 DMAO/C 10 OH=5.1(T=25 o C)<br />

C 16 EO 7 (T=45 o C)<br />

CTAB+ NaClO 3<br />

10 1<br />

10 1 10 2 10 3<br />

c/mM<br />

Figure 11.10: Double logarithmic plot of zero shear viscosity Z 0 vs. the total surfactant concentration<br />

for several surfactant systems with the same power law exponent of 1.3.<br />

in Fig. 11.10. The chemistry <strong>and</strong> also the viscosity values for these systems are completely<br />

different, yet the slope is the same. The shear moduli for these systems are very similar for<br />

given surfactant concentrations <strong>and</strong> they also scale with the same exponent. The low exponent<br />

for the viscosity therefore comes about by the structural relaxation time which scales<br />

with an exponent of –1 according to<br />

t /<br />

c 1<br />

: …9†<br />

c <br />

11.3.3 Mechanisms for the different scaling behaviour<br />

All studied surfactant systems show the same qualitative behaviour. The viscosity rises<br />

abruptly at a characteristic concentration c* which is the lower the longer the chain length<br />

of the surfactant is. At the concentration c* the rotational volumes of rod-like micelles start<br />

to overlap <strong>and</strong> form a network. This network can be an entanglement network, as in polymer<br />

solutions, or the micelles can be fused together, or can be held together by adhesive contacts.<br />

All these types have been proposed <strong>and</strong> it is conceivable that they can really exist<br />

[31]. Theoretical treatments assume that the cylindrical micelles are worm-like <strong>and</strong> flexible.<br />

This can be the case for some of the presented systems, but certainly not for the binary surfactant<br />

systems. For example, C 16 DMAO <strong>and</strong> ODMAO have very low c* values <strong>and</strong> if the<br />

micelles would be flexible they would have to be coiled below c*. But both, electric birefringence<br />

<strong>and</strong> dynamic light scattering determinations, show that at c 7c* the lengths of the rods<br />

are comparable with the mean distances between them. Hence, the rods must be rather stiff<br />

with persistence lengths of some 1000 Å. Similar results have been obtained by electron micrographs.<br />

The abrupt increase of Z 0 at c* is difficult to underst<strong>and</strong> for stiff rods even taking<br />

into account further growth of the rods with increasing concentration. Many systems with<br />

rod-like particles show that the rotational time constant for the rods is very little affected<br />

around c* <strong>and</strong> the solutions do not become viscoelastic above c*. We therefore have to assume<br />

that other interactions than just hard core repulsion between the rods must exist which<br />

209


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

are responsible for the formation of the network at c*. It is conceivable that the rods form<br />

adhesive bonds or that they actually form a connected network of fused rods, as has been<br />

proposed by Cates [32]. In such situations two different types of networks have to be distinguished,<br />

namely saturated <strong>and</strong> interpenetrating networks. In the first case the mean distance<br />

between the knots or entanglement points is equivalent to the meshsize but in the second<br />

case this distance can be much larger than the meshsize between neighbouring rods.<br />

The viscosities above c* increase abruptly following the power law (Eq. 8) with an exponent<br />

x>3.5, while the exponent of the power law for the shear modulus is always about<br />

2.3. This behaviour has been treated in detail by Cates et al. [33]. The structural relaxation<br />

times are affected by both, reptation <strong>and</strong> bond breaking processes. Cates treats three different<br />

kinetic mechanisms. The first consists of the break of a rod with the formation of two<br />

new end caps. In the recombination step the rods have to collide at the ends in order to fuse<br />

into a new rod. In the second mechanism the end cap of one rod collides with a second rod<br />

<strong>and</strong> in a three-armed transition state a new rod <strong>and</strong> a new end cap is formed. In the third<br />

mechanism two rods collide <strong>and</strong> form two new rods through a four armed transition state. It<br />

is obvious that stress can release all three mechanisms. These mechanisms lead to somewhat<br />

different power laws for the kinetic time constant t according to<br />

t /<br />

c x<br />

: …10†<br />

c <br />

But in all cases x is between 1 <strong>and</strong> 2. Mechanism 3 is less likely in systems with low<br />

c* values <strong>and</strong> stiff rods. For this argument it is likely that mechanism 2 or 3 is effective in<br />

the more concentrated region of the pure C x DMAO solutions.<br />

For C 16 DMAO <strong>and</strong> ODMAO solutions the slope of the log (Z 0 )-log (c) plots suddenly<br />

changes at a characteristic concentration c**. For both regions the same scaling law for the<br />

shear modulus with an exponent of about 2.3 is found while the power exponent for the relaxation<br />

times changes from 1 to zero. From the constant exponent for G 0 it can be concluded<br />

that the structures in both concentration regions are the same. The change in the<br />

slope must therefore be due to a new mechanism which becomes effective above c**. The<br />

independence of t of the concentration makes it likely that in this region the dynamics are<br />

governed by a purely kinetically controlled mechanism <strong>and</strong> that a reptation process is no<br />

longer possible. This situation has not yet been treated theoretically. Cates mentioned, however,<br />

that there might be situations where the reptation loses its importance. The more effective<br />

mechanism in this range could be the bond interchange mechanism.<br />

For the C 14 DMAO solutions with C 10 OH <strong>and</strong> the CPySal system at the minimum<br />

viscosity the extremely low power law exponent of 1.3 for the viscosity <strong>and</strong> an exponent of<br />

–1 for the structural relaxation times are found. A detailed explanation for this behaviour<br />

which has also been described by other authors [30] has not yet been given. The explanation<br />

for this behaviour could be that the cylindrical micelles for systems with such low exponents<br />

are very flexible. In such a situation the persistence length would be much shorter than the<br />

contour length between two neighbouring entanglement points. Furthermore, the persistence<br />

length should be independent of the concentration. The diffusion of the rods can therefore<br />

be described by a constant diffusion coefficient D. For two arms to collide they have to diffuse<br />

a distance x <strong>and</strong> for two neighbouring rods to undergo a bond exchange process they<br />

have to diffuse at least the average distance x between two arms. The time constant t D for<br />

210


11.4 Viscoelastic solutions with multilamellar vesicles<br />

the diffusion should be proportional to x 2 /D. Since the meshsize x decreases with the square<br />

root of the concentration one obtains for the structural relaxation time t the observed law<br />

t !1/D7c, which is identical with Eq. 9. We therefore conclude that for systems with the<br />

low exponent 1.3 the viscosity is governed by a diffusion-controlled bond interchange mechanism.<br />

The absolute values of Z 0 <strong>and</strong> t can still vary from system to system because the<br />

persistence length l p of the rods should depend on the particular conditions of the systems.<br />

With increasing chain length l p should decrease <strong>and</strong> D increase. For such situations we<br />

would expect to find the lowest activation energies for the viscosity.<br />

A similar mechanism could be based on the assumption of connected or fused threadlike<br />

micelles as crosslinks. These crosslinks could be visualized as disc-like micelles from<br />

which the rods extend. This means that the transient intermediate species in the various<br />

bond interchange mechanisms are now assumed to be stable. In this situation all end caps<br />

could be connected. The resulting network could be in the saturated or unsaturated state.<br />

The crosslink points could then slide along the thread-like micelles like a one-dimensional<br />

diffusion process with a concentration-independent diffusion coefficient. A knot can be dissolved<br />

if two network points meet on their r<strong>and</strong>om path. If the structural relaxation time is<br />

determined by this r<strong>and</strong>om movement a similar equation t !1/c can be derived. Both models<br />

can describe the low exponent of 1.3 for the scaling law for Z 0 <strong>and</strong> for both models reptation<br />

is no longer necessary for the release of stress. The mechanisms could probably be<br />

distinguished by the concentration dependence of the self-diffusion coefficients D s of the<br />

surfactant molecules. In a solution with a connected network a surfactant molecule should<br />

be in the same situation as in a L 3 phase for which it has been shown that D s is independent<br />

of the surfactant concentration [34]. As Kato et al. [35] showed that the D s values increase<br />

with concentration for a corresponding system with rods a diffusion limited bond interchange<br />

mechanism is more likely for the explanation of the scaling law of the structural relaxation<br />

time than the assumption of connected networks of thread-like micelles.<br />

11.4 Viscoelastic solutions with multilamellar vesicles<br />

11.4.1 The conditions for the existence of vesicles<br />

Thermodynamically stable vesicles have been found recently in many solutions with various<br />

types of surfactants [36–39]. Such vesicles occur especially in ternary systems of zwitterionic<br />

or non-ionic surfactants, aliphatic alcohols as cosurfactants, <strong>and</strong> water. As explained in<br />

detail in Section 11.4.2, this is due to the spontaneous curvature of the micellar surface<br />

which is continuously lowered with increasing amounts of the cosurfactant because of its<br />

very small headgroup area [3]. The systems try to come as close as possible to the spontaneous<br />

mean curvature without causing much bending energy by adjusting the two main curvatures<br />

on the micellar aggregates. As a consequence the micellar system undergoes several<br />

phase transitions with increasing cosurfactant concentration.<br />

211


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

The vesicle phase occurs within a wide range of the total surfactant <strong>and</strong> cosurfactant<br />

concentration but only within a small range of the molar ratio of cosurfactant/surfactant<br />

around 1 :1. If the surfactant bilayers are charged by the addition of an ionic surfactant the<br />

phase diagram becomes somewhat simpler because some mesophases are suppressed by the<br />

charge. As can be seen from the Fig. 11.11, the vesicle phase is still found under these conditions<br />

but it is shifted towards higher cosurfactant concentrations. Thermodynamically<br />

stable vesicles have also been found in ternary systems of non-ionic alkylpolyglycol surfactants,<br />

cosurfactants, <strong>and</strong> water [10] <strong>and</strong> also in the binary system of the double-chain surfactant<br />

didodecyldimethylammoniumbromide (DDABr) <strong>and</strong> water [42].<br />

600<br />

500<br />

400<br />

L 1<br />

+L 2<br />

L α<br />

L 1<br />

+L 3<br />

300<br />

L 3<br />

L 3<br />

L 1<br />

+L α<br />

L αh<br />

L 1<br />

*<br />

200<br />

L αl<br />

L * 1<br />

+L 1<br />

L*<br />

1<br />

100<br />

L 1<br />

+L*<br />

1 L 1<br />

L 1<br />

0,0 0,2 0,4 0,6 0,8 1,0 0<br />

X C 14 TMABr<br />

c C6 OH /mM<br />

Figure 11.11: Section of the phase diagram of the quaternary system 100 mM C 14 DMAO/C 14 TMABr/<br />

C 6 OH/H 2 Oat258C. For details see Refs. [9, 13, 40, 41].<br />

11.4.2 Freeze fracture electron microscopy<br />

The vesicles can be made visible by freeze fracture transmission electron microscopy (FF-<br />

TEM). Figure 11.12 shows the vesicles in a system of 90 mM C 14 DMAO, 10 mM tetradecyltrimethylammoniumbromide<br />

(C 14 TMABr), 220 mM n-hexanol (C 6 OH), <strong>and</strong> water. The<br />

cationic surfactant can also be replaced by the anionic surfactant SDS without causing a<br />

change of the vesicles or of the rheological properties. From this electron micrograph it is<br />

possible to recognize some general features which are of relevance for the properties of the<br />

212


11.4 Viscoelastic solutions with multilamellar vesicles<br />

Figure 11.12: Electron micrograph of vesicles in the system of 90 mM C 14 DMAO, 10 mM<br />

C 14 TMABr, 220 mM C 6 OH, <strong>and</strong> water (the bar represents 1 mm).<br />

systems. The vesicles have a rather high polydispersity; some seem to be rather small unilamellar<br />

vesicles, while others consist of about 10 bilayers. The interlamellar spacing is fairly<br />

uniform <strong>and</strong> is in the range of 800 Å. The vesicles are very densely packed <strong>and</strong> the whole<br />

volume of the system is completely filled with them. They have a spherical shape even<br />

though the outermost shell can have a radius of several 1000 Å. Some of them do not consist<br />

of concentric shells but have defects. The larger vesicles have typical sizes in the range<br />

of 1 mm <strong>and</strong> the wedges, which result from the dense packing, are completely filled with<br />

smaller vesicles. Thus, each vesicle is sitting in a cage from which they cannot escape by a<br />

simple diffusion process without deforming their shells. Therefore, the system must have<br />

viscoelastic properties.<br />

11.4.3 Rheological properties<br />

In Fig. 11.13 the viscoelastic properties of a vesicle phase of 90 mM C 14 DMAO, 10mM<br />

C 14 TMABr, 220 mM C 6 OH, <strong>and</strong> water are demonstrated by plots of the storage modulus G',<br />

the loss modulus G@, <strong>and</strong> the magnitude of the complex viscosity |Z*| as a function of the oscillation<br />

frequency f. G' is much larger than G@ <strong>and</strong> almost independent of f in the whole frequency<br />

range. The system behaves like a soft solid <strong>and</strong> must have a distinct yield stress. This<br />

can be seen from the plot (Fig. 11.14) of the shear rate _g as a function of the applied shear<br />

stress s. As can be seen from Fig. 11.15 both, shear modulus G 0 , the frequency-independent<br />

value of G', <strong>and</strong> yield stress s y, increase with increasing total surfactant concentration. Both<br />

quantities disappear at total concentrations below 1 wt%. This means that the vesicles are no<br />

longer densely packed <strong>and</strong> they can move around each other easily under shear flow. For<br />

higher concentrated systems s y varies linearly with G 0 <strong>and</strong> is about one tenth of G 0 . This<br />

means the vesicles must be deformed by about 10% before they can pass each other under<br />

shear. Similar results are obtained for the vesicles in the binary system of DDABr <strong>and</strong> water.<br />

213


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

10 2 G´<br />

G´´<br />

|η * |<br />

10 2<br />

G´, G´´ / Pa<br />

10 1<br />

10 1<br />

10 0<br />

|η*| / Pas<br />

10 0<br />

10 -1<br />

10 -2 10 -1 f/Hz<br />

10 0 10 1<br />

Figure 11.13: Double logarithmic plot of storage modulus G', loss modulus G@, <strong>and</strong> the magnitude of<br />

the complex viscosity |Z*| vs. frequency f for a vesicle phase of 90 mM C 14 DMAO, 10 mM<br />

C 14 TMABr, 220 mM C 6 OH, <strong>and</strong> water at 25 8C.<br />

4<br />

3<br />

σ/Pa<br />

2<br />

1<br />

0<br />

0,0 0,1 0,2 0,3 0,4 0,5<br />

γ<br />

.<br />

/s -1<br />

Figure 11.14: Plot of the shear rate _g vs. applied shear stress s for the same vesicle phase as in<br />

Fig. 11.13, showing a distinct yield stress value s y at 1.1 Pa.<br />

G´ / Pa<br />

50<br />

40<br />

30<br />

20<br />

10<br />

σ y<br />

G´<br />

5<br />

4<br />

3<br />

2<br />

1<br />

σ y /Pa<br />

0<br />

50 100 150 200<br />

c surfactant /mM<br />

0<br />

Figure 11.15: Plot of shear modulus G 0 <strong>and</strong> yield stress s y vs. surfactant concentration c surfactant for a<br />

vesicle phase of C 14 DMAO <strong>and</strong> C 14 TMABr with a molar ratio of 9 : 1 <strong>and</strong> C 6 OH at 25 8C.<br />

Figures 11.16 a <strong>and</strong> 11.16 b demonstrate the influence of charge density on the bilayers<br />

on G 0 . The modulus increases with increasing amounts of ionic surfactant <strong>and</strong> saturates at<br />

about 10 mol% of the ionic compound (Fig. 11.16 a). On addition of electrolyte the modulus<br />

decreases again linearly with the square root of the ionic strength (Fig. 11.16 b).<br />

214


11.4 Viscoelastic solutions with multilamellar vesicles<br />

25<br />

100 mM Surfactant (C 14 DMAO + C 14 TMABr), 220 mM C 6 OH<br />

G´ / Pa<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 5 10 15 20<br />

c C 14 TMABr<br />

Figure 11.16a: Plot of shear modulus G 0 vs. the concentration of the ionic surfactant C 14 TMABr for a<br />

vesicle phase with a total concentration of 100 mM C 14 DMAO + C 14 TMABr, <strong>and</strong> 220 mM C 6 OH at<br />

25 8C.<br />

G´(0.1 Hz) / Pa<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0,0 0,5 1,0 1,5 2,0 2,5 3,0 3,5 4,0<br />

Figure 11.16b: Plot of shear modulus G 0 vs. the square root of the concentration of added NaCl for a<br />

vesicle phase with 85 mM C 14 DMAO, 15mMC 14 TMABr, <strong>and</strong> 300 mM C 6 OH at 25 8C.<br />

(c NaCl /mM) 1/2 215<br />

The effect of the chain length of the surfactant compound can be seen from Fig. 11.17. For<br />

systems with the same concentrations the modulus increases with increasing chain length from<br />

C 10 to C 16 <strong>and</strong> decreases for C 18 . This means that the modulus is not only determined by the electrostatic<br />

repulsion between the bilayers but also depends on the thickness of the bilayers.<br />

The headgroup of the surfactant in the vesicle phase has only a moderate effect on<br />

rheological properties. For example, a vesicle phase of 90 mM C 12 E 6 ,10mMSDS, <strong>and</strong><br />

250 mM C 6 OH shows a very similar behaviour as the corresponding phase of 90 mM<br />

C 14 DMAO, 10mMC 14 TMABr, <strong>and</strong> 220 mM C 6 OH. The absolute values of G', G@, <strong>and</strong><br />

|Z*| are also very similar for both systems. The same can be found for systems where the<br />

concentration of the cosurfactant (Fig. 11.18) or the chain length of the cosurfactant is changed<br />

(Fig. 11.19). The figures show that the moduli <strong>and</strong> the yield stress increase slightly with<br />

the cosurfactant concentration <strong>and</strong> the chain length of the cosurfactant. But this increase is<br />

small in comparison with the strong dependence of these values on the concentration <strong>and</strong><br />

the chain length of the surfactant compounds. The temperature has only a very small effect<br />

on both, G 0 <strong>and</strong> s y , between 10 8C <strong>and</strong> 60 8C.<br />

An interesting effect is shown in Fig. 11.20 where the shear viscosity Z as a function of<br />

the shear rate _g <strong>and</strong> the magnitude of the complex viscosity |Z*| as a function of the angular


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

10 3 x=10 x=12<br />

x=14 x=16<br />

x=18<br />

G´ / Pa<br />

10 2<br />

10 1<br />

10 -2 10 -1 10 0 10 1<br />

f/Hz<br />

Figure 11.17: Double logarithmic plot of storage modulus G' vs. frequency f for vesicle phases of<br />

90 mM CxDMAO, 10mMC 14 TMABr, <strong>and</strong> 220 mM C 6 OH at 25 8C with various chain lengths x of<br />

the zwitterionic surfactant.<br />

Figure 11.18: Plot of storage modulus G', loss modulus G@, <strong>and</strong> yield stress s y against the cosurfactant<br />

concentration for a vesicle phase of 90 mM C 14 DMAO, 10mMSDS, <strong>and</strong> varying amounts of C 6 OH at<br />

25 8C.<br />

Figure 11.19: Double logarithmic plot of storage modulus G' vs. frequency f for a vesicle phase of<br />

90 mM C 14 DMAO, 10mMC 14 TMABr, <strong>and</strong> 160 mM C n OH for cosurfactants with various chain<br />

lengths at 25 8C.<br />

216


11.4 Viscoelastic solutions with multilamellar vesicles<br />

|η*|, η/Pas<br />

|η*|(ω)<br />

η(γ )<br />

10 2 90 mM C 12 E 6 ,10mMC 14 TMABr,<br />

280 mM C 6 OH<br />

|η*|(ω)<br />

η(γ )<br />

90 mM C 14 DMAO, 10 mM C 14 TMABr,<br />

10 1<br />

220 mM C 6 OH<br />

10 0<br />

10 -1 10 0 10 1 10 2 10 3<br />

Figure 11.20: Double logarithmic plot of shear viscosity Z vs. shear rate _g <strong>and</strong> of the magnitude of the<br />

complex viscosity |Z*| vs. angular frequency o for two different vesicle phases at 25 8C.<br />

frequency o are compared for two vesicle phases. The diagram shows the important difference<br />

to viscoelastic solutions of thread-like micelles [43]. They do not always fulfill the Cox-<br />

Merz rule, stating that for all values of _g the shear viscosity Z is equal to |Z*| at the corresponding<br />

value of o in the shear thinning region [44]. At low shear rates or frequencies both<br />

viscosities have the same value, while at higher shear rates Z (_g) is larger than |Z*|(o = _g). The<br />

curve for the zwitterionic system shows two breaks at characteristic shear rates. For shear<br />

rates above these characteristic values it is likely that the multilamellar vesicles undergo transformations<br />

to new structures as has recently been proposed by Roux et al. [45].<br />

11.4.4 Model for the shear modulus<br />

γ /s -1 , ω /rads -1 217<br />

In previous publications the shear modulus for the multilamellar phases was considered to<br />

be the result of the interactions of hard sphere particles [46–48]. In this picture each<br />

charged multilamellar vesicle is treated as a hard sphere. The theoretical treatment of the<br />

samples would then be similar to latex systems. The modulus of the systems depends on the<br />

chainlength of the surfactants that are used for the preparation of the systems if all other<br />

parameters like charge density, salinity, <strong>and</strong> concentration of surfactants <strong>and</strong> cosurfactants<br />

are kept constant. It can be argued that the differences of the moduli result from a change of<br />

the particle density of the vesicles. But these values are not known exactly. Systems with different<br />

chainlengths have similar conductivities which suggests that the particle density is<br />

also similar <strong>and</strong> therefore not responsible for the different shear moduli.<br />

Furthermore the birefringence looks the same, too. If the particle density of the vesicles<br />

decreases <strong>and</strong> if the mean size of the vesicles increases then the birefringence should increase.<br />

But this is not the case. The different moduli must therefore have a different origin. We propose,<br />

consequently, a different model for the explanation of the magnitude of the shear moduli. For<br />

the treatment of multilamellar vesicle phases <strong>and</strong> L a phases this model was proposed by E. v. d.


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

Linden [49]. To our knowledge the theory has not yet been applied to experimental results. E. v.<br />

d. Linden assumes that multilamellar vesicles (droplets) are deformed in shear flow from a<br />

spherical to an elliptical shape. Turning into the deformed state the energy of closed shells is<br />

shifted because their curvature as well as their interlamellar distance D are changed. Due to the<br />

interaction of the bilayers, expressed by the bulk compression modulus B, the inner shells are<br />

deformed <strong>and</strong> the total deformation energy E of the lamellar droplet gets minimized. Assuming<br />

that the volume of a droplet is not modified by the deformation, the surface A must increase.<br />

One can define an effective surface tension s eff = E/DA. E. v. d. Linden obtains:<br />

ef f ˆ 1<br />

2 …KB†1 = 2 ; …11†<br />

where K is the bulk rigidity which is correlated to the bilayer’s bending constant k by K = k/D.<br />

We can relate this effective surface tension to the shear modulus G of a vesicle with radius R.<br />

Using the identity G =2s eff /R yields:<br />

1 =2<br />

k<br />

G ˆ … KB†1<br />

= 2 D B<br />

ˆ<br />

R R<br />

: …12†<br />

Both, bulk compression modulus <strong>and</strong> bending constant, depend on the charge density<br />

of the bilayers <strong>and</strong> the shielding of the charges with excess salt.<br />

This means the theory of E. v. d. Linden results in a calculation of the geometrical<br />

average of the compression E B <strong>and</strong> bending energy E K per unit volume.<br />

The expression (Eq. 12) can be squared to<br />

G 2 ˆ<br />

<br />

k<br />

D B<br />

R 2 : …13†<br />

With n = R/D which denotes the number of bilayers in a vesicle we obtain:<br />

G 2 ˆ nk<br />

R 3 B ˆ E k E B :<br />

…14†<br />

Now we can try to find adequate expressions <strong>and</strong> values for the quantities B <strong>and</strong> K by<br />

other theories.<br />

For the bending constant as a function of the charge density we can use the expression<br />

(Eq. 15) that has been given by H. Lekkerkerker [50],<br />

K el ˆ kBT …q 1†…q ‡ 2†<br />

2pQk …q ‡ 1†q<br />

…15†<br />

p<br />

with q ˆ p 2 ‡ 1<br />

218<br />

<strong>and</strong> p ˆ 2pQjsj<br />

.<br />

ke


11.4 Viscoelastic solutions with multilamellar vesicles<br />

k B is Boltzmann’s constant, T the absolute temperature, Q the Bjerrum length, k the reciprocal<br />

Debye length, <strong>and</strong> s the surface charge density.<br />

For the bulk compression modulus we can use the expression that is often used to describe<br />

the interaction between two charged particles<br />

K ˆ f a f s n p k 2 d 2 V…d† ;<br />

…16†<br />

where V(d) is the energy of interaction between a pair of spherical particles<br />

V…d† ˆz2 e 2 <br />

exp…ka† 2<br />

exp… kd†<br />

; …17†<br />

4p" …1 ‡ ka† d<br />

d denotes the separation of particles <strong>and</strong> a their radius. f a <strong>and</strong> f s are numerical factors [51].<br />

There is a further possibility to get an expression for the compression modulus B.<br />

This quantity may be simply given by the osmotic pressure between the bilayers. According<br />

to a theory of Dubois et al. [52] we have calculated the osmotic pressure using the equation<br />

P ˆ c m kT :<br />

…18†<br />

where c m is the concentration of ion particles at the midplane between the bilayers, calculated<br />

by the Poisson-Boltzmann equation for the current conditions. In a previous paper we<br />

have discussed the possibility to identify the shear modulus with the osmotic pressure. This<br />

seems to be obvious because the osmotic pressure qualitatively increases like the shear modulus<br />

with increasing charge density of the bilayers <strong>and</strong> decreases on salt addition. But this<br />

attempt failed because the calculated values, which were in the order of several thous<strong>and</strong><br />

pascals, were too large. In the current context, however, it looks reasonable to identify the<br />

osmotic pressure with the compression modulus B from Eqs. 11 <strong>and</strong> 12.<br />

Now there is a concept which is appropriate to reproduce the characteristic features of<br />

our experimental results. These are the influence of the charge density <strong>and</strong> the influence of<br />

salinity on the shear modulus.<br />

The growth of the shear modulus with the charge density can mainly be attributed to<br />

the increase of the bending constant <strong>and</strong> the osmotic pressure with the charge density. For<br />

small salinities the decrease with the salinity seems to come from the pair potential V (d)<br />

whereas the linear decrease at higher salinities seems to come from the bending constants.<br />

However, we should keep in mind that for both changes the vesicular structures do not remain<br />

constant. Therefore we cannot give a complete quantitative interpretation of the experimental<br />

results <strong>and</strong> the data cannot be fitted precisely to an exact theoretical model.<br />

At the end of this paragraph we check whether Eq. 12 gives reasonable results for our<br />

primary data. If we assume d = 80 nm for the interlamellar distance, R = 0.5 mm for the radius<br />

of a vesicle, 10% for the charging degree (i. e. a surface charge density on the bilayers<br />

of e 0 /500 H 2 , where e 0 is the elementary charge), k = 0.56 k B T for the electrical contribution<br />

of the bending modulus (calculated according to Eq. 15) [47], <strong>and</strong> B = P = 3000 Pa for<br />

the compression modulus (calculated according to Eq. 18) [46] then Eq. 12 yields the shear<br />

modulus G&18 Pa. This value is very close to the experimental one <strong>and</strong> demonstrates that<br />

a correct approach for a theoretical description of the rheological properties of the solutions<br />

seems to be found.<br />

219


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

11.5 Ringing gels<br />

11.5.1 Introduction<br />

For a large variety of amphiphilic compounds cubic phases are a particular class of surfactant<br />

systems normally observed in the more concentrated region of the phase diagram [53].<br />

Such phases may occur in binary systems (surfactant <strong>and</strong> water) as well as in ternary or<br />

pseudo ternary systems, where usually a hydrocarbon is the third component. In addition<br />

they may contain a long alcohol chain as a fourth component (cosurfactant). Such a composition<br />

reminds one of that of microemulsions <strong>and</strong> for that reason they are also called microemulsion<br />

gels [54]. However, the denomination cubic phase is insofar more instructive since<br />

it relates directly to the structure of the corresponding systems.<br />

These phases are thermodynamically stable, optically isotropic, transparent, <strong>and</strong> highly<br />

viscous which distinguishes them easily from conventional L 1 or L 2 phases, i. e. micellar or<br />

reverse micellar phases. Often they possess a yield stress <strong>and</strong> exhibit elastic properties for<br />

not too large deformations [55]. In particular these elastic properties are responsible for the<br />

other name – ringing gels – which sometimes is used for these systems because samples of<br />

these phases usually show an acoustical resonance phenomenon (ringing sound) after being<br />

tapped with a soft object [56]. This effect is not necessarily associated with a cubic phase<br />

but quite frequently observed within this class of surfactant systems. The ringing phenomenon<br />

is observed for all the cubic phases studied by us.<br />

It might be mentioned here that such phases have been used already for pharmaceutical<br />

<strong>and</strong> cosmetical applications without having detailed structural information regarding the<br />

corresponding systems at that time [57]. More recently it has been claimed that cubic phases<br />

may play a key role in the fusion process of biological membranes [58, 59] <strong>and</strong> they are frequently<br />

formed by lipids, obtained from membrane extracts.<br />

The main perspective of our investigations was first to determine the detailed microstructure<br />

of the corresponding systems <strong>and</strong> then to relate them to the observed macroscopic<br />

properties. For that purpose two surfactant systems were chosen. Both have a cubic phase<br />

but at different locations in the phase diagram. In the first system – tetradecyldimethylaminoxide<br />

(C 14 DMAO)/hydrocarbon/H 2 O – the cubic phase is located between the isotropic L 1<br />

phase <strong>and</strong> the hexagonal phase at a constant surfactant/hydrocarbon ratio (Fig. 11.21 a) [55].<br />

In the second system – bis-(2-ethylhexyl)sulfosuccinate (AOT)/1-octanol/H 2 O – it is situated<br />

between the lamellar phase, at lower octanol content, <strong>and</strong> an isotropic L 2 phase, at higher<br />

octanol <strong>and</strong> lower water content, or a reverse hexagonal phase at higher AOT content<br />

(Fig. 11.21 b) [60].<br />

These systems are insofar similar as they both are next to an isotropic phase. Moreover<br />

a binary system containing triblock copolymers of the polyethyleneoxide/polypropyleneoxide/polyethyleneoxide<br />

(PEO/PPO/PEO) type has been studied for which we also<br />

have found a formation of a cubic phase with location in the phase diagram similar to the<br />

aminoxide case.<br />

220


11.5 Ringing gels<br />

Decane<br />

20<br />

5<br />

80<br />

T=25°C<br />

50<br />

1<br />

2<br />

3<br />

4<br />

50<br />

2<br />

+<br />

S<br />

80<br />

20<br />

H 1<br />

L α S<br />

G<br />

H L 2<br />

O<br />

1<br />

C 14<br />

DMAO<br />

0 20 N c 40 60 80 100<br />

Figure 11.21a: Phase diagram of the ternary system C 14 DMAO/decane/H 2 Oat258C. Isotropic water<br />

continuous phase (L 1 ), nematic phase (N c ), cubic phase (G), hexagonal phase (H 1 ), lamellar phase (L a ),<br />

crystals (S), other phase regions (1, 2, 3, 4, 5).<br />

Octanol<br />

20<br />

80<br />

50 50<br />

L 2<br />

80<br />

I 2<br />

D<br />

H AOT<br />

2<br />

O<br />

0 20 40 60 80 100<br />

L 1<br />

Figure 11.21b: Phase diagram of the ternary system AOT/1-octanol/H 2 Oat258C (in wt%). Isotropic<br />

water continuous phase (L 1 ), lamellar phase (D), (bicontinuous) cubic phase (I 2 ), isotropic oil continuous<br />

phase (L 2 ), reverse hexagonal phase (F), other phase regions without symbols.<br />

F<br />

20<br />

11.5.2 The aminoxide system<br />

For the aminoxide system investigations were done along a line of constant C 14 DMAO/decane<br />

ratio. Upon going along such a line close to the solubilization capacity of the surfactant one can<br />

cross from the L 1 phase into the cubic phase by increasing the surfactant concentration. Samples<br />

located on such a dilution line contain spherical aggregates of constant size. This was shown by<br />

means of static <strong>and</strong> dynamic light-scattering experiments as well as by SANS measurements<br />

[55, 61]. The interactions are very well described by a hard sphere model. In Fig. 11.22 ex-<br />

221


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

3000<br />

2000<br />

1<br />

I(q) in a.u.<br />

1000<br />

2<br />

3<br />

6 54<br />

0<br />

0,00 0,05 0,10 0,15 0,20 0,25<br />

qin1/Å<br />

Figure 11.22: SANS intensity curves as a function of the magnitude of the scattering vector q for samples<br />

of constant C 14 DMAO : decane weight ratio of 5.4 : 1 <strong>and</strong> different total concentrations of<br />

C 14 DMAO plus decane were 8 wt% for 1; 16 wt% for 2; 24 wt% for 3, 28 wt% for 4, 32 wt% for 5,<br />

37.1 wt% for 6. Samples 1–5 are in the L 1 phase whereas sample 6 is located in the cubic phase.<br />

perimental SANS curves for various total concentrations (constant weight ratio of<br />

C 14 DMAO :decane = 5.4 :1) are given which can be fitted curves in good agreement with the<br />

hard sphere model. Furthermore the SANS experiments show that the microemulsion droplets<br />

are fairly monodisperse, i.e. possess a polydispersity index of about 0.1 [62].<br />

For the cubic phase the SANS experiment shows still the same type of aggregates but<br />

more concentrated. However, here the packing exceeds the critical volume fraction of<br />

53 vol%, which is typical for a hard sphere crystallization [63, 64]. In the cubic phase the<br />

spherical aggregates are packed similar to metal atoms in a cubic lattice. From SANS investigations<br />

of samples in the cubic phase it seems that the packing is not primitive cubic but<br />

either face-centred cubic (fcc) or body-centred cubic (bcc). Between these two possibilities<br />

an experimental distinction was not feasible [65].<br />

Samples of the cubic phase show a very distinct SANS behaviour when the scattering<br />

intensity is detected two-dimensionally. Then an isotropic pattern is no longer observed.<br />

One finds more or less pronounced spikes superimposed on a symmetric correlation ring.<br />

These spikes are distributed r<strong>and</strong>omly on this ring. The occurrence of such scattering patterns<br />

is commonly observed with cubic phases as demonstrated by a typical example in<br />

Fig. 11.23 [61]. These relatively sharp peaks indicate a long-range ordering in the cubic<br />

phase. This scattering pattern can be explained by the presence of relatively large crystalline<br />

domains [66].<br />

A similar series of samples as in the SANS experiments was studied in cooperation<br />

with the group of Prof. Wokaun by NMR self-diffusion experiments. The pulsed field gradient<br />

spin echo (PGSE) method [67, 68] allows the determination of the self-diffusion coefficient<br />

of each of the individual constituent components in particular water, surfactant, <strong>and</strong><br />

hydrocarbon. Here, in order to obtain simpler NMR spectra the hydrocarbon was cyclohexane.<br />

The molar ratio of C 14 DMAO :cyclohexane was chosen to be 1:1.2, with three samples<br />

in the L 1 phase <strong>and</strong> three samples in the cubic phase.<br />

The obtained self-diffusion coefficient of water shows a continuous, approximately<br />

linear decrease with increasing volume fraction of the micellar aggregates <strong>and</strong> with no dis-<br />

222


11.5 Ringing gels<br />

Figure 11.23: Three-dimensional plot of the scattered intensity in a SANS experiment for 32.2 wt%<br />

C 14 DMAO/6.0 wt% decane/61.8 wt% D 2 O.<br />

continuity at the phase transition: L 1 phase ? cubic phase. Such a behaviour is in good<br />

agreement with the model of a continuous aqueous phase where the diffusion of the water<br />

molecules is simply hindered by the steric restrictions imposed by the presence of micellar<br />

aggregates [69].<br />

A much different picture is obtained for the diffusion of the surfactant <strong>and</strong> the hydrocarbon<br />

(Fig. 11.24). Within the L 1 phase both diffuse with the same coefficient. This is in<br />

good agreement with the diffusion coefficient of the micellar aggregates calculated from<br />

their size, determined by scattering methods. Again some decrease of the diffusion coeffi-<br />

100<br />

10<br />

D s /10 -12 m 2 /s<br />

1<br />

0,1<br />

a)<br />

b)<br />

c)<br />

d)<br />

e)<br />

L 1 -phase<br />

cubic<br />

phase<br />

0,01<br />

0,0 0,1 0,2 0,3 0,4 0,5<br />

Φ<br />

Figure 11.24: Self-diffusion coefficient D s of surfactant <strong>and</strong> hydrocarbon in logarithmic representation<br />

as a function of the micellar volume fraction F for the system C 14 DMAO/cyclohexane/D 2 O. The molar<br />

ratio of C 14 DMAO : cyclohexane was always 1:1.2. a) D s of (N(CH 3 ) 2 ); b) D s of (CH 2 ); c) D s of (CH 2 )<br />

for samples with deuterated cyclohexane; d) D s of H determined by means of a different instrument<br />

[69]; e) D s of H for samples with deuterated cyclohexane [69].<br />

223


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

cient with increasing micellar volume fraction occurs, as the movements of the aggregates<br />

get hindered increasingly. However, upon crossing into the cubic phase the situation is dramatically<br />

altered. The self-diffusion coefficient of surfactant <strong>and</strong> hydrocarbon drops suddenly<br />

by more than a factor of 20 for the cyclohexane <strong>and</strong> a factor of 200 for the surfactant.<br />

This is consistent with the picture from above, i. e. now the aggregates are frozen in <strong>and</strong> are<br />

no longer able to move in the cubic lattice which is formed. Hence the diffusion coefficient<br />

does not describe any more the motion of aggregates but is due to the diffusion of the individual<br />

molecules. This explains why the much smaller cyclohexane molecule now diffuses<br />

much faster than the larger C 14 DMAO molecule. Furthermore the very low diffusion coefficient<br />

excludes the possibility of a bicontinuous system since for such a structure a much faster<br />

diffusion would be expected [70].<br />

In summary, this experiment clearly demonstrates that in the L 1 phase <strong>and</strong> in the cubic<br />

phase the same water continuous structure is present. But while in the L 1 phase the micellar<br />

aggregates are freely mobile their translational mobility is blocked in the cubic phase, i. e.<br />

the microemulsion droplets are condensed into a glass-like liquid crystalline state of high<br />

elasticity.<br />

In general, from the investigation of the more dilute L 1 phase one may already conclude<br />

the microstructure of the cubic phase. This concept can be very useful for the explanation<br />

of macroscopic properties which are related to the aggregate size, like the shear modulus<br />

G 0 . By means of oscillatory rheological experiments one may determine G 0 which is typically<br />

in the range of 10 5 –10 6 Pa for cubic phases. Experimentally one finds for aminoxide<br />

systems that G 0 decreases with increasing size of the respective aggregates [71]. These experimental<br />

values may now be compared to theoretical calculations for hard sphere crystals<br />

[72]. In general, this ansatz will predict a proportionality of the elastic moduli to the particle<br />

density N of the aggregates. Furthermore, one can calculate for a given volume fraction the<br />

factor which relates the particle density to the elastic constants, like G 0 . The calculated values<br />

are in reasonable qualitative agreement with the experimental data [65]. This means<br />

that it is possible to deduce mechanical properties from the knowledge of the microstructure.<br />

Furthermore, this explains the strong dependence on the particle size since the particle density<br />

is proportional to 1/R 3 , i. e. the microstructure of the cubic phase directly determines the<br />

macroscopic elastic properties of the system.<br />

11.5.3 The bis-(2-ethylhexyl)sulfosuccinate system<br />

As stated above, the cubic phase of the AOT system is located differently in the phase diagram<br />

than the aminoxide system, which is due to the fact that AOT as a double-chain surfactant<br />

has a tendency to form reverse phases. Again it was of great interest to investigate the<br />

relation between the isotropic L 2 phase <strong>and</strong> the cubic phase. Furthermore, this cubic phase<br />

is interesting since here the surfactant concentration can be varied over a large range (30–<br />

76 wt%).<br />

The transition from the L 2 phase into the cubic phase has been studied by a variety of<br />

methods [73]. This can be done easily on a line of constant AOT content (in our case mostly<br />

35 wt%). Then increasing the octanol/H 2 O ratio one crosses from the cubic phase into the<br />

224


11.5 Ringing gels<br />

L 2 phase. Interestingly, measurements of the electric conductivity showed no discontinuity<br />

of the equivalent conductivity (specific conductivity/AOT concentration) upon crossing the<br />

phase boundary. Instead one observes a continuous increase of L with increasing H 2 O content.<br />

The value in the cubic phase is about 40% of that of the free Na + ions, which should<br />

mainly be responsible for the ionic conductivity (since the surfactant counterion will be largely<br />

immobilized being fixed in the amphiphilic film), which indicates that the structure<br />

must be water continuous.<br />

NMR PGSE self-diffusion studies on similar samples also showed a continuous increase<br />

of the water self-diffusion coefficient with rising water content. The value in the cubic<br />

phase is again about 40–50% of the bulk water diffusivity, confirming the water continuous<br />

structure. For the alkyl chains of octanol <strong>and</strong> AOT a separation was not possible, but<br />

the values of (2–4)610 –7 cm 2 /s are relatively large, more than one order of magnitude larger<br />

as in the case of the aminoxide, <strong>and</strong> show that the structure ought to be bicontinuous<br />

<strong>and</strong> similar to the neighbouring L 2 phase [73].<br />

SANS spectra of the cubic phase show the typical spikes on the isotropic diffraction<br />

ring (compare Fig. 11.22). From the position of the peaks <strong>and</strong> from the total scattering intensity,<br />

it can be concluded that the size of the structural units decreases with increasing AOT<br />

(water) content, at constant octanol/water (AOT/octanol) ratio.<br />

Measurements of the shear modulus G 0 on a series of samples with constant ratio of<br />

octanol/water showed that G 0 is proportional to c(AOT) 3.1 , i. e. again, as in the case of the<br />

aminoxide system, the elastic modulus increases with decreasing size of the structural units,<br />

since the SANS experiments show that the size of the structural units decreases with increasing<br />

AOT concentration. Correspondingly the particle density N of these units increases <strong>and</strong>,<br />

as stated above, G 0 should be proportional to N. This means that the determination of the<br />

microstructure already allows the prediction of the elastic properties, as in the case of aminoxide.<br />

Upon heating the cubic phase will melt yielding a lowly viscous L 2 phase. This phase<br />

transition was studied in some detail by differential scanning calorimetry (DSC). It was<br />

found that the higher the melting temperature the higher the surfactant content is, with melting<br />

temperatures in the range of 50 to 95 8C. The melting enthalpies DH show the same<br />

trend with typical values of 100–300 mJ/g. Similar values were also found for the aminoxide<br />

system. The maximum value, DH = 1034 mJ/g, was found for the cubic phase in the binary<br />

system of 75.8 wt% AOT <strong>and</strong> 24.2 wt% H 2 O at the melting temperature of 89.5 8C. For<br />

AOT concentrations close to the minimal value (around 28 wt%) necessary for forming the<br />

cubic phase, however, the situation becomes somewhat more interesting. Here, DH can be<br />

extremely small, below 0.5 mJ/g. This means that the macroscopically observable phase<br />

transition from the solid-like cubic phase to the lowly viscous L 2 phase is associated with almost<br />

no enthalpy difference. Furthermore, for samples with c (AOT) < 37 wt% another very<br />

striking phenomenon occurs. Here a transition from the cubic phase to the L 2 phase can be<br />

achieved by heating as well as by cooling! For the transition at lower <strong>and</strong> at higher temperature<br />

the transition enthalpy is very similar. Both cases are endothermic processes. Such a reverse<br />

melting process is quite unique for this type of cubic phase <strong>and</strong> has formerly only<br />

been reported for triblock copolymers of PEO/PPO/PEO-type (Section 11.5.4) [74–76] <strong>and</strong><br />

PEO/PPO diblock copolymers [77]. However, for the block copolymers the situation is insofar<br />

different as the cubic phase is made up from normal micellar aggregates, which are<br />

transformed into a cubic phase upon heating since here the effective volume fraction of the<br />

225


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

spherical aggregates increases [75]. In contrast to the cubic phase the AOT system is bicontinuous<br />

<strong>and</strong> transformed into a bicontinuous L 2 phase by a different <strong>and</strong> still unknown mechanism.<br />

11.5.4 PEO/PPO/PEO block copolymers<br />

PEO/PPO/PEO triblock copolymers exhibit properties similar to typical surfactants, i. e.<br />

they reduce surface <strong>and</strong> interfacial tension of aqueous solutions <strong>and</strong> form micellar aggregates<br />

above a critical micellar concentration [74, 78]. For some compounds of this type, like<br />

P104 (EO 18 PO 58 EO 18 ), P123 (EO 20 PO 70 EO 20 ), <strong>and</strong> F127 (EO 106 PO 70 EO 106 ), a similarly<br />

located cubic phase like the one of the aminoxide systems has been found in the binary aqueous<br />

system [74, 79].<br />

Structural investigations on these systems have shown that the block copolymers form<br />

spherical micellar aggregates in the L 1 phase close to the cubic phase. Here the PPO block<br />

is the hydrophobic moiety. It forms the core of the micellar aggregate <strong>and</strong> is surrounded by<br />

the more hydrophilic PEO blocks which act as the hydrophilic part of the polymeric surfactant<br />

molecules. The temperature dependence of this system is interesting. At low temperatures<br />

the block copolymers dissolve in water as unimers. Upon increasing the temperature<br />

the PO groups are dehydrated rendering them more hydrophobic. This increased hydrophobicity<br />

is responsible for the aggregation of the monomers into micellar aggregates. As has<br />

been determined by DSC measurements this micellization process occurs over a fairly large<br />

temperature range of typically 15–30 K [79]. For this dehydration process the enthalpy<br />

changes are about 3 kJ/mol per PO group. This uncommon temperature behaviour is also responsible<br />

for the interesting, already mentioned, reverse melting transition, i. e. in a corresponding<br />

concentration range the lowly viscous L 1 phase is transformed into the solid-like<br />

cubic phase upon heating. The simple explanation for this effect is that upon raising the<br />

temperature more <strong>and</strong> more monomers will aggregate in micelles. If the volume fraction F<br />

of the micellar aggregates exceeds the required one for the formation of a cubic phase<br />

(F = 0.53) then the gelification process occurs <strong>and</strong> a cubic phase, composed of individual<br />

micelles, is formed [64]. However, this phase transition is associated with an enthalpy about<br />

two orders of magnitude smaller than the heat of micellization – typically between 25 <strong>and</strong><br />

100 mJ/g, i. e. similar in extent as for the other cubic phases described above. Its nature is<br />

endothermic, i. e. the transition is associated with an increase of entropy. This transition can<br />

also be nicely monitored by SANS measurements. For the micellar solution an isotropic<br />

scattering ring is found where the typical spikes of the cubic phase can be observed if the<br />

transition temperature is crossed [74].<br />

More recently it has been found that the formation of the cubic phase can be suppressed<br />

by the admixture of a simple surfactant, such as SDS [80]. This effect is due to cooperative<br />

binding of SDS molecules on the block copolymer molecules. By doing so the micelles<br />

are dissolved in favour of monomeric units until the volume fraction of the micelles<br />

gets to small to form a cubic phase, i. e. one can melt the cubic phase by adding surfactant.<br />

For example, for transforming the cubic phase of 25 wt% F127 a SDS concentration of<br />

100 mM is required.<br />

226


11.6 Lyotropic mesophases<br />

Summarizing, cubic phases are a type of surfactant systems that can be observed for a<br />

large variety of different surfactants. Depending on the molecular structure of the surfactant<br />

they may be located in different places of the phase diagram with correspondingly different<br />

microstructures. They can be composed of an array of individual micellar aggregates or be<br />

of bicontinuous structure. However, even such a bicontinuous structure still will be characterized<br />

by a very well-defined typical size of the structural units with long-range ordering.<br />

At this point it is interesting to note that this size can already be obtained by studying isotropic<br />

phases close to the cubic phase. This is insofar advantageous since they are often much<br />

more amenable to experimental studies than the corresponding cubic phase itself. The<br />

knowledge of the size of these units (no matter whether they are individual aggregates or<br />

only structural repeat units of a bicontinuous structure) enables the prediction of the elastic<br />

moduli which will be proportional to the particle density of these units. For a given structural<br />

build-up theoretical relations can be used to calculate the moduli quantitatively. Therefore<br />

a detailed knowledge of the microstructure of the corresponding systems allows a quantitative<br />

underst<strong>and</strong>ing of macroscopic properties. Cubic phases are a nice example for a system<br />

where by now such structure property relations are well established.<br />

11.6 Lyotropic mesophases<br />

11.6.1 Introduction<br />

In general, at higher concentrations surfactant aggregates show a nearest neighbour order under<br />

the influence of the intermicellar interaction. Upon further increasing the concentration<br />

(mostly above 30–40 wt%) a long-range order between the micelles takes place <strong>and</strong> lyotropic<br />

mesophases are formed. Depending on the kind of micelles cubic phases are formed<br />

from globular micelles, rod-like micelles form hexagonal phases, <strong>and</strong> disc-like micelles<br />

form lamellar phases. At surfactant concentrations significantly above 50 wt% inverse hexagonal<br />

<strong>and</strong> cubic phases can also be built. Generally speaking, with increasing concentration<br />

the phase sequence: cubic > hexagonal > lamellar > inverse hexagonal > inverse cubic is<br />

found (Fig. 11.25) [81]. Figure 11.25 shows also schematically the structures of the different<br />

lyotropic mesophases.<br />

Figure 11.25 shows that between the hexagonal <strong>and</strong> the lamellar phases further isotropic<br />

phases, which also are cubic, can exist, although they do not consist of globular particles but<br />

show bicontinuous structures [73, 82]. Of course, not all these phases must be present in each<br />

surfactant solution.In most of the systems the first cubic phase is missing up to high surfactant<br />

concentrations because globular micelles are only present in few cases. Double-chain surfactants<br />

often have a lamellar phase as the first mesophase [83], but all systems show principally<br />

the given sequence of the mesophases. This is also valid for more-component systems, where<br />

according to the constitution of the further components, certain phases may be favoured or<br />

suppressed. Hydrocarbons, for example, favour globular micelles <strong>and</strong> hence cubic phases [55,<br />

227


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

Figure 11.25: Schematic phase diagram of a binary system surfactant/water. Inverse hexagonal phase<br />

H II , lamellar phase L a , hexagonal phase H I , isotropic (cubic) phases a, b, c, d.<br />

61, 84]. But cosurfactants, like aliphatic alcohols, with an intermediate chain length support<br />

the formation of disc-like micelles <strong>and</strong> thus of lamellar phases [85].<br />

The aggregates of the phases in Fig. 11.25 show a long-range order with respect to<br />

their orientation <strong>and</strong> to their centre of masses. Nevertheless, also lyotropic mesophases exist<br />

which are built up by rod-like or disc-like micelles, respectively. They show only a longrange<br />

order of their orientation, while their mass centres are statistically distributed in the<br />

solutions. These phases are called nematic calamitic (N c ) or nematic discotic (N d ) phases<br />

<strong>and</strong> exist at lower concentrations than the corresponding hexagonal or lamellar phases.<br />

Those nematic phases are very interesting because they can be uniformly oriented by weak<br />

external fields due to the lack of long-range order of their mass centres. This gives rise to<br />

the development of single crystals instead of the polycrystalline mesophases with differently<br />

oriented domains. On addition of chiral components the nematic phases can be transformed<br />

to lyotropic cholesteric phases with a helical twist of the orientation, as has been found for<br />

the corresponding thermotropic phases [86].<br />

11.6.2 Nematic phases <strong>and</strong> their properties<br />

Lyotropic nematic phases have been discovered a long time after the finding of the lyotropic<br />

mesophases. In 1967 first evidence of such a phase was published [87] but it took more than<br />

10 years to prove unambiguously the existence of these phases [88]. The first nematic<br />

228


11.6 Lyotropic mesophases<br />

phases have been found by fortune in complicated ternary <strong>and</strong> quaternary systems consisting<br />

of surfactants, cosurfactants, electrolytes, <strong>and</strong> water. Systematic studies on different surfactant<br />

systems have finally shown that lyotropic nematic phases do not occur so rarely as<br />

could be concluded from their late discovery. At the same time criteria could be established<br />

which allow a fairly precise prediction of the existence of a nematic phase <strong>and</strong> its position<br />

in the phase diagram [89]. This also allows the direct preparation of different types of the<br />

nematic phases <strong>and</strong> the study of their properties. Important conditions for the existence of<br />

nematic phases are the voluminous hydrophobic part <strong>and</strong> the headgroup area of the hydrophilic<br />

group which must be within certain values, for example 60–90 Å 2 for double-chain<br />

ionic surfactants with a N c phase under these conditions. The headgroup area can be lowered<br />

by adding cosurfactants (aliphatic alcohols with an intermediate chain length). This leads to<br />

the formation of a N d phase. Both phases border to the corresponding hexagonal or lamellar<br />

phase from the side of lower concentrations. For smaller headgroup areas a lamellar phase is<br />

formed first while for larger headgroup areas a hexagonal or a cubic phase is the first mesophase<br />

[89]. For example, numerous nematic phases have been found in solutions of anionic<br />

perfluoro surfactants with their voluminous hydrophobic groups which mostly have been<br />

identified as N d phases [90, 91]. Figure 11.26 shows a typical phase diagram of the binary<br />

system perfluoro-surfactant/water. It can be seen that the nematic phase exists only in a<br />

small concentration <strong>and</strong> temperature range which was probably the reason for the late detection<br />

of these phases. From Figure 11.26 the marked thermotropic behaviour of the nematic<br />

<strong>and</strong> the lamellar phases can also be seen. Thus it is possible by increasing the temperature<br />

to go through the phase sequence: lamellar > lamellar/nematic > nematic > nematic/<br />

isotropic > isotropic. These thermotropic phase transitions are reversible. In most cases DSC<br />

measurements have shown that the phase transitions are of first order with very small heat<br />

transitions. Further experiments have shown that the nematic phases can be oriented uniformly<br />

in a magnetic field according to the anisotropy of the diamagnetism of the surfactant<br />

molecules. Isolated anisotropic aggregates cannot be oriented in micellar solutions because<br />

the energy of the magnetic field is much smaller than the thermal energy which achieves a<br />

r<strong>and</strong>om orientation of the micelles in this case. The lamellar phases, on the other h<strong>and</strong>, also<br />

cannot be oriented by the magnetic field because this would require the simultaneous orientation<br />

of a sheet of lamellae for which the energy of the magnetic field is not sufficient.<br />

70<br />

60<br />

nematic<br />

50<br />

isotropic<br />

T/°C<br />

40<br />

30<br />

lamellar<br />

20<br />

crystalline<br />

10<br />

0 10 20 30 40 50<br />

wt%<br />

Figure 11.26: Section of the phase diagram of the system C 8 F 17 CO 2 N(CH 3 ) 4 /D 2 O.<br />

229


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

But it is possible to orient the phases uniformly in the nematic region <strong>and</strong> to freeze<br />

the orientation by cooling the system to the lamellar region. This orientation remains without<br />

the magnetic field [92]. This allows a simple study of such phases <strong>and</strong> of the type of<br />

particles present in the phases. It is also possible to solubilize anisometric dye molecules in<br />

the aggregates <strong>and</strong> to orient them together with these micelles in the nematic phase by a<br />

magnetic field, leading to a dichroism of the systems [93]. The oriented nematic <strong>and</strong> lamellar<br />

phases show also a strong anisotropy of their electric conductivity, i. e. increase of the<br />

conductivity parallel <strong>and</strong> decrease of the conductivity perpendicular to the large axis of the<br />

aggregates. N c phases have also been found in solutions of the cationic double-chain surfactants<br />

of the series C x C y N(CH 3 ) 2 Br with x =14or16<strong>and</strong>y = 1–4 which fulfill the established<br />

conditions with respect to the headgroup area <strong>and</strong> the volume of the hydrophobic alkyl<br />

chain. On the other h<strong>and</strong>, surfactants with y = 6 do not form mesophases up to concentrations<br />

of 75 wt%, while surfactants with y>8 show first a lamellar mesophase [89].<br />

These mesophases also exist only in a small concentration range near the hexagonal<br />

phase, as can be seen, for example, in Fig. 11.27. Above a characteristic temperature they<br />

show a reversible phase transition of first order into an isotropic phase. Figure 11.27 also<br />

shows that a thermotropic transition hexagonal > nematic cannot be observed in these systems.<br />

Furthermore it could be shown for these phases that on addition of a cosurfactant a<br />

transition of the N c into a N d phase can take place. For one system it was found that both nematic<br />

phases can be present in equilibrium. From orientation experiments of these nematic<br />

phases in a magnetic field it could be concluded that the sign of the anisotropy of their diamagnetism<br />

is positive. If the Br counterion was substituted by benzene sulfonate, the nematic<br />

phases are kept but the sign of the anisotropy of the diamagnetism changes. Hence it<br />

is possible to prepare <strong>and</strong> to study all four possible nematic phases N + c,N – c,N + d, <strong>and</strong> N – d with<br />

these surfactants.<br />

Figure 11.27: Section of the phase diagram of the system CTAB/H 2 O.<br />

The identification of the N c <strong>and</strong> N d phases can be archived by polarization microscopy<br />

based on their different textures which are presented in Fig. 11.28 a <strong>and</strong> 11.28 b. 2 H-NMR<br />

spectroscopic studies <strong>and</strong> orientation experiments on the nematic phases in a magnetic field<br />

are suitable for their detection, especially if SANS measurements are done on oriented samples.<br />

For these experiments the phases have to be prepared with D 2 O instead of H 2 O. This<br />

230


11.6 Lyotropic mesophases<br />

Figure 11.28a: Texture of a calamitic nematic (N c ) phase in the binary system of 25 wt% hexadecyltrimethylammoniumbromide<br />

(CTAB)/water at 35 8C.<br />

Figure 11.28b: Texture of a discotic nematic (N d ) phase in the binary system of 47.5 wt% tetradecylpyridiniumheptanesulfonate<br />

(C 14 PyC 7 SO 3 )/water at 26 8C.<br />

does not affect the nematic phases except a small shift of their existence region towards<br />

lower surfactant concentrations. Furthermore, SANS <strong>and</strong> SAXS measurements show a high<br />

degree of orientation of the nematic phases by the appearance of a third order scattering<br />

peak.<br />

<strong>Final</strong>ly it shall be mentioned that until now no inverse nematic phases could be detected.<br />

Also the existence of nematic phases in non-polar solvents could not be proven for<br />

any studied system. Thus it is not yet possible to replace the thermotropic nematic phases in<br />

displays by lyotropic systems because lyotropic nematics cannot be orientated in an electric<br />

field with low voltages due to the rather high conductivity of the water solvent compared to<br />

non-polar organic compounds.<br />

231


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

11.6.3 Cholesteric phases <strong>and</strong> their properties<br />

For thermotropic <strong>and</strong> also for lyotropic nematic phases it is known that the presence of<br />

chiral compounds leads to a transition of the nematic phase to a cholesteric one. This phase<br />

has also only a long-range order with respect to the orientation of the aggregates, but this orientation<br />

shows a twist within a domain, <strong>and</strong> the pitch depends on the concentration of the<br />

chiral compound. Cholesteric phases can be built up with chiral components (intrinsic cholesteric<br />

phases). The transition nematic > cholesteric can also take place by adding chiral<br />

samples (induced cholesteric phases) [86]. Intrinsic cholesteric phases require the use of<br />

chiral surfactants. For this purpose the group of the non-ionic sugar surfactants has been<br />

chosen. As the sugar head group is mostly very voluminous, it could be expected that only a<br />

few systems could fulfill the conditions for the formation of a nematic or cholesteric phase<br />

with respect to the headgroup area <strong>and</strong> the volume of the hydrophobic group. Further problems<br />

arose from the poor solubility of sugar surfactants with sufficiently small headgroups<br />

<strong>and</strong> because many of these surfactants are strongly sensitive to hydrolysis. Nevertheless, the<br />

first intrinsic cholesteric phase in a binary system surfactant/water could be detected for an<br />

alkylpolyglucoside with 12–14 C atoms <strong>and</strong> 1,1 glucose units on an average [94]. By orientation<br />

in a magnetic field this phase could be identified as a cholesteric discotic phase with<br />

negative anisotropy of its diamagnetism. On the other h<strong>and</strong>, it is no problem possible to<br />

transform N c <strong>and</strong> N d phases in a ternary system of the zwitterionic surfactant tetradecyldimethylaminoxide<br />

(C 14 DMAO), a cosurfactant (an aliphatic alcohol with 7–10 C atoms),<br />

<strong>and</strong> water into the corresponding induced cholesteric phases by adding chiral compounds<br />

[95]. This transition could be observed with a non-polar chiral additive (cholesterol), which<br />

can be solubilized only in the micellar aggregates, <strong>and</strong> with a polar chiral compound (tartaric<br />

acid), which remains dissolved in the aqueous phase.<br />

The cholesteric phases could be identified by their fingerprint textures in the polarization<br />

microscope, which are characteristic for non-oriented cholesteric phases. Uniformly oriented<br />

cholesteric phases show characteristic stripe textures. Both textures are presented in<br />

Figs. 11.29 a <strong>and</strong> 11.29 b. From the distance of the stripes the pitch of the phase can be cal-<br />

Figure 11.29a: Texture of a non-oriented cholesteric phase in the binary system of 60 wt% alkylpolyglucoside<br />

(APG 1)/water at 25 8C.<br />

232


11.6 Lyotropic mesophases<br />

Figure 11.29b: Texture of a oriented cholesteric phase (aligned in a magnetic field of 2 T) in the binary<br />

system of 60 wt% alkylpolyglucoside (APG 1)/water at 25 8C.<br />

culated. The experiments showed a linear relation between the reciprocal pitch <strong>and</strong> the concentration<br />

of the chiral additive. Until now, the reason for the pitch caused by the chiral additive<br />

is not understood. An interesting fact about these cholesteric phases is that above a<br />

certain temperature the pitch disappears <strong>and</strong> a nematic phase develops reversibly. This transition<br />

temperature is lower than the temperature for the transition nematic > isotropic. It<br />

could not be detected by DSC experiments, very likely due to its small enthalpy change.<br />

11.6.4 Vesicle phases <strong>and</strong> L 3 phases<br />

From Fig. 11.25 it can be seen that lyotropic lamellar phases normally exist at surfactant<br />

concentrations above 50 wt%. According to the theory of surfactant aggregation [4] surfactants<br />

with small headgroup areas <strong>and</strong> big hydrophobic groups are known to form lamellar<br />

phases already at concentrations far below 10 wt%. Such systems are perfluorosurfactants<br />

with strongly binding counterions [96] or double-chain surfactants [83]. In mixtures of surfactants<br />

with small headgroup areas like zwitterionic or non-ionic surfactants <strong>and</strong> cosurfactants<br />

– for example aliphatic n-alcohols with intermediate chain lengths – the geometrical<br />

conditions for the formation of lamellar phases in highly dilute systems are also fulfilled.<br />

Furthermore it is possible in these systems to change the natural curvature <strong>and</strong> the flexibility<br />

of surfactant lamellae simply by variation of the mixing ratio surfactant : cosurfactant. According<br />

to this, these ternary systems exhibit a typical phase behaviour with lamellar phases<br />

at the water rich corner, which has already been explained in detail in Section 11.4.1.<br />

The most powerful method for the identification of the different phases <strong>and</strong> for the determination<br />

of their microstructures is the transmission electron microscopy. This method requires<br />

a sample preparation using the freeze fracture <strong>and</strong> etching technique. With these measurements<br />

the microstructures of the various phases can be visualized directly. This is shown<br />

for three examples in Fig. 11.12 (Section 11.4.2) <strong>and</strong> in Figs. 11.30 a <strong>and</strong> 11.30 b [13, 41, 97].<br />

233


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

Figure 11.30a: Electron micrograph of a lamellar phase in the system of 100 mM C 14 DMAO, 190 mM<br />

C 7 OH, <strong>and</strong> water (the bar represents 2.5 mm).<br />

Figure 11.30b: Electron micrograph of a L 3 phase in the system of 100 mM C 12 DMAO, 220 mM<br />

C 6 OH, <strong>and</strong> water (the bar represents 140 nm).<br />

With these technique it can also be demonstrated in the two phase regions that both phases<br />

with their different structures are really coexisting. This result can be supported by self-diffusion<br />

measurements with the pulsed NMR field gradient method from which the existence<br />

of bicontinuous phases <strong>and</strong> their topology can be concluded [73, 98].<br />

The electron micrograph in Fig. 11.12 shows the existence of thermodynamically<br />

stable vesicles in the L al phase which can be formed as small unilamellar vesicles besides<br />

large multilamellar vesicles (liposomes). Depending on the composition of the ternary system<br />

the vesicles can also show structural faults like holes (perforated vesicles) [37]. The vesicle<br />

phases show a significantly reduced electric conductivity because a part of the water<br />

phase together with the ions is included in the interior of the vesicles. With stopped-flow experiments<br />

<strong>and</strong> optical or conductivity readout it is thus possible to determine the permeability<br />

of the vesicle membranes for dissolved compounds [99]. Vesicle phases show often high<br />

viscosities <strong>and</strong> viscoelasticity due to the mutual hindrance of the vesicles in sheared solu-<br />

234


11.6 Lyotropic mesophases<br />

tions. Especially if the vesicle membranes are charged by incorporation of certain amounts<br />

of added ionic surfactants, the phases show with increasing charge of the surfactant film an<br />

increasing yield stress [40, 100]. The vesicle phases usually are not or only weakly birefringent<br />

<strong>and</strong> thus cannot be identified in the polarization microscope.<br />

From Fig. 11.30 a <strong>and</strong> for the L ah phases the presence of flat surfactant lamellae like in<br />

normal lamellar phases is evident. These phases are birefringent but they often do not develop<br />

their characteristic texture in the polarization microscope due to their low concentration. Until<br />

now it could not be proven whether there is a phase boundary between the L a1 <strong>and</strong> the L ah<br />

phase or a continuous transition from L al to L ah with increasing cosurfactant concentration.<br />

This problem is very difficult due to the high viscosity <strong>and</strong> turbidity of the phases.<br />

The bicontinuous structure of the L 3 phase can also be seen directly from Fig. 11.30 b.<br />

This phase shows a low viscosity without elasticity <strong>and</strong> is optically isotropic but shows a<br />

strong flow birefringence. According to the similarity with normal micellar or reversed micellar<br />

solutions (L 1 <strong>and</strong> L 2 phases) these phases have been called L 3 phases. The surfactant<br />

lamellae are arranged as branched tubes similar to a sponge. The low viscosity results from<br />

the very high flexibility of the lamellae which break <strong>and</strong> reform easily.<br />

As mentioned in Section 11.2, the phase sequence in Fig. 11.1 can be understood regarding<br />

the properties of the surfactant films. With low amounts of cosurfactant the natural<br />

curvature of the film is convex <strong>and</strong> the flexibility is usually low. Hence micellar aggregates<br />

are present in the L 1 phase. With increasing amounts the curvature of the films decreases<br />

<strong>and</strong> their flexibility increases due to the small headgroup area of the cosurfactant. Thus transition<br />

into a vesicle phase <strong>and</strong> with more cosurfactant into a normal lamellar phase can take<br />

place. Further increase of the cosurfactant increases the flexibility of the films <strong>and</strong> permits<br />

the formation of the L 3 phase. Adding even more cosurfactant finally leads to the separation<br />

of the cosurfactant phase. This concept also allows to underst<strong>and</strong> the influence of further additives<br />

on the phase behaviour of the surfactant/cosurfactant/water systems. Adding an ionic<br />

surfactant to the L 3 phase leads to a reduction of the flexibility <strong>and</strong> to a convex curvature of<br />

the surfactant films due to the electrostatic repulsion between the ionic headgroups. Thus a<br />

transition of the L 3 phase to a vesicle phase is observed [38]. These vesicles have rather stiff<br />

membranes <strong>and</strong> the phase often shows a yield stress, as demonstrated in Section 11.4.3. But<br />

addion of an electrolyte to these systems shields the charge of the headgroups <strong>and</strong> gives rise<br />

to an increased flexibility <strong>and</strong> decreased curvature of the films. Hence the vesicle phase is<br />

transformed again into a L 3 phase. Addition of an ionic surfactant to phospholipid vesicles,<br />

which normally must be prepared by sonification of the aqueous phospholipid dispersions<br />

[101], also leads to a destruction of the vesicles <strong>and</strong> to a transition to rod-like micelles above<br />

a certain concentration of the ionic surfactant [25].<br />

<strong>Final</strong>ly, it is worth mentioning that thermodynamically stable vesicles can also be<br />

found in the binary systems surfactant/water if the hydrophobic group is sufficiently voluminous.<br />

As already mentioned, this has been observed for the double-chain surfactant didodecyldimethylammoniumbromide<br />

DDABr in water [102]. Recently also a L 3 phase has been<br />

found in the binary system of a polyoxyethylene-polyoxypropylene-polyoxyethylene triblock<br />

copolymer <strong>and</strong> water at high temperatures. This block copolymer has a composition<br />

EO 5 PO 30 EO 5 <strong>and</strong> hence a very small hydrophilic headgroup. It forms a lamellar phase at<br />

room temperature which is transformed into the L 3 phase above a certain temperature [103].<br />

The reason for this transition is probably the increased flexibility of the block copolymer<br />

film at the high temperature.<br />

235


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

11.7 Shear induced phenomena<br />

11.7.1 General<br />

In this Section we discuss a remarkable <strong>and</strong> puzzling phenomenon that, because of its complicated<br />

nature, is not yet completely understood. The phenomenon has, however, considerable potential<br />

for technical applications where it is necessary to control the flow behaviour of aqueous<br />

solutions. In all applications where large amounts of water have to be circulated for cooling or<br />

heating purposes the energy expense for the pumping is a major economic factor. Usually one is<br />

interested in pumping as fast as feasible so that the flow in the water pipes is generally in the turbulent<br />

flow region. Under these conditions it is possible to reduce the friction coefficient by<br />

polymer additives or by drag-reducing surfactants (Fig. 11.31). In recycling operations polymers<br />

have the big disadvantage that they deteriorate under shear because the molecules break<br />

under shear forces. Surfactants do not have this disadvantage because the micellar structures<br />

which produce this effect are self-healing. Pilot operations in Europe have been running for<br />

months without loss of efficiency. The energy costs have been cut to less than a half.<br />

pressure drop [hPa]<br />

360<br />

300<br />

240<br />

180<br />

120<br />

60<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 1 2 3 4 5<br />

0<br />

0,0 0,2 0,4 0,6 0,8<br />

flow velocity [m/s]<br />

Figure 11.31: Plot of the pressure drop vs.the flow velocity in a capillary in the laminar <strong>and</strong> turbulent<br />

flow regions for water (solid line) <strong>and</strong> for a drag reducing surfactant solution (750 ppm<br />

C 14 TABr + NaSal at 27.5 8C, dashed line).<br />

11.7.2 Under what conditions do we find drag-reducing surfactants?<br />

The phenomenon occurs often in surfactant solutions in which small rod-like micelles are<br />

formed that are charged <strong>and</strong> the charge is not screened by excess salt [104]. Many surfactant<br />

systems have been found where such conditions exist. Typically the length of the rods is<br />

236


11.7 Shear induced phenomena<br />

smaller than the mean distance between them. From this point of view the micellar solutions<br />

can be considered dilute even though there is repulsive interaction between the rods. Because<br />

of this repulsion the micelles try to be as far away from each other as possible <strong>and</strong> set<br />

up what is called a nearest neighbour order. The result of this order is a correlation peak in<br />

scattering experiments. In typical conditions the surfactant concentration is a few mM (about<br />

0.1–0.2 wt%), the rods are a few hundred angstroms long <strong>and</strong> their mean separation is<br />

somewhat larger. Because there is no steric hindrance between the rods they can undergo<br />

Brownian rotations with rotation times of a few microseconds. Such conditions can easily be<br />

set up when charged surfactants are mixed with zwitterionic surfactants. Usually mixtures of<br />

ionic (10–30 mol%) <strong>and</strong> zwitterionic surfactants are favourable for the effect. These features<br />

of the micelles are the necessary prerequisites for the occurrence of the effect. They are,<br />

however, not the only ones, as will become clear.<br />

In shear measurements one expects the described solutions behave like normal Newtonian<br />

aqueous solutions. This is in fact the case for small shear rates (Fig. 11.32). In Fig. 11.32<br />

the shear viscosity, which was measured in a capillary viscometer, is plotted vs.the shear rate.<br />

One observes a sudden rise of the viscosity at a characteristic shear rate _g c <strong>and</strong> for _g > _g c the<br />

solutions show some shear thickening behaviour. Obviously something dramatic has happened<br />

to the micelles in the solutions. Some conclusions about what has happened can be drawn<br />

from flow birefringence measurements. Some typical results of flow measurements from a<br />

Couette system are shown in Fig. 11.33. We note a sudden increase of the flow birefringence<br />

at a critical shear rate. For _g < _g c no flow birefringence could be detected.<br />

In flow experiments, besides the birefringence it is also easy to measure the angle of<br />

extinction, which is the angle between the direction of flow <strong>and</strong> the mean orientation of the<br />

rods. In normal flow orientation this angle decreases smoothly from 458 to zero with increasing<br />

flow rate because the rods become more <strong>and</strong> more aligned. In the drag-reducing solution<br />

the situation is very different. In the Newtonian region for _g < _g c the solution remains isotropic<br />

<strong>and</strong> no preferential alignment can be detected. However, if _g > _g c the angle of extinction<br />

is close to zero, i. e. the new structures which are produced by the shear are completely<br />

aligned. Obviously the newly found structures must be much larger than the original small<br />

rods which were not aligned.<br />

η[mPas]<br />

8<br />

6<br />

4<br />

2<br />

60 mM<br />

80 mM<br />

100 mM<br />

120 mM<br />

T=25°C<br />

0<br />

10 1 10 2 10 3<br />

.<br />

γ[s -1 ]<br />

Figure 11.32: The shear viscosity Z vs. the shear rate _g for mixtures of C 14 DMAO : SDS = 6 : 4 with<br />

various total concentrations in a capillary viscometer at 25 8C.<br />

237


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

-∆n/10 -6<br />

10<br />

8<br />

6<br />

4<br />

2<br />

30 mM<br />

60 mM<br />

80 mM<br />

100 mM<br />

0<br />

0 100 200 300 400 500 600 700 800 900<br />

.<br />

γ /s -1<br />

Figure 11.33: The flow birefringence Dn vs. the shear rate _g for the same solutions as in Fig. 11.32.<br />

It is believed now that the small rod-like micelles undergo collisions in the shear flow<br />

<strong>and</strong> they stick together for some time because of their interfacial properties. In this way long<br />

necklace-type structures are formed under shear <strong>and</strong> at the same time get aligned in the<br />

shear flow. This situation is schematically sketched in Fig. 11.34. These necklace-type structures<br />

act like high molecular weight polymers <strong>and</strong> give rise to drag reduction.<br />

These results show that shear is an important variable for micellar structures. We are<br />

aware that generally temperature, ionic strength, concentration, <strong>and</strong> cosurfactants can change<br />

micellar structures <strong>and</strong> hence the properties of surfactant solutions. We also should be aware<br />

that shear can change <strong>and</strong> influence micellar structures <strong>and</strong> even mesophases. It has been<br />

observed that micellar solutions can be transformed into liquid crystalline phases <strong>and</strong> single<br />

clear phase solutions become turbid as well as biphasic under shear. On the other h<strong>and</strong>, biphasic<br />

solutions can turn into a single phase under shear. A very striking example of a<br />

shear-produced transition is the transformation of a dilute L 3 phase into a L a phase with<br />

bright iridescent colours [105]. All these effects are based on the fact that the shape <strong>and</strong> size<br />

of the micelles depend to some degree on the intermicellar interaction energy, which itself<br />

depends on the mutual orientation of the micelles. When the interaction energy is changed<br />

the system responds with a change of structure <strong>and</strong> properties. These changes can be unexpected<br />

<strong>and</strong> large. They can, however, be used to our advantage.<br />

Figure 11.34: Model for the explanation of the shear induced micellar structures. The small rod-like<br />

micelles can form long necklace-type structures under shear.<br />

238


11.8 SANS measurements on micellar systems<br />

11.8 SANS measurements on micellar systems<br />

A large variety of our surfactant systems has been investigated by means of small angle neutron<br />

scattering (SANS) experiments mainly in cooperation with the group of Prof. Kalus<br />

(Universität Bayreuth). SANS is a method particularly suited for the study of self-aggregating<br />

colloids since its spatial resolution is typically in the range of 10–1000 Å, which is the<br />

size range of micellar aggregates. In the course of the investigation a shear apparatus was<br />

constructed in the group of Prof. Kalus which allows SANS experiments under shear. This<br />

has the advantage that one can align anisometric aggregates in the shear field <strong>and</strong> from the<br />

scattering curves of the aligned particles one can deduce more detailed information regarding<br />

their structure <strong>and</strong> their dynamic behaviour in the shear field. In addition, systems can<br />

be studied that exhibit shear induced structures, which are very interesting since they show<br />

drag-reducing behaviour. In the following we give just a short, exemplary overview over our<br />

large number of SANS experiments (some of them are discussed in the corresponding chapters,<br />

e. g. the work on cubic phases).<br />

The SANS is a method for the detailed determination of size <strong>and</strong> shape of the corresponding<br />

surfactant assemblies. For instance, from such experiments we found that tetramethylammoniumperfluoroctanesulfonate<br />

(TMAFOS) forms rod-like structures with a radius<br />

of 22 Å <strong>and</strong> a length of 200–300 Å [106]. For perfluorated surfactants the SANS<br />

method is only particularly advantageous because the refractive index of water <strong>and</strong> perfluorated<br />

compounds are very close. Therefore for these systems the powerful light scattering<br />

experiments for studying micellar systems will not work. Furthermore, detailed structural information<br />

might not be accessible for perfluorated surfactants, which are an interesting class<br />

of surfactants since normally they are even more surface-active than their hydrocarbon counterparts.<br />

However, elongated micelles are by no means restricted to perfluoro surfactants but<br />

also commonly found with conventional hydrocarbon alkyl surfactants. An interesting system<br />

is the mixed cationic/anionic surfactant tetradecylpyridinium-heptanesulfonate. Here SANS<br />

measurements have shown that below 160 mM charged rod-like aggregates are present which<br />

get shorter with increasing concentration. For these samples always a correlation peak is observed<br />

in the spectra but for more concentrated samples this peak will disappear. The reason<br />

is that with increasing concentration the degree of dissociation is decreasing. Therefore the<br />

charge density on the micelles decreases thus decreasing the electrostatic repulsion between<br />

the aggregates until they are effectively uncharged at high concentration [107].<br />

Another interesting type of surfactant are double-chain amphiphiles. As an example of<br />

this surfactant-type hexadecyloctyldimethylammoniumbromide (C 16 C 8 DMABr) has been<br />

investigated. Here just above the cmc globular aggregates are formed whereas above a second<br />

transition concentration relatively short (150 Å) rod-like micelles are formed [108].<br />

However, anisometry of the micellar aggregates is a necessary (yet not sufficient) prerequisite<br />

of liquid crystal formation. Indeed, at higher concentrations many perfluoro surfactants<br />

are known to form nematic, lyotropic, liquid, <strong>and</strong> crystalline phases [109]. One such<br />

system is the tetramethylammoniumperfluorononanoate (TMAPFN) which exhibits a nematic<br />

phase over a relatively large concentration range but only in a small temperature interval<br />

[91]. This system has been studied in detail by SANS [92]. The experiments showed that<br />

239


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

the nematic phase is build up from disc-like micellar aggregates, i. e. the local structure of<br />

the amphiphilic film is planar. This phase can be oriented in an external magnetic field,<br />

thereby allowing for a more detailed analysis of the underlying micellar structures. The<br />

thickness of the disk was found to be 37 Å <strong>and</strong> from a rocking experiment the order parameter<br />

S & 0.85 was deduced. In addition, it is interesting to note that only the nematic phase<br />

can be oriented in an external magnetic field. But upon cooling the samples are transformed<br />

into the corresponding lamellar phase. This phase could not be oriented but it is highly ordered<br />

<strong>and</strong> also keeps its orientation without applying a magnetic field.<br />

As just discussed, lyotropic nematic phases can easily be aligned in an external magnetic<br />

field <strong>and</strong> it is of course of interest to study this alignment process as a function of<br />

time, i. e. to study the dynamics of the nematic phase. This has been done for a system<br />

made up from an aqueous solution of 10.2 wt% hexadecyltrimethylammoniumbromide<br />

(CTAB), 9.9 wt% hexadecyltrimethylammoniumbenzenesufonate, <strong>and</strong> 2.45 wt% decanol in<br />

D 2 O. This system forms a nematic phase that consists of disc-like micelles. Such a sample<br />

had been prealigned in a magnetic field of 7 T. The highly ordered phase gives a strongly<br />

anisotropic scattering pattern with two sharp peaks (Fig. 11.35). The preoriented sample<br />

which was in the neutron beam was then exposed to a magnetic field (1.4 T) oriented perpendicular<br />

to the axis of the originally employed magnetic field as well as to the neutron<br />

beam. Of course, this new magnetic field wants to reorient the texture of the sample <strong>and</strong><br />

this reorientation process has been monitored by SANS. In Fig. 11.35 we see the time evolution<br />

of the scattering pattern: upon turning on the perpendicular field the two original peaks<br />

start to disintegrate into four smaller peaks which move on a ring to form finally again two<br />

narrow peaks which are then oriented perpendicular to the original peaks. The typical time<br />

for reorientation is about 60 min [110].<br />

a) b)<br />

c) d)<br />

Figure 11.35: SANS curves of a prealigned nematic phase of a CTAB system (composition see text).<br />

the original aligned sample (a), the sample in a perpendicular magnetic field B 1 after 20 (b), 50 (c), <strong>and</strong><br />

90 min (d). (B 0 =7T,B 1 = 1,4 T).<br />

240


11.8 SANS measurements on micellar systems<br />

Of course, anisometric micelles can also be oriented by shear <strong>and</strong> this again allows a<br />

more detailed study of the micellar structure as well as the dynamic alignment process. One<br />

such system has been cetylpyridiniumsalicylate in 20 mM NaCl D 2 O solution which at 20 mM<br />

concentration has been shown to contain rod-like micelles with a radius of 21.5 Å <strong>and</strong> a length<br />

of 500–750 Å [111]. Under shear an anisotropic scattering pattern is observed. It relaxes to<br />

the isotropic pattern after switching off the applied shear field. Time resolved measurements<br />

(time resolution of 250 ms) showed that the scattering curves during this relaxation process<br />

can be described by a single parameter, namely the rotational diffusion coefficient D rot .Itwas<br />

observed that D rot is time-dependent <strong>and</strong> decreases with time due to the interaction between<br />

the charged rod-like aggregates, i. e. at the beginning the aligned rods have the largest electrostatic<br />

repulsion giving a large driving force for the disorientation because for this aligned arrangement<br />

the electrostatic potential energy has its highest value. The less aligned the system<br />

the smaller becomes this driving force (since the electrostatic interaction becomes weaker)<br />

<strong>and</strong> correspondingly the rotational diffusion coefficient becomes smaller [112].<br />

The experimental scattering curves can be explained by an orientation distribution function<br />

of the rods depending on the applied shear gradient [113]. Under the experimental conditions<br />

the product of shear gradient _g <strong>and</strong> structural relaxation time t, which is increasing with<br />

the length of the aggregates, was always much larger than 1, i. e. a high orientational ordering<br />

was achieved because the ordering force becomes stronger than the diffusive force.<br />

Generally all anisometric aggregates can become oriented in the shear field but systems<br />

with shorter relaxation times t require higher shear gradients. This effect has been studied<br />

with the above mentioned C 16 C 8 DMABr at a concentration of 50 mM. Here the unsheared<br />

solution shows a correlation ring which becomes increasingly anisotropic with increasing<br />

shear rate. Figure 11.36 shows the scattering patterns for various shear gradients. In<br />

addition a higher order peak becomes visible. <strong>Final</strong>ly for shear rates above _g = 2000 s –1 the<br />

scattering pattern hardly changes any more. In this system not only the relatively short rodlike<br />

aggregates become weakly aligned but also a second type of micelle is formed in the<br />

shear field, where their relative concentration increases with rising shear rate. Therefore in<br />

this system a shear-induced transformation of micellar aggregates is observed [114].<br />

A similar behaviour has also been observed for the system tetradecyltrimethylammoniumsalicylate<br />

(C 14 TMA-Sal) [115] <strong>and</strong> seems to be quite common for viscoelastic surfactant<br />

systems. A more detailed analysis of this system indicates that the sharp peaks that occur<br />

upon applying the shear are due to a hexagonal array of cylindrical micelles, i. e. again now<br />

2 types of micelles are present: the originally present short rod-like micelles <strong>and</strong> the very long<br />

rod-like aggregates that are formed in the flow field [116]. This lattice-type array is formed<br />

within about 2 min as could be seen from time-resolved shear experiments [116]. This result<br />

agrees also well with flow birefringence experiments. At a given shear rate an equilibrium between<br />

the short rod-like micelles, which normally are only weakly aligned, <strong>and</strong> the long rodlike<br />

aggregates is present. This equilibrium depends on the shear rate. The higher the shear<br />

rate the more the equilibrium is shifted in favour of the long aggregates [117].<br />

The growth process of the large micellar structures, which are strongly aligned, has<br />

been studied in more detail by transient SANS experiments. In these experiments the shear<br />

rate for the samples was raised stepwise from zero to a certain finite value. These experiments<br />

showed that the large micelles grow according to the Avrami law<br />

c…t† ˆc inf …1 exp… kt †† : …19†<br />

241


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

a) b)<br />

c)<br />

d)<br />

Figure 11.36: SANS patterns of C 16 C 8 DMABr in D 2 O (50 mM) for various shear gradients (the momentum<br />

transfer is given in units of nm –1 ): _g =0s –1 (a), _g = 100 s –1 (b), _g = 400 s –1 (c), _g = 2000 s –1 (d).<br />

The neutron beam is perpendicular to the direction of shear.<br />

Originally this equation was used to describe nucleation <strong>and</strong> growth in metals <strong>and</strong> alloys.<br />

For the given system (C 14 TMA-Sal) the exponent n was found to be between 2 <strong>and</strong> 2.5<br />

[118]. The exponent n should be i + 1 for i-dimensional growth, which means that the<br />

growth process observed here is close to a one-dimensional one as should be expected since<br />

the micelles grow in length without change of dimension.<br />

Such shear-induced structures (SIS) are an interesting phenomenon in particular for<br />

self-aggregating systems like micelles where the equilibrium structure often depends very<br />

subtly on small energetic changes. Of course, these structural changes have a profound influence<br />

on the properties of these systems, especially on their flow behaviour. For instance one<br />

may observe a shear-thickening behaviour that is coupled to drag-reducing properties of the<br />

system. Shear-induced transitions have been found in a variety of micellar systems [105]. Of<br />

particular interest in our studies have been systems that in the unsheared state contain small<br />

rod-like micelles <strong>and</strong> systems which show a strongly anisotropic behaviour beyond a critical<br />

threshold shear rate _g c . Above this shear rate a strong increase of flow birefringence <strong>and</strong><br />

viscosity together with a large anisotropy of the electric conductivity are observed. At the<br />

same time the scattering patterns of the SANS experiments exhibit also a strong anisotropy.<br />

This shear-induced effect will already occur at _g t rot P 1, where t rot is the rotational time<br />

constant of the small type of micelles, i. e. in a range where the shear field should not be<br />

able to orient the small rod-like aggregates significantly. This means that the observed anisotropy<br />

is not due to the orientation of these originally present micelles but that larger oriented<br />

micellar aggregates have to be present in the solutions. So far the mechanism for formation<br />

of the SIS is not fully understood <strong>and</strong> several different mechanisms have been postulated<br />

[119–121].<br />

242


11.9 A new rheometer<br />

The tetradecyldimethylaminoxide/sodiumdodecylsulfate (C 14 DMAO/SDS) system has<br />

been studied in much detail. This system shows a pronounced SIS formation around a molar<br />

mixing ratio of 7 : 3 for C 14 DMAO/SDS [122] – <strong>and</strong> it might also be noted here that for this<br />

composition the nematic phase, which is found for those systems, extends to the lowest surfactant<br />

concentration [123]. In order to find relations between the macroscopic behaviour of<br />

the system <strong>and</strong> the structure of the micellar aggregates, SANS study was performed [124].<br />

Changing the contrast condition in the micellar aggregate by using both hydrogenated <strong>and</strong><br />

deuterated SDS, yields detailed information regarding the structure of the micelles. The<br />

SANS experiments show that at the mixing ratios where the length of these aggregates reduces<br />

with increasing SDS content <strong>and</strong> where SIS is observed, elongated micelles are present<br />

which are best described by a three-axes prolate ellipsoid. From a contrast variation experiment<br />

using both, deuterated <strong>and</strong> hydrogenated SDS, it could be concluded that the buildup<br />

of these micelles is homogeneous <strong>and</strong> no internal segregation of the surfactant molecules<br />

within the aggregate could be deduced [124]. Such an internal segregation by having an enrichment<br />

of SDS at the end caps (here the relative area per molecule is necessarily larger because<br />

of the larger curvature <strong>and</strong> one could imagine that the SDS with its larger hydrophilic<br />

head group would preferentially be located at this position) would have been conceivable<br />

<strong>and</strong>, of course, such a build-up that would contain two more highly charged ends would<br />

have been an important factor to consider for the explanation of the SIS. However, this evidently<br />

is not the case <strong>and</strong> SIS formation has to be explained starting from originally short,<br />

homogeneously charged, elongated aggregates.<br />

If we summarize we find that SANS is a very powerful tool to investigate self-aggregating<br />

structures. It is perfectly adapted with the size range to be observed <strong>and</strong> in addition it<br />

allows a detailed observation of anisotropic samples which might be oriented by a magnetic<br />

field, shear field, order effects close to the wall of the cells, etc. Scattering patterns of such<br />

oriented samples contain even more information regarding the intra <strong>and</strong> interparticle structure.<br />

Furthermore, the contrast conditions in the samples can in principle be changed because<br />

they consist to a large degree of hydrogen <strong>and</strong> because the scattering properties of H<br />

<strong>and</strong> D nuclei differ largely. Therefore isotopic substitutions, that normally have only a very<br />

minor effect on the properties of the respective systems, will enable us to get much more<br />

structural information than would be accessible via other methods. This method is called the<br />

contrast variation. Those experiments have shed a lot of light on structural properties of micellar<br />

systems with respect to some of their dynamic properties.<br />

11.9 A new rheometer<br />

From electric birefringence measurements it is known that surfactant solutions, like those of<br />

tetraethylammoniumperfluorooctanesulfonate, show different time constants in the range<br />

from microseconds up to some seconds [24, 94, 125]. The shortest time t 1 is attributed to<br />

the free rotation of rod-like micelles <strong>and</strong> amounts to 10 –5 to 10 –6 s. It can be found in almost<br />

the whole concentration range. The second time t 2 is the birefringence which has a dif-<br />

243


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

ferent sign <strong>and</strong> comes to 10 –4 s. It can only be detected in the narrow concentration region<br />

where the rods begin to overlap (7 mM–15 mM). The longest time constant t 4 (about some<br />

seconds) represents the structural relaxation time. It can be found in the electric birefringence<br />

<strong>and</strong> rheological measurements at sufficiently high concentrations (c >30 mM) when<br />

a network is formed. The remaining time constant t 3 (in the range of milliseconds) could<br />

not yet be detected by rheological measurements because the frequency range of commercial<br />

rheometers ends typically at frequencies of about 10 Hz. Therefore it is not possible to determine<br />

rheological time constants shorter than about 0.1 s.<br />

In order to overcome this problem, it was necessary to improve the frequency range<br />

for the dynamic rheological measurements. For this purpose a dynamic rheometer (HF rheometer)<br />

with a frequency range from 1 Hz up to 1 kHz was built. For samples with a high<br />

modulus (G&1 kPa) the measurements can even be extended to 2 kHz.<br />

With this apparatus measurements on viscoelastic surfactant solutions were carried<br />

out. As expected, these solutions show short rheological time constants t 3 . They correspond<br />

well to those determined with the electric birefringence.<br />

The new HF rheometer is based on a prototype that was acquired from a group of the<br />

Universität Ulm, Abteilung Angew<strong>and</strong>te Physik, Prof. W. Pechhold. The sensitivity of the<br />

apparatus could be notably improved by numerous technical modifications. A schematic<br />

drawing of the mechanical part of the apparatus is shown in Fig. 11.37.<br />

Figure 11.37: Schematic drawing of the mechanical part of the HF rheometer.<br />

244


11.9 A new rheometer<br />

The rheometer works with a concentric cylinder geometry. The gaps between the cylinders<br />

are 50 mm or 100 mm, respectively. The inner cylinder is driven by an electromechanical<br />

converter (shaker) <strong>and</strong> performs linear harmonic oscillations. The force is transferred<br />

by the sample to the outer cylinder <strong>and</strong> is detected by a very sensitive piezoelectric force<br />

transducer producing a signal which is amplified by a charge amplifier. The amplitude of<br />

the inner cylinder is determined by an inductive displacement transducer coupled with a carrier<br />

frequency measuring amplifier. A lock-in amplifier is used for the measurement of the<br />

force signal, the displacement signal, <strong>and</strong> the phase angle. In the present state measurements<br />

on viscoelastic samples with a modulus of 50 Pa, 100 Pa, <strong>and</strong> 1 kPa are possible up to<br />

900 Hz (50 mM gap), 1.3 kHz <strong>and</strong> 2 kHz, respectively.<br />

The reliability of the apparatus was tested <strong>and</strong> the frequency range was experimentally<br />

determined by measurements on Newtonian liquids (silicon oils, glycerol/water mixtures).<br />

The viscosities of Newtonian liquids can be well detected down to 50 mPas. In this case the<br />

frequency range extends to 1.6 kHz.<br />

Solutions of tetraethylammoniumperfluorooctanesulfonate (C 8 F 17 SO 3 NEt 4 ) were studied<br />

with the HF rheometer. A typical rheogram is shown in Fig. 11.38. The measurement<br />

was carried out on a 90 mM solution with the 50 mM gap in the frequency range above<br />

10 Hz. Below 10 Hz a Bohlin CS 10 rheometer with a cone plate geometry was employed.<br />

10 2 G´<br />

G´, G´´ / Pa, |η*| / Pas<br />

10 1<br />

10 0<br />

10 -1<br />

10 -2<br />

10 -3<br />

G´´<br />

|η*|<br />

10 -4<br />

10 -2 10 -1 10 0 10 1 10 2 10 3<br />

f/Hz<br />

Figure 11.38: Dynamic rheogram of a 90 mM solution of C 8 F 17 SO 3 NEt 4 at T =208C. Below 10 Hz<br />

the measurement was performed with a Bohlin CS 10 rheometer, above 10 Hz with the HF rheometer<br />

<strong>and</strong> the 50 mM gap (with 1% deformation).<br />

In the lower frequency range the samples show Maxwell behaviour. G' rises with the<br />

slope 2, G@ with slope 1. At frequencies above the crossover of G' <strong>and</strong> G@, G' levels out <strong>and</strong><br />

reaches a plateau value, whereas G@ first decreases <strong>and</strong> then increases again. At high frequencies<br />

both, G' <strong>and</strong> G@, increase. Even though a second plateau value of G' at high frequencies<br />

could not be found it was possible to fit the data by a Burger model (four parameter<br />

Maxwell model):<br />

G 0 o 2 t 2 4 o 2 t 2 3<br />

…o† ˆG 1<br />

1 ‡ o 2 t 2 ‡ G 2<br />

4<br />

1 ‡ o 2 t 2 3<br />

<strong>and</strong> G 0 ot 4 ot 4<br />

…o† ˆG 1<br />

1 ‡ o 2 t 2 ‡ G 2<br />

4<br />

1 ‡ o 2 t 2 3<br />

: …20†<br />

245


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

The structural relaxation time t 4 (the indices have been chosen for correspondence<br />

with previous results) decrease with rising concentration from 25 ms (70 mM) to 0.2 ms<br />

(300 mM) <strong>and</strong> the short time constant t 3 decrease from 0.35 ms (70 mM) to 0.1 ms<br />

(250 mM). Therefore, they correspond well with the time constants determined by dynamic<br />

electric birefringence measurements.<br />

For each concentration the minimum value of G@ is about a factor of 2.5 lower than<br />

the plateau value of G'. According to the theory of Granek <strong>and</strong> Cates [126] this yields a value<br />

of 2.5 for the ratio of the mean contour length of the micelles to the entanglement length<br />

(Eq. 7)). A decrease of the mean micellar length with rising concentration can be concluded<br />

on the basis of this value <strong>and</strong> from the decreasing structural relaxation times.<br />

Furthermore, mixtures of tetraethylammoniumperfluorooctanesulfonate <strong>and</strong> the pure<br />

perfluorooctanesulfonic acid (C 8 F 17 SO 3 H) were studied with the HF rheometer. In Fig. 11.39<br />

a rheogram is shown of the 150 mM solution with the salt : acid ratio of 7 : 3.<br />

10 3 G´<br />

G´, G´´ / Pa, |η*| / Pas<br />

10 2<br />

10 1<br />

10 0<br />

10 -1<br />

10 -2<br />

G´´<br />

|η*|<br />

10 -3<br />

10 -2 10 -1 10 0 10 1 10 2 10 3<br />

Figure 11.39: Dynamic rheogram of a 150 mM solution of C 8 F 17 SO 3 NEt 4 /C 8 F 17 SO 3 H = 7:3 at<br />

T =218C. The deformation was 1% again. The minimum of G@ is lower than in the case of the pure<br />

C 8 F 17 SO 3 NEt 4 .<br />

f/Hz<br />

From a qualitative point of view the rheogram does not differ too much from the one<br />

presented in Fig. 11.4a. It can be noticed that after substitution of salt by acid the minimum<br />

of G@ is more pronounced. This means a larger ratio of the mean contour length to the entanglement<br />

length (here about 4), calculated by Eq. 7.<br />

Generally one observes that up to a mole fraction of 40% of the acid the structural relaxation<br />

time t 4 increases by a factor 10 (from 7 ms to 70 ms). From this result the growth<br />

of the micellar aggregates can be concluded. The short time constant t 3 does not notably<br />

change <strong>and</strong> amounts to about 0.1 ms. At a molar ratio of more than 50 % of the acid the<br />

viscosity breaks down due to a decreasing length of the micelles. In this case it is not possible<br />

anymore to measure the samples with the HF rheometer.<br />

The high frequency increase of the moduli can be interpreted on basis of the following<br />

models:<br />

a) In addition to the continuous network there also exist unentangled shorter rods due to<br />

the equilibrium distribution of the micellar lengths. These shorter aggregates do not significantly<br />

influence the rheological behaviour of the samples at lower frequencies. At higher frequencies,<br />

however, they contribute to the moduli.<br />

246


References<br />

b) At angular frequencies higher than the reciprocal value of the Rouse time the chain segment<br />

of the entanglement length cannot relax anymore by Rouse diffusion <strong>and</strong> a glass process<br />

starts causing an increase of the moduli.<br />

References<br />

1. K.L. Mittal: ACS-Symp. Ser., 8, Am. Chem. Soc, Washington DC, (1975)<br />

2. C. Tanford: The Hydrophobic Effect, Wiley & Sons, New York, (1980)<br />

3. H. Hoffmann: Progr. Colloid Polym. Sci., 83, 16 (1990)<br />

4. J.N. Israelachvili, D.J. Mitchell, B.W. Ninham: J. Chem. Soc. Faraday Trans. II, 72, 1525 (1976)<br />

5. M. Borkovec: J. Chem. Phys., 91, 6268 (1989)<br />

6. K.S. Chan, D.O. Shah: J. Dispersion Sci. Technol., 1, 55 (1980)<br />

7. W.H. Wade, H. Seang, S.H. Yiv: J. Colloid Interface Sci., 92, 561 (1983)<br />

8. O. Abillon, P.B. Bings, D. Langevin, J. Meunier: Progr. Colloid Polym. Sci., 76, 84 (1988)<br />

9. G. Platz, C. Thunig, H. Hoffmann: Ber. Bunsenges. phys. Chem., 96, 667 (1992)<br />

10. M. Jonströmer, R. Strey: J. Phys. Chem., 96, 5994 (1992)<br />

11. O. Ghosh, C.A. Miller: J. Phys. Chem., 91, 4528 (1987)<br />

12. E.W. Kaler, A.K. Murthy, B.E. Rodriguez, J.A.N. Zasadzinski: Science, 245, 1371 (1989)<br />

13. H. Hoffmann, C. Thunig, U. Munkert, H.W. Meyer, W. Richter: Langmuir, 8, 2629 (1992)<br />

14. T.M. Clausen, P.K. Vinson, J.R. Minter, H.T. Davis, Y. Talmon, W.G. Miller: J. Phys. Chem., 96,<br />

474 (1992)<br />

15. S. Toshiyuki, H. Hirata, T. Kodaka: Langmuir, 3, 1081 (1987)<br />

B.K. Mishra, S.D. Samant, P. Pradhan, S.B. Mishra, C. Manohar: Langmuir, 9, 894 (1993)<br />

16. S.J. C<strong>and</strong>au, E. Hirsch, R. Zana, M. Delsanti: Langmuir, 5, 1225 (1989)<br />

17. H. Hoffmann, A.K. Rauscher, M. Gradzielski, S.F. Schulz: Langmuir, 8, 2140 (1992)<br />

18. A. Khatory, F. Kern, F. Lequeux, J. Appell, G. Porte, N. Morie, A. Ott, W. Urbach: Langmuir, 9,<br />

933 (1993)<br />

19. H.A. Barnes, J.F. Hutton, K. Walters: in: An Introduction to Rheology, Rheol. Ser., 3, Elsevier,<br />

New York, (1989)<br />

20. M.E. Cates: in: Proc. ACS-Symp. on Polymeric Microemulsions <strong>and</strong> Polymer-Microemulsion Interaction,<br />

New Orleans, (1987)<br />

21. A.K. Rauscher, H. Rehage, H. Hoffmann: Progr. Colloid Polym. Sci., 84, 99, (1991)<br />

22. A. Khatory, F. Lequeux, F. Kern, S.J. C<strong>and</strong>au: Langmuir, 9, 1456 (1993)<br />

23. H. Hoffmann, K. Reizlein, H. Rehage, H. Thurn: ACS-Symp. Ser., 27, 44 (1985)<br />

H. Hoffmann, U. Krämer, H. Thurn: J. Phys. Chem., 94, 2027 (1990)<br />

24. M. Angel, H. Hoffmann, U. Krämer, H. Thurn: Ber. Bunsenges. phys. Chem., 93, 184 (1989)<br />

25. H. Hoffmann, E. Schönfelder: Ber. Bunsenges. phys. Chem., 98, 842 (1994)<br />

26. F.C. McKintosh, S.A. Safran, P.A. Pincus: J. Phys. Condens. Matter, 2, 359 (1990)<br />

S.A. Safran, P.A. Pincus, M.E. Cates, F.C. McKintosh: J. Phys. France, 51, 503 (1990)<br />

27. H. Hoffmann, M. Löbl, H. Rehage, I. Wunderlich: Tenside Surf. Det., 22, 290 (1985)<br />

H. Thurn, M. Löbl, H. Hoffmann: J. Phys. Chem., 89, 517 (1985)<br />

28. H. Hoffmann, M. Löbl, H. Rehage: Suppl. Rheol. Acta, 26, 246 (1988)<br />

29. M. Valiente, C. Thunig, U. Munkert, U. Lenz, H. Hoffmann: J. Colloid Interface Sci., 160, 39<br />

(1993)<br />

30. T. Imae, M. Sasaki, S. Ikeda: J. Colloid Interface Sci., 127, 511 (1989)<br />

S.J. C<strong>and</strong>au, A. Khatory, F. Lequeux, F. Kern: in: Proc. of the Workshop on Complex Liquid Systems,<br />

Polistena (1992)<br />

247


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

31. J. Appell, G. Porte, A. Khatory, F. Kern, S.J. C<strong>and</strong>au: J. Phys. II, 2, 1045 (1992<br />

32. T.J. Drye, M.E. Cates: J. Chem. Phys., 96, 1367 (1992)<br />

33. M.S. Turner, C. Marques, M.E. Cates: Langmuir, 9, 695 (1993)<br />

34. A. Ott, W. Urbach, D. Langevin, H. Hoffmann: Langmuir, 8, 1045 (1992)<br />

35. T. Kato, T. Terao, M. Tsukada, T. Seimiya: J. Phys. Chem., 97, 3910 (1993)<br />

36. E.W. Kaler, K.L. Herrington, A. Murthy, J.A.N. Zasadzinski: J. Phys. Chem., 96, 6698 (1992)<br />

K. Fontell, A. Ceglie, B. Lindman, B.W. Ninham: Acta Chem. Sc<strong>and</strong>., A40, 247 (1986)<br />

T. Zemb, D. Gazeau, M. Dubois, T. Gulik-Krzywicki: Europhys. Lett., 21, 759 (1993)<br />

37. U. Munkert, H. Hoffmann, C. Thunig, H.W. Meyer, W. Richter: Progr. Colloid Polym. Sci., 93, 137<br />

(1993)<br />

38. H. Hoffmann, U. Munkert, C. Thunig, M. Valiente: J. Colloid Interface Sci., 163, 217 (1994)<br />

39. Y. Okahata, T. Kunitake: J. Am. Chem. Soc., 101, 5231 (1979)<br />

B.W. Ninham, D.F. Evans, G.J. Wei: J. Phys. Chem., 87, 5020 (1983)<br />

D. Andersson, H. Wennerström, U. Olsson: J. Phys. Chem., 93, 4243 (1989)<br />

R. Strey, W. Jahn, G. Porte, P. Bassereau: Langmuir, 6, 1635 (1990)<br />

C.A. Miller, O. Ghosh, W. Benton: Colloids Surf., 19, 197 (1986)<br />

S. Szönyi, H.J. Watzke, A. Cambon: Progr. Colloid Polym. Sci., 89, 149 (1992)<br />

40. C. Thunig, G. Platz, H. Hoffmann: Springer Proc. in Physics, 66, 266 (1992)<br />

41. H. Hoffmann, C. Thunig, M. Valiente: Colloids Surf., 67, 223 (1992)<br />

42. M. Dubois, T. Zemb: Langmuir, 7, 1352 (1991)<br />

43. H. Rehage, H. Hoffmann: Mol. Phys., 74, 933 (1991)<br />

44. W.P. Cox, E.H. Merz: J. Polym. Sci., 28, 619 (1958)<br />

45. D. Roux, F. Nallet, O. Diat: Europhys. Lett., 24, 53 (1993)<br />

D. Roux, F. Nallet, O. Diat: J. Phys. II, 3, 1427 (1993)<br />

46. H. Hoffmann, C. Thunig, P. Schmiedel, U. Munkert: Langmuir, 10, 3972 (1994)<br />

47. H. Hoffmann, C. Thunig, P. Schmiedel, U. Munkert: Nuovo Cimento, 16D, 1373 (1994)<br />

48. H. Hoffmann, C. Thunig, P. Schmiedel, U. Munkert, W. Ulbricht: Tenside Surf. Det., 31, 389<br />

(1994)<br />

49. E. van der Linden, J.H.M. Dröge: Physica A, 193, 439 (1993)<br />

50. H.N.W. Lekkerkerker: Physica A, 140, 319 (1989)<br />

51. T. Palberg, J. Kottal, F. Bitzer, R. Simon, M. Würth, P. Leiderer: J. Colloid Interface Sci., 169, 85<br />

(1995)<br />

52. M. Dubois, T. Zemb, L. Belloni, A. Delville, P. Levitz, R. Setton: J. Chem. Phys., 96, 2278 (1992)<br />

53. K. Fontell: Colloid Polym. Sci., 268, 264 (1990)<br />

54. W. Jettka: PhD thesis, Universität Braunschweig (1984)<br />

55. G. Oetter, H. Hoffmann: Colloids Surf., 38, 225 (1989)<br />

56. H.N. Hoffmann, E.F. Paulus: Fette, Seifen, Anstrichmittel, 71, 399 (1969)<br />

57. E.S. Lower, W.H. Harding: Manuf.Chem., 37, 52 (1966)<br />

58. J.S. Patton: M.C. Carey, Science, 204, 145 (1979)<br />

59. D.P. Siegel: Biophys.J., 49, 1155 (1986)<br />

60. D.W. Osborne, A.J.I. Ward, K.J. O’Neill: in: D.W. Osborne <strong>and</strong> A.H. Amann (eds.): Topical Drug<br />

Delivery Formulations, Marcel Dekker, New York, (1990)<br />

61. M. Gradzielski, H. Hoffmann, G. Oetter: Colloid Polym. Sci., 268, 167 (1990)<br />

62. M. Gradzielski, H. Hoffmann: Adv. Colloid Interface Sci., 42, 149 (1992)<br />

63. P.N. Pusey, W. van Megen: Nature, 320, 340 (1986)<br />

64. K. Mortensen, J.S. Pedersen: Macromolecules, 26, 805 (1993)<br />

65. M. Gradzielski: PhD thesis, Universität Bayreuth (1992)<br />

66. H. Jousma, J.G.H. Joosten, G.S. Gooris, H.E. Junginger: Colloid Polym. Sci., 267, 353 (1989)<br />

67. E.O. Stejskal, J.E. Tanner: J. Chem. Phys., 42, 288 (1965)<br />

68. P. Stilbs: Progr. NMR Spectrosc., 19, 1 (1987)<br />

69. K.L. Walther, M. Gradzielski, H. Hoffmann, A. Wokaun, G. Fleischer: J. Colloid Interface Sci.,<br />

153, 272 (1992)<br />

70. A. Ceglie, K.P. Das, B. Lindman: Colloids Surf., 28, 29 (1987)<br />

248


References<br />

71. M. Gradzielski, H. Hoffmann, in: D.M. Bloor <strong>and</strong> E. Wyn-Jones (eds.): The Structure, Dynamics<br />

<strong>and</strong> Equilibrium <strong>Properties</strong> of Colloidal Systems, Kluwer Academic Publishers, Dordrecht, p. 427<br />

(1990)<br />

72. B.B. Laird: J. Chem. Phys., 97, 2699 (1992)<br />

73. M. Gradzielski, H. Hoffmann, J.C. Panitz, A. Wokaun: J. Colloid Interface Sci., 169, 103 (1995)<br />

74. G. Wanka, H. Hoffmann, W. Ulbricht: Colloid Polym. Sci., 268, 101 (1990)<br />

75. K. Mortensen, W. Brown, B. Norden: Phys. Rev. Lett., 13, 2340 (1992)<br />

76. W. Brown, K. Schillen, S. Hvidt: J. Phys. Chem., 96, 6038 (1992)<br />

77. Y. Deng, G.E. Yu, C. Price, C. Booth: J. Chem. Soc. Faraday Trans. I, 88, 1441 (1992)<br />

78. Z. Zhou, B. Chu: J. Colloid Interface Sci., 126, 171 (1988)<br />

79. G. Wanka, H. Hoffmann, W. Ulbricht: Macromolecules, 27, 4145 (1994)<br />

80. E. Hecht, K. Mortensen, M. Gradzielski, H. Hoffmann: J. Phys. Chem., 99, 4866 (1995)<br />

81. M. Seddon: Biochim. Biophys. Acta, 1031, 1 (1990)<br />

82. J. Charvolin, J.F. Sadoc: J. Phys. France, 47, 683 (1987)<br />

83. J. Rogers, P.A. Winsor: J. Colloid Interface Sci., 30, 247 (1969)<br />

H. Kunieda, K. Shinoda: J. Phys. Chem., 82, 1710 (1978)<br />

84. G. Oetter, H. Hoffmann, in: D.M. Bloor <strong>and</strong> E. Wyn-Jones (eds.): The Structure, Dynamics <strong>and</strong><br />

Equilibrium <strong>Properties</strong> of Colloidal Systems, Kluwer Academic Publishers, Netherl<strong>and</strong>s, p. 427<br />

(1990)<br />

85. P. Ekwall, L. M<strong>and</strong>ell: Acta Polytech. Sc<strong>and</strong>., 74, 92 (1968)<br />

F.C. Larche, J. Appell, G. Porte, P. Bassereau, J. Marignan: Phys. Rev. Lett., 56, 700 (1986)<br />

86. M. Boidart, A. Hochapfel, P. Peretti: Mol. Cryst. Liq. Cryst., 172, 147 (1989)<br />

87. K.D. Lawson, T.J. Flautt: J. Am. Chem. Soc., 89, 5490 (1967)<br />

88. T. Haven, K. Radley, A. Saupe: Mol. Cryst. Liq. Cryst., 75, 87 (1981)<br />

B.J. Forrest, L.W. Reeves: Chem. Rev., 81, 1 (1981)<br />

89. G. Hertel: PhD thesis, Universität Bayreuth (1989)<br />

G. Hertel, H. Hoffmann: Progr. Colloid Polym. Sci., 76, 123 (1989)<br />

G. Hertel, H. Hoffmann: Liq. Cryst., 5, 1883 (1989)<br />

D. Tezak, G. Hertel, H. Hoffmann: Liq. Cryst., 10, 15 (1991)<br />

90. K. Reizlein: PhD thesis, Universität Bayreuth (1983)<br />

91. K. Reizlein, H. Hoffmann: Colloid Polym.Sci., 69, 83 (1984)<br />

92. L. Herbst, H. Hoffmann, J. Kalus, K. Reizlein, U. Schmelzer, K. Ibel: Ber. Bunsenges. phys.<br />

Chem., 89, 1050 (1985)<br />

93. M. Angel, H. Hoffmann, B. Schw<strong>and</strong>ner, R. Weber: in: H. Sackmann (ed.): Proc. of the Liquid<br />

Crystal Conference in Halle, Kongreß- und Tagungsberichte der Martin-Luther-Universität Halle-<br />

Wittenberg, p. 120 (1985)<br />

94. U. Krämer: PhD thesis, Universität Bayreuth (1990)<br />

95. G. Bartusch, H.D. Dörfler, H. Hoffmann: Progr. Colloid Polym. Sci., 89, 307 (1992)<br />

96. B. Lindman, K. Fontell: J. Phys. Chem., 87, 3289 (1983)<br />

97. U. Munkert: PhD thesis, Universität Bayreuth (1994)<br />

98. M. Gradzielski, H. Hoffmann, J.C. Panitz, A. Wokaun: J. Phys. Chem., 98, 6812 (1994)<br />

99. S. Kaiser, H. Hoffmann: J. Colloid Interface Sci., 184, 1 (1996)<br />

100. H. Hoffmann, A.K. Rauscher: Colloid Polym. Sci., 271, 390 (1993)<br />

101. C. Huang: Biochemistry, 8, 344 (1969)<br />

102. U. Lenz: PhD thesis, Universität Bayreuth (1991)<br />

103. E. Hecht, K. Mortensen, H. Hoffmann: Macromolecules, 28, 5465 (1995)<br />

104. D. Ohlendorf, W. Interthal, H. Hoffmann: Rheol. Acta, 25, 468 (1986)<br />

I. Wunderlich, H. Hoffmann, H. Rehage: Rheol. Acta, 26, 532 (1987)<br />

105. H. Hoffmann, S. Hofmann, A.K. Rauscher, J. Kalus: Progr. Colloid Polym. Sci., 84, 24 (1991)<br />

106. H. Hoffmann, J. Kalus K. Reizlein, W. Ulbricht, K. Ibel: Colloid Polym. Sci., 260, 435 (1982)<br />

107. H. Hoffmann, J. Kalus, B. Schw<strong>and</strong>ner: Ber. Bunsenges. phys. Chem., 91, 99 (1987)<br />

108. G. Neubauer, H. Hoffmann, J. Kalus: B. Schw<strong>and</strong>ner: Chem. Phys., 110, 247 (1986)<br />

109. M. Boidart, A. Hochapfel, M. Laurent: Mol. Cryst. Liq. Cryst., 154, 61 (1988)<br />

249


11 The Micellar Structures <strong>and</strong> the <strong>Macroscopic</strong> <strong>Properties</strong> of Surfactant Solutions<br />

110. J. Baumann, G. Hertel, H. Hoffmann, K. Ibel, V. Jindal, J. Kalus, P. Lindner, G. Neubauer, H.<br />

Pilsl, W. Ulbricht, U. Schmelzer: Prog. Colloid Polym. Sci., 81, 100 (1990)<br />

111. H. Hoffmann, J. Kalus, H. Thurn, K. Ibel: Ber. Bunsenges. phys. Chem., 87, 1120 (1983)<br />

H. Thurn, J. Kalus, H. Hoffmann: J. Chem. Phys., 80, 3440 (1984)<br />

112. L. Herbst, H. Hoffmann, J. Kalus, H. Thurn, K. Ibel, R.P. May: Chem. Phys., 103, 437 (1986)<br />

113. J. Kalus, H. Hoffmann: J. Chem. Phys., 87, 714 (1987)<br />

114. J. Kalus, H. Hoffmann, S.H. Chen, P. Lindner: J. Phys. Chem., 93, 4267 (1989)<br />

115. J. Kalus, H. Hoffmann, K. Ibel: Colloid Polym. Sci., 267, 818 (1989)<br />

116. V.K. Jindal, J. Kalus, H. Pilsl, H. Hoffmann, P. Lindner: J. Phys. Chem., 94, 3129 (1990)<br />

117. C. Münch, H. Hoffmann, J. Kalus, K. Ibel, G. Neubauer, U. Schmelzer: J. Appl. Cryst., 24, 740<br />

(1991)<br />

118. C. Münch, H. Hoffmann, K. Ibel, J. Kalus, G. Neubauer, U. Schmelzer, J. Selbach: J. Phys.<br />

Chem., 97, 4514 (1993)<br />

119. M.E. Cates, M.S. Turner: Europhys. Lett., 11, 681 (1990)<br />

120. S.Q. Wang: J. Phys. Chem., 94, 8381 (1990)<br />

121. R. Bruinsma, W.M. Gelbart, A. Ben-Shaul: J. Chem. Phys., 96, 7710 (1992)<br />

122. S. Hofmann, A.K. Rauscher, H. Hoffmann: Ber. Bunsenges. phys. Chem., 95, 153 (1991)<br />

123. H. Hoffmann, S. Hofmann, J.C. Illner: Prog. Colloid Polym. Sci., 97, 103 (1994)<br />

124. H. Pilsl, H. Hoffmann, S. Hofmann, J. Kalus, A.W. Kencono, P. Lindner, W. Ulbricht: J. Phys.<br />

Chem., 97, 2745 (1993)<br />

125. W. Schorr, H. Hoffmann: J. Phys. Chem., 85, 3160 (1981)<br />

126. R. Granek, M.E. Cates: J. Chem. Phys., 96, 4758 (1992)<br />

250


12 Photophysics of J Aggregates<br />

Herrmann Pschierer, Hauke Wendt, <strong>and</strong> Josef Friedrich<br />

12.1 Introduction<br />

Dye molecules of the pseudoisocyanine (PIC)-type form, under certain conditions, linear<br />

long chain aggregates [1, 2]. The conditions concern the solvent, temperature, <strong>and</strong> freezing<br />

procedure <strong>and</strong> depend, in addition, on the dye molecule <strong>and</strong> its counterion [3]. At sufficiently<br />

low concentration <strong>and</strong> quick freezing aggregate formation is hampered <strong>and</strong> one obtains<br />

the monomer. The spectral properties of the monomer show the usual features of dye<br />

molecules in frozen solution, namely a rather broad structureless inhomogeneously broadened<br />

absorption b<strong>and</strong>. It is centred around 18850 cm –1 (Fig. 12.1). On the other h<strong>and</strong>, if<br />

aggregates are formed, the spectral properties change dramatically. The wavenumber is<br />

shifted from the monomer b<strong>and</strong> by about 1450 cm –1 to the red <strong>and</strong> one or several extremely<br />

sharp b<strong>and</strong>s appear (Fig.12.1), whose intensities depend on the formation procedure. These<br />

sharp b<strong>and</strong>s are the so-called J b<strong>and</strong>s [2]. They have exciton-like character with large coherence<br />

lengths. It is the large coherence length which determines – with respect to localized<br />

states – the unusual optical properties, namely the narrowing phenomena in the inhomogeneous<br />

b<strong>and</strong>, the molecular superradiative properties [4–7], the characteristic temperature de-<br />

Figure 12.1: Low temperature (4.2 K) absorption spectra of the pseudoisocyaninechloride (PIC-Cl) aggregate<br />

<strong>and</strong> monomer.<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

251


12 Photophysics of J Aggregates<br />

pendence of the homogeneous line width [7, 8], <strong>and</strong> non-linear optical features [9–12]. In<br />

this paper we focus on the narrowing phenomena.<br />

A physically transparent interpretation of this narrowing phenomenon is based on fast<br />

moving excitons which average the inhomogeneities of the host glass to a large degree. The<br />

magnitude of the b<strong>and</strong> narrowing depends on the coherence length.<br />

p<br />

If the number of coherently<br />

coupled monomers is N c the narrowing is approximately 1/ N c . From the ratio of the<br />

inhomogeneous b<strong>and</strong> widths of monomer <strong>and</strong> aggregate, N c is estimated to be of the order<br />

of 100. The problem with this estimation is that the inhomogeneous broadening cannot be<br />

determined accurately enough since phonons <strong>and</strong> molecular vibrations may be hidden below<br />

the inhomogeneous envelope. Another problem may arise from correlation effects in the disorder,<br />

see below. Even the nature of the inhomogeneity in the J b<strong>and</strong> is not known exactly.<br />

There are various suggestions in the literature: chain length distribution [13], conformational<br />

inhomogeneity [14], <strong>and</strong> coherence length distribution [15, 16].<br />

In this paper we present a comparative pressure <strong>and</strong> electric-field tuning hole burning<br />

study between PIC monomers <strong>and</strong> aggregates. It is our goal to gain information on<br />

the coherence length of the excitons on the aggregate chains from these comparative experiments.<br />

12.2 Basic aspects of pressure <strong>and</strong> electric-field phenomena in hole<br />

burning spectroscopy of J aggregates<br />

Pressure has a twofold influence on spectral holes: it shifts the hole <strong>and</strong>, in addition, broadens<br />

it. The pressure shift of spectral holes in aggregates is quite strong. This strong pressure<br />

shift is also related to the coherence length [17], because it scales with the polarizability of<br />

the exciton chain. In this paper, however, we focus on the pressure broadening which has a<br />

very direct relation to the coherence length. In small molecules, like the PIC monomer,<br />

pressure broadening of spectral holes arises because the local configuration of host molecules<br />

surrounding the probe is changed a little bit when the lattice is compressed. These<br />

configurational changes signal a lack of spatial correlation among the glass-forming molecules.<br />

The respective broadening is inhomogeneous in nature. Hence, it will be motionally<br />

narrowed by a fast moving exciton, very much in the same way as the inhomogeneous b<strong>and</strong><br />

is narrowed. As a consequence, by comparing the pressure broadening s M of the monomer<br />

with the respective s A of the aggregate, we directly get the number N c of molecules within<br />

the coherence length [4–6, 11],<br />

r<br />

s M<br />

ˆ 2…N c ‡ 1†<br />

: …1†<br />

s A 3<br />

The advantage of a hole burning experiment is that the change in line width under<br />

pressure can be measured much more accurately than the inhomogeneous b<strong>and</strong>s. No correc-<br />

252


12.3 Experimental<br />

tion due to underlying phonons, or vibrations, or b<strong>and</strong>shape asymmetries is necessary. Hole<br />

burning in J aggregates was demonstrated first by de Boer et al. [18].<br />

As to the influence of an electric field the situation is somewhat different: the monomer<br />

has a dipole moment which gives rise to a first order Stark effect. The result is a splitting<br />

<strong>and</strong> a broadening of the hole. In addition, the local environment determined by the host<br />

glass may induce a dipole moment which can make a significant contribution to the field<br />

broadening as well [19–21].<br />

In the aggregate, a variety of things can happen: the structure of the chain may have<br />

an inversion center. As a consequence, the broadening will be reduced. In addition, similar<br />

to the pressure phenomena, the field broadening due to the environmentally induced dipole<br />

moments may be motionally narrowed. In this case motional narrowing can be interpreted in<br />

the sense that the influence of the matrix fields averages to zero within a scale given by the<br />

coherence length.<br />

12.3 Experimental<br />

So far, we performed comparative pressure experiments for two different PIC systems,<br />

namely PIC-Cl <strong>and</strong> PIC-I in an ethyleneglycol-water glass with a mixing volume ratio of<br />

1:1. In the following, PIC st<strong>and</strong>s for 1,1'-diethyl-2,2'-cyanine. Hole burning was performed<br />

with a ring dye laser pumped by an argon ion laser. The laser b<strong>and</strong> width was about 1 MHz.<br />

The respective scan range covered 30 GHz. The holes were detected in transmission. In the<br />

J b<strong>and</strong>, typical burning times were of the order of one minute. The respective power levels<br />

varied between 0.01 <strong>and</strong> 0.1 mW. The hole burning efficiency is very low in the monomers.<br />

Hence, burning times of the order of 20 minutes were used at power levels of about 100 mW.<br />

The aggregate spectra were measured at 4.2 K, whereas the monomer spectra were<br />

measured at 1.5 K. The reasons are the following: in the monomer, the holes are subject to<br />

thermal line broadening, hence are much broader at 4.2 K than at 1.5 K. This lowers the accuracy.<br />

As mentioned above, the photoreactive quantum yield in the monomer is extremely<br />

low. Hence, it is much harder to burn a hole at 4 K than at 1.5 K. However, at 1.5 K the<br />

pressure range is rather limited because He solidifies at a level of 2.4 MPa. The narrow<br />

pressure range at 1.5 K, on the other h<strong>and</strong>, limits the accuracy for the aggregates. In this<br />

case pressure broadening is so low that one needs a larger range to get significant results.<br />

Consequently, the respective experiments were performed at 4.2 K. We note that thermal<br />

line broadening between 1.5 K <strong>and</strong> 4.2 K is of no concern for the aggregate. Due to the molecular<br />

superradiance, the line width is not affected by phonons. Hence, there is no significant<br />

change with temperature in this range [8].<br />

For the Stark experiments on J aggregates holes were burned with the ring dye laser<br />

but scanned with a pulsed dye laser system because of its larger scan range. The aggregate<br />

solution was dispersed on an indium tin oxide (ITO) coated glass substrate. Maximum field<br />

strengths were about 300 kV/cm.<br />

253


12 Photophysics of J Aggregates<br />

The monomers were investigated in a glass cuvette placed between two electrodes. In<br />

this case, maximum field strengths were about 10 kV/cm <strong>and</strong> both, burning <strong>and</strong> scanning of<br />

the holes, were performed with a pulsed dye laser system.<br />

12.4 Results<br />

In Fig. 12.1, we compare the inhomogeneous long wavelength absorption b<strong>and</strong>s of PIC-Cl<br />

monomer <strong>and</strong> aggregate. Figure 12.2 shows for both cases as holes deform under isotropic<br />

pressure conditions.<br />

Figure 12.2: Behaviour of spectral holes in the aggregate (left) <strong>and</strong> monomer (right) under different<br />

pressures. Burn-frequencies: 17334 cm –1 for the aggregate, 18727 cm –1 for the monomer. Temperature<br />

4.2 K (aggregate) <strong>and</strong> 1.5 K (monomer). Lorentzian fit curves are indicated.<br />

Figures 12.3 <strong>and</strong> 12.4 show the influence of pressure on the line width of monomer<br />

<strong>and</strong> aggregate. In the insets, it is demonstrated for the monomers on an exp<strong>and</strong>ed scale that,<br />

despite the small pressure range, line broadening is linear with pressure <strong>and</strong> that the respective<br />

slope can be determined with significant accuracy. For the monomer of PIC-Cl we measured<br />

(0.18 +0.02) cm –1 /MPa. For PIC-I a value of (0.21 +0.01) cm –1 /MPa was found. We<br />

stress that values of the same order of magnitude have been found for a series of similar<br />

sized molecules in organic glasses [22–24].<br />

Going from the monomer to the aggregate a dramatic decrease in pressure broadening<br />

occurs. We measured (0.020 +0.001) cm –1 /MPa <strong>and</strong> (0.026+0.001) cm –1 /MPa for PIC-Cl<br />

<strong>and</strong> PIC-I aggregates, respectively. The respective ratios in the broadening per pressure are<br />

(9+1) <strong>and</strong> (8 +1).<br />

Figure 12.5 shows the broadening of a spectral hole for the PIC-I monomer (a) <strong>and</strong><br />

aggregate (b) in an electric field. For the monomer the field-induced broadening is quite<br />

254


12.4 Results<br />

Figure 12.3: Pressure broadening of spectral holes for PIC-Cl monomer <strong>and</strong> aggregate. The inset shows<br />

the data for the monomer on a smaller scale.<br />

Figure 12.4: Pressure broadening of spectral holes for PIC-I monomer <strong>and</strong> aggregate. The inset shows<br />

the data for the monomer on a smaller scale.<br />

strong (about 1 GHz cm/kV) <strong>and</strong> linear with the applied Stark field E St . In the aggregate the<br />

field-induced broadening is dramatically reduced. The fitted curve in Fig. 12.5 has a strong<br />

quadratic component but there is a linear contribution as well. Comparing the linear regime<br />

of the aggregate with the respective one of the monomer, it is obvious that the field-induced<br />

broadening is reduced by a factor of about 150.<br />

255


12 Photophysics of J Aggregates<br />

Figure 12.5: Electric field-induced broadening (Stark effect) of spectral holes for pseudoisocyanineiodide<br />

monomer (a) <strong>and</strong> aggregate (b). Note the different scales of the applied Stark field for monomer<br />

<strong>and</strong> aggregate.<br />

12.5 Discussion<br />

12.5.1 Pressure phenomena<br />

First, let us address broadening of the holes in the monomer b<strong>and</strong>. The reason for pressure<br />

broadening is based on the fact that a large variety of microscopic environments in amorphous<br />

host materials can correspond with the same absorption energy of the probe molecule.<br />

This degeneracy is partly lifted under pressure <strong>and</strong> is reflected in pressure broadening. The<br />

magnitude of this broadening is determined through the magnitude of the probe solvent interaction<br />

<strong>and</strong> through the similarity of the microscopic probe-lattice interaction configurations<br />

with their respective change under pressure [25]. This similarity is expressed in a specific<br />

degree of correlation. In a rather perfectly ordered crystal, for instance, this degree of<br />

correlation is close to 1 <strong>and</strong>, correspondingly, the pressure broadening is close to zero as has<br />

been observed [26]. In amorphous solids, on the other h<strong>and</strong>, the respective correlation is<br />

low <strong>and</strong> broadening is rather large. Typical values are of the order of 0.1 to 0.2 cm –1 /MPa<br />

[22–24]. As is obvious from the discussion above, pressure broadening in solids is inhomogeneous<br />

in nature.<br />

In the aggregates the wavefunction is delocalized over N c monomer units. Then, we<br />

can consider two cases:<br />

a) the microenvironments of the individual monomers in the aggregate are statistically independent.<br />

In this case the site energy of molecule n is completely independent from molecule<br />

n +1;<br />

b) there is a finite correlation length l 0 , i. e. within l 0 the site energies of the molecules<br />

forming the aggregate are correlated to some degree.<br />

256


12.5 Discussion<br />

For case a), with statistically independent microenvironments, it was shown by several<br />

authors that the inhomogeneous line width of the monomers q forming the aggregate is narrowed<br />

according to Eq. 1, i. e. roughly by a factor N c<br />

1 . This narrowing is known as exchange<br />

narrowing <strong>and</strong> it can be interpreted in a way that a fast moving exciton on an aggregate<br />

chain averages over the local inhomogeneities. The amount of averaging is determined<br />

by the coherence length of the exciton. Hence, the coherence length can be determined from<br />

the ratio of the inhomogeneous b<strong>and</strong> widths of aggregate <strong>and</strong> monomer. The exact determination<br />

of these b<strong>and</strong> widths is a problem, because of hidden states <strong>and</strong> b<strong>and</strong> asymmetries.<br />

However, we stressed above that pressure broadening of spectral holes is inhomogeneous in<br />

nature. It reflects in a specific way the disorder in the local environments <strong>and</strong>, hence, should<br />

be narrowed by a fast moving exciton in quite the same way as the inhomogeneous b<strong>and</strong>. If<br />

so, we can determine the number N c of monomers within the coherence length from Eq. 1,<br />

by comparing the pressure broadening in the monomer with the respective one in the aggregate.<br />

We found (120+32) for PIC-Cl <strong>and</strong> (101+24) for PIC-I. These numbers may<br />

slightly depend on the solvent <strong>and</strong> freezing procedure. Generally speaking they fit to what is<br />

known of J aggregates.<br />

For case b), with correlated site energies of the monomers in the chain, the situation<br />

becomes quite different. Equation 1 has to be substituted by<br />

s<br />

s M<br />

ˆ 1 b<br />

s A 1 ‡ b 2…N c ‡ 1†<br />

: …2†<br />

3<br />

Equation 2 is approximative <strong>and</strong> holds only for sufficiently small b values.<br />

Again, b is a degree of correlation between the site energies of the monomers in the<br />

chain. It can be written as [27]<br />

<br />

b ˆ exp<br />

<br />

1<br />

l 0<br />

…3†<br />

with l 0 being the correlation length, i. e. beyond l 0 the site energies of the two monomers are<br />

statistically independent. As b goes to zero, we have case a) discussed above. As b goes to<br />

unity, s A approaches s M <strong>and</strong> there is no motional narrowing at all. This latter case cannot be<br />

directly derived from Eq. 2, since b = 1 is beyond the respective validity range.<br />

The main point is that, if there were correlation in the site energies, it would not be<br />

possible any more to determine N c from a linear optical experiment alone because Eq. 2<br />

would contain two unknown quantities. There are indications from non-linear experiments<br />

that l 0 is finite [28]. Another way to estimate b is via the ratio of the radiative rates of<br />

monomer <strong>and</strong> aggregate. This ratio is also determined by the number N c of coherently<br />

coupled molecules [7, 29] but is not influenced by correlation effects in the site energies,<br />

g A ˆ N c g M ;<br />

…4†<br />

where g A is the radiative rate of the aggregate <strong>and</strong> g M the radiative rate of the monomer. In<br />

an earlier paper we measured the homogenous line width down to 0.300 K [8]. The mea-<br />

257


12 Photophysics of J Aggregates<br />

sured value corresponded with a lifetime of about 30 ps. Since the fluorescence quantum<br />

yield in the J b<strong>and</strong> is supposed to be close to unity, the aggregate lifetime is governed by radiative<br />

processes. With g M –1 = 3.7 ns [30], we get from Eq. 4 an N c value of about 123 which<br />

is close to what we found from our pressure tuning experiments. Hence, according to this estimation,<br />

b should be close to zero.<br />

We stress however, that the results for N c in the literature, as obtained with various<br />

techniques, are subject to large variations. Clearly, more work must be done to achieve unambiguous<br />

results.<br />

12.5.2 Electric field-induced phenomena<br />

The electric field-induced phenomena in the J b<strong>and</strong> are much more complex than the pressure<br />

phenomena <strong>and</strong> not yet understood. Firstly we note that the reduction of the linear component<br />

in the Stark effect is in agreement with what we expected from motional narrowing<br />

effects. What was not expected at all, however, is the order of magnitude which is more than<br />

a factor of 10 larger as compared to the pressure experiments. We conclude that additional<br />

mechanisms must play a role. One such mechanism is the loss of the permanent dipole moment<br />

upon aggregate formation. The absence of a dipole moment has severe structural implications<br />

which are not clear at the present time. Our result is also at variance with experiments<br />

by Misawa et al. [31], who reported huge changes in the static dipole moment of J aggregates,<br />

which were attributed to their enhanced size. However, since these experiments<br />

were performed with spin-coated J aggregates, it may possibly be that geometric effects play<br />

an important role.<br />

Acknowledgements<br />

The authors thank the Deutsche Forschungsgemeinschaft (Sonderforschungsbereich 213,<br />

B15) <strong>and</strong> the Fonds der Chemischen Industrie for financial support.<br />

258


References<br />

References<br />

1. G. Scheibe: Kolloidzeitschrift, 82, 1 (1938)<br />

2. E. E. Jelley: Nature, 139, 631 (1937)<br />

3. E. Daltrozzo, G. Scheibe, K. Gschwind, F. Haimerl: Photographic Science <strong>and</strong> Engineering, 18,<br />

441 (1974)<br />

4. E. W. Knapp: Chem. Phys., 85, 73 (1984)<br />

5. E. W. Knapp: P. O. J. Scherer, S. F. Fischer, Chem. Phys. Lett., 111, 481 (1984)<br />

6. F. C. Spano, S. Mukamel: Phys. Rev. A, 40, 5782 (1989)<br />

7. H. Fidder, J. Knoester, D. A. Wiersma, Chem. Phys. Lett., 171, 529 (1990)<br />

8. R. Hirschmann, J. Friedrich: J. Chem. Phys., 91, 7988 (1989)<br />

9. S. Kobayashi, F. Sasaki: Nonlinear Optics, 4, 305 (1993)<br />

10. F. C. Spano, S. Mukamel: Phys. Rev. Lett., 66, 1197 (1991)<br />

11. F. C. Spano, S. Mukamel: J. Chem. Phys., 91, 683 (1989)<br />

12. F. C. Spano, J. Knoester: in: Advances in Magnetic <strong>and</strong> Optical Resonance, Vol. 18, 117 (1994)<br />

13. B. Kopainsky, W. Kaiser: Chem. Phys. Lett., 88, 357 (1982)<br />

14. R. Hirschmann, W. Köhler, J. Friedrich, E. Daltrozzo: Chem. Phys. Lett., 151, 60 (1988)<br />

15. H. Fidder, D. A. Wiersma: Phys. Rev. Lett., 66, 1501 (1991)<br />

16. H. Fidder: Thesis, University of Groningen, (1993)<br />

17. R. Hirschmann, J. Friedrich: JOSA B, 9, 813 (1992)<br />

18. S. de Boer, K. J. Vink, D. A. Wiersma: Chem. Phys. Lett., 137, 91 (1987)<br />

19. M. Maier: Appl. Phys. B, 41, 73 (1986)<br />

20. A. J. Meixner, A. Renn, S. E. Bucher, U. P. Wild: J. Phys. Chem., 90, 6777 (1986)<br />

21. L. Kador, D. Haarer, R. Personov: J. Chem. Phys., 86, 5300 (1987)<br />

22. G. Gradl, J. Zollfrank, W. Breinl, J. Friedrich: J. Chem. Phys., 94, 7619 (1991)<br />

23. J. Zollfrank, J. Friedrich: J. Phys. Chem., 96, 7887 (1992)<br />

24. H. Pschierer, J. Friedrich, H. Falk, W. Schmitzberger: J. Phys. Chem., 97, 6902 (1993)<br />

25. B. B. Laird, J. L. Skinner: J. Chem. Phys., 90, 3274 (1989)<br />

26. P. Schellenberg, J. Friedrich, J. Kikas: J. Chem. Phys., 100, 5501 (1994)<br />

27. J. Knoester: J. Chem. Phys., 99, 8466 (1993)<br />

J. Knoester: J. Lumin., 58, 107 (1994)<br />

28. J. R. Durrant, J. Knoester, D. A. Wiersma: Chem. Phys. Lett., 222, 450 (1994)<br />

29. J. Grad, G. Hern<strong>and</strong>ez, S. Mukamel: Phys. Rev. A, 37, 3835 (1988)<br />

30. H.-P. Dorn, A. Müller: Chem. Phys. Lett., 130, 426 (1986)<br />

31. K. Misawa, K. Minoshima, H. Ono, T. Kobayashi: Chem. Phys. Lett., 220, 251 (1994)<br />

259


13 Convection Instabilities in Nematic Liquid Crystals<br />

Lorenz Kramer <strong>and</strong> Werner Pesch<br />

13.1 Introduction<br />

Pattern formation in hydrodynamic instabilities has been studied intensely over the last decades<br />

[1, 2]. Although Rayleigh-Bénard convection (RBC) in simple fluids has been the<br />

prime example [3], the rich variety of scenarios found in nematic liquid crystals (LCs) has<br />

attracted increased attention.<br />

LCs are materials made up of highly anisotropic organic molecules in a phase that reflects<br />

this anisotropy. The class of nematic LCs (nematics) are fully liquid without longrange<br />

translational but with long-range uniaxial orientational ordering of the molecules. As<br />

a result of the coupling of the molecular alignment axis (described by the director ^n) with<br />

the (mass) flow electric or thermal current the hydrodynamic equations involve numerous<br />

non-linearities (Section 13.2), which easily lead to instabilities when a state of non-equilibrium<br />

is maintained [4]. Convective flow can be driven electrically through space charges<br />

that naturally arise in an anisotropic conductor in the presence of spatial variations, the electrohydrodynamic<br />

convection (EHC), or thermally through buoyancy forces, the Rayleigh-Bénard<br />

convection (RBC). EHC has attracted more attention <strong>and</strong> will play a major role in this<br />

review.<br />

The study of EHC by Williams [5], Kapustin <strong>and</strong> Larinova [6] in 1963 initiated extensive<br />

experimental work in the typical thin-layer geometry shown in Fig. 13.1a. The nematic<br />

is s<strong>and</strong>wiched between glass plates (separation d * 10–100 µm) with transparent electrodes.<br />

The surfaces are treated to provide (ideally) uniform anchoring of the director, in most<br />

cases along the x direction (planar or homogeneous alignment) but sometimes also in the z<br />

direction (homeotropic alignment) 1 . Above an applied voltage V c > 10 V (typically low-frequency<br />

ac) convection rolls appear with associated director distortions, which are easily detected<br />

optically. The spacing of the rolls is of order d except in the higher frequency dielectric<br />

range. Figure 13.1b shows a typical pattern with the rolls in the y direction normal to<br />

the undistorted director.<br />

1 We will concentrate here on the planar case. Some remarks about other alignments are made in<br />

Section 13.5.<br />

260 Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


13.1 Introduction<br />

E || z^<br />

^ y<br />

z^<br />

^ z<br />

x^ v<br />

Director n<br />

Figure 13.1: a) Cell geometry with section of a roll pattern for EHC (planar configuration). E = electric<br />

field, v=velocity; b) Normal roll pattern for EHC with a dislocation.<br />

The mechanism for instability in EHC based mainly on space charges generated by<br />

preferential conduction along ^n (charge focusing, see Section 13.2) was suggested by Carr<br />

[7] <strong>and</strong> then incorporated into a first one-dimensional model by Helfrich [8]. Subsequently<br />

the linear theory, giving the onset of the instability <strong>and</strong> in principle the pattern up to degeneracy,<br />

was generalized to include the common case of ac driving [9, 10], a rigorous two-dimensional<br />

analysis [11], <strong>and</strong> finally a full three-dimensional treatment, signalizing the beginning<br />

of a renewed interest in the subject. For references see e. g. [12], for a comparison between<br />

the experimental <strong>and</strong> theoretical threshold see Fig. 13.2 a. Unfortunately already the linear<br />

theory is a numerical problem, but useful analytic approximations are possible [12–15].<br />

The need for the full three-dimensional theory became particularly evident from the<br />

experimental work of Ribotta <strong>and</strong> co-workers [16, 17], where three-dimensional structures<br />

(oblique rolls) were observed at threshold, <strong>and</strong> from systematic measurements by Kai <strong>and</strong><br />

co-workers [18, 19] under well-defined conditions. Figure 13.2b shows the oblique rolls,<br />

100<br />

Threshold Voltage [V]<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 10 20 30 40<br />

Driving Frequency f [Hz]<br />

Figure 13.2: a) Threshold curve for EHC as function of frequency. Experimental points from [20];<br />

b) Zigzag pattern after increasing the voltage in Fig. 13.1b.<br />

261


13 Convection Instabilities in Nematic Liquid Crystals<br />

which in this case bifurcated from normal rolls after an increase of V. Note the appearance<br />

of grain boundaries separating domains with the two symmetry degenerate directions (zig<br />

<strong>and</strong> zag). At threshold the boundary would become less sharp.<br />

As the theory progressed exp<strong>and</strong>ing into a weakly non-linear analysis providing the<br />

Ginzburg-L<strong>and</strong>au amplitude equation description [12, 14] <strong>and</strong> eventually also including<br />

mean flow effects that allow to capture the transition from ordered periodic to weakly turbulent<br />

patterns [21–24] there was also a revival of experimental activity, For a phenomenological<br />

treatment, see Refs. [25, 26]. Some milestones were the identification of the Eckhaus instability<br />

(Section 13.3) that gives limits of the stable wave number b<strong>and</strong> [27–29], the characterization<br />

of the instability that may lead to pattern turbulence [36, 37] (Fig. 13.3b), <strong>and</strong><br />

the structure <strong>and</strong> dynamics of single dislocations <strong>and</strong> their interaction [30–33]. For a comparison<br />

of the experiments [32, 34] with the results obtained from the Ginzburg-L<strong>and</strong>au theory<br />

[35] see Fig. 13.3a.<br />

8<br />

5<br />

7<br />

6<br />

exp.<br />

exp. ξ 1 , ξ 2 , τ<br />

theo. ξ 1 , ξ 2 , τ<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

-0.01 0.01 0.02 0.03 0.04<br />

0.00<br />

∆q [µm -1 ]<br />

Figure 13.3: a) Climb velocity of a single defect vs. wave number mismatch. Parameters of the GLE<br />

are either from experiments [32] (solid curve) or from hydrodynamical calculations [12] (broken line);<br />

b) Snapshot of a defect turbulent pattern.<br />

Further highlights were the identification of thermal noise slightly below threshold by<br />

Rehberg et al. [38–40] <strong>and</strong> finally the clear identification of a Hopf bifurcation leading to<br />

travelling rolls or waves in sufficiently thin layers, below about 50 µm <strong>and</strong> clean material<br />

(low conductivity) by Refs. [18, 41–43]. It is not possible to explain the Hopf bifurcation<br />

within the conventional theoretical framework the st<strong>and</strong>ard model (SM), see Section 13.2,<br />

where the LC is treated as an anisotropic ohmic conductor. Indeed some of the theoretical<br />

effort, in particular the inclusion of the rather complicated flexoelectric terms [14, 44–46],<br />

was aimed primarily at resolving this problem. The situation is further complicated by the<br />

fact that the bifurcation is often observed to be slightly subcritical, i. e. with a very small<br />

hysteresis [38–40, 47], whereas the theory predicts a supercritical bifurcation.<br />

Very recently an extension of the st<strong>and</strong>ard theory, the weak-electrolyte model (WEM),<br />

has been worked out where electric transport in the nematic is described in terms of two mobile<br />

ion species of opposite charge which are coupled via a slow dissociation-recombination<br />

reaction <strong>and</strong> whose densities are treated as dynamic variables [48]. Some results of the<br />

262


13.1 Introduction<br />

WEM will be discussed in Section 13.5 together with experiments in the material I52 [49].<br />

Whether the WEM can also capture the subcritical bifurcation remains unclear.<br />

In almost all the measurements the st<strong>and</strong>ard reference material 4-methoxybenzylidene-4'-n-butyl-aniline<br />

(MBBA) or a mixture, Merck Phase V, have been used, sometimes<br />

doped with an ionic substance. MBBA is the only room-temperature nematic with dielectric<br />

anisotropy e a < 0 where all the material parameters have been measured. For tabulated values<br />

see e. g. Ref. [12]. Unfortunately, it is a Schiff base <strong>and</strong> rather unstable when exposed<br />

to moisture. Therefore, it is difficult to control the long-time conductivity in situ. Thus the<br />

recent successful introduction of the very stable material 4-ethyl-2-fluoro-4'-[2-(trans-4-npentylcyclohexyl)-ethyl]-biphenyl<br />

(I52) doped with iodine is very promising [50, 51]. This<br />

material exhibits at low external frequencies strongly oblique travelling rolls which bifurcate<br />

supercritically, leading to a particularly interesting scenario. 2<br />

In RBC the traditional instability mechanism by buoyancy forces is enhanced considerably<br />

by preferential conduction of heat parallel to the director leading to a heat focusing<br />

effect, as was first shown by Dubois-Violette in 1971 [52]. After some amount of experimental<br />

<strong>and</strong> theoretical work, for a summary see Ref. [53], Feng et al. [54] recently presented a<br />

fully three-dimensional treatment including the weakly non-linear analysis. Subsequent experiments<br />

in thick layers (some millimetres) with an additional stabilizing magnetic field H x<br />

in the x direction (Fig. 13.1a) have substantiated several predictions [55]. In Fig.13.3 a a zigzag<br />

pattern is shown which arises with increasing magnetic field after normal rolls (n is horizontal).<br />

Further increase of H x leads to a larger angle of obliqueness with superposition of<br />

the two roll systems (Fig. 13.3b) <strong>and</strong> eventually to rolls which are parallel to n. The very<br />

common nematic 4-n-pentyl-4'-cyanobiphenyl (5CB) was used, whose relevant material<br />

parameters have all been measured. 3<br />

Figure 13.4: a) Oblique roll pattern for RBC in the planar configuration [55]; b) Superposition of zig<br />

<strong>and</strong> zag for RBC [55].<br />

2 In the note added at the end of this Chapter some very recent developments in EHC have been summarized.<br />

3 Some very recent progress is summerized at the end of Section 13.4.<br />

263


13 Convection Instabilities in Nematic Liquid Crystals<br />

In this review we will address mostly the developments of the past twelve years, characterized<br />

by substantial progress in qualitative as well as quantitative underst<strong>and</strong>ing of the<br />

various instabilities. After introducing <strong>and</strong> explaining the basic equations (Section 13.2) the<br />

theoretical concepts of the linear <strong>and</strong> weakly non-linear analysis will be presented<br />

(Section 13.3). In Section 13.4 <strong>and</strong> Section 13.5 the results for the two systems, introduced<br />

above, are discussed in particular in the light of recent experiments. <strong>Final</strong>ly, we will list topics<br />

which are omitted due to space limitations <strong>and</strong> we comment on some perspectives for<br />

future work (Section 13.6).<br />

For a classical review of convective instabilities in LCs see Ref. [56] <strong>and</strong> for EHC one<br />

may consult the books of Blinov [57], Pikin [58], <strong>and</strong> recent review articles [14, 15, 59–63].<br />

RBC has been reviewed by Barrat [53] <strong>and</strong> very recently by Ahlers [64].<br />

13.2 Basic equations <strong>and</strong> instability mechanisms<br />

13.2.1 The director equation<br />

The macroscopic nematodynamic equations describe the dynamics of the slowly relaxing<br />

variables, which usually are either connected with conservation laws or with the Goldstone<br />

modes of the spontaneously broken symmetries. To formulate them we will follow the traditional<br />

approach [65–67] rather than the one based more directly on the principles of hydrodynamics<br />

<strong>and</strong> irreversible thermodynamics [68]. In the nematic state isotropy is spontaneously<br />

broken <strong>and</strong> the averaged molecular alignment singles out an axis whose orientation<br />

defines the director ^n, i. e. an object that has the properties of a unit vector with ^n =–^n.<br />

The static properties are conveniently expressed in terms of a free energy density whose orientational<br />

elastic part is given by [69]<br />

f elast ˆ 1<br />

2 k 11…r ^n† 2 ‡ 1 2 k 22…^n …r^n†† 2 ‡ 1 2 k 33…^n …r^n†† 2 : …1†<br />

The elastic constants k 11 , k 22 , <strong>and</strong> k 33 pertain to the three basic deformations splay,<br />

twist, <strong>and</strong> bend, respectively. For typical nematics with prolate molecules one has<br />

k 33 > k 11 > k 22 <strong>and</strong> k ii * 10 –11 N.<br />

Electric <strong>and</strong> magnetic fields E <strong>and</strong> H exert torques on ^n that can be derived from an<br />

additional contribution to the free energy<br />

f em ˆ 1<br />

2 m 0w a …^n H† 2 1<br />

2 e 0 e a …^n E† 2 E P f lexo : …2†<br />

Here w a = w k – w k <strong>and</strong> e a = e k – e k are the relative diamagnetic <strong>and</strong> dielectric anisotropies,<br />

so that the uniaxial susceptibility <strong>and</strong> dielectric tensors can be written in Cartesian coordinates<br />

in the form<br />

264


13.2 Basic equations <strong>and</strong> instability mechanisms<br />

w ij ˆ w ? d ij ‡ w a n i n j ; e ij ˆ e ? ij ‡ e a n i n j …3†<br />

<strong>and</strong> the flexoelectric polarization is [70]<br />

P f lexo ˆ e 1 ^n…r ^n†‡e 3 …^n r†^n :<br />

…4†<br />

Static director configurations are obtained by minimizing the total free energy<br />

F = R dV ( f elast + f em ) under the condition n j n j = 1 with suitable boundary conditions, which<br />

results in equating to zero the torque density G on the director. The generalization to dynamic<br />

situations is usually done by defining a vector<br />

S ˆ S rate ‡ S cons ‡ S diss<br />

…5†<br />

<strong>and</strong> requiring the balance of torques [65, 71, 72]<br />

ˆ ^n S ˆ 0 :<br />

…6†<br />

S has the form<br />

S rate ˆ g 1 N;<br />

<br />

<br />

1<br />

N ˆ d^n=dt …r v†^n ;<br />

2<br />

…7†<br />

S cons ˆ dF=d^n ; …8†<br />

S dissji<br />

ˆ g 2 A ij n j ; A ij ˆ 1 <br />

uv i<br />

‡ uv <br />

j<br />

:<br />

2 ux j ux i<br />

…9†<br />

Here d/dt = u t + v 7r denotes the usual material derivative <strong>and</strong> in N the rigid rotation<br />

part x = 1 2 …r v† of the fluid has been subtracted out. A notation like ux j(n i )=u j n i = n i,j is<br />

used freely. g 1 <strong>and</strong> g 2 are called rotational viscosities, which couple the flow field v to the director.<br />

To confirm that Eq. 6 is indeed only a rate equation for the director one may take the<br />

vector product of Eq. 6 with ^n leading to<br />

g 1 ‰ d^n=dt 1=2…r v†^n Š ˆ …1 ^n^n T †…S cons ‡ S diss † : …10†<br />

The typical relaxation time of the director in the thin-layer geometry of EHC is easily<br />

seen to be<br />

t d ˆ g 1 d 2 =…p 2 k 11 † ;<br />

…11†<br />

where we chose the splay elastic constant as a representative; obviously g 1 > 0 has to hold.<br />

t d is typically in the order of 1 s.<br />

It can be seen that Eq. 6 <strong>and</strong> Eq. 10 have only two independent components which<br />

can be made explicit by transforming into one of the two local coordinate systems<br />

^n; ^x ^n; ^n …^x ^n† ;<br />

…12†<br />

265


13 Convection Instabilities in Nematic Liquid Crystals<br />

^n; ^z ^n; ^n …^z ^n† :<br />

…13†<br />

The first one becomes singular when ^n is parallel to ^x <strong>and</strong> is thus not suitable for planar<br />

alignment. Similarly, the other one is not practical for homeotropic alignment.<br />

13.2.2 The velocity field<br />

The generalized Navier-Stokes equation for the velocity field v follows from momentum<br />

balance<br />

r m<br />

dv<br />

dt ˆ f ‡rT ;<br />

…14†<br />

where r m is the mass density, f the bulk force to be discussed below, <strong>and</strong> T the stress tensor<br />

with components<br />

T ij ˆ pd ij<br />

uF<br />

un k;i<br />

<br />

Eˆ0<br />

n k;j ‡ t ij ;<br />

…15†<br />

<strong>and</strong> p is pressure. The viscous stress tensor for an incompressible nematic contains the six<br />

Leslie shear viscosity coefficients a i [72],<br />

t ij ˆ a 1 n k n m A km n i n j ‡ a 2 n i N j ‡ a 3 n j N i ‡ a 4 A ij ‡ a 5 n i n k A kj ‡ a 6 n j n k A ki :<br />

…16†<br />

It is instructive to consider the three simple geometries for plane parallel shear flow,<br />

the Miesovicz geometries [73], corresponding to the orientations of the director relative to<br />

the flow axis <strong>and</strong> the shear gradient. Choosing v = u (z) ^x one has the effective shear viscosities<br />

4<br />

1) for ^n = ^z along the shear gradient ?Z 1 =(a 4 + a 5 – a 2 )/2, G y = a 2 u z ;<br />

2) for ^n = ^x along the flow axis ?Z 2 =(a 3 + a 4 – a 6 )/2, G y = a 3 u z ;<br />

3) for ^n = ^y along the shear gradient ?Z 3 = a 4 /2, G = 0.<br />

The value of the non-vanishing component of the torque G on the director is also given.<br />

Another positive effective viscosity is Z 0 = a 1 + a 4 + a 5 + a 6 . All shear viscosities are<br />

typical of order 10 –1 kg m –1 s –1 .<br />

From the Onsager reciprocity relations one finds [74]<br />

a 6 a 5 ˆ a 2 ‡ a 3 …17†<br />

4 Unfortunately the definitions of Z 1 <strong>and</strong> Z 2 are not universally accepted. Our definition is adopted by<br />

Blinov [57] whereas the definitions in the review articles [67] <strong>and</strong> [73] are in the reverse order.<br />

266


13.2 Basic equations <strong>and</strong> instability mechanisms<br />

<strong>and</strong> also<br />

g 1 ˆ a 3 a 2 ; g 2 ˆ a 3 ‡ a 2 : …18†<br />

Note that the flow-alignment parameter l =–g 2 /g 1 is a reversible quantity. For l >1<br />

the director can align in the flow plane at a fixed angle b with respect to the velocity (xaxis),<br />

where tan 2 b =(l–1)/(l +1)=a 3 /a 2 .Forl < 1 there is tumbling [73].<br />

The pressure is to be determined from the incompressibility condition<br />

rv ˆ 0 :<br />

…19†<br />

An efficient way to implement this relation is to represent the velocity field in terms<br />

of a toroidal <strong>and</strong> a poloidal potential g <strong>and</strong> f, respectively [75]<br />

v…x; y; z† ˆr^zg ‡r…r^zf †ˆeg ‡ df<br />

…20†<br />

with<br />

e T ˆ…@ y ; @ x ; 0† ; d T ˆ…@ 2 xy ;@2 yz ; @2 xx @ 2 yy † ; …21†<br />

where z corresponds to the coordinate perpendicular to the plane of the layer. Applying e<br />

<strong>and</strong> d on Eq. 14 gives two equations for g <strong>and</strong> f eliminating the pressure.<br />

The typical relaxation time for a velocity field in a thin-layer geometry is<br />

t visc = r m d 2 /(p 2 Z 1 ). This time is usually of the order of 10 –5 s, much shorter than the others.<br />

Thus the velocity field can usually be treated adiabatically (neglect of inertial terms).<br />

13.2.3 Electroconvection<br />

13.2.3.1 The st<strong>and</strong>ard model<br />

Now we come to the additional equations that are specific to the processes driving the instability.<br />

In EHC the bulk force in the Navier-Stokes Equation 14 is derived from the Maxwell<br />

stress tensor, which here reduces to<br />

f ˆ r el E ‡…P r†E; P ˆ D E : …22†<br />

The equation determining the charge density r el is obtained from charge conservation<br />

<strong>and</strong> the Poisson law<br />

dr el<br />

dt<br />

‡rj el ˆ 0; j el ˆ s E ;<br />

r el ˆrD; D ˆ e E ‡ P f lexo :<br />

(23)<br />

267


13 Convection Instabilities in Nematic Liquid Crystals<br />

In the SM the usual assumption of an anisotropic but fixed ohmic conductivity is<br />

made. The conductivity tensor r has the same form as the other material tensors with<br />

s a = s k – s k , see Eq. 3.<br />

Equations 23 are easily seen to lead to charge relaxation with the time scale<br />

t q ˆ e 0 e ? =s ?<br />

…24†<br />

which is typically of the order of 10 –3 s. Sometimes a dopant is added to the LC to obtain<br />

sufficient or well controlled conductivity (not much is needed).<br />

Now we are in a position to discuss the basic driving mechanism for EHC. The important<br />

point is that in almost all nematics s a is substantially positive, typically s a /s k & 0.3–1.<br />

Choosing materials with negative or only slightly positive dielectric anisotropy e a , here the<br />

materials show great diversity, one easily sees that in the presence of an applied field E 0 <strong>and</strong><br />

with a small spatial variation (fluctuation) of the director n a space charge r el results. For r el =0<br />

there is no solution of Eqs. 23. Roughly speaking, the charges are focused at locations where<br />

the director bends. The bulk force in the Navier-Stokes equation may then overcome viscous<br />

stresses <strong>and</strong> drive a velocity field v. Via the viscous coupling this may enhance the spatial variation<br />

of the director <strong>and</strong> thus generate an instability. The threshold voltage is for low frequency<br />

<strong>and</strong> for materials with not too large dielectric anisotropy of order V c 2 &p 2 k 11 /(s a t q )<br />

<strong>and</strong> the introduction of the reduced control parameter R = V 2 (s a t q )/(p 2 k 11 ) is often useful.<br />

13.2.3.2 The weak electrolyte model<br />

The SM fails in particular to describe the Hopf bifurcation leading to travelling rolls, which<br />

is observed quite frequently. For a Hopf bifurcation two processes that compete on a comparable<br />

time scale are necessary. In our case, however, the director relaxation is much slower<br />

than the other processes <strong>and</strong> thus determines the dynamics. Charge relaxation can usually<br />

also be treated adiabatically. In addition, director relaxation <strong>and</strong> charge relaxation do not<br />

compete but rather support each other usually, which also excludes a Hopf bifurcation.<br />

Thus another slow process appears to be operating. A recently proposed model assumes<br />

that this process is a relaxation of the mobile ion densities n + <strong>and</strong> n – on the time scale<br />

t rec , which may result from a dissociation-recombination reaction. One then gets if singlecharged<br />

ions are assumed<br />

where<br />

r el ˆ e…n ‡ n †; s ij ˆ ss 0 ij ; …25†<br />

s ˆ e…m ‡ ? n‡ ‡ m ? n †;<br />

s 0 ij ˆ d ij ‡ s a<br />

s ?<br />

n i n j ;<br />

…26†<br />

Here m + k, m + k are the ionic mobilities perpendicular <strong>and</strong> parallel to the director, respectively.<br />

For simplicity the anisotropies were assumed to be the same for both types of<br />

ions so that s a /s k = m + k /m + k.Sonows is an additional variable. From the balance equations<br />

for n + <strong>and</strong> n – one easily recovers Eqs. 23 which now read<br />

268


13.3 Theoretical analysis<br />

ds<br />

dt ‡r <br />

m‡ ? ‡ m <br />

? s ‡ m<br />

‡<br />

? m ? r el<br />

" #<br />

ˆ seq s ‡ m<br />

1<br />

? r el s m? r el<br />

1 <br />

m ‡<br />

s s eq ‡<br />

? m ?<br />

2t rec t rec 2<br />

s 2 eq<br />

r el<br />

<br />

; …27†<br />

where s eq = e (m + k + m – k)n eq contains the equilibrium ion density n eq . The last expression is<br />

obtained by linearization in the quantities n + – n eq <strong>and</strong> n – – n eq .<br />

Thus, in this model ion accumulation effects are included whereas ionic diffusion is<br />

neglected as in the SM. The charge accumulation counteracts the st<strong>and</strong>ard (Helfrich) mechanism<br />

of generation of space charges. If t rec is sufficiently slow one can find an oscillatory<br />

behaviour of the system at threshold, i. e. a Hopf bifurcation (Section 13.5).<br />

13.2.4 Rayleigh-Bénard convection<br />

In RBC the bulk force in the Navier-Stokes equation is f = rg. In the spirit of the Boussinesq<br />

approximation one has for the mass density r = r m [1 – a (T– T 0 )], where g is the gravitational<br />

acceleration <strong>and</strong> a the thermal expansion coefficient. One needs in addition the heat<br />

conduction equation<br />

dT<br />

dt ‡rj T ˆ 0; j T ˆ krT ; …28†<br />

with k ij = k k d ij + k a n i n j , k a = k || – k k (k *10 –7 m 2 s –1 ). As first pointed out by Dubois-<br />

Violette [52, 76] the conventional instability mechanism operative in isotropic fluids is here<br />

enhanced considerably by preferential conduction of heat parallel to the director (k a >0)<br />

leading to a focusing effect in the presence of out-of-plane fluctuations of the director. For<br />

anisotropies of order one the reduction of the threshold is of order F = t d /t therm &10 3 ,<br />

where t therm = d 2 /(p 2 k k ) is the vertical thermal diffusion time. Other important dimensionless<br />

quantities besides F are the Pr<strong>and</strong>tl number Pr = t therm /t visc &10 3 <strong>and</strong> the Rayleigh<br />

number R = ag t visc /t therm (T/d), which is the traditional control parameter.<br />

13.3 Theoretical analysis<br />

We here present the general methods used to extract the relevant information from the basic<br />

equations. It is convenient to introduce the notation V=(n,v, …) for the collection of all<br />

field variables involved in the specific problem. We choose the fields in such a way that<br />

269


13 Convection Instabilities in Nematic Liquid Crystals<br />

V = 0 corresponds to the non-convecting basic <strong>and</strong> primary state. Then the set of macroscopic<br />

equations, as presented in the previous Section, can be written in the following symbolic<br />

form:<br />

LV ‡ N 2 …VjV†‡N 3 …VjVjV†‡ ˆ…B 0 ‡B 1 …V†‡B 2 …VjV†† uV<br />

ut :<br />

…29†<br />

The vector operators N 2 , N 3 … denote quadratic, cubic … operators in V <strong>and</strong> its spatial<br />

derivatives, whereas the operators L <strong>and</strong> B i represent matrix differential operators of the<br />

indicated order on V. Computer algebra can be used to perform the expansion.<br />

A direct simulation of the coupled system of the partial differential equations, Eq. 29,<br />

with appropriate boundary conditions (v = 0, n prescribed at the confining plates, etc.) is at<br />

the limits of the supercomputers of today. It will turn out that in the liquid crystal systems a<br />

rich scenario of patterns, including spatio temporal chaos, develops already near threshold<br />

so that perturbational calculations are useful.<br />

The onset of the instability is obtained from a st<strong>and</strong>ard linear stability analysis of the<br />

basic (primary) state. The problem can be diagonalized with respect to horizontal coordinates<br />

x=(x, y) associated with the directions of idealized infinite extent by a Fourier transform,<br />

V…x;z;t† ˆ R d 2 qe iq<br />

x U q …†e z<br />

l… q†t<br />

;<br />

with q=(q x , q y ). From Eq. 29 with boundary conditions at z=± d/2 one arrives at an eigenvalue<br />

problem,<br />

lB 0 …iq; u z ; R†U q …z† ˆL…iq; u z ; R†U q …z† ;<br />

…30†<br />

where R=R, S, … are the control parameters of the system. 5<br />

In the typical scenario the real part s of one of the eigenvalues l (q, R) =<br />

s (q, R) ± io (q, R) crosses zero upon increase of the main control parameter R at fixed q<br />

beyond a value R 0 (q) while the real parts of all other eigenvalues remain negative. Thus the<br />

neutral surface R=R 0 (q), which separates the unstable (R > R 0 (q)) from the stable<br />

(R < R 0 (q)) modes, is given by the condition of vanishing growth rate s(q, R) = 0. Minimizing<br />

R 0 (q) with respect to q gives the threshold R c = R 0 (q c ) with the critical wave-vector<br />

q c =(q c , p c ) <strong>and</strong> the critical frequency o c = o (q c ), which is zero for a stationary bifurcation,<br />

which is the more common case, while non-zero for a Hopf (oscillatory) bifurcation.<br />

We will encounter situations where different minima of R 0 (q) coincide. This multicritical<br />

behaviour is either accidentally or caused by some symmetry <strong>and</strong> calls for a special treatment.<br />

In isotropic systems q c is even continuously degenerate on a circle.<br />

In planarly aligned nematics, i. e. in an axially anisotropic system, one can distinguish<br />

two fundamental cases.<br />

5 We will treat R as the main control parameter whose increase carries the system across the instability,<br />

i. e. the squared voltage in the case of EHC <strong>and</strong> DT in the case of RBC (both non-dimensionalized).<br />

270


13.3 Theoretical analysis<br />

a) If q c is parallel to one of the symmetry directions, p c =0orq c = 0, one speaks of normal<br />

(Fig. 13.1 b) or parallel rolls, respectively.<br />

b) If q c is at an oblique angle, one speaks of oblique rolls (Figs. 13.2 b <strong>and</strong> 13.4a).<br />

Clearly, we get the two symmetry-degenerate directions zig <strong>and</strong> zag, which may superpose<br />

to give rectangles (Fig. 13.4 b). In the case of a Hopf bifurcation one has degeneracy<br />

between travelling waves in opposite directions, which may also superpose to give st<strong>and</strong>ing<br />

waves. Oblique rolls arising via a Hopf bifurcation give four degenerate modes.<br />

In EHC, for the usual case of driving with a pure ac field of frequency o =2/pf, the<br />

eigenvector U q of Eq. 30 inherits the additional periodic time dependence <strong>and</strong> the eigenvalue<br />

l becomes a Floquet coefficient. Then there is an additional discrete symmetry<br />

(z, t) ? (–z, t + 1/(2f )) <strong>and</strong> each component of U q has a definite parity. Generally the conductive<br />

mode (for even parity the out-of-plane component of director <strong>and</strong> v z ; for odd parity<br />

the in-plane components of velocity) destabilizes first at low frequencies f. For materials<br />

with negative dielectric anisotropy, e a < 0, there exists a cut-off frequency f c so that the dielectric<br />

mode with the other parity destabilizes first for f > f c , where f c ?? for e a ?0. The<br />

existence of these two regimes was first pointed out by Orsay’s group [9, 10]. For further details<br />

see Refs. [14, 61].<br />

The linear problem (Eq. 30) with realistic boundary conditions has to be solved numerically,<br />

often by treating the z-dependence by a Galerkin expansion with a suitable cutoff.<br />

In particular for stationary bifurcations a cut-off at lowest non-trivial order, i. e. one trial<br />

function for each component of U q gives approximate analytic expressions for the neutral<br />

surface <strong>and</strong> the growth rate, from which the interplay of the many material parameters becomes<br />

transparent [12–15, 48, 54, 77].<br />

The basic idea of the weakly non-linear analysis [1, 78–80] in its rather general form<br />

[54, 81, 82] is to reduce the phase-space dimension of the system by choosing an appropriate<br />

basis of states, characterized as the dynamically active ones [80]. We demonstrate the<br />

method only for stationary bifurcations, where the relevant time scale becomes slow near<br />

threshold but a generalization to Hopf bifurcations is straightforward. One exp<strong>and</strong>s V in<br />

Eq. 29 at lowest order as a wave packet of the eigenmodes U q of Eq. 30 at the neutral surface<br />

R 0 (q) <strong>and</strong> with the growth rate s =0,<br />

V V 1 ˆ R<br />

dq A…q†U q …z†e iqx ‡ c:c: ;<br />

…31†<br />

D ‡<br />

where A (q) denotes the amplitude order parameter, which vanishes at threshold <strong>and</strong> c.c. is<br />

the complex conjugate of the preceeding expression. The integration domain consists of<br />

small areas D ± centred at ±q c which need not be specified precisely at this point. We will arrive<br />

at an order parameter equation for A (q) by exp<strong>and</strong>ing Eq. 29 up to order A 3 . In an intermediate<br />

step, one needs the contributions V 2 * A 2 , generated by quadratic interaction from<br />

V 1 . They are determined from the relation LV 2+ N 2 (V 1 |V 1 ) = 0 derived from Eq. 29. Here<br />

terms from the right-h<strong>and</strong> side can be neglected, since they contain in addition a (slow) time<br />

derivative.<br />

The solution V 2 contains separate contributions with wave-vectors near ±2q c <strong>and</strong> 0. In<br />

the sector near q=0the so called mean flow or mean drift can be isolated systematically<br />

[24, 83]. The mean flow arises from long wavelength variations – on a scale much larger<br />

271


13 Convection Instabilities in Nematic Liquid Crystals<br />

than the spacing of the rolls – along the roll axis (undulations) leading to a lateral pressure<br />

gradient, whose spatial average across the cell is non-zero [1, 2]. A second amplitude B is<br />

introduced to describe the resulting Hagen-Poisseulle-like shear flow, also characterized by<br />

the non-zero vertical vorticity (curl v) z . The equations can be closed at order A 3 by inserting<br />

V 2 in Eq. 29 <strong>and</strong> projecting onto the subspace spanned by the linear modes V 1 . One arrives<br />

at two coupled integral equations for the amplitudes A (q) <strong>and</strong> B (s):<br />

a 1<br />

dA…q;t†<br />

dt<br />

Z Z<br />

ˆ a 2 A…q†‡ dq 0 dq 00 a 3 A…q 0 †A…q 00 †A…q q 0 q 00 †<br />

‡ R dsb 1 B…s†A…q s† ; …32†<br />

<br />

c 1 s 2 x ‡ c 2 s 2 y B…s† ˆR dqb 2 A…q†A…s q† : …33†<br />

The q-integrations are confined to the regions D = D + & D – <strong>and</strong> s is near zero. The<br />

coefficient functions a i (q), b i (q, s) <strong>and</strong> the constants c i depend on the material parameters.<br />

They involve z-integrations <strong>and</strong> have to be calculated numerically.<br />

The above procedure guaranties that all coefficients are smooth functions of the wavevectors.<br />

Since the field B satisfies an anisotropic Poisson equation, transformed to Fourier<br />

space, its long-range character is evident. In nematics its effect turns out to enhance transverse<br />

modulations in distinct contrast to isotropic fluids in most cases. As will become clear<br />

below, the mean flow contributions can be neglected in the immediate vicinity of threshold.<br />

Stationary periodic roll solutions with wave-vector q 0 can be calculated by the ansatz<br />

A r (q) = c r d (q – q 0 )+c r * d (q + q 0 ). Then the double integral on the right h<strong>and</strong> side of<br />

(Eq. 32) becomes trivial <strong>and</strong> one finds that B : 0. One also easily sees that |c r | 2 is proportional<br />

to (R – R 0 (q 0 ))/R 0 (q 0 ), the reduced distance from the neutral surface. A subcritical bifurcation<br />

is signalized by a negative proportionality factor. Then higher powers in A would<br />

have to be included.<br />

The stability analysis of the periodic roll solution [54] is performed by introducing a<br />

small perturbation dA (q, t) of the amplitude A r with a modulation wave-vector s in Eqs. 32<br />

<strong>and</strong> 33,<br />

dA…q;t†ˆ…c 1 …q; s†d…q q 0 s†‡c 2 …q; s†d…q ‡ q 0 s††e Lt ; …34†<br />

corresponding to a perturbation dV 1 in Eq. 31.<br />

Roll solutions with wave-vector q 0 are unstable if the maximum of Re (L (q 0 , s)) with<br />

respect to s is positive. Well-known long wavelength destabilization mechanisms involve:<br />

a) local dilation <strong>and</strong> compression of the roll pattern, the Eckhaus process (E) [84], i. e.<br />

s || q 0 ,<br />

b) undulations along the roll axis, the zigzag process (ZZ) [86], i. e. skq 0 ,<br />

c) combinations of both processes, the skewed varicose process (SV) [86].<br />

In the last case mean flow is decisive. In addition one has short wavelength instabilities<br />

where |s| is of the order of q c . They are well-known in systems that are isotropic in the<br />

plane, where they can lead from rolls to squares, hexagons or bimodal structures. A different<br />

272


13.3 Theoretical analysis<br />

version is found in our anisotropic systems (planar director alignment) when oblique rolls<br />

become unstable with respect to a superposition of zig <strong>and</strong> zag (Fig. 13.4 b).<br />

Many features become more transparent when formulated in real (position) space in<br />

terms of amplitude (envelope) or Ginzburg-L<strong>and</strong>au equations (GLE). Then one sees that the<br />

important information is really condensed in a few parameters <strong>and</strong> the universal aspects of<br />

the systems become apparent. By model calculations, which can often be performed analytically,<br />

stability boundaries <strong>and</strong> secondary bifurcation scenarios are traced out. The real space<br />

formulation is essential when it comes to the description of more complex spatio-temporal<br />

patterns with disorder <strong>and</strong> defects, which have been studied extensively in EHC slightly<br />

above threshold (Figs. 13.1b, 13.3b). One introduces a modulation amplitude A(x) defined<br />

as<br />

A…x† ˆ R<br />

dqA…q†e i…q qc†x : …35†<br />

D ‡<br />

Near threshold one expects that only a small region D + (i. e. |q-q c | P q c ) is relevant in<br />

Eq. 35 <strong>and</strong> that correspondingly the amplitude A(x) varies on a slow scale. The various coefficients<br />

of the order parameter equations (Eqs. 32 <strong>and</strong> 33) can now be exp<strong>and</strong>ed into Taylor<br />

series around q c with respect to q <strong>and</strong> around zero with respect to s. Since all non-analyticities<br />

have been absorbed in the amplitude B the expansion is smooth <strong>and</strong> powers of the components<br />

of (q-q c ) <strong>and</strong> of s can obviously be identified with spatial derivatives of A(x) <strong>and</strong><br />

B(x), which is constructed in analogy to Eq. 35.<br />

If the q-dependence is taken into account only in the linear coefficient a 2 of Eq. 32 –<br />

all other coefficients taken at q=q c <strong>and</strong> s=0, where b 1 = b 2 = 0 – one ends up with the famous,<br />

slightly generalized Ginzburg L<strong>and</strong>au equation [87]<br />

u t A ˆ l…q c ir; e†A gjAj 2 A; …36†<br />

where l is the linear growth rate (Eq. 30). Clearly l is zero at threshold e =(R – R c )/R c =0<br />

<strong>and</strong> r = 0 <strong>and</strong> should be exp<strong>and</strong>ed in both arguments. For our purposes at lowest order it is<br />

sufficient to keep the following terms<br />

l…q c ir; e† e ‡ x 2 1 u2 x ‡ 2ax 1x 2 u x u y ‡ x 2 2 Wu2 y iZx 1 x 2 2 u xu 2 y x 4 2 u4 y : …37†<br />

The reason for keeping higher powers in u y than in u x will become clear shortly. On<br />

this level the expansion can be cast into an overall expansion scheme in terms of e 1/2 ,or<br />

equivalently A. In the anisotropic case, where there is no continuous degeneracy of the critical<br />

mode(s), one may in general assume e *A 2 *u t *u 2 x *u 2 y, so that the higher order terms<br />

*u x u 2 y <strong>and</strong> u 4 y drop out. Then Eq. 36 becomes uniformly of order e 3/2 .<br />

The remaining mixed derivative term in Eq. 37, that vanishes anyway for normal as<br />

well as for parallel rolls, can always be transformed away by rotating the coordinate system.<br />

Moreover, by rescaling x <strong>and</strong> y the differential operator becomes proportional to the Laplacian,<br />

so that Eq. 36 finally reduces to<br />

tu t A ˆ eA ‡ x 2 DA gjAj 2 A: …38†<br />

273


13 Convection Instabilities in Nematic Liquid Crystals<br />

This constitutes the simplest GLE [88, 89]. Until now we have not specified explicitly<br />

the new spatial scaling. By appropriate scaling the time, length, <strong>and</strong> A all parameters can be<br />

scaled away. Note that the conventional way to derive the GLE starts from a multi-scale analysis<br />

in space <strong>and</strong> time, exp<strong>and</strong>ing systematically in powers of e 1/2 .<br />

Going back to Eqs. 36 <strong>and</strong> 37 with a = 0 it is easy to see that changing W from positive<br />

to negative – by changing some secondary control parameter like the frequency in EHC<br />

– describes a normal pitch fork bifurcation from normal to oblique rolls. 6 For W < 0 the<br />

maximum growth rate of plane wave solutions of Eq. 36 occurs at wave-vectors with nonzero<br />

y component. Details of this transition, which is the analogue of an LP in the theory of<br />

equilibrium phase transitions, have been discussed elsewhere [12, 21, 89, 90]. The corresponding<br />

uniform scaling *e 3/2 as in Eq. 38 is recovered with W*e 1/2 *u x <strong>and</strong> now<br />

u y *e 1/4 . This corresponds to the scaling adopted in isotropic media <strong>and</strong> in fact the wellknown<br />

Newell-Whitehead-Segel amplitude equation for isotropic systems [91, 92] can now<br />

be obtained as the special case W=0 <strong>and</strong> Z=±2 in Eq. 37.<br />

The general form of the amplitude equations is known a priori from symmetry, translation<br />

in space <strong>and</strong> time as well as rotation <strong>and</strong> reflexion symmetry within the plane of the<br />

layer, which manifests itself in the linear growth rate function. The coefficients for a specific<br />

problem have to be determined only for quantitative comparison with experiments.<br />

So far we concentrated on the simplest version of a stationary bifurcation where only<br />

a single mode becomes critical. One may have (near) degeneracy due to symmetry or accidentally<br />

by tuning of a secondary control parameter. If there is a n-fold degeneracy of marginal<br />

modes one is led to a coupled system of n amplitudes. Here degeneracy due to symmetry<br />

occurs in the oblique roll case (n =2). From the ratio of the two non-linear coefficients<br />

for self <strong>and</strong> cross coupling one then deduces if rolls zig or zag with the possibility of domain<br />

boundaries (Fig. 13.2b) or if their superposition (Fig. 13.4b) is stable.<br />

In the case of a Hopf bifurcation, as observed in EHC, one has degeneracy due to reflection<br />

symmetry with two modes, corresponding to left <strong>and</strong> right travelling waves (rolls).<br />

Also, the coefficients of Eq. 38 become complex <strong>and</strong> from the linear dispersion relation one<br />

gains a term ± itv g 7rA (the signs ± pertain to left <strong>and</strong> right travelling waves), which describes<br />

a group velocity v g . From a phenomenological point of view changes of complex<br />

coefficients <strong>and</strong> group velocity stem from the absence of reflection symmetry in the solutions.<br />

Destroying reflection symmetry by an external perturbation has a similar effect. Depending<br />

on the ratio of the real parts of the non-linear coefficients there are again two possibilities:<br />

either travelling waves with the possibility of domain boundaries (sources <strong>and</strong> sinks)<br />

or a superposition of the two wave systems leading to st<strong>and</strong>ing (oscillating) rolls. In the first<br />

case one is essentially left with the celebrated complex Ginzburg-L<strong>and</strong>au equation (CGLE),<br />

which exhibits transitions, e. g. at the Benjamin-Feir instability, to various forms of spatiotemporal<br />

chaos <strong>and</strong> is presently studied intensely. For general reviews see Refs. [2, 80]. The<br />

CGLE is applicable also to systems that show a Hopf bifurcation leading to a spatially homogeneous<br />

state (q c = 0), which is mainly found in oscillatory chemical reactions. The point defects<br />

of CGLE correspond to the famous spirals. In our system there is also the possibility of<br />

oblique travelling waves, which have in fact been observed in Merck Phase V [47] <strong>and</strong> in<br />

I52 [49, 51] <strong>and</strong> are known to lead to an interesting type of four-wave interaction [93].<br />

6 Now the higher powers in u y become essential; if we wanted to describe the transition between parallel<br />

<strong>and</strong> oblique rolls, higher powers in u x would have to be retained.<br />

274


13.4 Rayleigh-Bénard convection<br />

For the description of modulated roll patterns away from threshold (possibly only<br />

slightly) the amplitude B must be included, now also transformed to real space. A uniform<br />

scaling in e is then no more possible <strong>and</strong> coupling terms like Au y B appear in Eq. 36 as well<br />

as derivative terms in the cubic non-linearities. The gradient expansion starting from Eqs. 32<br />

<strong>and</strong> 33 is systematically truncated in such a way that all O(s 2 ) contributions to the growth<br />

rate (Eq. 34) are included [24, 83]. The additional equation for B is of the form<br />

…c 1 u 2 x ‡ c 2u 2 y †B ˆ q 1u x u y jAj 2 ‡ q 4 u y …iA u 2 y A ‡ c:c:†‡...<br />

…39†<br />

The terms occurring are those allowed by symmetry. Due to the anisotropy more terms<br />

appear than in isotropic systems [94]. The clue for the characteristic appearance of the zigzag<br />

instability as a secondary bifurcation is that q 4 is typically negative in nematics [23, 24]<br />

leading, in contrast to isotropic fluids, to amplification of transverse modulations of roll patterns.<br />

Model calculations that include this feature [25, 26] were quite successful in describing<br />

qualitatively the secondary instability <strong>and</strong> the behaviour beyond.<br />

13.4 Rayleigh-Bénard convection<br />

In the following two Sections we will discuss <strong>and</strong> compare theoretical <strong>and</strong> experimental studies<br />

on RBC <strong>and</strong> EHC. We often use non-dimensionalized units. Thus we write wave numbers<br />

as q i = q i 'p/d, where the prime is sometimes omitted, <strong>and</strong> magnetic fields as H i = h x H f<br />

with the splay Freédericksz transition field H f =(p/d)[k 11 /(m 0 w a )] 1/2 . Other quantities have<br />

been introduced before. Note that for the Cartesian components of the wave-vector we use<br />

two symbols, q=(q x , q y )=(q, p).<br />

In the case of RBC we will mainly present an analysis of recent experimental investigations<br />

[55] in comparison with theoretical results [54]. In Fig. 13.5a the critical Rayleigh<br />

number R c (continuous curve) normalized to the value R c0 = 1707.37 for isotropic fluids is<br />

shown as a function of a stabilizing magnetic field h x = H x /H f (R c /R c0 ? 1 for h x 2 /F p1).<br />

The symbols are experimental results for the nematic 5CB. The linear theory predicts for increasing<br />

field a lower Lifshitz point (LP), where the roll orientation changes from normal<br />

( p c = 0) to oblique <strong>and</strong> at higher field an upper LP with a transition from oblique to parallel<br />

rolls (q c = 0). In Fig. 13.5b the experimental results for the squared wave numbers (q c , p c )<br />

together with |q c | 2 are plotted <strong>and</strong> compared with the prediction of the theory (solid curves).<br />

The agreement is quite remarkable considering that there is no adjustable parameter.<br />

Weakly non-linear theory predicts two tricritical points (TP) with a subcritical bifurcation<br />

in the field range 5 < h x < 26. Note that the upper TP at h x = 26 is slightly above the<br />

lower LP. One may expect rather complex non-linear behaviour in that range, which has not<br />

been worked out in detail for 5CB so far. The situation should be simpler for MBBA where<br />

the TPs (h x = 4.15 <strong>and</strong> 31) are below <strong>and</strong> well-separated from the Lifshitz points (h x =36<br />

<strong>and</strong> 62).<br />

275


13 Convection Instabilities in Nematic Liquid Crystals<br />

Figure 13.5: a) Rayleigh number as a function of the stabilizing field h x in the planar configuration.<br />

b) The squared components of q c <strong>and</strong> |q c | 2 as a function of h x .<br />

In Figs. 13.6a, b stability diagrams of MBBA in the e-q plane (normal rolls, p =0)<br />

<strong>and</strong> in the e-p plane (oblique rolls, q=q c ) are shown for a magnetic field h x = 34 between<br />

the upper TP <strong>and</strong> the lower LP (e =(R – R c )/R c ). Rolls exist inside the neutral curves (NC)<br />

<strong>and</strong> are stable in a region bounded by the E, SV, <strong>and</strong> ZZ lines. The lowest-order theory<br />

(GLE) would only give the Eckhaus instability. Including the higher order terms with mean<br />

flow produces the skewed varicose instability SV, which however is hard to distinguish from<br />

the Eckhaus instability E. More important is that the stability regime for normal rolls is now<br />

limited from above by a ZZ line. When the ZZ line for normal rolls is extended on the left<br />

it joins the neutral curve at a point L which one may call a LP on the neutral curve. To the<br />

left of that point the growth rate for oblique rolls becomes larger than for normal rolls, or<br />

equivalently, the curvature of the neutral surface in the p-direction becomes negative. With<br />

0.02<br />

0.02<br />

ε=(R-R c<br />

)/R c<br />

0.01<br />

E<br />

L<br />

ZZ<br />

SV<br />

NC<br />

0.00<br />

–0.2 –0.1 0.0 0.1 0.2 0.3<br />

q-q c<br />

P<br />

ε=(R-R c<br />

)/R c<br />

0.01<br />

ZZ<br />

SV<br />

NC<br />

E<br />

0.00<br />

0.0 0.5 1.0 1.5<br />

p<br />

Figure 13.6: a) Stability diagram for normal rolls slightly below the lower LP; b) Stability diagram in<br />

the transverse (oblique) direction for q=q c .<br />

276


13.4 Rayleigh-Bénard convection<br />

increasing h x the point L moves down along the neutral curve until it reaches q c , e =0at<br />

h x = 36. This is the lower LP beyond which oblique rolls appear at threshold. In the vicinity<br />

of the LP the scenario can be described by the extended GLE as discussed in the last Section.<br />

The parameter W in Eq. 37 changes sign when the LP is crossed. From this equation<br />

without higher order terms one obtains a ZZ line that goes vertically up [89]. Coupling to<br />

the mean flow enhances the ZZ instability (Section 13.6) <strong>and</strong> consequently tilts the curve to<br />

the right. By contrast, in simple fluids, where the ZZ line always emanates from q c , mean<br />

flow effects, which are important for small to medium Pr<strong>and</strong>tl numbers, turn the line to the<br />

left. Thus anisotropy is responsible for moving the LP away from q c along the neutral curve<br />

<strong>and</strong> mean flow tilts the line to the right. This seems to be a general scenario which also applies<br />

to EHC.<br />

Let us turn attention to the point P on the neutral curve NC in Fig. 13.6a, where several<br />

stability limits merge. P corresponds to a tricritical point on the neutral curve at q tri .<br />

For q6q tri the solution bifurcates subcritically from the neutral curve, i. e. a small amplitude<br />

solution exists only outside the neutral curve. With decreasing magnetic field P moves<br />

down along the neutral curve <strong>and</strong> at h x = 31 it reaches e = 0 <strong>and</strong> q = q c , signalizing the<br />

change to a subcritical bifurcation with further decrease of h x . Interestingly, at small magnetic<br />

fields, where the bifurcation becomes supercritical again, the tricritical point moves upwards<br />

on the left branch of the neutral curve.<br />

Figure 13.6 b shows that normal rolls can escape the ZZ instability by undergoing a<br />

secondary transition to oblique rolls. Their stability range is bounded by two roughly parabolic<br />

curves. This transition has indeed been observed [55] <strong>and</strong> is quite well-known in EHC.<br />

There exists in addition a tertiary bifurcation (dotted lines), where the oblique rolls become<br />

unstable against a short wavelength mode which appears to lead to oblique rolls with p<br />

roughly reversed. At present it cannot be predicted into which state the system evolves. It<br />

could be a superposition of zig <strong>and</strong> zag rolls, as observed in the experiments, or a complex<br />

dynamic state where the system oscillates between the two states separated by grain boundaries<br />

(<strong>and</strong> maybe other defects) generated persistently. Also, in the light of the results for<br />

EHC (see below), one may expect that under some conditions in large aspect-ratio systems<br />

spatio-temporal chaos develops when the stability limit of normal rolls is exceeded.<br />

The non-linear scenarios have not been systematically studied in experiments so far.<br />

Actually investigations at small fields (h x ^ lower TP) would be very interesting. According<br />

to the theory the upper limitation of the stability regime is now of the SV-type, beyond<br />

which the system cannot evade into stable oblique rolls <strong>and</strong> complex behaviour seems unavoidable.<br />

Very recently a generalized weakly non-linear analysis including homogeneous twist<br />

of the director has been worked out [171]. Then the above mentioned SV instability changes<br />

into a ZZ instability in agreement with experiments [167]. In the non-linear regime one has<br />

a transition to abnormal rolls <strong>and</strong> bimodal convection as is also found in EHC, see note<br />

added at the end.<br />

277


13 Convection Instabilities in Nematic Liquid Crystals<br />

13.5 Electrohydrodynamic convection<br />

The theoretical results are obtained with the SM (Section 13.2) unless otherwise stated.<br />

13.5.1 Linear theory <strong>and</strong> type of bifurcation<br />

As pointed out before, most experiments were done with the material MBBA, sometimes<br />

doped with an ionizing substance. Typical material parameters can be found in the literature<br />

[12] which look fairly reliable except the flexoelectric coefficients. Unfortunately in most experiments<br />

the conductivity, which may vary strongly, has not been measured. One therefore<br />

has to determine the charge relaxation time t q , e. g. by fitting the cut-off frequency f c where<br />

the crossover between the low-frequency p conductive <strong>and</strong> the higher frequency dielectric regime<br />

occurs. The formula 2 pf c t q = C applies, where C is a function of s|| /s k , s || /s k , <strong>and</strong><br />

the viscosity ratio [12]. For MBBA C is about 6.3. In Fig. 13.2 a we show the threshold curve<br />

for a rather thick (d = 100 µm) <strong>and</strong> clean cell showing quite good agreement between theory<br />

<strong>and</strong> experiment. In the calculations the flexoelectric coefficients were taken considerably<br />

smaller than those estimated from some measurements. Otherwise one would find in contrast<br />

to the observations oblique rolls at a threshold in the dielectric regime. The conductivity was<br />

chosen as s k = 0.28610 –8 O –1 m –1 <strong>and</strong> s || /s k = 1.65. For more details see Ref. [14]. The theory<br />

correctly describes many qualitative features like the occurrence of oblique rolls at threshold<br />

at sufficiently low frequencies for materials with not too negative e a . MBBA (e a &–0.5)<br />

is a border-line case [14], Merck Phase V (e a & –0.3) has well developed oblique rolls <strong>and</strong><br />

the newly introduced material I52 (e a & 0) exhibits strongly oblique rolls [50, 51].<br />

For materials with e a < 0 the above behaviour with the conductive threshold apparently<br />

diverging at f c is a reasonable approximation if the charge relaxation time t q is much smaller<br />

than the director relaxation time t d . Actually the conductive threshold curve always becomes<br />

vertical at a frequency f m ^ f c <strong>and</strong> bends back at higher voltages giving rise to restabilization<br />

of the conductive mode at high voltages. The effect becomes apparent in thinner <strong>and</strong> cleaner<br />

specimens. The fact that the conductive mode becomes ineffective for high voltages as well as<br />

for large ac frequencies is a consequence of the competition of the (destabilizing) electrohydrodynamic<br />

torques <strong>and</strong> the dielectric torques, which are stabilizing when e a 0<br />

the homogeneous (q c = 0) Freédericksz transition competes with EHC [12] <strong>and</strong> in fact always<br />

preempts the latter at sufficiently high frequencies. For a recent investigation see [95].<br />

Actually the quantitative agreement between theory <strong>and</strong> experiment over a large frequency<br />

range including the dielectric regime, suggested from Fig. 13.2a, is somewhat deceptive.<br />

For the more common thinner specimens we are not always able to fit experimental<br />

curves satisfactorily over the whole frequency range by adjusting the conductivity, the flexocoefficients,<br />

<strong>and</strong> (to a lesser extent) the ratio s || /s k . When adjusting the conductivity so that<br />

f c (<strong>and</strong> thereby essentially the whole conductive range) is described correctly, the dielectric<br />

threshold tends to come out too low. One could increase the conductivity to fit the dielectric<br />

threshold, which is roughly proportional to s k d 2 , but then f c becomes too large.<br />

278


13.5 Electrohydrodynamic convection<br />

This discrepancy is in line with other observed quantitative discrepancies such as in<br />

the threshold behaviour when a stochastic component is present in the applied voltage [96].<br />

As already mentioned, the most drastic non-st<strong>and</strong>ard behaviour, which is not understood<br />

from the SM, is the observed Hopf bifurcation in sufficiently thin <strong>and</strong> clean specimens [18,<br />

41–43, 49–51] <strong>and</strong> the very small hysteresis sometimes observed at threshold [38–40, 47].<br />

In fact a little further above threshold (but still near it) the amplitude of the pattern does appear<br />

to coincide reasonably with the results of the weakly non-linear theory [59].<br />

Since the measurements on RBC exhibit such beautiful quantitative agreement with<br />

theory one has concluded that the electrodynamic part of the basic equations needs improvement.<br />

This was the motivation for introducing the WEM (Section 13.2). The linear stability<br />

calculations for the conductive mode have been carried out within this model [48, 49] using<br />

the same approximations which led to the analytic threshold formulas within the SM [12–<br />

15]. It is found that there is an upward shift in the threshold, which may be quite small, <strong>and</strong>,<br />

more importantly, a Hopf bifurcation with critical frequency<br />

s<br />

o H t d ˆ 2pf H t d ˆ Rc ~aC<br />

1<br />

1 ‡ o 02<br />

<br />

…1 ‡ o 02 †l s t 2<br />

d<br />

R c ~aC<br />

; …40†<br />

if the expression under the square root is positive. Here R = V 2 s a t q<br />

p 2 k 11<br />

is the reduced control parameter,<br />

~a 2 = m km + kg – 1 p 2 /(s a d 2 ) is proportional to the geometric means of the mobilities, <strong>and</strong><br />

o' = ot q b, with b = …s jj=s k † q 02 ‡p 02 ‡1<br />

…e jj =e k<br />

, is a reduced frequency.<br />

† q 02 ‡p 02 ‡1<br />

Moreover, l s =–[t –1 rec + t –1 d R~a 2 b/(1 + o' 2 )] < 0 is the damping rate of the (new) WEM<br />

mode. Its dominant contribution is usually just determined by the ion recombination rate<br />

1/t rec . Thus for the Hopf bifurcation to occur the quantity t d /(~at rec ) must not be too large.<br />

This requires that the recombination of ions is sufficiently slow <strong>and</strong> that the layer is sufficiently<br />

thin <strong>and</strong> clean. Note, the Hopf condition is always satisfied near the cut-off for materials<br />

with negative dielectric anisotropy, where R c diverges at the cut-off frequency (in the<br />

approximation used). But the Hopf frequency, which is then just given by the prefactor of<br />

the square root from Eq. 40, becomes large there. This appears to be consistent with the experiments<br />

[18, 41, 42]. Moreover, the prediction of the theory that for materials with vanishing<br />

dielectric anisotropy the Hopf condition <strong>and</strong> Hopf frequency become essentially independent<br />

of the external frequency has been verified experimentally using the material I52<br />

[49]. I52 has the property that e a changes from negative to positive when the temperature<br />

passes through T&60 8C.<br />

13.5.2 Results of Ginzburg-L<strong>and</strong>au equation<br />

The large aspect-ratios, which can be achieved comparatively easily in EHC, make this system<br />

particularly well suited to test predictions of GLEs. Experimental results for the existence<br />

<strong>and</strong> stability of normal roll patterns in the q-e plane (Busse balloon) are shown in<br />

Fig. 13.7 [28, 36]. Similar experiments were reported by Ref. [29]. The symbols give the experimental<br />

points <strong>and</strong> the curves represent parabolic fits for the neutral curve (N) <strong>and</strong> the<br />

279


13 Convection Instabilities in Nematic Liquid Crystals<br />

Defect Turbulence<br />

ε<br />

0.2<br />

E<br />

II<br />

ZZ Pattern<br />

0.1<br />

0.0<br />

–0.2 –0.1 0.0 0.1 0.2<br />

q-q c<br />

Figure 13.7: Experimental stability diagram (see text).<br />

N<br />

Normal Rolls<br />

I<br />

Eckhaus stability limit (E). The secondary <strong>and</strong> tertiary instabilities that limit the regions<br />

from above will be discussed further below.<br />

The wave number q can be controlled within certain limits on the small wave number<br />

side by the so-called frequency-jump technique [32]. Since q c increases with external frequency<br />

one can prepare first a state for the desired wave number by choosing the appropriate<br />

frequency <strong>and</strong> then by jumping to the frequency <strong>and</strong> e values. In the course of the experiment<br />

e can still be varied. Outside of the neutral curve the rolls decay whereas inside they<br />

grow up to their non-linear saturation. The neutral point is then determined by extrapolation.<br />

In this way also the GL relaxation time t was determined.<br />

The coherence length x 1 in Eq. 37 can be obtained from the curvatures of the neutral<br />

surface at the minimum. x 1 pertains to the direction parallel to the wave-vector of the pattern<br />

<strong>and</strong> is therefore often denoted as x || . Both, x 1 <strong>and</strong> x 2 (x 2 = x k for normal rolls) have been measured<br />

by imaging the core of defects (see below) [32]. In the same paper comprehensive measurements<br />

on MBBA are presented for all parameters of the GLE as a function of external frequency<br />

in the conductive range. The comparison with theory shows quite good agreement.<br />

Another elegant method for measuring the linear parameters of the GLE makes use of thermal<br />

fluctuations very slightly below threshold [38, 39, 97, 98]. Then the dynamic structure function<br />

is exploited. The agreement with theoretical results [12] is generally, maybe except some<br />

cases of thin layers, satisfactory including the predicted variations with ac frequency.<br />

Slightly above onset the pattern should be dominated by the critical mode at q = q c<br />

<strong>and</strong> according to the<br />

p<br />

weakly non-linear analysis (Section 13.3) the pattern amplitude should<br />

grow proportional to e . This behaviour has been confirmed in experiments, where the optical<br />

contrast of the roll pattern was monitored as function of e [59] using the shadowgraph<br />

method [34, 38, 99]. The proportionality factor, which is determined by the non-linear (cubic)<br />

coefficient g in the GLE, agreed satisfactorily with theoretical results [12].<br />

In a next step the Eckhaus instability boundaries (curves E in Fig. 13.7), which probe<br />

non-linear aspects of the system, could be identified by observing the destabilization of the<br />

pattern via longitudinal modulations after a frequency jump [29, 36, 37] 7 . Subsequently the<br />

7 The forcing of a pattern into a wave number state with q(q c can also be done by the use of digitized electrodes<br />

<strong>and</strong> led to the first detailed investigation of the Eckhaus process in a remarkable experiment [27].<br />

280


13.5 Electrohydrodynamic convection<br />

pattern evolves into a state with wave number in the stable range by spontaneous creation of<br />

defects either at the surface or in pairs. p The prediction of the GLE that the Eckhaus curve is<br />

narrower by the universal factor 1/ 3 than the neutral curve, both in parabolic approximation,<br />

has been confirmed. Actually also the full evolution process from an unstable to a<br />

stable wave number can be analyzed with the GLE [100].<br />

A particularly beautiful example for the usefulness of the GLE is the description of<br />

the structure <strong>and</strong> dynamics of point defects (dislocations), see Fig. 13.1 b. A defect is characterized<br />

by a zero of the amplitude A (x) where it behaves as |x| exp (±if), where f is the polar<br />

angle <strong>and</strong> the topological charge is ±1 in the isotropic scaling. A stationary, isolated defect<br />

solution exists only for a pattern with the background wave-vector q = q c , i. e. with<br />

modulation wave-vector Q = q – q c = 0 in the GL description. Otherwise the defect moves<br />

with constant velocity V perpendicularly to Q. Each defect crossing the system carries away<br />

one periodic unit <strong>and</strong> brings the system nearer to its globally stable state Q=0(or q = q c )<br />

by an amount 2 p/L, where L is the system length. Since the GLE has a minimizing potential<br />

one can speak in such terms <strong>and</strong> the force on the defect is in fact the analogue of the Peach-<br />

Kohler force in solids [101, 102]. Due to the two-dimensional nature of the problem V (Q) is<br />

non-analytic at V=Q= 0. One has V log (V t/(3.29 x)) = 2 (x 2 /t)Q for small Q [88, 103].<br />

The full universal curve V (Q) can be obtained numerically [14] <strong>and</strong> is compared with detailed<br />

experimental results in EHC for motion of defects along the rolls [32] (Fig. 13.3a).<br />

Also the interaction of defects, which is repulsive for equal topological charge <strong>and</strong> attractive<br />

otherwise, was investigated <strong>and</strong> compared with experiments [33, 104].<br />

The theoretical results discussed above pertain to the normal-roll regime with wave<br />

number changes confined to the longitudinal component, i. e. the rolls remain normal. Some<br />

rather qualitative experiments with transverse wave number changes, showed that defects<br />

moved perpendicular to the rolls, as theoretically expected [30]. One also gets a neutral<br />

curve <strong>and</strong> a (generalized) Eckhaus stability limit roughly as expected [19]. In the latter experiments<br />

a magnetic field, applied in the plane of the layer at an oblique angle, was used to<br />

influence q c .<br />

13.5.3 Beyond the Ginzburg-L<strong>and</strong>au equation<br />

From the foregoing discussion we know that the weakly non-linear behaviour of EHC in the<br />

range of validity of the GLE is understood fairly well. The situation is not quite as satisfactory<br />

when it comes to secondary bifurcations, which confine the q- range <strong>and</strong> which are captured<br />

in the theory only if corrections to GLE are included (Section 13.3) or a full evaluation<br />

is done to the hydrodynamic equations.<br />

13.5.3.1 Experimental results<br />

From the most recent experiments it is quite clear that, beginning with normal rolls at<br />

threshold <strong>and</strong> increasing e, the ZZ instability, where long wavelength modulations with<br />

wave-vector s = (0, s y ) perpendicular to q =(q, 0) start growing, provides the most impor-<br />

281


13 Convection Instabilities in Nematic Liquid Crystals<br />

tant destabilization mechanism, which sets in at quite low e (^0.2). This has been observed<br />

for MBBA (Fig. 13.7) [29, 36, 37, 105], where there are usually normal rolls at threshold<br />

for all frequencies. Furthermore, in Merck Phase V, where one has an LP at moderately low<br />

frequencies [105], <strong>and</strong> in I52 too, where oblique rolls at threshold extend to higher frequencies<br />

[50, 51]. Actually in the first clear observation of oblique rolls [106] they appeared after<br />

a secondary bifurcation.<br />

It also seems fairly clear now that under sufficiently ideal conditions – very homogeneous<br />

large aspect-ratio system, slow increase of e – beyond the ZZ instability the system<br />

can settle down in an (ideally) stationary oblique roll state. Usually zig <strong>and</strong> zag domains<br />

seem to persist. For MBBA the angle of obliqueness remains small, below about 108. The<br />

necessity to change sufficiently slowly has been stressed in particular by Ribotta [167].<br />

Otherwise the system tends to remain in a dynamic defect turbulent state, which was called<br />

fluctuating Williams domains by Kai <strong>and</strong> co-workers [19]. It is characterized by continuous<br />

generation <strong>and</strong> annihilation of dislocation pairs modulating the (normal) roll pattern, see<br />

e. g. the snapshot shown in Fig. 13.3b. Thus one seems to have two coexisting attractors.<br />

The transition to normal roll defect turbulence takes place at e above e zz (diamonds in<br />

Fig.13.7) [36, 37]. At even higher values of e a transition to a rectangular or grid pattern is<br />

typically observed before the onset of the strong turbulence types with creation <strong>and</strong> annihilation<br />

of disclinations in the director field [17, 62].<br />

According to some measurements of (doped) MBBA the SV instability also plays a<br />

significant role, particularly in the higher frequency part of the conductive range, where it<br />

appeared to replace the ZZ instability as the first roll destabilization mechanism [29, 37].<br />

When increasing e slowly beyond the SV instability the system settled down at first in an interesting<br />

regular defect-lattice structure [37, 107, 108]. At higher values of e a transition to<br />

defect turbulence again sets in. For an ac driving o around the crossover frequency o cr between<br />

the two types of behaviour a (fairly) direct transition from normal rolls to defect turbulence<br />

is observed. Below o cr<br />

the SV instability manifested itself as transients in experiments<br />

where e was increased suddenly, starting out from below the ZZ line. Whereas immediately<br />

above the ZZ line the first destabilization took place via the ZZ instability as expected.<br />

It changed over to a SV destabilization at higher values of e thus defining a SV line<br />

in the q-e plane. With increasing o the SV line moved closer to the ZZ line <strong>and</strong> apparently<br />

joined it at o cr . Although defect-lattice structures of the above type have apparently been<br />

observed by other researchers [168], they did not characterize quantitatively the existence<br />

range. Possibly their cells with undoped MBBA behaved somewhat differently.<br />

We only mention in passing that various undulated structures have been reported by<br />

Ribotta <strong>and</strong> co-workers [17, 105] but could not always be reproduced by others [169]. Actually<br />

such structures can be obtained near the LP even within the GLE approximation, but<br />

they then turn out to be metastable since their (generalized) potential is higher than that of<br />

straight (usually oblique) rolls [14, 89, 90, 109, 110]. This may explain why they are difficult<br />

to observe.<br />

13.5.3.2 Theoretical results <strong>and</strong> discussion<br />

The experimental scenario can be partly understood with the SM [24, 111, 112]. One finds<br />

that the general shape of the measured ZZ instability limit in the q-e plane, shown in<br />

282


13.5 Electrohydrodynamic convection<br />

Fig. 13.7 (triangles), is reproduced by the theory in a rather robust fashion, whereas the actual<br />

position depends crucially on the material parameters <strong>and</strong> is very sensitive to approximations.<br />

Thus using st<strong>and</strong>ard MBBA parameters the ZZ instability comes out too high<br />

within the order-parameter approach, which involves an expansion up to cubic order in the<br />

pattern amplitude (Section 13.3) [24, 111]. One needs the full Galerkin computations<br />

(Section 13.3), which lead to rather good quantitative agreement (triangles <strong>and</strong> broken line<br />

in Fig. 13.8 a) [112]. We conclude that higher order terms in the amplitude become important<br />

already at quite low values of e.<br />

In any case corrections to the GLE approximation in particular the coupling of the<br />

pattern amplitude to the mean flow [24] (Eq. 39), which is made explicit in the order parameter<br />

approach, has the effect of reinforcing the ZZ instability with increasing amplitude, in<br />

distinct contrast to the effect in isotropic fluids 8 . This results in a ZZ line roughly horizontal<br />

in the q-e plane in the range where it determines the stability boundaries. A convenient characterization<br />

of the ZZ line is its position at q = q c , which is called e zz . On the low-q side one<br />

expects the ZZ line to extend to the neutral curve <strong>and</strong> join it at a point L, which is sometimes<br />

called an LP on the neutral curve, because below normal <strong>and</strong> above oblique rolls become<br />

critical on the neutral curve. This point can be calculated from the linear theory <strong>and</strong> is<br />

also shown in Fig. 13.8a. Following the ZZ line through the Eckhaus unstable range on the<br />

low q side there is a delicate numerical problem <strong>and</strong> the curve may exhibit rather unexpected<br />

variations (Fig.13.8 a).<br />

Figure 13.8: a) Stability diagram for normal rolls for a rather thick cell with MBBA at ot q = 0.2. Full<br />

Galerkin computation. For details see text; b) Stability diagram for rolls escaping in the oblique direction<br />

at fixed q=q c . The unstable bubble was tested by using Galerkin expansion with more terms (Z).<br />

For details see text.<br />

In any case, if decreasing the external frequency the LP on the neutral curve typically<br />

moves down <strong>and</strong> may eventually cross the minimum at q c leading to oblique rolls at threshold.<br />

This is the case for the materials Merck Phase V <strong>and</strong> I52. Then, simultaneously e zz<br />

moves down to zero. Of course shifts of e zz <strong>and</strong> the LP on the neutral curve can also be effected<br />

by changing the material parameters. Figures 13.9a,b give an impression of the de-<br />

8 In technical terms it is the opposite sign of the coefficient q 4 of Eq. 39 which is responsible for the<br />

different behaviour<br />

283


13 Convection Instabilities in Nematic Liquid Crystals<br />

0.08<br />

0.06<br />

ωτ 0<br />

= 0.5<br />

ε a<br />

= –0.53<br />

d = 55µm<br />

0.070<br />

ε ZZ<br />

0.065<br />

0.060<br />

ε ZZ<br />

–1.0 –0.8 –0.6 –0.4 –0.2 0.0<br />

σ a<br />

= 0.5<br />

d = 55µm<br />

0.04<br />

0.02<br />

0.055<br />

0.050<br />

0.045<br />

0.00<br />

0.1 0.3 0.5<br />

σ a<br />

0.7 0.9<br />

Figure 13.9: Onset of the ZZ instability at b<strong>and</strong> centre as a function of (a) the anisotropy of the conductivity<br />

s a <strong>and</strong> (b) the anisotropy of the dielectric constant e a .<br />

0.040<br />

ε a<br />

pendence of e zz on the anisotropies of the conductivity <strong>and</strong> the dielectric tensor. The trend<br />

shown in Fig. 13.9b is in qualitative agreement with experiments when comparing MBBA<br />

to Merck Phase V <strong>and</strong> I52, which have e a & –0.5, –0.3, 0, respectively. 9<br />

The mean flow also has the effect of transforming the Eckhaus instability on the large<br />

q side into a SV instability. This effect becomes noticeable only in the upper part as the ZZ<br />

line is approached because at smaller e the ratio s y /s x is very small, see the broken line in<br />

Fig. 13.8a. In fact the SV instability then turns around <strong>and</strong> joins the ZZ line smoothly. This<br />

effect indicates that the SV instability may become relevant <strong>and</strong> could be a clue to the observations<br />

quoted above [29, 37, 108].<br />

In order to show that the system can escape at e zz into an oblique roll state, as is actually<br />

observed, the stability of rolls with q, p&0 was investigated. An example of the stability<br />

diagram for MBBA is shown in Fig. 13.8b, where q is fixed at q c . The point ZZ on the<br />

axis corresponds to e zz . It can be seen that the smallest roll angle arctan ( p/q c ), where rolls<br />

are stable, first increases with increasing e but saturates at about 88. This is in agreement<br />

with experiments [36, 37, 113]. With increasing e the minimal angle decreases again <strong>and</strong>,<br />

surprisingly, beyond the point denoted by SV normal rolls could restabilize. It is interesting<br />

to note that the order parameter approach gives the ZZ destabilization qualitatively correct<br />

but fails to reproduce the restabilization of normal rolls [24]. We conclude that at e zz a forward-type<br />

bifurcation to oblique rolls can occur as has been observed.<br />

From the curve CR in Fig. 13.8b we see that at larger e a short wavelength instability<br />

comes into play, i. e. a roll system with a different wave-vector, particular in different orientation,<br />

starts growing. This may saturate the often-observed rectangular patterns or, for a<br />

non-symmetric superposition, lead to the sometimes-observed oblique modulated structures<br />

[17, 105]. In order to examine this possibility one would have to test the stability of such<br />

patterns by a suitable Galerkin procedure, which was done for normal rolls. However, since<br />

there are other possibilities, in particular turbulent states, this approach is not exhaustive<br />

<strong>and</strong> has to be complemented by simulations of the dynamics.<br />

9 Recently most material parameters of Merck Phase V have been measured, see note added at the<br />

end. The material parameters of I52 have either been measured directly or fitted to EHC measurements<br />

[49].<br />

284


13.5 Electrohydrodynamic convection<br />

Unfortunately, solving the full hydrodynamic equations is impossible. Therefore the<br />

order parameter approach, applied here, becomes useful. As mentioned before, this approach<br />

[24] (Section 13.3) captures all features of the secondary bifurcation scenarios found in rigorous<br />

calculations, though the ZZ destabilization line comes out too high for realistic material<br />

parameters. This may be corrected by using a larger value of s a /s k . When formulated<br />

in real space one is led to coupled amplitude equations which have been simulated numerically<br />

[21, 24] 10 . Below the ZZ line the system approaches rapidly the stable normal roll attractor<br />

when starting from r<strong>and</strong>om initial conditions. Beyond the ZZ line the situation becomes<br />

complicated. Starting slightly above the ZZ line, from small r<strong>and</strong>om initial conditions,<br />

one could either end up in a stationary zigzag pattern (Fig. 13.10 a) for s a /s k , e = 0.1,<br />

<strong>and</strong> ot q = 0.5) or in states with shorter <strong>and</strong> more sinusoidal undulations (Fig. 10 in<br />

Ref. [24]). Then a few slowly moving defects appeared to persist. There is no clear-cut separation<br />

between undulations <strong>and</strong> zigzag, analogue to the situation near the LP without the<br />

mean flow mentioned above.<br />

For larger e (e = 0.2), on the other h<strong>and</strong>, the pattern evolves into a state similar to the<br />

one shown in Fig. 13.10 b. This is a time-dependent state with alternating zigs <strong>and</strong> zags<br />

where defects are continuously generated in pairs <strong>and</strong> is reminiscent of some cases of the<br />

observed defect turbulence, although in the experiments the rolls often tend to be more<br />

aligned in the normal direction. This discrepancy could be an artefact of the order parameter<br />

approximation, where normal rolls do not restabilize at higher e (see above), so the roll orientation<br />

remains more oblique. The generation <strong>and</strong> annihilation of defects can be understood<br />

from an advection of the roll pattern by the mean flow, which amplifies small undulations.<br />

Because the anisotropy counteracts the bending of rolls the stress is released by<br />

straightening the rolls <strong>and</strong> dislocations are left behind.<br />

The situation is reminiscent of Rayleigh-Bénard convection in isotropic fluids where<br />

stable roll attractors apparently compete with complex patterns, spiral defect turbulence<br />

[115–117]. It has been shown very recently that if anisotropy is introduced into this system<br />

by inclining the convection cell, a normal roll pattern with dislocation defect turbulence occurs,<br />

which looks quite similar to patterns observed in EHC [118].<br />

Figure 13.10: (a) Stationary zigzag pattern for e = 0.1. (b) Snapshot of a defect turbulent zigzag pattern<br />

at e = 0.2.<br />

10 Similar results have been obtained from quite simple model equations if certain parameters are adjusted<br />

ad hoc [25, 114].<br />

285


13 Convection Instabilities in Nematic Liquid Crystals<br />

13.6 Concluding remarks<br />

Clearly this review could not be comprehensive, so we first list here some of the omitted<br />

material. Recently some experiments in planarly aligned samples with an additional destabilizing<br />

magnetic field in the z-direction (parallel to the electric field) were performed [119,<br />

120]. Some earlier work is discussed in [60]. For not too high magnetic fields the EHC<br />

threshold is merely reduced in accordance with the st<strong>and</strong>ard theory. For fields that are larger<br />

than a critical value, however, the Freédericksz transition comes first <strong>and</strong> then EHC sets in<br />

as a secondary instability, similarly to case F in homeotropic alignment, see below. The voltage<br />

threshold should then increase with increasing field, because destabilization of the<br />

Freédericksz distorted state becomes increasingly difficult, which is in qualitative agreement<br />

with experiments. Also the change to a subcritical bifurcation appears to be born out by theory<br />

[121].<br />

In the experiments with I52 oblique travelling rolls were observed. In this four-wave<br />

scenario (left-right <strong>and</strong> zigzag) a number of superpositions are possible, depending on the<br />

non-linear coefficients [93]. The left <strong>and</strong> right travelling roll systems always appear to separate,<br />

whereas in some parameter range the zigs <strong>and</strong> zags make a transition from separation<br />

to coexistence [51]. However, the superposed state is disordered (weakly turbulent). One can<br />

show theoretically that just beyond this codim-2 point a Benjamin-Feir-type modulational instability<br />

(Section 10.3) becomes generic, depending only on the sign of coefficients <strong>and</strong> not<br />

on their magnitude [122]. Presumably the above experiments were done in that parameter<br />

range.<br />

EHC is particularly suited to apply space or time-modulated forcing. The effect of<br />

spatially periodic forcing of stationary patterns in a quasi one-dimensional situation was studied<br />

by using structured electrodes [27, 123]. In particular, commensurate-to-incommensurate<br />

transitions were observed as predicted by theory based on phenomenological amplitude<br />

equations [124, 125]. Additional two-dimensional effects in spatially forced isotropic systems<br />

were studied theoretically [126]. Experiments appear to be lacking at this time. The effect<br />

of resonant time-periodic forcing of travelling (normal) rolls was first discussed theoretically<br />

using normal-form equations [127, 128] <strong>and</strong> was then confirmed experimentally [43].<br />

Interesting scenarios including a transition to st<strong>and</strong>ing, oscillating rolls can be induced in<br />

this way. Similar effects for travelling oblique rolls, where one has four-wave mixing, have<br />

also been considered [47, 129]. The effect of stochastic driving was investigated theoretically<br />

[96, 130] <strong>and</strong> experimentally [131, 132].<br />

Some statistical properties of defect turbulent states in EHC were studied experimentally<br />

[42, 133, 134] <strong>and</strong>, as pointed out before, the role of thermal fluctuations slightly below<br />

threshold was measured <strong>and</strong> analyzed. Of the numerous investigations of far-off threshold<br />

effects we only mention the study of phase waves in the oscillatory bimodal state [135, 136]<br />

<strong>and</strong> the transition between strongly turbulent states [60]. 11<br />

Besides the case of planar surface alignment of the director, considered here, one can<br />

have homeotropic alignment where by appropriate treatment of the confining plates the di-<br />

11 Very recently a multifractal analysis of electroconvective turbulence has been performed [170].<br />

286


13.6 Concluding remarks<br />

rector is oriented perpendicular to the plane. Then the system is isotropic in the plane of the<br />

layer <strong>and</strong> therefore from a symmetry point of view more related to convection in simple<br />

fluids. Both, RBC <strong>and</strong> EHC, have recently been investigated theoretically using the methods<br />

described before [81, 137, 138]. In RBC, when heated from below, a subcritical Hopf bifurcation<br />

is predicted, which at high stabilizing magnetic fields first transforms into a stationary<br />

bifurcation <strong>and</strong> subsequently becomes supercritical. The results are consistent with experiments,<br />

mostly without visualization of the patterns, by Guyon et al. [139]. One also obtains<br />

convection when heating from above leading to stationary squares. Application of a<br />

weak planar magnetic field renders the system anisotropic <strong>and</strong> then one has rolls in a small<br />

range above threshold. Experiments were performed in [140].<br />

In homeotropic EHC one has to distinguish two cases. For materials with positive or<br />

only very slightly negative dielectric anisotropy, like e. g. the newly introduced material I52<br />

in an appropriate temperature range [49–51], theory predicts a direct transition to convection<br />

(case C) in the form of stationary squares, except at large e a where one finds rolls. The<br />

linear results can be described in good approximation by an analytic threshold formula<br />

[138]. For materials with more negative e a , like MBBA, one has first a bend Freédericksz<br />

transition (case F) leading to a planar component of the director, which (ideally) singles out<br />

spontaneously a direction in the plane. Subsequently one has a transition to EHC, which on<br />

the linear level exhibits similar scenarios as in the planar case. Some experimental results<br />

have been reported in [107]. The tendency towards oblique rolls is more pronounced in the<br />

homeotropic case, <strong>and</strong> in fact MBBA should show oblique rolls at low frequency. Experiments<br />

without [141] <strong>and</strong> with a stabilizing magnetic field [142] have verified these predictions.<br />

On a deeper (non-linear) level there is, however, a very essential difference that makes<br />

this system rather unique. The direction singled out by the Freédericksz transition in the<br />

plane of the layer is the result of a spontaneously broken symmetry in the absence of an applied<br />

planar magnetic field. This makes itself known in experiments, of course, by the fact<br />

that the planar component is not really uniform over the cell but varies slowly in space being<br />

pinned by inhomogeneities. For a description (under ideal conditions) one has to couple the<br />

EHC mode from the beginning on with the Freédericksz Goldstone mode. For normal rolls<br />

this situation may often not be very important, although it could lead to a destabilization of<br />

the pattern from the very beginning. But for oblique rolls there is certainly a drastic consequence.<br />

One easily underst<strong>and</strong>s that oblique rolls exert a mean torque on the director via the<br />

velocity field. This torque cannot be balanced. So the director has to react by a rotation<br />

which in turn acts back on the rolls reducing their amplitude. Clearly there cannot be a static<br />

state even under ideal conditions <strong>and</strong> indeed in the experiments a pattern with slow dynamics<br />

shows up [141]. The weakly non-linear theory is being worked out now. Hopefully<br />

there will be more experiments to study the transition between order <strong>and</strong> chaos as the magnetic<br />

field is decreased <strong>and</strong> a better characterization of the chaotic state.<br />

Besides the planar <strong>and</strong> homeotropic alignment there is also the possibility of hybrid<br />

anchoring, i. e. planar on one side, homeotropic on the other. Then, in the basic state the director<br />

bends by p/2 when going from one plane to the other <strong>and</strong> there are in fact two symmetry<br />

equivalent directions for the bend to occur. Usually the preparation procedure of the<br />

probe, i. e. the filling with the LC, will single out one direction. Thus the basic state is not<br />

reflection symmetric <strong>and</strong> the pattern is expected to drift. Recently, this has been shown theoretically<br />

[143]. Under dc driving drift can also be imposed by non-ideal planar alignment,<br />

287


13 Convection Instabilities in Nematic Liquid Crystals<br />

i. e. by a pretilt [144, 145]. Indeed a small pretilt is hard to avoid. There is some similarity<br />

to systems with an externally imposed drift that has been studied recently in Taylor vortex<br />

flow in simple fluids [146, 147].<br />

<strong>Final</strong>ly we point out that the uniaxial symmetry of planarly oriented cells can be perturbed<br />

in at least two ways: by applying a planar magnetic field at an oblique angle to the<br />

alignment direction, which has been utilized to some extent in [19], or by imposing a twist<br />

on the director by having different planar alignment directions on the two plates. In this case<br />

one also has to break reflection symmetry around the midplane of the layer, which may be<br />

done by applying a dc voltage (the flexoelectric effect is linear in the field). As a consequence<br />

one loses the sharp separation between normal <strong>and</strong> oblique rolls <strong>and</strong> in fact the roll<br />

direction should turn smoothly by changing any parameter, e. g. the voltage. This has indeed<br />

been demonstrated experimentally [148]. Actually, twisted cells have interesting properties<br />

also under ac excitation [149]. We point out that twisted cells are employed in the most<br />

common type of LC displays. But the materials used there have strongly positive dielectric<br />

anisotropy that leads to a Freédericksz transition instead of EHC.<br />

Although during the last 12 years much progress could be achieved, there still remain<br />

many open problems. For planar alignment there exist some quantitative discrepancies with<br />

the SM which are possibly not explained by the WEM. It is also not yet clear if that model<br />

will explain the (very weakly) subcritical bifurcation, observed under some conditions. But<br />

this should be resolved in the near future.<br />

A different source of space charges is operative in the very beautiful electroconvection<br />

experiments in thin freest<strong>and</strong>ing smectic films [150, 151]. Here the free surface charges <strong>and</strong><br />

the inhomogeneity of the applied field certainly play a crucial role. Very recently a linear<br />

<strong>and</strong> weakly non-linear theory has been worked out [152].<br />

Acknowledgements<br />

We have benefited from discussions with G. Ahlers, A. Buka, W. Decker, M. Dennin,<br />

A. Hertrich, S. Rasenat, I. Rehberg, A. Rossberg, H. Richter, M. Treiber, <strong>and</strong> W. Zimmermann.<br />

A. Buka, I. Rehberg, H. Richter, <strong>and</strong> A. Rossberg have also kindly provided graphs<br />

of their results. We are grateful to W. Decker, A. Hertrich, <strong>and</strong> F. Schmögner for help in preparing<br />

the manuscript. Financial support by Deutsche Forschungsgemeinschaft (Sonderforschungsbereich<br />

213) is gratefully acknowledged. L. Kramer wishes to thank the Center<br />

Emile Borel, Institut Henry Poincare, <strong>and</strong> the ECM, Universidad de Barcelona, <strong>and</strong> W. Pesch<br />

the LASSP, Cornell University, where part of this work was performed, for their hospitalities.<br />

288


Note added<br />

Note added<br />

We here list some recent developments in EHC. Firstly, on the level of the SM, the weakly<br />

non-linear theory for homeotropic surface alignment in the usual case F, where one first encounters<br />

a bend Freédericksz transition, has been worked out [153, 154]. One obtains in the<br />

normal roll regime one GLE for the patterning mode (two equations in the oblique roll<br />

range) coupled to a dynamic equation for the Goldstone mode that signifies the broken continuous<br />

symmetry. In the normal roll case the equations essentially scale uniformly in e. It<br />

turns out that in the absence of any external symmetry breaking (no stabilizing magnetic<br />

field) no stable periodic roll solutions exist <strong>and</strong> one seems to have a disordered state in all<br />

cases. Theory suggests that the state is always dynamic, <strong>and</strong> the type of spatio-temporal defect<br />

chaos has been termed soft-mode turbulence [155, 156], some experiments appear to<br />

show in the normal roll range for very small e static (frozen) disorder [141, 142]. The destabilization<br />

of normal rolls is brought about by the important fact that, although the growth<br />

rate is maximal for the rolls oriented perpendicular to the (undistorted) director, they exert<br />

an abnormal torque around the z-axis on the director that amplifies fluctuation-induced misalignments<br />

away from the normal orientation. The abnormal torque is a signature of the<br />

non-variational nature of the system, even at onset. In the presence of a (weak) stabilizing<br />

magnetic field H there are stable rolls up to e c *H 2 . For normal rolls the destabilization occurs<br />

in a frequency range o L < o < o AR at e ZZ by a ZZ instability, <strong>and</strong> for higher frequencies<br />

by a spatially homogeneous transition (in the x-y plane) at e AR to abnormal rolls, where<br />

the director is rotated away from its normal position, either to the left or to the right. The<br />

term abnormal rolls has been introduced earlier in the context of experiments without magnetic<br />

field [166]. The rotation is again a manifestation of the abnormal torque. The abnormal<br />

rolls destabilize at 1.5 e c (in the weak-field limit) <strong>and</strong> then one has defect turbulence.<br />

Interestingly, at even higher values of e, the coupled GLEs describe in a substantial parameter<br />

range a new spatial superstructure where defects of equal topological charge tend to<br />

assemble along chains aligned parallel to the magnetic field. The chains form a periodic<br />

stripe pattern <strong>and</strong> the topological charge alternates from chain to chain. Between the chains<br />

the rolls (<strong>and</strong> the director) are rotated alternatively to the right <strong>and</strong> to the left. The roll angle<br />

is related to the defect density along a chain by a simple topological constraint. These structures<br />

are reminiscent of the chevrons, observed for more than 25 years, in the dielectric range<br />

of EHC in planarly aligned cells – for photographs see any of the st<strong>and</strong>ard textbooks [57, 65,<br />

66] – <strong>and</strong> it appears that a similar mechanism is operative there. The formation of chevrons<br />

can be modelled as a Turing-like instability arising in the defect turbulent state [157].<br />

Secondly, it has been realized recently that also in the conductive range of planarly<br />

aligned cells the transition to abnormal rolls occurs generically at e = e AR * 0.1 [158]. As in<br />

the homeotropic case the transition occurs at high frequencies in the stable range <strong>and</strong> can be<br />

preceded for lower frequencies (o L < o < o AR ) by a ZZ instability. In contrast to the (nearly)<br />

isotropic homeotropic case one finds for o < o AR restabilization of abnormal rolls above a<br />

certain e ARstab . Thus, in the oblique roll range <strong>and</strong> in the normal roll range below o AR normally<br />

oriented rolls reappear as abnormal rolls. The restabilization had been found before in<br />

the full Galerkin computations, described in Section 13.5.3 (Fig. 13.8b), without underst<strong>and</strong>ing<br />

the physical basis. Although in planarly aligned systems abnormal rolls cannot be distin-<br />

289


13 Convection Instabilities in Nematic Liquid Crystals<br />

guished from normal rolls by the usual optical techniques, there exists indirect evidence for<br />

their occurrence [158]. Increasing e further, the abnormal rolls are destabilized by a short<br />

wavelength bimodal varicose instability that transforms at large frequencies into a long wavelength<br />

instability of the SV-type.<br />

<strong>Final</strong>ly, we mention some recent progress in the application <strong>and</strong> analysis of the WEM<br />

model. The linear analysis was shown to describe quantitatively recent systematic measurements<br />

on Merck Phase V of V c , q c , <strong>and</strong> the Hopf frequency o c as a function of o for cells<br />

of different thickness [159]. This agreement is particularly noteworthy because it is based on<br />

recent measurements of the viscosities a 2 – a 6 [160] as well as on independent measurements<br />

of the elastic constants. The only remaining SM parameter a 1 was fitted to give the correct<br />

Lifshitz frequency. So it appears that the WEM can describe at least the three systematically<br />

studied materials MBBA [161], Merck Phase V, <strong>and</strong> I52 [49]. Meanwhile it has also become<br />

clear that the WEM model can explain the weakly subcritical nature of the bifurcation<br />

observed under some conditions in MBBA <strong>and</strong> Merck Phase V [38–40, 47] <strong>and</strong>, more recently,<br />

also in I52 [162–164]. Whereas the results for I52 show that the bifurcation is subcritical<br />

only in the stationary regime near the crossover to the Hopf bifurcation, in agreement<br />

with the predictions [165], one observes hysteretic behaviour in the other materials<br />

also in the Hopf range. This can be understood in terms of the subcritical stationary branch<br />

persisting in the non-linear range even in the region where the Hopf bifurcation precedes<br />

slightly the stationary one.<br />

References<br />

1. P. Manneville: in: Dissipative Structures <strong>and</strong> Weak Turbulence, Academic Press, New York, (1990)<br />

2. M.C. Cross, P.C. Hohenberg: Rev. Mod. Phys., 65, 851 (1993).<br />

3. F.H. Busse: Fundamentals of thermal convection, in: W.R. Peltier (ed.): Mantle convection, Plate<br />

Tectonics <strong>and</strong> Global Dynamics, Gordon <strong>and</strong> Breach, p. 23 (1989)<br />

4. A. Buka, L. Kramer: in: Pattern formation in liquid crystals, Springer, New York, (1995)<br />

5. R. Williams: J.Chem.Phys., 39, 384 (1963)<br />

6. A. Kapustin, L. Larinova: Kristallografya, 9, 297 (1963)<br />

7. E.F. Carr: Mol. Cryst. Liq. Cryst., 7, 253 (1969)<br />

8. W. Helfrich: J. Chem. Phys., 51, 4092 (1969)<br />

9. Orsay Liquid Crystal Group: Phys. Rev. Lett., 26, 1642 (1970)<br />

10. E. Dubois-Violette, P.G. de Gennes, O.J. Parodi: J. Phys. (Paris), 32, 305 (1971)<br />

11. P.A. Penz, G.W. Ford: Phys. Rev. A, 6, 414 (1972)<br />

12. E. Bodenschatz, W. Zimmermann, L. Kramer: J. Phys. (Paris), 49, 1875 (1988)<br />

13. W. Zimmermann, L. Kramer: Phys. Rev. Lett., 55, 402 (1985)<br />

14. L. Kramer, E. Bodenschatz, W. Pesch, W. Thom, W. Zimmermann: Liquid Crystals, 5(2), 699<br />

(1989)<br />

15. L. Kramer, W. Pesch: Electrohydrodynamic instabilities in nematic liquid crystals, in: A. Buka,<br />

L. Kramer (eds): Pattern formation in liquid crystals, Springer, New York, (1995)<br />

16. R. Ribotta, A. Joets: Defects <strong>and</strong> interactions with the structure in ehd convection in nematic liquid<br />

crystals, in: J. E. Wesfreid, S. Zaleski (eds): Cellular Structure in Instabilities, Springer, Berlin,<br />

p. 249 (1984)<br />

290


References<br />

17. A. Joets, R. Ribotta: J. Phys. (Paris), 47, 595 (1986)<br />

18. S. Kai, K. Hirakawa: Prog. Theor. Phys. Suppl., 64, 212 (1978)<br />

19. S. Kai, N. Chizumi, M. Kohno: Phys. Rev. A, 40, 6554 (1989)<br />

20. S. Kai: unpublished (1988)<br />

21. E. Bodenschatz, M. Kaiser, L. Kramer, W. Pesch, A. Weber, W. Zimmermann: Patterns <strong>and</strong> defects<br />

in liquid crystals, in: P. Coullet, P. Huerre (eds): New Trends in Nonlinear Dynamics <strong>and</strong> Pattern<br />

Forming Phenomena: The Geometry of Nonequilibrium, Plenum Press, NATO ASI Series, (1990)<br />

22. W. Pesch, W. Decker, Q. Feng, M. Kaiser, L. Kramer, A. Weber: Weakly nonlinear analysis of pattern<br />

formation in nematic liquid crystals, in: J. M. Coron, F. Helein, J. M. Ghidaglia (eds): Nematics:<br />

Mathematical <strong>and</strong> Physical Aspects, Kluwer Academic Publishers, Dordrecht, NATO ASI<br />

Series, p. 291 (1991)<br />

23. M. Kaiser, W. Pesch, E. Bodenschatz: Physica D, 59, 320 (1992)<br />

24. M. Kaiser, W. Pesch: Phys. Rev. E, 48, 4510 (1993)<br />

25. S. Sasa: Prog. Theor. Phys., 83, 824 (1990)<br />

26. S. Sasa: Prog. Theor. Phys., 84, 1009 (1990)<br />

27. M. Lowe, J. Gollub: Phys. Rev. Lett., 55, 2575 (1985)<br />

28. S. Rasenat, E. Braun,V. Steinberg: Phys. Rev. A, 42, 5728 (1991)<br />

29. S. Nasuno, S. Kai: Europhys. Lett., 14(8), 779 (1991)<br />

30. S. Nasuno, S. Takeuchi, Y. Sawada: Phys. Rev. A, 40, 3457 (1989)<br />

31. G. Goren, I. Procaccia, S. Rasenat, V. Steinberg: Phys. Rev. Lett., 63, 1237 (1989)<br />

32. S. Rasenat, V. Steinberg, I. Rehberg: Phys. Rev. A, 42, 5998 (1990)<br />

33. E. Braun,V. Steinberg: Europhys. Lett., 15(2), 167 (1991)<br />

34. S. Rasenat: PhD thesis, Universität Bayreuth, (1991)<br />

35. L. Kramer, E. Bodenschatz, W. Pesch: Phys. Rev. Lett., 64, 2588<br />

36. E. Braun, S. Rasenat, V. Steinberg: Europhys. Lett., 15(6), 597 (1991)<br />

37. S. Nasuno, O. Sasaki, S. Kai, W. Zimmermann: Stability diagram in electrohydrodynamic convection<br />

of nematics electrohydrodynamic convection in nematics: the homeotropic case, in: S. Kai<br />

(ed.): Pattern formation in complex dissipative systems <strong>and</strong> Global Dynamics, World scientific,<br />

p. 275 (1992)<br />

38. I. Rehberg, S. Rasenat, M. de la Torre Juarez, W. Schöpf, F. Hörner, G. Ahlers, H. Br<strong>and</strong>: Phys.<br />

Rev. Lett., 67, 596 (1991)<br />

39. I. Rehberg, F. Hörner, L.Chiran, H. Richter, B. Winkler: Phys.Rev. A, 44, R7885 (1991)<br />

40. I. Rehberg, F. Hörner, G. Hartung: J. Stat. Phys., 64, 1017 (1991)<br />

41. A. Joets, R. Ribotta: Phys. Rev. Letters, 60, 2164 (1988)<br />

42. I. Rehberg, S. Rasenat, V. Steinberg: Phys. Rev. Lett., 62, 756 (1989)<br />

43. I. Rehberg, S. Rasenat, M. de la Torre-Juarez, V. Steinberg: Phys. Rev. Lett., 61, 2448 (1988)<br />

44. N.V. Madhusudana,V.A. Raghunathan, K.R. Sumathy: Pramana J. Phys., 28, L311 (1987)<br />

45. V. Raghunathan, P.N. Madhusudana: Pramana J. Phys., 131, 163 (1988)<br />

46. W. Thom, W. Zimmermann, L. Kramer: Liq. Cryst., 4, 309 (1989)<br />

47. M. de la Torre Juarez, I. Rehberg: Phys.Rev. A, 42, 2096 (1990)<br />

48. M. Treiber, L. Kramer: Mol. Cryst. Liq. Cryst., 261, 951/311 (1995)<br />

49. M. Dennin, M. Treiber, L. Kramer, G. Ahlers, D. Cannell: Phys. Rev. Lett., 76, 319 (1995)<br />

50. M. Dennin, G. Ahlers, D. Cannell: Measurement of material parameters of the nematic crystal I52,<br />

in: P. Cladis, P. Palffy-Muhoray (eds): Spatio-temporal patterns in nonequilibrium complex systems,<br />

Santa Fe Institute Studies in the Sciences of Complexity XXI, Addison-Wesley, New York,<br />

(1994)<br />

51. M. Dennin, D. Cannell, G. Ahlers: Mol. Cryst. Liq. Cryst., 261, 977/337 (1995)<br />

52. E. Dubois-Violette: C. R. Acad. Sci., 21, 923 (1971)<br />

53. P.J. Barratt: Liquid Crystals, 4(3), 223 (1989)<br />

54. Q. Feng, W. Pesch, L. Kramer: Phys. Rev. A, 45, 7242 (1992)<br />

55. L.I. Berge, G. Ahlers, D. S. Cannell: Phys. Rev. E, 48, 3236 (1993)<br />

56. E. Dubois-Violette, G. Dur<strong>and</strong>, E. Guyon, P. Manneville, P. Pieranski: Instabilities in nematic liquid<br />

crystals, in: L. Liebert (ed): Solid state physics, Suppl. 14, Academic Press, New York, p. 147 (1978)<br />

291


13 Convection Instabilities in Nematic Liquid Crystals<br />

57. L.M. Blinov: in: Electrooptical <strong>and</strong> Magnetooptical <strong>Properties</strong> of Liquid Crystals, John Wiley,<br />

New York, (1983)<br />

58. A. Pikin: in: Structural Transformations in Liquid Crystals, Gordon <strong>and</strong> Breach Science Publishers,<br />

New York, (1991)<br />

59. I. Rehberg, B.L. Winkler, M. de la Torre, S. Rasenat, W. Schöpf: Adv. Solid State Physics, 29, 35<br />

(1989)<br />

60. S. Kai, W. Zimmermann: Prog. Theor. Phys. Suppl., 99, 458 (1989)<br />

61. W. Zimmermann: MRS Bulletin, XVI, 46 (1991)<br />

62. S. Kai: Forma, 7, 189 (1992)<br />

63. L. Kramer, W. Pesch: Annu. Rev. Fluid Mech., 27, 515 (1995)<br />

64. G. Ahlers: Experiments on thermally driven convection, in: A. Buka, Kramer (eds.): Pattern formation<br />

in liquid crystals, Springer, New York, (1995)<br />

65. P. de Gennes, J. Prost: in: The Physics of Liquid Crystals, Clarendon Press, Oxford, (1993)<br />

66. S. Ch<strong>and</strong>rasekhar: in: Liquid Crystals, University Press, Cambridge (1992)<br />

67. M.J. Stephen, J. P. Straley: Rev. Mod. Phys., 46, 617 (1974)<br />

68. H. Pleiner, H. Br<strong>and</strong>t: Hydrodynamics <strong>and</strong> electrohydrodynamics of liquid crystals, in: A. Buka, L.<br />

Kramer (eds.): Pattern formation in liquid crystals, Springer, New York, (1995)<br />

69. F.C. Frank: Discuss. Faraday Soc., 25, 19 (1958)<br />

70. R.B. Meyer: Phys. Rev. Lett., 22, 918 (1969)<br />

71. J.L. Ericksen: Arch. Ration. Mech. Anal., 4, 231 (1960)<br />

72. F.M. Leslie: Quart. J. Mech. Appl. Math., 19, 357 (1966)<br />

73. E. Dubois-Violette, P. Manneville: Flow instabilities in nematics, in: A. Buka, L. Kramer (eds.):<br />

Pattern formation in liquid crystals, Springer, New York, (1995)<br />

74. O. Parodi: J. Phys. (Paris), 31, 581 (1970)<br />

75. F.H. Busse, E.W. Bolton: J. Fluid Mech., 146, 115 (1984)<br />

76. E. Dubois-Violette: J. Physique, 33, 95 (1972)<br />

77. A. Buka, L. Kramer: Theory of transient patterns in the splay freedericksz transition of nematics,<br />

in: S. Kai (ed.): Pattern formation in complex dissipative systems, World Scientific, Kitakyushu,<br />

Japan, (1991)<br />

78. H. Haken: in: Synergetics, Springer, (1978)<br />

79. M.C. Cross: Phys. Fluids, 23, 1727 (1980)<br />

80. A.C. Newell, T. Passot, J. Lega: Annu. Rev. Fluid Mech., 25, 399 (1993)<br />

81. Q. Feng, W. Decker, W. Pesch, L. Kramer: J. Phys. France, II 2, 1303 (1992)<br />

82. W. Pesch, L. Kramer, General mathematical description of pattern forming instabilities, in: A.<br />

Buka, L. Kramer (eds.): Pattern formation in liquid crystals, Springer, New York, (1995)<br />

83. W. Decker, W. Pesch: J.Phys. France, II 4, 493 (1994)<br />

84. W. Eckhaus: in: Studies in Nonlinear Stability Theory, Springer, New York (1965)<br />

85. F.H. Busse: Rep. Prog. Phys., 41, 1929 (1978)<br />

86. F.H. Busse, R.M. Clever: J. Fluid Mech., 91, 319 (1979)<br />

87. V. Ginzburg, L. L<strong>and</strong>au: Zh. Eksp. Teor. Fiz., 20, 1064 (1950)<br />

88. E. Bodenschatz, W. Pesch, L. Kramer: Physica D, 32, 135 (1988)<br />

89. W. Pesch, L.Kramer: Z.Phys. B, 63, 121 (1986)<br />

90. L. Kramer, E. Bodenschatz, W. Pesch, W. Zimmermann: in: W. Güttinger, G. Dangelmayr (eds.):<br />

The Physics of Structure Formation, Springer, Berlin, (1987)<br />

91. A.C. Newell, J. A. Whitehead: J. Fluid Mech., 38, 279 (1969)<br />

92. L.A. Sege: J. Fluid Mech., 38, 203 (1969)<br />

93. M. Silber, H. Riecke, L. Kramer: Physica D, 61, 260 (1992)<br />

94. A. Zippelius, E. Siggia: Phys. Fluids, 26, 2905 (1983)<br />

95. P. Kishore, T. Raj, A. Iqbal, S. Sastry, G. Satyan<strong>and</strong>am: Liquid Crystals, 14, 1319 (1993)<br />

96. Müller, U. Behn: Z. Phys. B – Condensed Matter, 69, 185 (1987)<br />

97. F. Hoerner, I. Rehberg: Using thermal noise to measure the correlation lengths of electroconvection,<br />

in: S. Kai (ed.): Pattern formation in complex dissipitative systems <strong>and</strong> Global Dynamics,<br />

World scientific, p. 429 (1992)<br />

292


References<br />

98. B. Winkler, W. Decker, H. Richter, I. Rehberg: Physica D, 61, 284 (1992)<br />

99. S. Rasenat, G. Hartung, B. Winkler, I. Rehberg: Exp. in Fluids, 7, 412 (1989)<br />

100. L. Kramer, H. Schober, W. Zimmermann: Physica D, 31, 212 (1988)<br />

101. G. Tesauro, M. C. Cross: Phys. Rev. A, 34, 1363 (1986)<br />

102. E.D. Siggia, A. Zippelius: Phys. Rev. A, 24, 1036 (1981)<br />

103. L.M. Pismen, J. D. Rodriguez: Phys. Rev. A, 42, 2471 (1990)<br />

104. E. Bodenschatz, A. Weber, L. Kramer: J. Stat. Phys., 64, 1007 (1991)<br />

105. A. Joets, R. Ribotta: Electro-hydrodynamical convective structures <strong>and</strong> transitions to chaos in liquid<br />

crystals, in: J. E. Wesfreid, S. Zaleski (eds.): Cellular Structure in Instabilities, Springer,<br />

Berlin, p. 294 (1984)<br />

106. C. Hilsum, F. Saunders: Mol. Cryst. Liq. Cryst., 64, 25 (1980)<br />

107. S. Kai, Y. Adachi, S. Nasuno: Stability diagram, defect turbulence, <strong>and</strong> new patterns in electroconvection<br />

in nematics, in: P.E. Cladis, P. Palffy-Muhoray (eds.): Spatio-Temporal Patterns in Nonequilibrium<br />

Complex Systems, SFI Studies in the Sciences of Complexity, Addison-Wesley, (1994).<br />

108. S. Nasuno, O. Sasaki, S. Kai, W. Zimmermann: Phys. Rev. A, 46, 4954 (1992)<br />

109. L Kramer, W. Zimmermann, E. Bodenschatz, W. Pesch, in: J.E. Wesfried, R. Br<strong>and</strong>, P. Manneville,<br />

G. Albinet, N. Boccara (eds.): Propagation in Systems far from Equilibrium, Springer, Berlin,<br />

(1988)<br />

110. E. Bodenschatz: PhD thesis, Universität Bayreuth, (1989)<br />

111. M. Kaiser: PhD thesis, Universität Bayreuth, (1992)<br />

112. W. Decker: PhD thesis, Universität Bayreuth, (1995)<br />

113. V. Steinberg: private communication.<br />

114. S. Sasa: Defect chaos in 2d anisotropic systems, in: S. Kai (ed.): Pattern formation in complex<br />

dissipitative systems <strong>and</strong> Global Dynamics, World scientific, p. 336 (1992)<br />

115. S. Morris, E. Bodenschatz, D.S. Cannell, G. Ahlers: Phys. Rev. Lett., 71, 2026 (1993)<br />

116. M. Assenheimer,V. Steinberg: Phys. Rev. Lett., 70, 3888 (1993)<br />

117. W. Decker, W. Pesch, A. Weber: Phys. Rev. Lett., 73, 648 (1994)<br />

118. B.B. Plapp, E. Bodenschatz: Bull. APS., 40, 765 (1995)<br />

119. J. Gleeson, D. Todorovic-Marinic: Mol. Cryst. Liq. Cryst., 261, 327 (1995)<br />

120. A. Hertrich, W. Pesch, J. Gleeson: Europhys. Lett., 34, 417 (1996)<br />

121. A. Hertrich: PhD thesis, Universität Bayreuth, (1995)<br />

122. H. Riecke, L. Kramer: unpublished (1995)<br />

123. M. Low, B. Albert, J. Gollub: J. Fluid. Mech., 173, 253 (1986)<br />

124. P. Coullet: Phys. Rev. Lett., 56, 724 (1986)<br />

125. P. Coullet, P.Huerre: Physica D, 63, 123 (1986)<br />

126. W. Zimmermann, A. Ogawa, S. Kai, K. Kawasaki, T. Kawakatsu: Europhys. Lett., 24, 217 (1993)<br />

127. H. Riecke, J. Crawford, E. Knobloch: Phys. Rev. Lett., 61, 1942 (1988)<br />

128. D. Walgraef: Europhys.Lett., 7, 495 (1988)<br />

129. H. Riecke, M. Silber, L. Kramer: Phys.Rev., E49, 4100 (1994)<br />

130. A. Lange, R. Müller, U. Behn: Z. Phys. B, 100, 477 (1996);<br />

U. Behn, A. Lange, T. John: preprint (1997)<br />

131. S. Kai, in: S. Moss, P. McClintock (eds.): Noise in Nonlinear Dynamical Systems, volume 3,<br />

Cambridge University Press, p. 22 (1989)<br />

132. U. Behn, A. Buka, N. Kloepper: unpublished (1995)<br />

133. L. Gil, J. Lega, J. Meunier: Phys. Rev. A, 41, 1138 (1990)<br />

134. S. Kai, M. Kohno, M. Andoh, M. Imizaki, W. Zimmermann: Mol. Cryst. Liq. Cryst., 198, 247<br />

(1991)<br />

135. M. Sano, H. Kokubo, H. Janiaud, K. Sato: Prog. Theor. Phys., 90, 1 (1993)<br />

136. V. Delev, O. Scaldin, A. Chuvyrov: Liquid Crystals, 12, 441 (1992)<br />

137. A. Hertrich, W. Decker, W. Pesch, L. Kramer: J. Phys. (France), II 2, 1915 (1992)<br />

138. L. Kramer, A. Hertrich, W. Pesch: Electrohydrodynamic convection in nematics: the homeotropic<br />

case, in: S. Kai (ed.): Pattern formation in complex dissipitative systems <strong>and</strong> Global Dynamics,<br />

World scientific, p. 238 (1992)<br />

293


13 Convection Instabilities in Nematic Liquid Crystals<br />

139. E. Guyon, P. Pieranski, J. Salan: J. Fluid Mech., 93, 65 (1979)<br />

140. J. Salan, E. Guyon: J. Fluid Mech., 126, 13 (1983)<br />

141. H. Richter, A. Buka, I. Rehberg: Phys. Rev. E, 51, 5886 (1995)<br />

142. H. Richter, N. Klöpper, A. Hertrich, A. Buka: Europhys. Lett., 30, 37 (1995)<br />

143. A. Hertrich, A.P. Krekhov, W. Pesch: J. Phys. (France), II 5, 773 (1995)<br />

144. V. Raghunathan, P.M. Murthy, N. Madhusudana: Mol. Cryst. Liq. Cryst. 199, 48 (1992)<br />

145. W. Zimmermann: On travelling waves in nematics, in: J.M. Coron, F. Helein, M. Ghidaglia<br />

(eds.): Nematics: Mathematical <strong>and</strong> Physical Aspects, Kluwer Academic Publishers, NATO ASI<br />

Series, Dordrecht, pages 401–426 (1991)<br />

146. K. Babcock, G. Ahlers, D. Cannell: Phys. Rev. Lett., 67, 3388 (1991)<br />

147. A. Tsameret, V. Steinberg: Europhys. Lett., 14, 331 (1991)<br />

148. S. Frunza, R. Moldovan, T. Beica, M. Giurgea, D. Stoenescu: Europhys. Lett., 20, 407 (1992)<br />

149. A. Hertrich, A. P. Krekhov, O. A. Scaldin: J. Phys. (France), II 4, 239 (1994)<br />

150. Z.A. Daya, S. Morris, J.R. de Bruyn, A. May: Phys. Rev. A, 44, 8146 (1991)<br />

151. S.S. Mao, J.R. de Bruyn, S.W. Morris: Phys. Rev. E, 54, R1048 (1996)<br />

152. Z.A. Daya, S.W. Morris, J.R. de Bruyn: Phys. Rev. E, 55, 2682 (1997); V.B. Deyirmenjian, Z.A.<br />

Daya, S. Morris: preprint (1997)<br />

153. A.G. Rossberg, A. Hertrich, L. Kramer, W. Pesch: Phys. Rev. Lett., 76, 4729 (1996)<br />

154. A.G. Rossberg, L. Kramer: Physica Scripta, T67, 121 (1996)<br />

155. S. Kai, K Hayashi, Y. Hidaka: J. Phys. Chem., 100, 19007 (1996)<br />

156. Y. Hidaka, J.-H. Huh, K. Hayashi, M. Tribelski, <strong>and</strong> S. Kai, Phys. Rev. E56, R6256 (1997)<br />

157. A.G. Rossberg, L. Kramer: Physica D, 115, 19 (1998)<br />

158. E. Plaut, W. Decker, A. Rossberg, L. Kramer, W. Pesch: Phys. Rev. Lett., 79, 2367 (1997)<br />

159. M. Treiber, N. Eber, Á. Buka, L. Kramer: J. Phys. (France), II 7, 649 (1997)<br />

160. H.H. Graf, H. Kneppe, Schneider: Molecular Physics, 77, 521 (1992)<br />

161. M. Treiber: PhD thesis, Universität Bayreuth, (1996)<br />

162. M. Dennin, G. Ahlers, D.S. Cannell: Phys. Rev. Lett., 77, 2475 (1996)<br />

163. M. Dennin, G. Ahlers, D.S. Cannell: Science, 272, 388 (1997)<br />

164. M. Dennin, D.S. Cannell, G.Ahlers: Phys. Rev. E, 57, 638 (1998)<br />

165. M. Treiber, L. Kramer: Phys. Rev. E, 58, 1973 (1998)<br />

166. H. Richter, A. Buka, <strong>and</strong> I. Rehberg, in: P.E. Cladis <strong>and</strong> P. Palffy-Muhoray (eds.): Spatio-Temporal<br />

Patterns in nonequilibrium Complex Systems, Addison-Wesley Publishing Company, (1994)<br />

167. Ribotta, private communication<br />

168. S. Rasenat, private communication<br />

169. I. Rehberg, private communication<br />

170. V. Carbone, N. Scaramuzza, <strong>and</strong> C. Versace, Physica D, 106, 314 (1997)<br />

171. E. Plaut <strong>and</strong> W. Pesch<br />

294


14 Preparation <strong>and</strong> <strong>Properties</strong> of Ionic <strong>and</strong> Surface<br />

Modified Micronetworks<br />

Michael Mirke, Ralf Grottenmüller, <strong>and</strong> Manfred Schmidt<br />

14.1 Introduction<br />

In recent years micronetworks have gained increasing interest in academic <strong>and</strong> industrial research.<br />

Like dendrimers microgels could principally serve as molecular containers in order<br />

to transport guest molecules (drug delivery) or as microscopic reaction sites for the formation<br />

of nanosized colloidal particles.<br />

In the present work we have<br />

a) investigated the preparation <strong>and</strong> surface modification of microgels via crosslinking reactions<br />

in microemulsion;<br />

b) utilized microgels as the core in core-shell <strong>and</strong> star-like structures;<br />

c) prepared <strong>and</strong> characterized ionic micronetworks.<br />

14.2 Polymerization in normal microemulsion<br />

14.2.1 Mechanism <strong>and</strong> size control<br />

Antonietti et al. [1] <strong>and</strong> Wu [2] have reported on the size control of the polymerized microemulsion<br />

in the system: styrene, di-isopropenylbenzene (cross-linker), cetyltrimethylammoniumchloride<br />

(CTMACl) or a similar surfactant, <strong>and</strong> water. According to a simple model<br />

developed by Wu [2] the resulting latex size is a function of the fleet ratio s = m s /m M with<br />

m s <strong>and</strong> m M , the masses of surfactant <strong>and</strong> monomer respectively, given by<br />

S 1 ˆ NA r<br />

3M s<br />

a 0 R ‡ C …1†<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

295


14 Preparation <strong>and</strong> <strong>Properties</strong> of Ionic <strong>and</strong> Surface Modified Micronetworks<br />

with N A the Avogadro number, M s the molar mass of the surfactant, a 0 the surface area per<br />

surfactant molecule, r the mean particle density, <strong>and</strong> the constant [3]<br />

C ˆ NA r<br />

3M s<br />

a 0 2d 1 …2†<br />

with d the surfactant microgel interpenetration depth.<br />

Whereas Eq. 1 is almost perfectly confirmed by the experimental data of Antonietti et<br />

al. <strong>and</strong> Wu, the present data do not follow Eq. 1 <strong>and</strong> show much more scatter than literature<br />

results (Fig. 14.1). Variation of the ionic strength of the continuous phase, of the reaction<br />

temperature, of the total monomer content, <strong>and</strong> of the concentration of crosslinking agent<br />

has essentially no effect on the resulting microemulsion.<br />

6<br />

5<br />

4<br />

S -1<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

0 5 10 15 20 25 30 35 40 45 50 55 60<br />

R h /nm<br />

Figure 14.1: Inverse fleet ratio S –1 versus size R h for AIBN initiated, polymerized latices: data from<br />

Ref. [1] (_), Ref. [2] (M), <strong>and</strong> present results (y).<br />

As shown in Fig. 14.2 differences, however, were observed for different initiators, e. g.<br />

AIBN, redox system (K 2 S 2 O 8 /K 2 S 2 O 5 ), <strong>and</strong> dibenzylketone. Although much effort was taken<br />

to improve the reproducibility <strong>and</strong> size control we could never reproduce the literature<br />

data for AIBN initiation particularly at small S. We presently cannot offer a satisfactory explanation<br />

for this discrepancy.<br />

Concerning the mechanism of particle formation the polymerization starts in a two<br />

phase or in a monophasic region of the phase diagram depending on the total monomer concentration<br />

<strong>and</strong> on the fleet ratio [4]. The biphasic regime is easily detected by a bimodal decay<br />

of the time correlation function g 1 (t) as recorded by a dynamic light scattering instrument<br />

whereas for the monophase an extremely fast, monomodal decay curve is observed.<br />

The hydrodynamic radius R h , deducted from the measured diffusion coefficient, is less than<br />

1 nm which is – within experimental error – identical to the radius of empty micelles formed<br />

by the pure surfactant in water at the respective concentration. It is tempting to interpret<br />

these results in terms of a bicontinuous monophase rather than in terms of monomer-filled<br />

microdroplets. Since similar investigations are currently performed in Hoffmann’s group [5],<br />

296


14.2 Polymerization in normal microemulsion<br />

6<br />

5<br />

4<br />

S -1<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

0 5 10 15 20 25 30 35 40 45 50 55 60<br />

R h /nm<br />

Figure 14.2: Inverse fleet ratio S –1 versus size R h for differently initiated polymerized lattices: AIBN<br />

(y), redox (o), <strong>and</strong> photo initiation by dibenzylketone ($).<br />

we have not further persued the phase diagram of the unpolymerized emulsion but concentrated<br />

on the changes during polymerization.<br />

As shown in Fig. 14.3, the bimodal decay of the correlation function has merged into a<br />

single, almost monoexponential curve after 90 min reaction time, which does not change with<br />

monomer conversion any more. The final latex size is already reached after 90 minutes or at<br />

about 50 % monomer conversion! Again, this observation is not easily understood, because it<br />

means that polymerized particles do not grow beyond a certain size <strong>and</strong> new particles are<br />

formed even during the later course of reaction. Surface tension measurements as function of<br />

the reaction time <strong>and</strong> as function of the fleet ratio only show the well-known surface activity<br />

of styrene as compared to polystyrene, which is still significant at 50 % monomer conversion.<br />

At this point also the rate of conversion exhibits a maximum as function of time which is<br />

1.0<br />

0.5<br />

g 1 (t)<br />

0.1<br />

0 100 200 300 400 500 600 700 800 900<br />

t/msec<br />

Figure 14.3: Time correlation function g 1 (t) for AIBN initiated latices at different polymerization<br />

times: 0 min (T), 30 min (y), 90 min (o), 150 min (M), <strong>and</strong> 2 days (6).<br />

297


14 Preparation <strong>and</strong> <strong>Properties</strong> of Ionic <strong>and</strong> Surface Modified Micronetworks<br />

usually interpreted in terms of a two stage mechanism. At the beginning of the polymerization<br />

more <strong>and</strong> more growing particles are formed until a certain number of particles is reached.<br />

Then the reaction rate decreases because the monomer concentration in the system decreases<br />

or the number of active polymerization sites is diminished. In view of the discussion above,<br />

the hypothesis of the constant number of growing particles appears questionable.<br />

Qualitatively similar results are obtained for emulsions which were initiated by light<br />

or redox initiators. Despite the extreme lack of underst<strong>and</strong>ing <strong>and</strong> reproducibility in most<br />

cases the resulting microemulsions were monodisperse <strong>and</strong> spherical. Therefore they are<br />

principally well suited as starting material for surface-functionalized spherical particles.<br />

14.2.2 Surface functionalization of microgels<br />

A suitable functional group on the surface of the microgel could be an azo-moiety which<br />

can be utilized to start a grafting from reaction in order to synthesize core-shell or star-like<br />

structures.<br />

Here, we have taken advantage of the cosurfactant properties of the methacrylic acid<br />

ester (Fig. 14.4) which is simply added with DTMACl to the reaction mixture. During the<br />

polymerization the methacrylic group is bound into the microgel thus forming the pendent<br />

azo group. Since we wish to preserve the azo group during microgel formation the microemulsion<br />

was redox initiated. Redox initiation with the system K 2 S 2 O 8 /K 2 S 2 O 5 introduces ionic<br />

SO – 3 groups to each primary chain, thus creating an ionic micronetwork. In fact, increasing<br />

the initiator to monomer ratio results in an increased conductivity of the microgels dissolved<br />

in DMF.<br />

O<br />

O<br />

N N CN<br />

CH CN<br />

3<br />

Figure 14.4: Cosurfactant containing the azo-moieties methacrylic acid [4-(1,1-dicyanoethyl)azo]benzylic<br />

ester.<br />

The presence of ionic groups makes the particle characterization somewhat ambigious,<br />

since now interparticle interactions might seriously influence the interpretation of the<br />

light scattering data. One example is shown in Tab. 14.1, where the measured apparent radius<br />

of gyration R g,app of the latex in water changes from negative values to about 17 nm with increasing<br />

salt concentration, whereas the hydrodynamic radius R h remains approximately constant<br />

at about 8.5 nm. This large ratio of R g /R h = 2 is not compatible with the value of<br />

0.775, expected for a spherical structure.<br />

Also, most of the redox-initiated structures do not dissolve in toluene after careful removal<br />

of the surfactant <strong>and</strong> the solution characterization was performed in DMF. Again, the<br />

experimental R g /R h values were larger than 1.5 indicating that no spherical structures were<br />

298


14.3 Polymerization in inverse microemulsion<br />

Table 14.1: Influence of added salt on the light scattering results of redox polymerized latices.<br />

c (NaCl)/mol/l R g,app /nm R h,app /nm R g /R h<br />

0 –10.5 8.1 –<br />

4 610 –3


14 Preparation <strong>and</strong> <strong>Properties</strong> of Ionic <strong>and</strong> Surface Modified Micronetworks<br />

Br<br />

O<br />

CH 3 CH 3<br />

O<br />

N<br />

+<br />

O<br />

O<br />

CH 3<br />

N<br />

CH 3<br />

O<br />

Monomer I<br />

CH 3 CH 3<br />

CH 3 Br<br />

O<br />

N Br -<br />

+ CH 3<br />

Monomer II<br />

O<br />

O<br />

O - Na + + O S<br />

O<br />

O<br />

O SO 3 - Na +<br />

Monomer III<br />

O<br />

O<br />

N<br />

H<br />

OH<br />

N<br />

H<br />

O<br />

Crosslinker I<br />

Figure 14.5a: Structure of monomers <strong>and</strong> cross-linking agents employed in this study: methacrylic<br />

acid-[ethyl-dimethyl-benzyl-ammonium bromide]ester (Monomer I),: methacrylic acid-[trimethylammonium<br />

bromide]ester (Monomer II), methacrylic acid-[propyl-3-sodium sulfonate]ester (Monomer III),<br />

bis-(acrylamido)acetic acid (Crosslinker I), <strong>and</strong> 1,2-dihydroxyethylene-bis-acrylamide (Crosslinker II).<br />

O<br />

N<br />

H<br />

OH<br />

OH<br />

H<br />

N<br />

Crosslinker II<br />

O<br />

O<br />

HO<br />

HO<br />

HO<br />

O<br />

O<br />

O<br />

O<br />

y<br />

O<br />

z<br />

O<br />

x<br />

O<br />

O<br />

Tween 80<br />

O<br />

HO<br />

O<br />

HO<br />

O<br />

Span 20<br />

HO<br />

Figure 14.5b: Structures of surfactants employed in this study: polyoxyethylene(20)-sorbitane-monooleate<br />

(Tween 80), sorbitane monolaurate (Span 20).<br />

In Fig. 14.6 the measured apparent particle radii are shown as function of the ionic<br />

strength. The ionic strength dependence does not clearly emerge from the plot because above<br />

a concentration of 0.01 m salt the particles start to aggregate <strong>and</strong> eventually precipitate, <strong>and</strong><br />

below 0.001 m salt electrostatic effects prohibit a proper size evaluation, as will be discussed<br />

in detail below. It appears, however, that due to the high cross-linking density the particles<br />

300


14.3 Polymerization in inverse microemulsion<br />

Table 14.2: Characterization of the ionic microgels in aqueous 0.01 m KBr solution.<br />

S –1 M w 710 6 g mol –1 R g /nm R h /nm R g /R h Monomer 1)<br />

0.3 3.97 2) 16.8 18.2 0.92 I<br />

0.3 1.48 12.9 16.4 0.78 II<br />

0.3 1.32 18.7 21.8 0.85 III<br />

0.35 3.97 2) 13.5 18.9 0.71 I<br />

0.35 1.44 16.0 18.3 0.87 II<br />

0.35 1.84 23.2 23.8 0.97 III<br />

0.4 5.99 2) 14.9 21.1 0.71 I<br />

0.4 1.45 13.5 18.5 0.73 II<br />

0.4 1.83 20.3 22.7 0.89 III<br />

0.45 6.26 2) 15.4 21.9 0.70 I<br />

0.45 2.22 15.4 21.2 0.73 II<br />

0.45 2.27 20.3 25.6 0.79 III<br />

0.50 7.42 2) 16.0 23.2 0.69 I<br />

0.50 2.76 15.8 19.8 0.80 II<br />

0.50 3.69 22.9 27.8 0.82 III<br />

0.55 10.2 2) 17.8 24.7 0.72 I<br />

0.55 5.45 25.4 25.6 0.99 II<br />

0.60 15.3 2) 20.0 25.4 0.79 I<br />

0.60 5.33 22.8 25.0 0.91 II<br />

1) Monomers as shown in Fig. 14.5: I: 15 mol % crosslinker I<br />

II: 10 mol % crosslinker II<br />

III: 10 mol % crosslinker I<br />

2) Approximate molar masses measured in methylformamide solution with an estimated dn/dc = 0.1 cm 3 /g<br />

radius [nm]<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0.0001 0.001 0.01 0.1<br />

log ( c (KBr)<br />

[mol/l ] )<br />

Figure 14.6: Ionic strength dependency of the microgel dimension for monomer II with 10 wt% of<br />

crosslinker I: R h (y), R g (T).<br />

do not swell much. This is a bit unexpected since conductivity measurements have indicated<br />

the salt concentration within the microgel to be in the order of 10 –1 m. Although this largely<br />

reduces the electrostatic expansion of the chains between the crosslinks, it induces an enormous<br />

osmotic pressure of water towards the inside of the microgel which should cause a<br />

301


14 Preparation <strong>and</strong> <strong>Properties</strong> of Ionic <strong>and</strong> Surface Modified Micronetworks<br />

0.05<br />

0.04<br />

η sp /c [l/g]<br />

0.03<br />

0.02<br />

0.01<br />

0.00<br />

0 1 2 3 4 5<br />

c[g/l]<br />

Figure 14.7: Reduced viscosity of the anionic microgel (monomer III, 10 wt% crosslinker I) with molar<br />

mass M w = 4.97610 6 g/mol. Salt-free water (y), 0.01 m potassium bromide solution (T).<br />

strong swelling. This point is still under investigation along with peculiar interparticle, electrostatic<br />

interaction effects affecting the solution properties of the micronetworks, as already<br />

discussed above.<br />

Here, we have investigated the interparticle effects in more detail. In Fig. 14.7 the reduced<br />

viscosity Z sp /c of the sulfonated microgel is shown as function of the microgel concentration<br />

c in pure water <strong>and</strong> in 0.01 m salt solution. For the salt-free solution a maximum<br />

in the reduced viscosity is observed, as already reported for linear polyelectrolytes <strong>and</strong> for<br />

charged spheres [7]. In Fig. 14.8 the reduced scattering intensities are plotted versus q 2 ,<br />

where q is the scattering vector, in a concentration regime similar to the viscosity measurements.<br />

The filled squares represent the measurements in 0.02 m salt solution <strong>and</strong> yield the<br />

true radius of gyration R g = 24.5 nm. This value is never observed in salt-free solution, even<br />

if the microgel concentration is chosen as low as 10 mg/ml, where R g = 21.0 nm is evaluated.<br />

It is to be noted that the size of the microgels is expected to increase with decreasing<br />

salt in contrast to the measured apparent values. Increasing the concentration to larger than<br />

0.2 mg/ml R g,app vanishes below the measurable value of 10 nm. This behaviour can be only<br />

interpreted in terms of the interparticle structure factor. The observed linearity of the scattering<br />

curves is somewhat surprising <strong>and</strong> suggests that the structure factor S (q) is able to precisely<br />

compensate the angular dependence of the intraparticle formfactor P (q), since for<br />

spheres<br />

K c<br />

ˆ 1 1<br />

<br />

R q M w P…q†S …q† :<br />

…3†<br />

In the limit of small q, P (q) = 1 – 1/3 R g 2 q 2 which leads to<br />

S (q) =S (0)/(1 – x 2 q 2 ) (4)<br />

if the scattering curve remains linear in q 2 as observed in Fig. 14.8. For large concentrations,<br />

obviously x 2 = 1/3R g 2 , which yields the angular-independent scattering intensity.<br />

302


14.4 Conclusion <strong>and</strong> relevance to future work<br />

K·c/R Θ<br />

[mol/g]<br />

3.6·10 -7<br />

3.4·10 -7<br />

3.2·10 -7<br />

3.0·10 -7<br />

2.8·10 -7<br />

2.6·10 -7<br />

2.4·10 -7<br />

2.2·10 -7<br />

2.0·10 -7<br />

0 2.0·10 10 4.0·10 10 6.0·10 10<br />

q 2 [cm -2 ]<br />

Figure 14.8: Light scattering results of the microgel described in Fig. 14.7 in 0.02 m potassium bromide<br />

solution (solid squares, c = 0.02g/l). Further in salt-free water at different concentrations for the<br />

microgel in g/l: 0.017 (_), 0.026 (E), 0.1 (l), 0.20 (:), 0.47 (o), 0.55 (M), 1.02 (y).<br />

These experiments support the recent claims [7] that the conformational change of the<br />

polyions is not responsible for the anomalous viscosity behaviour. The observed maximum<br />

in the reduced viscosity is most likely caused by a maximum in the electrostatic interaction<br />

energy. Light scattering experiments at concentrations larger than the concentration of the<br />

highest viscosity are currently performed <strong>and</strong> will provide a more detailed insight into such<br />

complex electrostatic phenomena.<br />

14.4 Conclusion <strong>and</strong> relevance to future work<br />

Despite a big lack of underst<strong>and</strong>ing polymerization in microemulsion leads to model micronetworks<br />

which are easily surface-functionalized to radical macroinitiators or to ionic microgels.<br />

The macroinitiators have been utilized to obtain core-shell or star-like structures.<br />

The ultimate goal of the investigations on ionic microgels is the preparation of a coreshell<br />

structure via electrostatic interaction: a cationic microgel is mixed with a telechelic<br />

oligo or polystyrene bearing one sulfonic acid group at one chain end. Since complex formation<br />

between the oppositely charged species will occur the ionic microgel will be sur-<br />

303


14 Preparation <strong>and</strong> <strong>Properties</strong> of Ionic <strong>and</strong> Surface Modified Micronetworks<br />

rounded by an unpolar polystyrene shell which is dense enough in order to provide perfect<br />

solubility in unpolar solvents such as toluene. Here, the microgel as a molecular container is<br />

locked because water (or any other polar species) is prevented to diffuse out of the microgel<br />

by the polystyrene layer. Again, in preliminary experiments the principle is shown to work<br />

perfectly. Any quantitative work, however, has to be subject of more extensive work in the<br />

near future.<br />

References<br />

1. M. Antonietti, W. Bremser, D. Müschenborn, Ch. Rosenauer, B. Schupp, M. Schmidt: Macromolecules,<br />

24, 6636 (1991)<br />

2. C. Wu: Macromolecules, 27, 298 (1994)<br />

3. C. Wu: Macromolecules, 27, 7099 (1994)<br />

4. V. H. Pérez-Luna, J.E. Puig, V.M. Castano, B.E. Rodriguesz, A.K. Murthy, E.W. Kaler: Langmuir, 6,<br />

1040 (1990)<br />

5. H. Hoffmann, G. Sühler: Vortrag anläßlich des Abschlußsymposiums zum BMBF-Projekt „Mesoskopische<br />

Polymer-Systeme“, Frankfurt (1995)<br />

6. F. C<strong>and</strong>au, J.M. Copart: Colloid Polym. Sci., 271, 1055 (1993)<br />

7. S. Förster, M. Schmidt: Adv. Polym. Sci., 120, 51 (1995)<br />

304


15 Ferrocene-Containing Polymers<br />

Oskar Nuyken,Volker Burkhardt, Thomas Pöhlmann, Max Herberhold,<br />

Fred Jochen Litterst, <strong>and</strong> Christian Hübsch<br />

15.1 Introduction<br />

Classical organic polymers consist of carbon <strong>and</strong> hydrogen. Some of them contain beside C<br />

<strong>and</strong> H additional atoms such as N, O, Si, S, P, or halogens. A relatively small number of publications<br />

deal with metal-containing polymers despite the fact that more than 40 metals are<br />

available [1]. However, metal-containing polymers have emerged as an important class of<br />

polymers in the last twenty years. One reason for this development is that many of them became<br />

available via conventional synthetic routes [2]. Further motivation for developing these<br />

materials is based on the premise that polymers are expected to posses properties significantly<br />

different from those of conventional organic polymers <strong>and</strong> that they show electrical conductivity,<br />

magnetic behaviour, thermal stability, non-linear optical properties etc. [3–5].<br />

This field of activity can be seen as crossing the conventional boundaries of organic,<br />

inorganic, <strong>and</strong> polymer chemistry as well as physics <strong>and</strong> material science.<br />

The aim of our work was the synthesis of tailor-made polymers containing ferrocene<br />

units at different positions of the polymer chain.<br />

From Mößbauer spectroscopy of ferrocene-labelled polymers one could expect valuable<br />

information about the polymer structure, diffusion, <strong>and</strong> fluctuation processes. In earlier<br />

experiments we have been able to show remarkable differences in Mößbauer spectra taken<br />

from solvent-free <strong>and</strong> swollen polymers [6].<br />

Basically there are two strategies for the synthesis of metal-containing polymers. One<br />

approach is to synthesize derivatives of organic polymers. The second pathway involves the<br />

conversion of monomeric ferrocenes into polymers. We have concentrated on the latter approach.<br />

A. Polymers with pendent ferrocene units<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

305


15 Ferrocene-Containing Polymers<br />

B. Polymers with ferrocene units in the main chain<br />

C. Polymers with ferrocenyl end groups<br />

15.2 Addition polymers<br />

Vinylferrocene (VFc) was synthesized in 1955 <strong>and</strong> its polymerization behaviour has been<br />

studied under radical [7, 8], cationic [9], <strong>and</strong> Ziegler-Natta conditions [9]. It was claimed<br />

that VFc is inert to anionic initiation [10].<br />

15.2.1 Radical polymerization<br />

VFc undergoes oxidation with peroxide initiators for which azoinitiators have been used almost<br />

exclusively. Unlike most vinyl monomers the molecular weight of poly(vinylferrocene)<br />

(PVFc) does not increase with the decrease of the initiator concentration. This is the consequence<br />

of an anomalously high chain-transfer constant [11].<br />

The experimental results can be described under the assumption that termination is an<br />

intramolecular reaction,<br />

P <br />

! k t<br />

P ;<br />

<strong>and</strong> that transfer processes play an important part in the radical polymerization of VFc,<br />

P ‡ M k tr;M<br />

!<br />

P ‡ Sol k tr;Sol<br />

!<br />

P ‡ M ;<br />

P ‡ Sol :<br />

306


15.2 Addition polymers<br />

In this case the degree of polymerization (DP n ) is given by the equation<br />

DP n ˆ<br />

k p c…VFc†<br />

k t ‡ k tr;M c…VFc†‡k tr;Sol c…Sol† :<br />

…1†<br />

Since the transfer to the solvents applied in these investigations is known to be small,<br />

one can simplify the last equation <strong>and</strong> write<br />

1 k t<br />

ˆ<br />

DP n k p c…VFc† ‡ C tr;M<br />

…2†<br />

C tr;M ˆ ktr;M<br />

k p<br />

; …3†<br />

where k t is the rate constant of termination, k tr,M the transfer constant to monomer, k tr,Sol the<br />

transfer constant to solvent, <strong>and</strong> k p the rate constant of propagation.<br />

Transfer to polymer was also considered unimportant because monomer conversion<br />

was kept below 10% in all experiments.<br />

From the following graph (Fig. 15.1) one can determine k t /k p = 0.056 mol l –1 <strong>and</strong><br />

the transfer constant C tr,M = 3.5610 –2 , which is about three times higher than found in literature<br />

[11].<br />

1/DP) 10 –2<br />

1/DP) * 10 -2<br />

*<br />

10<br />

8<br />

6<br />

4<br />

0 0.4 0.8 1.2<br />

1/[VFc] /l*mol -1 –1<br />

1 / / l *<br />

Figure 15.1: DP –1 n =f(c (VFc) –1 for the radical polymerisation of VFc in toluene at 70 8C, c (AIBN)<br />

= 1.5610 –2 mol l –1 .<br />

At present one can only speculate about the nature of the intramolecular termination.<br />

However, macroradicals can be considered as strong electrophiles, comparable to o-chloroanil<br />

or tetracyano ethene [12]. It is also known that ferrocene is oxidized by o-chloroanil or<br />

tetracyano ethene [13, 14]. Therefore it seems reasonable to assume termination via an intramolecular<br />

electron transfer step according to the following scheme:<br />

307


15 Ferrocene-Containing Polymers<br />

Scheme 15.1: Termination via intramolecular electron transfer.<br />

This view is supported by the detection of Fe 3+ species via Mößbauer spectroscopy<br />

[15–18]. Moreover, in order to get high conversion one had to add initiator portion by portion.<br />

The following scheme may explain the effect of the initiator on the basis of intermolecular<br />

electron transfer:<br />

Scheme 15.2: Intermolecular electron transfer between an initiator fragment <strong>and</strong> a deactivated chain end.<br />

The glass transition temperature of PVFc was determined by DSC <strong>and</strong> was found in the<br />

area of about 220 8C. From TGA studies we learned that this polymer is stable up to 400 8C.<br />

15.2.2 Radical copolymerization<br />

VFc has been copolymerized with common monomers such as styrene [19], methyl methacrylate<br />

[19], N-vinylpyrollidone [20], <strong>and</strong> acrylo nitrile [19]. The electron richness of VFc<br />

has been demonstrated in its copolymerization with maleic anhydride, in which an alternating<br />

composition of the copolymer was observed over a wide range of feed ratios [16, 19].<br />

The Q <strong>and</strong> e values of VFc were determined. The value of e = –2.1 again emphasizes the<br />

electron rich nature of the vinyl group of VFc [21]. Therefore it looked rather hopeless to<br />

apply anionic initiators for the polymerization of VFc.<br />

15.2.3 Anionic polymerization of VFc<br />

Many attempts were necessary in order to find conditions which allowed anionic polymerization.<br />

<strong>Final</strong>ly, we found n-BuLi, s-BuLi, <strong>and</strong> distyryl dianion in THF most suitable for the<br />

308


15.2 Addition polymers<br />

polymerization of VFc. However, polymerization was not possible in toluene, methylene<br />

chloride, or dioxane. The reason for these differences are due to the complexation of Li + by<br />

THF <strong>and</strong> the formation of highly reactive naked carbanions. This view is strongly supported<br />

by the fact that polymerization in dioxane became possible by addition of 12-crown-4, which<br />

functioned as complexing agent [16].<br />

15.2.3.1 Living polymerization<br />

The living polymerization is characterized by the following properties<br />

a) spontaneous initiation<br />

b) control of molar masses by the ratio c (monomer) : c (initiator)<br />

c) linear molar mass conversion relationship<br />

d) continuous growth of the molar mass when monomer is added portion by portion<br />

e) narrow molar mass distribution.<br />

They are fulfilled in the case of the polymerization of VFc in THF initiated by n-BuLi.<br />

Typical results are shown in Figs. 15.2 <strong>and</strong> 15.3.<br />

The accuracy of the control of molar masses is shown in Tab. 15.1.<br />

Molar masses of the samples were determined by field-desorption mass spectrometry<br />

(FD-MS), vapour pressure osmometry (VPO), <strong>and</strong> GPC. GPC was calibrated with PVFc.<br />

Both, calibration via an oligomer sample in which oligomers from n = 3 through n =11<br />

could be identified <strong>and</strong> via polymers produced by living anionic polymerization (characterized<br />

by VPO or FD-MS), gave similar results <strong>and</strong> excellent agreement with theoretical values.<br />

The FD-MS did not show only the oligomers from n = 3 through n = 11 but satellite<br />

peaks of M + 79 for each molar mass.<br />

Mn/g * mol -1<br />

8000<br />

7000<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

addition of monomer<br />

1000<br />

0<br />

0 5 10 15 20<br />

converted VFc/mmol<br />

Figure 15.2: Monomer addition experiment, M n = f (converted monomer), in THF, T = –25 8C, c (n-<br />

BuLi) = 0.015 mol l –1 (addition of monomer: after complete consumption of the first portion of monomer<br />

a second portion was added).<br />

309


15 Ferrocene-Containing Polymers<br />

IV<br />

III<br />

II<br />

I<br />

I: M n = 2450 g/mol –1<br />

M w =M n = 1.14<br />

II: M n = 4200 g/mol –1<br />

M w =M n = 1.15<br />

III: M n = 5650 g/mol –1<br />

M w =M n = 1.14<br />

IV: M n = 6450 g/mol –1<br />

M w =M n = 1.17<br />

M n is the number average of the molar<br />

mass <strong>and</strong> M w the weight average of the<br />

molar mass.<br />

21 24 27 30 33<br />

Elution volume /ml<br />

Figure 15.3: Polymerization of VFc with n-BuLi in THF at –25 8C, GPC of samples taken at different<br />

reaction times.<br />

Table 15.1: Polymerisation of VFc, initiated with n-BuLi. Monomer addition experiment.<br />

Time/min converted VFc M n (calc.) M n M w M w =M n<br />

/mmol /g mol –1 /g mol –1 /g mol –1<br />

7 0.8253 295 375 430 1.15<br />

30 2.6207 936 1070 1240 1.16<br />

80 5.5384 1979 2070 2360 1.14<br />

127 6.3357 2265 2340 2640 1.13<br />

178 6.5010 2324 2400 2740 1.14<br />

206 6.7242 2403 2400 2730 1.13<br />

248 6.8705 2455 2450 2810 1.14<br />

Addition of Monomer<br />

269 9.2446 3304 3110 3540 1.14<br />

292 12.2361 4373 4180 4800 1.15<br />

364 13.8921 4969 5030 5690 1.13<br />

446 15.1741 5423 5650 6440 1.14<br />

1122 18.4713 6601 6450 7546 1.17<br />

For these findings we suggest the following explanation:<br />

Scheme 15.3: Possible explanation for satellite peaks in the FD-MS with M + 79 (79: THF + Li).<br />

310


15.2 Addition polymers<br />

n=10 n= 9 n=11<br />

n= = 8<br />

n= = 7<br />

n= = 6<br />

n= = 5<br />

26 28 30<br />

Elution volume / ml<br />

Figure 15.4: Identification of oligomeric (VFc) n with n = 3 through n = 11 by GPC; polymerisation<br />

conditions: T =–408C, solvent: THF, M n = 2400 g mol –1 , M w =M n = 1.11.<br />

n= = 4<br />

n= = 3<br />

This again would strongly support the view of complexation of the gegenion by THF.<br />

All attempts to start an anionic polymerization of VFc initiated with naphthyl – K +<br />

failed. However, it was possible to apply distyryl dianion as initiator. The conversion vs.<br />

time is shown in Fig. 15.5.<br />

The molar masses are again controlled by the ratio of monomer :initiator <strong>and</strong> the molar<br />

mass distribution is narrow.<br />

80<br />

conversion/%100<br />

60<br />

40<br />

20<br />

0<br />

0 50 100 150 200 250 300<br />

Time/min<br />

Figure 15.5: Polymerisation of VFc initiated with naphthyl-K/styrene at –30 8C in THF with<br />

c (VFc) = 1 mol l –1 , c (naphthyl-K) = 0.1 mol l –1 , c (styrene) = 0.1 mol l –1 .<br />

311


15 Ferrocene-Containing Polymers<br />

IV<br />

III<br />

II<br />

I<br />

I: M n = 2400 g/mol –1<br />

M w =M n = 1.11<br />

II: M n = 3500 g/mol –1<br />

M w =M n = 1.12<br />

III: M n = 4200 g/mol –1<br />

M w =M n = 1.11<br />

IV: M n = 7500 g/mol –1<br />

M w =M n = 1.14<br />

24 26 28 30 32 34<br />

Elution volume / ml<br />

Figure 15.6: Polymerisation of VFc in THF, initiated by naphthyl-K/styrene at –30 8C, GPC for samples<br />

taken at different reaction times.<br />

15.2.3.2 Block copolymers<br />

One of the main reasons for the interest in living polymerization is its potential for the<br />

synthesis of block copolymers.<br />

ABA-type block copolymers are available if VFc is added after the polymerization of<br />

styrene (PSt) has reached 100%:<br />

Scheme 15.4: Synthesis of ABA-type block copolymers; A = poly(vinylferrocene) segment; B = poly(-<br />

styrene) segment.<br />

312


15.2 Addition polymers<br />

IV<br />

III<br />

II<br />

I<br />

I: M n = 7500 g/mol –1<br />

M w =M n = 1.20<br />

II: M n = 10500 g/mol –1<br />

M w =M n = 1.19<br />

III: M n = 18000 g/mol –1<br />

M w =M n = 1.22<br />

IV: M n = 25000 g/mol –1<br />

M w =M n = 1.21<br />

24 26 28 30 32 34<br />

Elution volume / ml<br />

Figure 15.7: GPC of ABA-type block copolymers; A = c (PVFc), B = c (PSt), solvent THF, T = –30 8C.<br />

Typical results are shown in Fig. 15.7.<br />

AB-type block copolymers can be synthesized by adding a second monomer (B) to<br />

preformed polyvinylferrocene anion (A), which was synthesized as described above. Styrene,<br />

methyl methacrylate, <strong>and</strong> propylenesulfide have been applied as monomer B. Typical results<br />

are given in Tab. 15.2 <strong>and</strong> Fig. 15.8.<br />

Table 15.2: Anionic block copolymerization in THF initiated by n-BuLi; T = –25 8C, c(VFc) = 0.173 mol/l,<br />

c(n-BuLi) = 0.015 mol/l, c (Monomer B) = 2 mol/l.<br />

Monomer A Monomer B Poly. temp./8C a Poly. time/min a M n /g mol –1 b M w /g mol –1 b MWD c<br />

VFc Styrene –70 30 30000 33900 1.13<br />

VFc MMA –70 30 25500 29800 1.17<br />

VFc Propylene- 0 15 11900 13100 1.11<br />

sulfide<br />

a) after addition of the second monomer; b) determined by GPC with internal st<strong>and</strong>ard PSt; c)M w /M n<br />

molecular weight distribution.<br />

It is interesting to note that the block copolymers show only one instead of two glass<br />

transition temperature (Tab. 15.3). This might be due to the fact that the molar mass of segment<br />

A is still too small for being able to establish its own phase.<br />

313


15 Ferrocene-Containing Polymers<br />

IV<br />

III<br />

II<br />

I<br />

I: M n = 2300 g/mol –1<br />

M w =M n = 1.10<br />

II: M n = 10000 g/mol –1<br />

M w =M n = 1.09<br />

III: M n = 20000 g/mol –1<br />

M w =M n = 1.08<br />

IV: M n = 30000 g/mol –1<br />

M w =M n = 1.07<br />

24 26 28 30 32 34 36<br />

Elution volume / ml<br />

Figure 15.8: Anionic block copolymerization of VFc with styrene in THF at –70 8C.<br />

Table 15.3: Glass transition temperature of homopolymer A, homopolymer B, <strong>and</strong> block copolymers AB.<br />

Glass transition temperature a (DSC) of<br />

Monomer A Monomer B Homopolymer A/8C Homopolymer B/8C Block copolymers AB/8C<br />

VFc Styrene 219 100 112<br />

VFc MMA b 219 105 127<br />

VFc Propylene sulfid 219 –40 –29<br />

a) constant heating rate with 10 K/min; b) methylmethacrylat.<br />

15.2.4 Polymeranalogeous reactions<br />

Polymeranalogeous reactions which do not need more than one modification step are an interesting<br />

alternative for the preparation of polymers with pendent ferrocene units. Therefore<br />

it is not surprising that a wide range of conventional organic polymers has been used for this<br />

purpose [22]. These reactions have the advantage of simplicity <strong>and</strong> variability concerning<br />

molar masses <strong>and</strong> composition, which we could demonstrate for the esterification of poly<br />

(MMA-co-MACl) [23] (Tab. 15.4).<br />

314


15.3 Polymers with ferrocene units in the main chain<br />

Scheme 15.5: Esterification of poly(MMA-co-MACI) with hydroxymethylferrocene.<br />

Table 15.4: Typical results of the esterification of poly(MMA-co-MACl) with hydroxymethylferrocene.<br />

MMA : MACl a ferrocene b M n g/mol c M w g/mol c<br />

mol ratio<br />

mol%<br />

90 : 10 6 70000 200000<br />

70 : 30 6.4 45000 61000<br />

60 : 40 10.7 20000 42000<br />

0 : 100 85 40000 66000<br />

a) from IR; b) from 1 H NMR; c) from GPC calibrated with poly(MMA) st<strong>and</strong>ards<br />

15.3 Polymers with ferrocene units in the main chain<br />

15.3.1 Polycondensation<br />

Numerous bifunctional ferrocene derivatives have been synthesized <strong>and</strong> used for polycondensation<br />

reactions [24]:<br />

Scheme 15.6: C<strong>and</strong>idates for polycondensation reactions.<br />

X = COOH, COCl, CH 2 OH, CH 2 CH 2 OH, CH 2 CH 2 NH 2<br />

315


15 Ferrocene-Containing Polymers<br />

Reaction of 1,1'-dichlorocarbonyl ferrocene with diamines such as 1,6-diamino-hexane<br />

<strong>and</strong> 1,8-diamino-octane yielded polyamides which were soluble in DMSO, DMF, <strong>and</strong> m-cresol<br />

[16].<br />

Reaction of 1,1'-dimercapto-ferrocene with succinc acid chloride <strong>and</strong> sebacinic acid<br />

chloride yielded polymers with molar masses of M n = 12000 g mol –1 which were determined<br />

by end group analysis assuming OCH 3 end groups <strong>and</strong> by GPC calibrated with a series<br />

of homologous oligomers identified in the GPC graph [25].<br />

15.3.2 Polymers by addition of dithiols to diolefins<br />

15.3.2.1 Radical reaction<br />

The addition of dithiols to diolefins is a well-established method for the synthesis of polymers<br />

[26–28]. Reaction of 1,1'-dimercapto-ferrocene with norbornadiene in toluene in the<br />

presence of AIBN as initiator <strong>and</strong> heating to 70 8C yielded a yellow powder after precipitation<br />

in methanol. The GPC shows a homologous series of oligomers in which the first<br />

members were identified by FD-MS with m/e = 682, 930, 1179. Instead of the expected alternating<br />

copolymers from ferrocene <strong>and</strong> norbornene, bridged by sulphur, only oligomers<br />

of ferrocenyl disulphides with norbornene or nortricyclyl end groups were detected<br />

(Scheme 15.7).<br />

Scheme 15.7: Reaction of norbornene <strong>and</strong> 1,1'-mercaptoferrocene.<br />

15.3.2.2 Base catalyzed reactions<br />

It has been known for a long time that thiols can react with activated olefins such as unsaturated<br />

ketones in the presence of a basic catalyst [29]. This reaction has also been used for<br />

the synthesis of polymers [30] (Scheme 15.8).<br />

316


15.3 Polymers with ferrocene units in the main chain<br />

22 24 26 28 30 32 34 36 38<br />

Elution volume / ml<br />

Figure 15.9: GPC diagram of oligo(ferrocene disulphide)s detected by UV absorption.<br />

Table 15.5: Molecular weight of the oligomers, calculated M n;theor <strong>and</strong> found M n;exp .<br />

Number of ferrocene units M n;theor g/mol M n;exp g/mol<br />

2 682 682<br />

3 930 930<br />

4 1178 1179<br />

5 1426 1420<br />

6 1674 1690<br />

7 1922 1930<br />

8 2170 2170<br />

9 2418 2420<br />

10 2666 2670<br />

11 2914 2920<br />

12 3162 3140<br />

Scheme 15.8: Base-catalyzed addition of thiols onto activated olefins.<br />

1,1'-dimercapto-ferrocene was used as dithiol <strong>and</strong> 1,4-butanol-dimethacrylate was applied<br />

as diolefin. From both, GPC <strong>and</strong> end group analysis with the assumption of olefinic<br />

end groups, M n = 4000 g mol –1 was determined. The structure of the polymer was confirmed<br />

by elemental analysis, IR, 1 H, <strong>and</strong> 13 C NMR. Similar results were found for divinylsulfone<br />

(M n = 2700 g mol –1 ) <strong>and</strong> dibenzylidene acetone (M n = 6850 g mol –1 ) [25].<br />

317


15 Ferrocene-Containing Polymers<br />

Scheme 15.9: Reaction of 1,1'-bis(mercaptopropylthio)ferrocene with divinyl sulfone.<br />

1,1'-bis(2-mercapto-propylthio)-ferrocene reacts with divinylsulfone according to the<br />

following scheme [31] (Scheme 15.9).<br />

Molar masses were determined by GPC (M n = 2900 g mol –1 ) <strong>and</strong> end group analysis<br />

(M n = 2400 g mol –1 ).<br />

15.3.2.3 Acid catalyzed reactions<br />

Markovnikov addition of thiols to olefins is observed, if acids are applied as catalysts according<br />

to the following mechanism:<br />

Scheme 15.10: Acid-catalyzed reaction of a thiol with an olefin.<br />

Olefins suitable for this reaction are activated by substituents such as OR, SR, NR 2 .<br />

Polymer formation is exemplified by the reaction of 1,1'-dimercapto ferrocene with a divinyl<br />

ether [25] (Scheme 15.11).<br />

Molar masses of M n = 8900 g mol –1 were determined by GPC <strong>and</strong> end group analysis,<br />

assuming OCH 3 end groups. These end groups are a clear indication that H + does not only<br />

function as catalyst for polymer forming but also for polymer degradation.<br />

318


15.3 Polymers with ferrocene units in the main chain<br />

Scheme 15.11: Reaction of 1,1'-dimercapto ferrocene with divinyl ether.<br />

15.3.3 1,1'-dimercapto-ferrocene as initiator<br />

Ring opening polymerization of propylene sulphide by 1,1'-dimercapto ferrocene yields<br />

polymers which contain one ferrocene unit in the center of the macromolecule (Scheme<br />

15.12).<br />

Until now there is no good explanation for the discrepancy between the GPC value<br />

(M n = 28,000 g mol –1 ) <strong>and</strong> the 1 H NMR analysis value (M n = 6000 g mol –1 ), which is very<br />

close to the theoretical value (M n = 6300 g mol –1 ). Maybe the calibration of GPC with polystyrene<br />

st<strong>and</strong>ards is not suitable in this particular case.<br />

Scheme 15.12: Ring opening of propylene sulfide initiated with ferrocene-1,1'-dithiolat.<br />

319


15 Ferrocene-Containing Polymers<br />

15.3.4 Reductive coupling<br />

Reductive polycondensation of 1,1'-diacetyl ferrocene with low-valent titanium, synthesized<br />

by a reaction of TiCl 4 with Zn powder [32, 33], yield polymers with ferrocene-vinylene<br />

repeating units [34]:<br />

Scheme 15.13: Reductive polycondensation of 1,1'-diacetyl ferrocene.<br />

Molar masses up to M n = 6800 g mol –1 were obtained in the low molar mass area of<br />

the GPC, which was calibrated with oligomers given above.<br />

15.4 Mößbauer studies of polymers containing ferrocene<br />

The study of motional dynamics close to the transition from viscous liquid to a glassy state<br />

is a highly important issue for underst<strong>and</strong>ing its physical origin <strong>and</strong> its applicational consequences.<br />

Recently the glass transitions in synthetic polymers <strong>and</strong> other glass forming systems<br />

have been extensively investigated via neutron scattering [35, 36]. We have employed<br />

57 Fe Mößbauer spectroscopy for the study of polymers containing ferrocenyl groups (Fc) [6,<br />

37, 38]. Apart from the vibrational behaviour as derived from the temperature dependent<br />

Mößbauer-Lamb factor f giving a measure of the mean square atomic displacement, these<br />

studies also allow to detect relaxation processes in a time window of 10 –9 –10 –6 s, i. e. close<br />

but complementary to neutron spectroscopy (10 –12 –10 –9 s).<br />

Ferrocene units have been used as side group, as part of the backbone, as terminating<br />

group within various homo <strong>and</strong> copolymers, crosslinked <strong>and</strong> swollen polymers, block polymers,<br />

<strong>and</strong> telecheles. For comparison also monomeric compounds <strong>and</strong> crystallized polymers<br />

were studied. Only a few examples may suffice to be mentioned:<br />

poly(vinyl Fc), poly(dimercapto Fc) poly(Fc dioxymethylene terephthalate), poly(itaconic<br />

anhydrid-co-vinyl Fc), poly(methylmethacrylate-co-vinyl Fc), polystyrene terminated<br />

by vinyl Fc, poly(vinyl Fc-block-propylene sulphide) (PROPS), poly(ferrocenelene-dimethylvinylene),<br />

telechel of norbonadiene, <strong>and</strong> 1,3-dimercapto-benzole with Fc end groups.<br />

The analysis of f below the glass transition T g reveals very similar behaviour for all the<br />

polymers in the temperature range between 50–170 K where nearly harmonic behaviour is<br />

found with a temperature y –2 & 40–50 K, corresponding to the second moment of the<br />

320


15.4 Mößbauer studies of polymers containing ferrocene<br />

vibrational frequency distribution. The motion of Fc is the same if included in the backbone or<br />

as side group thus reflecting the motion of the backbone. Fc at the chain end, however, reveals<br />

a reduction of y –2 to about 30 K. At lower temperatures the amorphous samples show anharmonicities,<br />

which are interpreted to be due to soft quasi-localized modes, which is also typical<br />

for other amorphous materials. The width of the local potentials is up to about 0.08 nm.<br />

Deviations from harmonic behaviour are also found above about 200 K, however, only<br />

for the amorphous samples. These high temperature anharmonicities occur often far below<br />

T g , which is typically around 300–350 K. They are supposed to be caused by residual solvents<br />

in the polymer matrix. We have also studied f (T) for some<br />

p<br />

polymers with a relatively<br />

low T g of 250–300 K. ln f (T) decreases rapidly following a T C T dependence as predicted<br />

by mode coupling theory (MCT). This is interpreted as the onset of local processes.<br />

T c represents the transition from non-ergodic to ergodic behaviour, which occurs typically<br />

30–150 K above the macroscopic glass transition temperature T g . In Fig. 15.10 we show<br />

f (T) for PROPS. The MCT fit is indicated by the broken line yielding T c = 306+12 K<br />

whereas T g &240 K. Simultaneously with the onset of anharmonic behaviour of f the Mößbauer<br />

resonance lines broaden <strong>and</strong> quasi-elastic lines appear close to T c .<br />

Figure 15.10: Temperature dependence of –ln f. Intramolecular contributions of Fc have been subtracted.<br />

We have performed an analysis of the entire Mößbauer line shape taking into account<br />

a density fluctuation correlation function of the Kohlrausch type y(t) =a exp (–|t |/t) b with<br />

b = 1/2 <strong>and</strong> the relaxation time t. We could thus also follow the onset of the structural relaxation<br />

a process related to the dynamic glass transition. The relaxation times decrease<br />

from about 10 –6 satT g to 5610 –8 s at 300 K. Details of these studies especially concerning<br />

the behaviour of swollen polymers will be given elsewhere.<br />

During the course of these studies, which concentrated on the motional dynamics, it<br />

was recognized that many polymers did not only reveal the well-known hyperfine interaction<br />

321


15 Ferrocene-Containing Polymers<br />

of Fc (a quadrupole doublet with a splitting of about 2.35 mm/s) but also have an additional<br />

doublet with a splitting of about 0.8 mm/s which is still a typical isomer shift for a Fe III species<br />

[39]. The relative intensity of the Fe III resonance increases at high temperatures, at<br />

lower temperatures it is usually hardly visible in the spectra (Fig. 15.11). Actually these two<br />

species have already been detected earlier in similar polymers <strong>and</strong> were attributed to Fe II<br />

(Fc) <strong>and</strong> Fe III with different f factors. Belov [40] related the Fe III to a Fc with both rings included<br />

as crosslink in the macromolecule. George <strong>and</strong> Hayes [41] explained it by Fe III<br />

formed in a special termination step during radical polymerization. Our investigations<br />

showed that the Fe III resonance occurs irrespective of the way of polymerization, e. g. also in<br />

polycondensation. In addition, the bonding of both rings to the backbone turned out to be<br />

neither a necessary nor sufficient condition for the formation of Fe III . A careful analysis reveals<br />

that at low temperatures Fe II is dominantly present whereas it is partly transformed to<br />

Figure 15.11: Mößbauer absorption spectra of poly(ferrocenelene-dimethyl-vinylene) at different temperatures.<br />

322


References<br />

Fe III at higher temperatures. This transformation is completely reversible. A quantitative description<br />

of this valence transformation is achieved within a model assuming a thermally activated<br />

excitation of the highest populated molecular orbital of Fc to an unoccupied orbital<br />

of lig<strong>and</strong> character, or a narrow b<strong>and</strong>, which may be supposed from the semiconducting<br />

properties of many polymers containing Fc. The relative population of the Fe III can then be<br />

expressed as<br />

p III ˆ<br />

2<br />

exp…=kT†‡1 :<br />

…4†<br />

For all the studied compounds p we obtain the excitation energy D = 1.50+4 meV.<br />

Notably y –2 is about 2 times that for Fe II . This indicates a duplication of the vibrating<br />

mass. We therefore propose that the Fe III species is related to crosslinks formed by two Fc<br />

units. Whether these are bound via a bridging oxygen or something else is still unclear. We<br />

note that the Fe III cannot be identified with ferrocenium. No relaxational averaging or broadening<br />

is observed. This means that the mixed valent state is fluctuating slowly (610 –6 s) unlike<br />

usual mixed valence systems with much faster fluctuations.<br />

The details of the electronic structure responsible for the mixed valence behaviour <strong>and</strong><br />

the relation to the electron delocalization found in several of these polymers is still unclear<br />

<strong>and</strong> need further investigation. In particular the excitation energy of about 1.5 meV has to<br />

be verified spectroscopically.<br />

Work was supported by Deutsche Forschungsgemeinschaft grant Li 244/8-1,2.<br />

References<br />

1. C.E. Carraher, C.U. Pittman, in: J.E. Sheats, C.E. Carraher, C.U.Pittman (eds.): Metal-containing<br />

Polymeric Systems, Plenum Press, New York, p. 1 (1985)<br />

2. E.W. Neuse: Adv. Macromol. Chem., 1, 1 (1968)<br />

3. K.E. Gonsalves, X. Chen, in: A. Togni, T. Hayashi (eds.): Ferrocenes, Verlag Chemie, Weinheim,<br />

p. 497 (1995)<br />

4. P.D. Hale, T. Inagaki, H.I. Karan, Y. Okamoto, T.A. Skotheim: J. Am. Chem. Soc., 111, 3482 (1989)<br />

5. M.E. Wright, E.G. Toplikar: Macromolecules, 25, 6050 (1992)<br />

6. F.J. Litterst, A. Lerf, O. Nuyken, H.A lcala: Hyperfine <strong>Interactions</strong>, 12, 317 (1982)<br />

7. F.S. Arimoto, A.C. Haven: J. Am. Chem. Soc., 77, 6295 (1955)<br />

8. J.C. Lai, T. Rounsefell, C.U. Pittman: J. Polym. Sci. A-1, 9, 651 (1971)<br />

9. C. Aso, T. Kunitake, T. Nakashima: Makromol. Chem., 124, 232 (1969)<br />

10. C.U. Pittman, C.C. Lin: J. Polym. Sci. Chem. Ed., 17, 271 (1979)<br />

11. Y. Sasaki, L.L. Walker, E.L. Hurst, C.U. Pittman: J. Polym. Sci. Polym. Chem. Ed., 11, 1213<br />

(1973)<br />

12. G. Henrici-Olivé, S. Olivé: Makromol. Chem., 68, 219 (1963)<br />

13. R.I. Collins, R. Pettit: J. Inorg. Nucl. Chem., 29, 5ß3 (1967)<br />

14. C.U. Pittman, P.L. Grube: J. Appl. Polym. Sci., 18, 2269 (1974)<br />

15. H.J. George, G.F. Hayes: J. Polym. Sci. Polym. Lett. Ed., 11, 471 (1973)<br />

323


15 Ferrocene-Containing Polymers<br />

16. V. Burkhardt: PhD thesis, Univerität Bayreuth (1992)<br />

17. R. Feyerherm, F.J. Litterst, V. Burkhardt, O. Nuyken: Solid State Commun., 82, 141 (1992)<br />

18. W. Wagener, M. Hilberg, R. Feyerherm, W. Stieler, F.J. Litterst, T. Pöhlmann, O. Nuyken: J. Phys.<br />

Condens. Matter, 6, L391 (1994)<br />

19. J.C. Lai, T.D. Rounsefell, C.U. Pittman: J. Polym. Sci., A1, 9 (1971)<br />

20. C. Simionescu, T. Lix<strong>and</strong>ru, T. Mazilu: Makromol. Chem., 163, 59 (1973)<br />

21. T. Alfrey, T.T. Price: J. Polym. Sci., 2, 101 (1947)<br />

22. I.R. Butler, W.R. Cullen, N.F. Han, F.G. Herring, N.R. Jagannathan, J. Li: Appl. Organomet.<br />

Chem., 2 (1988)<br />

23. O. Nuyken, V. Burkhardt, T. Pöhlmann, M. Herberhold: Makromol. Chem. Macromol. Symp., 44,<br />

195 (1991)<br />

24. K.E. Gonsalves, R.W. Lenz, M.D. Rausch: Appl. Organomet. Chem., 81 (1987)<br />

25. T. Pöhlmann. PhD thesis, Universität Bayreuth (1993)<br />

26. C. S.Marvel, A. Kotch: J. Am. Chem. Soc., 73, 481 (1951)<br />

27. O. Nuyken, T. Völkel, T. Pöhlmann: Makromol. Chem., 192, 1959 (1991)<br />

28. T. Völkel: PhD thesis, Universität Bayreuth (1990)<br />

29. T. Posner: Ber. dtsch. Chem. Ges., 34, 1395 (1901)<br />

30. O. Nuyken, T. Völkel: Makromol. Chem., 191, 2469 (1990)<br />

31. O. Nuyken, T. Pöhlmann, M. Herberhold: Macromol. <strong>Report</strong>s, A29, 211 (1992)<br />

32. T. Mukaiyama, T. Sato, J. Hanna: Chem. Lett., 1041 (1973)<br />

33. D. Lenoir: Synthesis, 883 (1989)<br />

34. R. Bayer, T. Pöhlmann, O. Nuyken: Makromol. Chem., Rapid Commun., 14, 359 (1993)<br />

35. W. Petry, E. Bartsch, F. Fujara, M. Kiebel, H. Sillescu, B. Farago: Z. Phys. B, 83, 175 (1991)<br />

36. D. Richter, R. Zorn, B. Farago, B. Frick, L. J. Fetters: Phys. Rev., 68, 71 (1992)<br />

37. R. Feyerherm, F.J. Litterst, V. Burkhardt, O. Nuyken: Solid State Commen., 82, 141 (1992)<br />

38. M. Hillberg, W. Stieler, F.J. Litterst, V. Burkhardt, O. Nuyken: Hyp. Int., 88, 137 (1994)<br />

39. W. Wagener, M. Hillberg, R. Feyerherm, W. Stieler, F.J. Litterst, Th. Pöhlmann, O. Nuyken: J. Phys.<br />

Cond. Matt., 6, L 391 (1994)<br />

40. V.F. Belov et al.: Dokl. Akad. Nauk SSSR, 159, 831 (1964)<br />

41. M.H. George, G.F. Hayes: J. Polym. Sci. Lett., 11, 471 (1973)<br />

324


16 Transfer of Vibrational Energy in Dye-Doped Polymers<br />

Johannes Baier, Thomas Dahinten, <strong>and</strong> Alois Seilmeier<br />

16.1 Introduction<br />

Photoexcitation or radiationless relaxation processes supply a lot of thermal excess energy<br />

to polymer systems. In particular, dye molecules dispersed in polymers reach transient excess<br />

temperatures of several 100 K by absorbing visible light [1]. The equilibration of the<br />

excess energy is of interest in the discussion of the photostability <strong>and</strong> of the kinetics of polymer<br />

dye interaction [2] of such systems. In the condensed phases the energy relaxation is determined<br />

by microscopic transfer processes proceeding on a picosecond time scale.<br />

In recent years many investigations have been performed on intra <strong>and</strong> intermolecular<br />

energy transfer in the gas phase <strong>and</strong> in liquids. Whereas in low-pressure gases time constants<br />

extending to the microsecond range have been reported, transfer times from a fraction of a<br />

picosecond to several picoseconds are observed in the liquid state [3]. There the considerably<br />

higher density results in larger collision frequencies <strong>and</strong> shorter relaxation times.<br />

In this paper, the transfer of vibrational energy in polymers doped with dye molecules<br />

is investigated. Similar experiments have been already discussed after exciting C-H stretching<br />

modes around ~n = 3000 cm –1 [4]. Here we present data for the first time where the<br />

more important skeletal modes between ~n = 1500 cm –1 <strong>and</strong> ~n = 1800 cm –1 are pumped.<br />

Two types of experiments are discussed:<br />

a) The dye molecules are vibrationally excited by an ultrashort infrared pulse. Intramolecular<br />

energy redistribution <strong>and</strong> intermolecular energy transfer to the surrounding polymer are<br />

studied.<br />

b) In experiments, where the infrared frequency is tuned to vibrational b<strong>and</strong>s of the polymer,<br />

equilibration of the excess energy in the polymer is investigated. In this case, the dye<br />

molecules serve as sensors for the transient temperature increase of the system.<br />

The experimental data are taken on polycarbonate doped with the dye coumarin 6.<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

325


16 Transfer of Vibrational Energy in Dye-Doped Polymers<br />

16.2 Experimental<br />

Excess energy is supplied to the system by an infrared light pulse either via vibrational excitation<br />

of the dye molecules (Fig. 16.1a) or via vibrational excitation of the polymer matrix<br />

(Fig. 16.1b). In the first case a first rapid energy redistribution occurs within the dye molecule<br />

with a time constant in the order of one picosecond, which is characteristic for the excited<br />

state. The energy is subsequently transferred to the polymer matrix via intermolecular<br />

processes. The instantaneous state of excitation is monitored by a second delayed light pulse<br />

which promotes the excited dye molecules into the fluorescing S 1 state. The frequency of<br />

the probing pulses is chosen in the long wavelength edge of the electronic transition. In this<br />

way only vibrationally excited molecules populated either by photoexcitation or by thermal<br />

excitation contribute to the signal. In the following we discuss the transiently enhanced<br />

fluorescence signal due to photoexcitation of the vibrational manifold. Generally, a biexponential<br />

decay curve is expected representing the two relaxation processes.<br />

In the second type of experiments the polymer is vibrationally excited by the infrared<br />

pulse (Fig. 16.1b). The excitation energy is redistributed giving rise to an increased temperature<br />

of the pumped volume [5]. The energy equilibration in the polymer system involves<br />

intermolecular transfer of energy to the dye molecules which serve as sensors for the equilibration<br />

process. The instantaneous temperature is monitored via electronic transitions <strong>and</strong><br />

the subsequently emitted fluorescence signal of the dispersed dye molecules as discussed<br />

above. Information is obtained on the time for complete energy equilibration but not on the<br />

individual relaxation steps which can be investigated e. g. by an IR double resonance technique.<br />

The experiments are performed with a modelocked Nd :YLF laser system of 10 Hz repetition<br />

rate. Tunable infrared pulses are generated by difference frequency mixing between<br />

Figure 16.1: Energy transfer processes involved in the energy equalization after vibrational excitation<br />

a) of the dispersed dye molecules <strong>and</strong> b) of the polymer matrix by an infrared pulse of frequency n pump .<br />

326


16.3 Results <strong>and</strong> discussion<br />

pulses at the fundamental frequency of the laser <strong>and</strong> the idler pulses of an optical parametric<br />

oscillator driven by the second harmonic of the Nd laser [6]. In this way infrared pulses tunable<br />

between 5 mm <strong>and</strong> 10 mm with energy of about 1 mJ are produced [7]. The probe pulses<br />

are generated by frequency doubling of a weak part of the laser pulses. The fluorescence signal<br />

is measured in a spectral region from 19,800 cm –1 to 20,100 cm –1 with the help of a<br />

spectrometer <strong>and</strong> a photomultiplier. The duration of both pulses amounts to approximately<br />

3 ps. The zero point of the time scale <strong>and</strong> the time resolution of the system is determined by<br />

simultaneously measured cross correlation curves of the two pulses.<br />

The experiments are performed on polycarbonate films with high optical quality that<br />

are prepared as follows. Commercially available 2,2-bis(4-hydroxy phenyl)propane-polycarbonate<br />

granulate (Bayer, m = 1.5 g) <strong>and</strong> coumarin 6 (m = 120 mg) are dissolved in trichloromethane<br />

(V = 10 ml). The solution is carefully spread over a cellulose acetate sheet<br />

with a thickness of 200 mm. After drying a film of about 3 mm results with a dye concentration<br />

of 0.3 M. The thin film is detached from the cellulose acetate substrate <strong>and</strong> used as a<br />

free st<strong>and</strong>ing film.<br />

16.3 Results <strong>and</strong> discussion<br />

Before time-resolved data are presented the infrared properties of our samples should be discussed.<br />

In Fig. 16.2a the infrared absorption of the investigated polycarbonate film is presented.<br />

The two strong lines are assigned to an aromatic C-C ring mode (~n = 1506 cm –1 )<br />

<strong>and</strong> to the C=O stretching mode (~n = 1774 cm –1 ) of the polycarbonate matrix. The weaker<br />

infrared absorption b<strong>and</strong>s can be identified as vibrational b<strong>and</strong>s of the dispersed coumarin<br />

molecules by comparing the spectrum with that of coumarin 6 in a C 2 Cl 4 solution<br />

(c = 2.5610 –3 M) in Fig. 16.2 b. The strongest vibrational lines of coumarin 6 in the spectral<br />

range presented in Fig. 16.2 are due to aromatic C-C ring modes (~n = 1516 cm –1 ,<br />

~n = 1590 cm –1 , <strong>and</strong> ~n = 1616 cm –1 ) <strong>and</strong> due to the C=O stretching mode (~n = 1717 cm –1 ).<br />

The observed absorbance of A *0.2 of the coumarin absorption lines in Fig. 16.2a nicely<br />

compares with estimates on the basis of the absorbance shown in Fig. 16.2b, taking into account<br />

the dye concentration <strong>and</strong> the sample thicknesses in the film <strong>and</strong> the C 2 Cl 4 solution,<br />

respectively.<br />

In time-resolved experiments of the first type (Fig. 16.1 a) vibrational b<strong>and</strong>s of the<br />

coumarin 6 molecules are resonantly excited. Experimental data for two pump frequencies<br />

~n pump = 1590 cm –1 <strong>and</strong> ~n pump = 1717 cm –1 are shown in Figs. 16.3 <strong>and</strong> 16.4, respectively. The<br />

curves are measured with probe pulses polarized parallel (I) <strong>and</strong> perpendicular (y) to the<br />

pump pulses. The signal plotted here is the enhancement of the fluorescence signal (F–F 0 )/F 0<br />

due to photoexcitation relative to the thermally induced fluorescence signal F 0 . The observed<br />

relaxation behaviour is very similar to that found in liquid solutions [8]. Figure 16.3 shows a<br />

first rapidly decaying signal with a time constant e1 ps followed by a longer lived component<br />

with a relaxation time of (7+1) ps. The C–C stretching mode of molecules is directly excited<br />

327


16 Transfer of Vibrational Energy in Dye-Doped Polymers<br />

Figure 16.2: a) Infrared absorbance of a polycarbonate film with 0.3 M coumarin 6. The strong lines<br />

are due to the polycarbonate matrix. b) Infrared absorption of coumarin 6 in a C 2 Cl 4 solution which<br />

clearly shows the absorption b<strong>and</strong>s of the dye molecules.<br />

Figure 16.3: Temporal evolution of the excess fluorescence signal versus delay time. A C–C stretching<br />

mode of coumarin 6 is excited. Results for a probe pulse polarized parallel <strong>and</strong> perpendicular are<br />

shown. The broken curve represents the cross correlation curve of both pulses.<br />

328


16.3 Results <strong>and</strong> discussion<br />

Figure 16.4: Change of the fluorescence signal as a function of the delay time. The C=O stretching<br />

mode of coumarin 6 is pumped. Data are shown for a parallel <strong>and</strong> perpendicularly polarized probe<br />

pulse. The broken line gives the time resolution of the experimental system.<br />

with the probe frequency ~n probe = 19,100 cm –1 , giving rise to a rapidly decaying signal in the<br />

first few picoseconds. Intramolecular processes lead to a new energy distribution within the<br />

molecule which can be described by an increased internal temperature of approximately<br />

350 K. The relaxation of this increased temperature with a time constant of about 7 ps via intermolecular<br />

energy transfer is monitored at longer delay times.<br />

At short delay times the signal amplitude is determined by the population of the C–C<br />

mode, by its Franck-Condon factor, <strong>and</strong> by the lifetime of the vibrational state. At longer delay<br />

times the decrease of the internal temperature of the vibrational manifold is observed.<br />

The probe pulse measures the absorption in the long wavelength edge of the S 1 absorption<br />

b<strong>and</strong> which is in good approximation proportional to the Boltzmann factor exp(hn/kT )in<br />

coumarin 6. Consequently the temperature decrease is measured via the fluorescence signal<br />

which is proportional to exp[h (n probe – n 00 )/kT (t)] <strong>and</strong> which does not depend on the primarily<br />

excited vibrational state [9].<br />

Figure 16.4 shows the situation after direct excitation of the C=O stretching mode.<br />

The observed signal at delay 0 is one order of magnitude smaller. This fact is due to an infrared<br />

absorption reduced by a factor of two <strong>and</strong> due to a considerably smaller Franck-Condon<br />

factor of the C=O mode [8]. The reduced signal amplitude of the fast component does<br />

not allow a clear separation of two relaxation components. However, a better fit of the experimental<br />

data is obtained with two time constants of about 2 ps <strong>and</strong> about 7 ps as compared<br />

to a monoexponential curve. Similar time constants have been identified in liquid solutions<br />

of coumarin 6 [8]. Whereas the signals at delay time 0 are strongly influenced by the<br />

Franck-Condon factors of the excited vibrations, the long-lived component is only dependent<br />

on the absorbed infrared energy which is roughly the same for ~n pump = 1590 cm –1 <strong>and</strong><br />

~n pump = 1717 cm –1 . Consequently the slowly relaxing component is of comparable magnitude<br />

in Figs. 16.3 <strong>and</strong> 16.4.<br />

329


16 Transfer of Vibrational Energy in Dye-Doped Polymers<br />

Of special interest is the polarization dependence of the transient signals. In Fig. 16.3a<br />

signal increased by approximately a factor of two is found for parallel polarization of the<br />

pump <strong>and</strong> the probe pulse as compared to perpendicular polarization. Such a behaviour is<br />

expected for a situation where the infrared transition moment of the C–C mode is oriented<br />

nearly parallel to the electronic dipol moment. In contrast in Fig. 16.4, where the C=O<br />

stretching mode is pumped, the signal is smaller for parallel polarization of the two pulses<br />

indicating an angle between the C=O dipole moment <strong>and</strong> the electronic dipole moment larger<br />

than the magic angle of 54.78 [10]. The polarization dependence will be discussed in<br />

more detail elsewhere.<br />

Next, we want to discuss experiments where the polymer matrix is vibrationally excited<br />

by resonant infrared absorption. Figure 16.5 shows time-resolved data for an infrared<br />

excitation frequency of ~n pump = 1774 cm –1 , where only the C=O stretching vibration of the<br />

carbonyl group of polycarbonate is populated. The fluorescence signal in Fig. 16.5 represents<br />

the momentary excess energy transferred to the probe molecules coumarin 6. We find<br />

a signal rising with a time constant of approximately 40 ps. After 150 ps the signal is nearly<br />

constant <strong>and</strong> a quasiequilibrium with a temperature increased by about 5 K is established in<br />

the excited volume. Obviously, the energy equalization process in the excited volume via intermolecular<br />

energy transfer processes is completed within this time constant. Relaxation of<br />

the increased temperature is expected to occur with a time constant in the millisecond range<br />

via macroscopic heat conduction.<br />

Of special interest is the origin of the time constant of 40 ps observed in Fig. 16.5.<br />

Several processes are involved in the energy equalization process:<br />

a) vibrational relaxation of the primarily excited C=O mode,<br />

b) energy redistribution within the polymer matrix, <strong>and</strong><br />

c) energy transfer from the excited matrix to the probed dye molecules.<br />

Figure 16.5: Enhancement of the fluorescence signal as a function of the delay time. The C=O stretching<br />

mode of polycarbonate is vibrationally excited. The rise of the signal monitors the energy transferred<br />

to the dye molecules. The solid curve is calculated for a transfer time of 40 ps.<br />

330


16.3 Results <strong>and</strong> discussion<br />

The latter step is believed to be much faster because energy transfer between dye molecules<br />

<strong>and</strong> the polymer was found to occur on a time scale shorter than 10 ps (Figs. 16.3<br />

<strong>and</strong> 16.4). Energy redistribution in <strong>and</strong> between the polymer chains is expected to be also relatively<br />

efficient due to the high densities of vibrational states. Consequently, there is indication<br />

that the observed time constant is governed by the depopulation of the primarily excited<br />

C=O stretching mode which obviously does not couple very efficiently to the vibrational<br />

manifold of the polymer. This fact nicely compares with the result on coumarin 6 (Fig. 16.4<br />

<strong>and</strong> Ref. [8]) where also a relatively long lifetime is found for the C=O stretching mode.<br />

The transient fluorescence signal due to vibrational excitation at ~n pump = 1516 cm –1 is<br />

depicted in Fig. 16.6. According to Fig. 16.2 both, the polymer matrix <strong>and</strong> the probe molecules,<br />

are excited at this frequency.<br />

Figure 16.6: Time dependence of the excess fluorescence signal after excitation at ~n = 1516 cm –1 .<br />

Both, the polycarbonate matrix <strong>and</strong> the coumarin molecules, are directly excited. Deconvolution of the<br />

two relaxation processes gives the dashed curves.<br />

Consequently, a relatively complex relaxation behaviour is observed. At short delay<br />

times the directly excited coumarin molecules are detected whereas at longer delay times the<br />

equalization of the excess energy in the matrix is monitored. We observe a first time constant<br />

in the order of a picosecond representing the intramolecular energy redistribution in<br />

coumarin 6 <strong>and</strong> a second time constant of about 7 ps which reflects the intermolecular transfer<br />

to the matrix. In contrast to Fig. 16.3, a residual signal remains at longer delay times.<br />

Here we observe the increased temperature of the excited volume due to the infrared energy<br />

supplied to the polymer matrix. From the fit curves a time constant shorter than about 20 ps<br />

is deduced for the equalization process which is clearly shorter than the time constant observed<br />

after excitation of the C=O mode. The corresponding dashed line in Fig. 16.6 is calculated<br />

for a time constant of 15 ps. Obviously, the energy transfer from the initially excited<br />

C-C stretching mode of polycarbonate to the vibrational manifold is more efficient.<br />

331


16 Transfer of Vibrational Energy in Dye-Doped Polymers<br />

16.4 Summary<br />

In this paper vibrational energy transfer in polymers with dispersed dye molecules is investigated.<br />

Energy is selectively supplied to the system via resonant excitation of skeletal modes<br />

around ~n = 1500 cm –1 . For the energy transfer, between the dye molecules <strong>and</strong> the polymer<br />

matrix, time constants of several picoseconds are measured. The relaxation of excess energy<br />

is governed by the depopulation of the initially pumped vibrational state or by the intermolecular<br />

energy transfer depending on the excited vibrational state. It is interesting to see that<br />

the energy can be stored for a few tens of picoseconds, e. g. in the C=O stretching mode, before<br />

the energy is redistributed. This fact is of relevance for structural relaxation processes<br />

in polymers.<br />

References<br />

1. W. Wild, A. Seilmeier, N. H. Gottfried, W. Kaiser: Chem. Phys. Lett., 119, 259 (1985)<br />

2. C. D. Eisenbach, K. Fischer: Amer. Chem. Soc. Polym. Prepr., Polym. Chem. Div., 29/1, 501<br />

(1988)<br />

3. A. Seilmeier, W. Kaiser: Ultrashort Intramolecular <strong>and</strong> Intermolecular Vibrational Energy Transfer<br />

of Polyatomic Molecules in Liquids, in: W. Kaiser (ed.): Ultrashort Laser Pulses <strong>and</strong> Applications,<br />

Springer, Berlin, p. 279 (1988)<br />

4. A. Seilmeier, J. P. Maier, F. Wondrazek, W. Kaiser: J. Phys. Chem., 90, 104 (1986)<br />

5. P. O. J. Scherer, A. Seilmeier, W. Kaiser: J. Chem. Phys., 83, 3948 (1985)<br />

6. R. Laenen, K. Wolfrum, A. Seilmeier, A. Laubereau: J. Opt. Soc. Am. B, 10, 2151 (1993)<br />

7. T. Dahinten, U. Plödereder, A. Seilmeier, K. L. Vodopyanov, K. R. Allakhverdiev., Z. A: Ibragimov:<br />

IEEE J. Quant. Electron., QE-29, 2245 (1993)<br />

8. H.-J. Hübner, M. Wörner, W. Kaiser, A. Seilmeier: Chem. Phys. Lett., 182, 315 (1991)<br />

9. F. Wondrazek, A. Seilmeier, W. Kaiser: Chem. Phys. Lett., 104, 121 (1984)<br />

10. H. Graener, G. Seifert, A. Laubereau: Chem. Phys., 175, 193 (1993)<br />

332


17 Picosecond Laser Induced Photophysical Processes of<br />

Thiophene Oligomers<br />

Dieter Grebner, Matthias Helbig, <strong>and</strong> Sabine Rentsch<br />

17.1 Introduction<br />

Within the conjugated polymers polythiophene is an important material for future technical<br />

applications, e. g. for LEDs, electronic devices, or optical switches [1]. In this context studies<br />

of light induced processes are of interest, especially details of the degradation of electronic<br />

energy within the ultrashort time scale of picoseconds <strong>and</strong> sub-picoseconds.<br />

These processes mostly proceed within the molecules <strong>and</strong> are presteps to transfer processes<br />

of energy or electrons between different molecules, which dominate in technical devices.<br />

A polymer molecule with about 100 monomer units is a complex system which corresponds<br />

to an inhomogeneous distribution of subsystems with slightly different optical properties,<br />

called effective conjugated segments within a polymer chain.<br />

In this study we had the opportunity to investigate systematically the spectroscopic<br />

properties of a series of linear thiophene oligomers [2–8]. The idea was to observe the fast<br />

light induced processes of these small constituents of a polymer as starting point for deeper<br />

underst<strong>and</strong>ing of processes in the polymer itself.<br />

We studied oligothiophenes from 2 to 6 monomers in dilute solution because in this<br />

case the interaction between the oligomer molecules is negligible. Therefore the dominance<br />

of intramolecular processes was to be expected.<br />

To carry out the intended time resolved studies on oligothiophene a picosecond laser<br />

spectrometer, based on a Nd :YLF oscillator with amplifier system was, build up. The spectrometer<br />

is described in the experimental part of the report. Moreover, the report contains a<br />

brief description of the steady state spectroscopic properties of thiophene oligomers <strong>and</strong> the<br />

results of the picosecond time resolved spectroscopic studies.<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

333


17 Picosecond Laser Induced Photophysical Processes of Thiophene Oligomers<br />

17.2 Experimental<br />

Our picosecond absorption spectrometer operates with optical single pulses generated by a<br />

passively mode locked Nd : YLF laser with a negative feedback loop for stabilisation of<br />

single pulse energy. Figure 17.1 shows the experimental setup of the Nd :YLF laser spectrometer.<br />

The resonator is formed by two flat mirrors, M1 (R = 99%) <strong>and</strong> M2 (R = 52%), <strong>and</strong><br />

by an AR-coated lens L1 ( f = 1250 mm). The total resonator length amounts to 1500 mm.<br />

A single flashlamp-pumped 76.2 mm long Nd : YLF rod with a diameter of 6.35 mm is used<br />

as the active element (AE). The aperture (A) ensures the TEM oo mode structure. The dye<br />

cell (DC) with non-linear absorber is placed in contact with the mirror M1 to achieve passive<br />

mode locking. To get higher pulse stability the non-linear absorber circulates slowly in<br />

a special circuit. We used the absorber dye No.3274y in methanol (relaxation time 6 ps).<br />

The resonator was designed for optimum beam intensity in the non-linear absorber <strong>and</strong> in<br />

the laser rod by variation of the lens position. All elements of the cavity are wedged to avoid<br />

additional resonators.<br />

Figure 17.1: Experimental setup of the Nd:YLF laser spectrometer.<br />

The negative feedback loop consists of a single crystal Pockels cell (PC), a polarizer<br />

(P1), a fast photodiode (PD1), <strong>and</strong> the feedback electronic (FE). A constant voltage (around<br />

l/8 voltage) is applied on the electrodes of the PC. The laser radiation has a circular polarisation<br />

after a double pass through the PC <strong>and</strong> therefore several tens of percent are decoupled<br />

by the polarizer <strong>and</strong> detected by PD1. The feedback electronic drives an additional voltage<br />

on the electrodes of the PC depending on the signal level. So the round trip losses are increased<br />

if the signal level rises <strong>and</strong> therefore the pulse intensity is nearly constant for many<br />

round trips. Figure 17.2 a shows a typical pulse train of the passive feedback controlled<br />

mode locked Nd : YLF laser. The pulse train has a duration of approximately 20 µs with<br />

single pulse energies smaller than 2 µJ.<br />

334


17.2 Experimental<br />

a) b)<br />

c)<br />

Figure 17.2: Pulse trains of a passive feedback controlled mode locked Nd : YLF laser: a) passive feedback<br />

controlled mode locked pulse train without Q factor control (5 µs/div); b) passive feedback controlled<br />

mode locked pulse train with Q factor control after 5 µs (5µs/div); c) nanosecond pulse train<br />

with decoupled single pulse (50 ns/div).<br />

A single energy pulse of several microjoules is too small for application in our picosecond-spectrometer.<br />

To increase the pulse energy we can actively control the Q factor of the<br />

resonator as follows:<br />

The electrical delay line (DL) is triggered by the second photodiode (PD2). The constant<br />

<strong>and</strong> dynamical voltages on the PC are switched off in several tens of nanoseconds after<br />

an adjustable delay time. Thereby the Q factor of the resonator abruptly increases <strong>and</strong> the<br />

pulse saturates the gain in a few round trips (Fig. 17.2b). A high intensity pulse train with<br />

nanosecond duration (Fig. 17.2 c) is the result. In our case the time delay for optimum pulse<br />

shortening is about 5 µs (500 pulse round-trips). We achieved a 80 ns pulse train (FWHM)<br />

with a total energy of 2.5 mJ.<br />

One of the first pulses in the train is extracted by an electro-optical pulse selector<br />

(PS). The single pulse has a duration of 7 ps (measured by autocorrelation) <strong>and</strong> a total energy<br />

of 130 µJ with a pulse to pulse st<strong>and</strong>ard deviation of 2%. Amplification of the selected<br />

pulse to an energy of 20 mJ is achieved in three single pass amplifiers. The laser system<br />

works with a repetition rate of 10 Hz.<br />

335


17 Picosecond Laser Induced Photophysical Processes of Thiophene Oligomers<br />

The second harmonic (SH) (l = 523 nm) <strong>and</strong> the third harmonic (TH) (l = 349 nm)<br />

of the fundamental wave (l = 1047 nm) with type I phase matching for SHG <strong>and</strong> type II<br />

phase matching for THG are generated in two non-linear KDP crystals. We accomplished<br />

conversion coefficients of 29% for SHG (E 2o = 5.8 mJ) <strong>and</strong> 11% for THG (E 2o = 2.2 mJ).<br />

The second or third harmonic radiation, used as the excitation pulse, is filtered by the<br />

filter 1 (F1) to remove other wavelengths <strong>and</strong> to adjust the intensity. Then the beam passes<br />

an optical DL <strong>and</strong> is focused by the lens 4 (L4) into the sample cell (SC).<br />

The unconverted part of the fundamental wave, separated from the harmonics by<br />

mirror 3 (M3) (HT 523,349 nm ,HR 1047 nm ), is focused by lens 2 (L2) into a 10 cm long D 2 O<br />

filled cell. The generated white light pulse (400–970 nm) passes the IR filter 2 (F2) <strong>and</strong> is<br />

recollimated by lens 3 (L3). It is necessary, that the probe beam is polarized under the magic<br />

angle (54.78) to the pump beam to neglect orientation relaxation effects in the probe. Therefore,<br />

the white light beam passes the polarizer 2 (P2) followed by the beam splitting device<br />

(BD). One of the two probe beams passes through the sample cell to monitor the absorbance<br />

change upon excitation. The other beam is used as reference. The pump pulse reaches the<br />

sample cell after passing the optical delay with a maximum time difference to the white<br />

light pulse of 2 ns. Test <strong>and</strong> reference beams are focused onto the entrance slit of a low dispersion<br />

polychromator. Both spectra are detected <strong>and</strong> recorded by an OMA system (OMA-<br />

Vision, EG & G) with a resolution of 4 nm. The used CCD array with 5126512 pixels<br />

(9.769.7 mm 2 ) allows to detect spectra in the range from 200 nm to 1100 nm with a selected<br />

wavelength interval of 200 nm.<br />

To obtain a high accuracy the detector channel matrix is cooled by a Peltier element<br />

to get a signal-to-noise ratio more than 100. Moreover, we average 40 test <strong>and</strong> reference<br />

spectra to determine the difference spectrum at each delay setting. From the transmission<br />

measurements with excitation (we) <strong>and</strong> with no excitation (ne) the difference of the optical<br />

density DD(l,t) is calculated by:<br />

D…l;t†ˆ<br />

2<br />

log6<br />

4<br />

P 40<br />

I t<br />

iˆ1<br />

P 40<br />

I r<br />

iˆ1<br />

3<br />

7<br />

5<br />

2<br />

‡log6<br />

4<br />

P 40<br />

I t<br />

jˆ1<br />

P 40<br />

I r<br />

we jˆ1<br />

3<br />

7<br />

5<br />

ne<br />

; …1†<br />

where I t , is the intensity of the test beam <strong>and</strong> I r of the reference beam. The lowest variation<br />

of the pulse energy from the fundamental pulse <strong>and</strong> average procedure gives an accuracy of<br />

DD(l,t) better than 0.01.<br />

The picosecond pump probe technique was build up during the running time of the<br />

project. Therefore, in parallel to the build up of the spectrometer, measurements were carried<br />

out with a version of a Nd : YLF laser without negative feedback. In this case we<br />

achieved a time resolution of 25 ps.<br />

336


17.4 Results<br />

17.3 Spectroscopic properties of oligothiophenes<br />

Oligomer molecules exhibit a benzoid structure with a single bond between the rings. This<br />

structure allows various conformers in solution at room temperature which show deviations<br />

from planarity but only small energy differences [9].<br />

These properties result in the observed wide structureless absorption spectra. In contrast<br />

to the absorption the fluorescence b<strong>and</strong> is well structured. This is explained by the planar<br />

quinoid structure of the fluorescence state. In quantum chemical calculations [6] the<br />

bonds between thiophene rings exhibit electron densities like double bonds.<br />

Both, absorption <strong>and</strong> fluorescence b<strong>and</strong>s, shift to longer wavelengths with increasing<br />

oligomer size as shown in Fig. 17.3.<br />

a) a)<br />

Figure 17.3: a) Steady-state absorption <strong>and</strong> b) fluorescence spectra of thiophene oligomers nT with<br />

n = 2–6 monomers in dioxane.<br />

It can be seen from the absorption spectra that all oligothiophenes (nT) except 2T can<br />

be excited well with the third harmonic of a Nd : YLF laser (349 nm). 2T was excited with<br />

the second harmonic of a sub-picosecond dye laser system at 308 nm [7].<br />

17.4 Results<br />

17.4.1 Picosecond-transient spectra of oligothiophenes in solution<br />

The transient spectra DD(l, t) ofnT in dioxane are shown in the Figs. 17.4 <strong>and</strong> 17.5. At first<br />

for 3T the transient absorption (A1) at 600 nm arises during the excitation. Moreover, the in-<br />

337


17 Picosecond Laser Induced Photophysical Processes of Thiophene Oligomers<br />

a) b)<br />

Figure 17.4: Transient spectra DD(l,t) of 3T (a) <strong>and</strong> 5T (b) in dioxane after excitation at 349 nm.<br />

Figure 17.5: A2 b<strong>and</strong>s of 3T, 4T, 5T, <strong>and</strong> 6T1 (substituted with alcyl groups) at 2 ns delay to excitation.<br />

duced fluorescence (F) is to be seen in the wavelength region between 430 <strong>and</strong> 480 nm.<br />

After excitation a new absorption (A2) arises at that spectral region. After about 400 ps the<br />

b<strong>and</strong> A2 dominates, whereas the A1 b<strong>and</strong> is vanished.<br />

For oligomers nT with n = 4, 5, 6, also fast arising A1 b<strong>and</strong>s occur at longer wavelengths.<br />

Thereby induced fluorescence was also registered. The induced fluorescence of 5T<br />

is very well seen in Fig. 17.4b.<br />

We compared the b<strong>and</strong> shape of the measured induced fluorescence b<strong>and</strong> with the<br />

well-known stationary fluorescence b<strong>and</strong>, shown in Fig. 17.3b. Thereby we found a good<br />

agreement of the maximum position but deviation in b<strong>and</strong> form. This difference is mainly<br />

caused by the different Einstein coefficients for spontaneous <strong>and</strong> stimulated emission.<br />

The b<strong>and</strong>s A2 also appear at the larger nT, but more delayed as at 3T. They shift more<br />

to longer wavelengths than the corresponding F spectra, so that the superposition of F <strong>and</strong><br />

A2 becomes weaker. On the other h<strong>and</strong>, the A1 b<strong>and</strong>s become wider <strong>and</strong> thereby they overlap<br />

the A2 b<strong>and</strong>s. At long delay the A2 b<strong>and</strong>s solely exist, whereas A1 <strong>and</strong> F decrease.<br />

These b<strong>and</strong>s are identical with triplet absorption b<strong>and</strong>s reported from nanosecond to<br />

microsecond laser flash experiments [10].<br />

338


17.4 Results<br />

17.4.2 Time behaviour of transient spectra<br />

The time behaviour of transient spectra is very essential for the assignment of these spectra<br />

to energy levels or species. If possible, independent experimental methods should be used to<br />

gain data of the time behaviour. Therefore, we performed time resolved fluorescence measurements<br />

on the oligothiophenes <strong>and</strong> obtained the fluorescence lifetimes of nT [3]. Today<br />

other references are also available [11].<br />

The kinetic curves DD(t) at selected wavelengths for 3T (Fig. 17.6) could be fitted<br />

very well with a single exponential function for wavelengths within the A1 absorption b<strong>and</strong>.<br />

Figure 17.7 shows semilogarithmic plots of DD of A1 versus time for 2T–6T.<br />

Figure 17.6: Time behaviour of the absorption A1, A2, <strong>and</strong> the fluorescence F at selected wavelengths<br />

of 3T in dioxane.<br />

Figure 17.7: Semilogarithmic plot of DD of A1 versus time t for nT, n =2–6.<br />

The obtained lifetimes for the A1 b<strong>and</strong> agree with the fluorescence lifetimes [3, 11],<br />

as demonstrated in Tab. 17.1.<br />

From these facts we conclude for all oligothiophenes that A1 is an absorption from<br />

the excited fluorescing state S 1 .<br />

The development of A2 could be observed at all the longer oligothiophenes (n = 3–6).<br />

The A2 spectra are detectable after the decay of A1 <strong>and</strong> F in times of about 100 ps (for 3T)<br />

339


17 Picosecond Laser Induced Photophysical Processes of Thiophene Oligomers<br />

Table 17.1: Decay time s A1 of the transient absorption A1 <strong>and</strong> fluorescence decay time s S for nT in dioxane.<br />

s A1 nT s S<br />

(51 ± 5) ps 2T 40/100 ps [11]<br />

(135 ± 13) ps 3T 150 ps [11]<br />

(531 ± 53) ps 4T (530 ± 53) ps<br />

(880 ± 88) ps 5T (870 ± 87) ps<br />

(1018 ± 102) ps 6T1 (1100 ± 110) ps<br />

<strong>and</strong> 500 ps (for 4T–6T) <strong>and</strong> rise to their terminal value during the used delay time. A2 is a<br />

long living absorption. The A2 spectra of 3T to 6T at a delay to excitation of 2 ns are shown<br />

in Fig. 17.5.<br />

Figure 17.8 shows the scheme of processes which proceed after occupation of S 1 by a<br />

one photon absorption process (EPA). The excited state absorption (ESA) indicate a transition<br />

from S 1 to S N <strong>and</strong> the triplet absorption (TTA) from T 1 to T R , respectively. We assume<br />

that the internal conversion (IC) within the triplet ladder is fast. The rate of S 1 depletion is<br />

k = k R + k ISC (internal conversion is neglectible at nT).<br />

Figure 17.8: Scheme of S 1 depletion <strong>and</strong> triplet formation after excitation.<br />

The formation of triplets is demonstrated for 3T. According to Fig. 17.8 we consider<br />

the decay processes after d-like excitation with simple rate equations (Eqs. 2 <strong>and</strong> 3) for the<br />

number density of the excited singlet S1 <strong>and</strong> triplet T1. The solutions (Eqs. 4 <strong>and</strong> 5) yield<br />

the time behaviour of the excited states.<br />

dc… S1†<br />

ˆ<br />

dt<br />

k S c… S1† …2†<br />

dc… T1†<br />

ˆ‡k ISC c… S1† …3†<br />

dt<br />

c…<br />

S1† ˆ c… S1† tˆ0 exp… k S t† …4†<br />

c…<br />

T1<br />

k T<br />

† ˆ c…<br />

S1† tˆ0 …1 exp… k S t†† ˆ c… S1† k tˆ0 F T …1 exp… k S t†† …5†<br />

S<br />

340


17.4 Results<br />

2,303 D<br />

ˆ<br />

z<br />

2,303 D<br />

z<br />

ˆ<br />

s F c…<br />

S1†‡s TTA c… T1† …6†<br />

s F c… S1† tˆ0 exp… k S t†‡s TTA c… S1† tˆ0 F T …1 exp… k S t†† …7†<br />

Both, number densities <strong>and</strong> their emission s F or absorption cross section s TTA , influence<br />

the DD spectra, as can be seen from Eqs. 6 <strong>and</strong> 7. If triplets originate directly from the<br />

S 1 state, only the S 1 depletion rate determines the DD(t) curve of superposed fluorescence<br />

<strong>and</strong> triplet absorption. This was proved in a fitting procedure as shown in Fig. 17.9.<br />

Figure 17.9: Experimental results (x) <strong>and</strong> fitting curve (- -) which describes fluorescence decay <strong>and</strong> triplet<br />

formation with one common rate constant (k = 1/135 ps).<br />

17.4.3 Size dependence of spectroscopic properties of oligothiophenes<br />

Size dependent steady state spectroscopic properties of thiophene oligomers have been reported<br />

by a series of authors. The goal of our studies was the investigation of the behaviour<br />

of excited levels or species in dependence of the oligomer size.<br />

In the literature mostly you can find a graphic representation of spectral energies in<br />

dependence of the inverse of the monomer or double bond number. This graphic yields a<br />

fast convergence of values (with 1/n) to polymer chains of infinite length. More subtle is<br />

the question at which finite number of monomers the optical properties do not show further<br />

alteration. This gives hints for practical preparation of optical devices <strong>and</strong> it allows to prove<br />

theoretical models. At our studies of the series of small oligomers, n = 2 … 6 we utilized<br />

various models for the representation of size dependence, e. g. the Lewis-Calvin formula for<br />

p<br />

the absorption wavelengths l = n , the energy gap dependence DE & 1/n, <strong>and</strong> the extended<br />

FEMO model of Kuhn [13] which yields for the energy gap,<br />

DE ˆ<br />

<br />

h2<br />

8mL 2 …2DB ‡ 1†‡V 0 1<br />

<br />

1<br />

; …8†<br />

2DB<br />

341


17 Picosecond Laser Induced Photophysical Processes of Thiophene Oligomers<br />

where h is Planck’s constant, m the electron mass, L the box length, DB the double bond<br />

number, <strong>and</strong> V 0 a constant value.<br />

This formula was estimated for a linear polyene chain. We tried to apply it for the oligothiophenes<br />

<strong>and</strong> found that this formula excellently fits our experimental results.<br />

Figure 17.10 shows our fitting results for the different models.<br />

Figure 17.10: Description of the HOMO-LUMO transition (fluorescence FL) by different models.<br />

Not only the stationary b<strong>and</strong> maxima of absorption <strong>and</strong> fluorescence match this formula<br />

but also the radical absorption <strong>and</strong> even the triplet <strong>and</strong> singlet excited state absorption.<br />

Thereby the fitting parameters have been altered.<br />

The extension to larger DB values of the curve DE (DB) allows a comparison with the<br />

well-known absorption <strong>and</strong> fluorescence b<strong>and</strong>s of polythiophene. These values match the<br />

curve at a number of 50 double bonds which corresponds to 25 thiophene rings. This means<br />

larger oligomers show no relevance to optical spectra.<br />

17.4.4 Size dependence of the kinetic behaviour of oligothiophenes<br />

We found excited state absorption <strong>and</strong> fluorescence b<strong>and</strong>s, both originating from the excited<br />

singlet level S 1 for all nT. During its decay a triplet absorption b<strong>and</strong> was observed. Both processes<br />

slow down with increasing oligomer size.<br />

In other words, the fluorescence quantum yield increases because the triplet quantum<br />

yield decreases. Other degradation processes except emission <strong>and</strong> intersystem crossing were<br />

not detectable in our experiments.<br />

342


References<br />

17.5 Discussion<br />

The laser pulse initiated processes as observed from sub-picosecond to nanaosecond time<br />

range on a series of oligothiophenes in solution started with the increase of the A1 b<strong>and</strong> during<br />

the absorption of the pulse energy. The A1 b<strong>and</strong>s are S n /S 1 absorptions, identified by<br />

their decay times which are identical to the fluorescence decay times measured by time resolved<br />

fluorescence measurements (Tab. 17.1). The observed induced fluorescence F roughly<br />

decays with the same rate constant as A1 if it is not superposed by absorption b<strong>and</strong>s of different<br />

time function.<br />

The nature of the delayed appearance of the b<strong>and</strong> A2 was enlightened. Thereby it was<br />

helpful to find recent works [10–12] on nanosecond laser flash experiments in which triplet<br />

spectra with microsecond to millisecond lifetimes have been identified. A comparison of<br />

these spectra with our A2 spectra showed that both types of spectra are identical in spectral<br />

shape <strong>and</strong> position.<br />

This means, in our picosecond experiments we have observed for the first time the<br />

formation process of triplets in the nT series.<br />

It seems to proceed an effective singlet-triplet intersystem crossing process in nT. Reasons<br />

for such a fast intersystem crossing may be the influence of the sulfur atom on the spinorbit<br />

coupling elements on one side. Moreover, we assume that there is a higher lying triplet<br />

level neighbouring the excited singlet state S 1 . This is likely accompanied by thermal initiated<br />

transitions from higher occupied vibronic levels of S 1 to the corresponding triplet states.<br />

References<br />

1. International Conference on Science <strong>and</strong> Technology of Synthetic Metals, Göteborg, 1992, Synthetic<br />

Metals, 55–57, (1993)<br />

2. H. Chosrovian, D. Grebner, S. Rentsch, H. Naarmann: Synthetic Metals, 52, 213 (1992)<br />

3. H. Chosrovian, S. Rentsch, D. Grebner, D. U. Dahm, E. Birckner: Synthetic Metals, 60, 23 (1993)<br />

4. D. V. Lap, D. Grebner, S. Rentsch, H. Naarmann: Chemical Physics Letters, 211, 135 (1993)<br />

5. D. Grebner, D. V. Lap, S. Rentsch, H. Naarmann: Chemical Physics Letters, 228, 651 (1994)<br />

6. R. Colditz, D. Grebner, M. Helbig, S. Rentsch: Chemical Physics, submitted<br />

7. D. Grebner, D. V. Lap, S. Rentsch: International Conference on Molecular Spectroscopy, Essen,<br />

1994, Journal of Molecular Structure, 348, 433 (1995)<br />

8. D. Grebner, M. Helbig, S. Rentsch: Journal of Physical Chemistry, 99, 16 991 (1995)<br />

9. G. Zerbi, L. Chierichetti, G. Ingänas: Journal of Chemical Physics, 94, 4637 (1991)<br />

10. V. Wintgens, P. Valat, F. Garnier: Journal of Physical Chemistry, 98, 228 (1994)<br />

11. P. Emele, D. U. Meyer, N. Holl, H. Port, H. C. Wolf, F. Würthner, P. Bäuerle, F. Effenberger: Chemical<br />

Physics, 181, 417 (1994)<br />

12. R. Rossi, M. Ciofalo, A. Carpita, G. Ponterini: Journal of Photochemistry <strong>and</strong> Photobiology, A70,<br />

59 (1993)<br />

13. H. Kuhn: Journal of Chemical Physics, 17, 1198 (1949)<br />

343


18 Topospecific Chemistry at Surfaces<br />

Hans Ludwig Krauss<br />

18.1 Introduction<br />

18.1.1 The problem<br />

Surfaces of solids exhibit different sites – a triviality. Even the simplest model, the ideal<br />

NaCl crystal, shows faces, edges, <strong>and</strong> corners. Different coordinative situations result in a<br />

different behaviour in all reactions. Of course, systems that are more complex contain a<br />

greater variety of different surface species. <strong>Final</strong>ly, we should expect that in amorphous solids<br />

the topological situation of the surface is almost chaotic – apart from general restrictions<br />

given by the stoichiometry, bond lengths, <strong>and</strong> bond angles of the specific compound.<br />

Very frequently the surface is chemically different from the bulk’s composition: highly<br />

unsaturated situations are more or less moderated by accommodating molecules (as a whole<br />

or in parts) from the surrounding media like H-OH from moist atmosphere. In case of substances<br />

with high specific surface area, these additional surface groups may form a substantial<br />

part of the whole solid. Silica, made from aqueous solutions of silicates for instance, is<br />

completely saturated at its surface by SiOH groups that react in the usual way, as silanol<br />

groups, with respect to their immobility <strong>and</strong> to their specific topology in the surface.<br />

Amorphous solids with large specific surface area play an important role in many chemical<br />

processes, especially as supports for heterogeneous catalysts. Therefore, the general<br />

problem is the behaviour of surface compounds as a function of their individual topological situation<br />

that provides information of great practical value, e. g. for the modelling of catalysts.<br />

18.1.2 Preparative <strong>and</strong> analytical methods<br />

The use of selected surfaces as educts for chemical reactions is very common, e. g. in semiconductor<br />

modelling. As another example the intercalation may be mentioned as a typical<br />

344 Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


18.1 Introduction<br />

reaction of layer compounds. The preparation of catalysts however usually does not include<br />

crystalline substrates. Here it is the main objective to provide a large specific surface, as offered<br />

e. g. by an amorphous substrate; that includes the inevitable presence of a variety of<br />

many topologically different surface sites or surface groups. When these groups react as anchoring<br />

sites for the specific catalytic agent, e. g. a metal compound, the topology of the<br />

support at a given site controls the physical <strong>and</strong> chemical behaviour of this individual surface<br />

compound. Vice versa the metal may act as a probe to allow a characterisation of the<br />

site in question.<br />

We must keep in mind that real surfaces are part of a usually disordered solid <strong>and</strong> in<br />

addition act as an inseparable ensemble of sites. Therefore, the analytical approach is restricted<br />

to methods that do not require free mobility of the target group or a long distance<br />

order of the solid. So the most powerful tools of the modern analytical chemistry, NMR (except<br />

MASNMR), <strong>and</strong> X-ray structure analysis, are ruled out. Others may be used with considerable<br />

limitations only.<br />

18.1.3 Industrial applications<br />

Surface chemistry can follow two different objectives: the production of a constructional<br />

matter with special properties on its border to the environment, e. g. a coated polymer, or<br />

the preparation of a material with special functions to the environment, e. g. a catalyst. While<br />

the former aspect has attracted the attention of chemists increasingly in the last years, the research<br />

in catalysis at surfaces is an old field. The advantages of heterogeneous catalysts<br />

compared with their homogeneous analogues are well-known <strong>and</strong> have caused an extensive<br />

search for such systems. On the other h<strong>and</strong>, in many cases the mechanisms of the reactions<br />

at the active surfaces are not well understood – even if the process is used on a large-scale<br />

production worldwide. The Haber-Bosch process e. g. finally could be explained after almost<br />

a century of research. The mechanism of the Ziegler-Natta process needs still further studies<br />

<strong>and</strong> the underst<strong>and</strong>ing of the Phillips process [1, 2] is insufficient yet.<br />

18.1.4 The st<strong>and</strong>ard procedures of the Phillips process<br />

The enormous variety of parameters in surface chemistry requires a restriction to typical<br />

systems. On the background of our earlier work on chromium <strong>and</strong> its redox chemistry [3],<br />

the Phillips catalyst for olefin polymerization seemed to be a good example for the general<br />

problem of reactions with topologically defined educts. Here we find an inert oxidic support<br />

with a large specific surface (usually silica), the chemical fixation of a metal compound<br />

(usually chromium) at this surface, a thermal or chemical treatment to convert some of the<br />

Cr surface centres to catalytically active species, <strong>and</strong> finally several uncleared catalytical<br />

processes at these species. In short, we find all the elements of a challenging chemical problem<br />

under the aspect of a topological regulation. Of course we varied the st<strong>and</strong>ard prepara-<br />

345


18 Topospecific Chemistry at Surfaces<br />

tion of the catalyst, i. e. impregnation of silica with CrO 3 in aqueous solution <strong>and</strong> thermal<br />

activation in air, as given in the first patents [4] almost 40 years ago – not to achieve better<br />

yields of the catalytic reactions but to underst<strong>and</strong> the influence of the topologic parameters.<br />

18.1.5 Earlier work<br />

Since its discovery in 1956, the Phillips process [5] was an object of research in many parts<br />

of the world. Our own activity started in 1965 with redox studies [6]. These lead us to the<br />

discovery of coordinatively unsaturated chromium(II) as a substantial constituent of the surface<br />

chromium in active catalysts [7, 8].<br />

The catalytic activity was of course one of the important properties to be determined<br />

as a function of the preparation conditions (<strong>and</strong> hence the topological situation) of the metal<br />

surface compounds. The variation of the catalyst as well as the variation of the reaction to<br />

be catalysed was within the scope of the continuation of earlier work. Our interest was <strong>and</strong><br />

still is focussed here on the mechanism of the catalytic reactions. With respect to the restricted<br />

analytical situation, the spectrum of the products was an important but often ambiguous<br />

source of information.<br />

In the next paragraphs our results of the last decade will be reported, arranged by following<br />

the steps of the catalyst’s preparation: starting from a selected support, preparing a<br />

variety of surface species, exploring the different reactions of these species. However, the<br />

chronological progress of the work did not follow this one-dimensional pattern. Actually, the<br />

variation of single parameters causes usually changes in a multidimensional network of responses.<br />

For drawing a coherent picture, old <strong>and</strong> new results should be put together – like<br />

pieces of a jigsaw puzzle just one as we hope.<br />

18.2 The support<br />

18.2.1 Unmodified silica<br />

Silica with its constituents Si, O, <strong>and</strong> H may be regarded as one of the simplest supports.<br />

The huge number of varieties however shows that the architecture of the material is manifold<br />

<strong>and</strong> sensitively dependent on the preparation. Attempts to construct silica or just an amorphous<br />

SiO 2 by a computer program gave unsatisfactory results [9]. Since the basic work of<br />

Peri <strong>and</strong> Hensley [10], a simplified model of cristobalite planes is used to explain the properties<br />

of the silica surface, neglecting the aspect that the fractal dimension of the latter may<br />

reach almost the value three. The thermal behaviour of geminal, vicinal, or isolated OH<br />

346


18.2 The support<br />

groups can be interpreted with this model. Experimentally it was shown earlier that the release<br />

of water from vicinal pairs around 680 8C (DTG experiment) is the most important dehydration<br />

process [11]. New work using MASNMR for 29 Si confirmed the existence of geminal<br />

pairs <strong>and</strong> predominantly single (vicinal, isolated?) OH groups in the untreated gels [12].<br />

After thermal treatment, unfortunately, no well-resolved spectra were obtained.<br />

By electron microscopy we found a general three-stage organisation of the gels: raspberries<br />

with 100–200 nm diameters were composed of small balls with ca 40 nm diameters,<br />

these again consisting of smaller irregular units with diameters of about 10 nm [13]. The<br />

compact particles of some of the gels exhibit surfaces looking like a cobble stone pavement<br />

of narrow-packed 40 nm spheres.<br />

Usually we worked with the supports Merck 7733, Grace 952, Grace XWP types, <strong>and</strong><br />

Degussa Aerosil 100 or 200. Different gels had to be used for different experiments. For instance<br />

rigid gels, like Merck 7733, give very good UV-VIS-NIR spectra in reflectance. However,<br />

they are not suitable for ethylene polymerization catalysts, which should easily undergo<br />

a structural degradation (popping) during the formation of the solid polymer. On the other<br />

h<strong>and</strong> they work well when the polymer is soluble: The polytest with 1-octene, see below,<br />

showed an almost linear dependency of the catalyst’s activity on the specific surface but no<br />

correlation with pore volume or pore diameter [14].<br />

18.2.2 Modified silica<br />

Although industrial applications recommend the use of silica supports with built-in ions of<br />

Ti(IV), Zr(IV), Al(III), Mg(II), <strong>and</strong> fluoride to increase the activity of polymerization catalysts,<br />

our polytests with 1-octene showed no influence in the cases of Ti, Zr, <strong>and</strong> Mg. Probably<br />

their effect is restricted to the polyethylene case, here disturbing <strong>and</strong> destabilising the<br />

structure of the support (easier popping!). The use of Al doped silicas, e. g. Grace 13–110,<br />

increased the activity of the catalyst under otherwise equal conditions by a factor of 2.5 to 3<br />

[14]. This effect may be explained by an electrostatic influence of the Lewis acid Al(III) enhancing<br />

the electronegativity of the gel. By 27 Al MASNMR experiments we could show that<br />

in these gels Al(III) occurs primarily in a tetrahedral configuration. Thus, Al-OH groups are<br />

not detected [12]. Hence, a later fixation of the transition metal as Al-O-M surface compound<br />

may be neglected. The effect of fluoride doping again might be due to an electrostatic<br />

effect. Unfortunately MASNMR measurements were not possible because of interference<br />

with Teflon parts.<br />

Because these modified systems did not promise new evidence in the basic question<br />

of topology but even increased the number of parameters, further work was restricted to the<br />

use of pure silicas as support.<br />

347


18 Topospecific Chemistry at Surfaces<br />

18.2.3 Others<br />

The use of other supports like gels of Al 2 O 3 , AlPO 4 (isoelectronic to SiO 2 ), <strong>and</strong> activated<br />

carbon was shown to be possible. In the latter case, the reduction of the added Cr(VI) compound<br />

to the active species is carried out by the support itself at higher temperatures [15].<br />

At present no significant advantage over the conventional systems was found.<br />

18.3 Transition metal surface compounds<br />

18.3.1 The metals<br />

The formation of surface compounds from genuine surface groups, e. g. Si-OH, <strong>and</strong> a metal<br />

compound usually is a condensation releasing water, HCl, NH 3 , <strong>and</strong> similar products. After<br />

impregnation <strong>and</strong> drying, this preparation step is performed mostly at higher temperatures<br />

(activation). The thermal treatment does not only allow a fixation of the metal compound at<br />

topologically <strong>and</strong> energetically different sites but simultaneously includes the condensation<br />

of Si-OH groups among each other, as stated above. Table 18.1 lists the metal applications<br />

used in this work.<br />

Table 18.1: Metal applications used in this work.<br />

Metal Compd. Medium T act 8C Atm. act Ox.Nr. Ref.<br />

Ti TiCl 3 water 500 oxygen &4 16, 17<br />

V (NH 4 ) 3 VO 3 water 800 oxygen &5 16, 18–21,<br />

VOCl 3 gas phase 800 oxygen &5 20<br />

Cr CrO 3 water ^800 oxygen &5.9 13, 22<br />

Cr 2 ac 4 72H 2 O NH 3 /H 2 O 800 oxygen &5.7 23<br />

Fe FeCl 2 , FeCl 3 water 800 vacuum &2.2 24<br />

Cu Cuac 2 7H 2 O water 800 oxygen &2 25<br />

Mo Mo-diglycolate water 800 oxygen &6 26<br />

(NH 4 ) 6 Mo 7 O 24 water 800 oxygen &6 27, 28<br />

Mo 2 ac 4 NH 3 /H 2 O 800 oxygen &6 23<br />

Mo/Cr CrMoac 4 NH 3 /H 2 O ^800 oxygen &6 23<br />

Ru “RuCl 3 7H 2 O” acetone ^600 oxygen ? 29<br />

(For Mn, Co <strong>and</strong> Ni see Ref. [30, 31])<br />

In the following, the systems with the metals Cr, V, <strong>and</strong> Mo will be discussed in more<br />

detail.<br />

348


18.3 Transition metal surface compounds<br />

18.3.2 Impregnation <strong>and</strong> activation<br />

As reported in Table 18.1, the impregnation was mostly carried out from aqueous suspensions<br />

where the metal compound was dissolved in a concentration suitable for the experiments.<br />

Usually for chemical work, the applied concentrations are smaller than 1 wt% of<br />

metal in the final product. Since the support offers just a limited number of sites for the<br />

accommodation of the metal a further increase of the metal compound concentration is not<br />

incorporated into the surface reaction. These free metal compounds may lead to a decomposition<br />

if heated during activation. In the case of CrO 3 an oxidation state of 6 in the activated<br />

product is reached only if c(Cr) ?0. In the case of 1% Cr <strong>and</strong> activation in oxygen<br />

the product has an oxidation number near 5.9. Free CrO 3 is decomposed to Cr 2 O 3 , as<br />

shown by REM, EDX [6, 58, 59]. (Only the surface bond Cr(VI) is a c<strong>and</strong>idate for the later<br />

reduction to the coordinatively unsaturated sites (c.u.s.) Cr(II), see below.) The average<br />

oxidation number approaches finally the value 3 for very high loads; there are no intermediate<br />

steps.<br />

Another information on the oxidised form of the surface compounds can be derived<br />

from the photoluminescence, which is observed in the cases of vanadium, chromium, <strong>and</strong><br />

molybdenum. Considering the best investigated case, surface chromium(VI), we find a drastic<br />

increase of luminescence intensity at an activation temperature around 600 8C (Fig. 18.1).<br />

The emission shows now a well-resolved fine structure while an unstructured background<br />

decreases in intensity [50].<br />

Figure 18.1: The influence of T act on different activities of surface Cr(VI) <strong>and</strong> the later c.u.s. system:<br />

T <strong>and</strong> solid curve: photoluminescence [50], & chemoluminescence [50], y polymerization activity (formation<br />

of poly-1-octene, 1/t [14]), _ Cr s /Cr(II) under st<strong>and</strong>ard conditions [72].<br />

349


18 Topospecific Chemistry at Surfaces<br />

Thermal treatment of products that were impregnated with metals in lower valences,<br />

e. g. Fe(II), results as well in a fixation of the metal – releasing now the protonated anion<br />

<strong>and</strong> water, see below. The impregnation with Cr(II)/Mo(II) [23] requires a pretreatment of<br />

the support with aqueous ammonia to increase the nucleophilic character of the SiO<br />

groups. The question of fixation of aggregated metal units, e. g. dichromate(VI), has not<br />

been generally settled yet. For the vanadium case there is evidence from UV, VIS, NIR,<br />

<strong>and</strong> Ra spectra that polymeric surface vanadates(V) are formed already at very low concentrations.<br />

Thus chain <strong>and</strong> b<strong>and</strong> structured vanadates can be found [21, 32]. This tendency<br />

continues up to the formation of separate V 2 O 5 phases at higher loading. It may not<br />

be excluded that the reduced species exhibit a certain mobility in the surface <strong>and</strong> perform<br />

a redistribution at higher temperatures to reach the thermodynamically most favourable<br />

situation.<br />

18.4 Coordinatively unsaturated sites<br />

18.4.1 Reduction of saturated surface compounds<br />

By the pretreatment with oxygen at high temperatures the surface compounds contain the<br />

metal in a high oxidation state <strong>and</strong> are coordinatively saturated. By reduction the situation is<br />

changed. Oxidation <strong>and</strong> coordination numbers drop drastically, the metal forms a c.u.s. For<br />

instance the surface chromate(VI) – anchored via two oxygens to the support <strong>and</strong> saturated<br />

by two other oxygens – is reduced to a surface chromium(II) with only two links to the support<br />

<strong>and</strong> four vacant coordination sites:<br />

surface(Si-O) 2 CrO 2 +2CO ! surface(SiO) 2 Cr W 4 +2CO 2<br />

This idealised formulation neglects the fact that even in the rather immobile situation<br />

of the metal in the surface there will be at least the chance of a partial saturation by interaction<br />

with neighbouring groups in this surface, e. g. SiOH or eventually SiOSi [33, 34]. The<br />

extent of this saturation will depend on the existence <strong>and</strong> the topological accessibility of<br />

such groups. The important parameter is therefore, besides the concentration of the metal in<br />

the surface, the relative concentration of SiOH groups next to the metal. Decisive are the activation<br />

<strong>and</strong> reduction temperatures which influence the amount of SiOH <strong>and</strong> the mobility of<br />

the Cr(II) ions respectively. At a given Cr concentration high activation temperatures of<br />

800 8C <strong>and</strong> low reduction temperatures of 350 8C (800 8C/350 8C) favour the formation of<br />

highly unsaturated metal sites. Low activation temperatures <strong>and</strong> high reduction temperatures<br />

(e. g. 500 8C/500 8C) on the other h<strong>and</strong> foster a coordination of surface groups to the metal,<br />

yielding a certain saturation. Catalysts prepared with these two different thermal combinations<br />

will be used in the following as st<strong>and</strong>ards.<br />

350


18.4 Coordinatively unsaturated sites<br />

Considering the amorphous structure of the support, a large multitude of different individual<br />

positions of the metal on the surface was to be expected. By spectroscopic studies [35–<br />

37] we found surprisingly that surface chromium(II) occurs in only two discrete topological<br />

situations, which we have called A <strong>and</strong> B. Our st<strong>and</strong>ard catalysts (800 8C/350 8C <strong>and</strong> 500 8C/<br />

500 8C) are dominated by the species A or B respectively. But there is no way to prepare samples<br />

that contain only one of these species. The determination of the population profile of A<br />

<strong>and</strong> B species in different Cr catalysts by TPD with innocent lig<strong>and</strong>s confirmed this result<br />

[14, 38]. Since there is a certain (energetic) b<strong>and</strong>width of both species they better should be<br />

seen as ensembles. Analogue observations were made with surface compounds of Cu, Fe, Mo<br />

[16]. Vanadium is a special case because of its agglomeration tendencies [32].<br />

The generally accepted interpretation at present is the assignment of the coordination<br />

numbers 2 for the Cr(II)A <strong>and</strong> 3 for the Cr(II)B, where the third site coordinates oxygen of<br />

neighbouring SiOH or SiOSi. A third species (ensemble) of surface chromium was reported<br />

by the Torino group [39] in 1987. If a reduced catalyst is heated in vacuo or in inert atmosphere<br />

to higher temperatures, say 600 8C, a substantial part of the Cr(II)A is converted into<br />

an almost inert Cr(II)C with a proposed coordination number of four. It is assumed by the<br />

authors that highly exposed Cr(II) ions may migrate to a thermodynamically more favourable<br />

position in the surface-near part of the bulk, reaching there a four-fold coordination.<br />

Consequently the new species behaves semisaturated <strong>and</strong> may be distinguished from A by<br />

its lack of reactivity vs. certain lig<strong>and</strong>s. Cr(II)B is not affected by the thermal treatment.<br />

In the light of new results we believe that this concept of Cr(II)(A,B,C) has to be reconsidered,<br />

as will be discussed below. Here we will just emphasise the surprising fact that<br />

the surface of amorphous silica offers only two or three topologically different positions for<br />

a two-valent metal ion. The origin from a covalent precursor – in the chromium case the<br />

mixed anhydride of poly-silicic acid <strong>and</strong> chromium(VI) acid – is not essential, since systems<br />

impregnated with Cr(II), Cu(II), <strong>and</strong> Fe(II) show the same phenomenon. We can argue that<br />

the different topology is inherently coming up in the activation step already (where we see<br />

the support attacked by a twotined fork of the metal compound). However, it needs the unsaturated<br />

character of the low valence ion for its chemical expression.<br />

How about the other metals? In the case of molybdenum a simultaneous irradiation<br />

with UV light is necessary to get the Mo(VI) attacked by CO [26, 27]. Even then, the average<br />

oxidation number reached was not lower than 2.7 (oxygen titration [27]). Spectroscopic<br />

evidence shows that the product contains Mo(IV) <strong>and</strong> Mo(II), formed by the reduction of octahedral<br />

<strong>and</strong> tetrahedral Mo(VI) [28]. Vanadium(V) requires a very long time <strong>and</strong> higher<br />

temperatures (600 8C) for the reduction by CO [19, 40] (slow de-aggregation?). For iron <strong>and</strong><br />

copper see Section 18.4.2.<br />

So far, we have formulated the reductions with CO as the reducing agent. Of course<br />

other means are as well applicable. Each has its specialities: CO for instance must be removed<br />

from the catalyst at the reduction temperature by vacuum or argon to avoid complex<br />

formation. Hydrogen does not form complexes but produces water which can hydrolyse the<br />

bond between metal <strong>and</strong> support of the not yet reduced units. The reductive effect does not<br />

surpass the one of CO. Other reducing agents like SO 2 do not offer any advantage either.<br />

The use of Li(AlH 4 ) or of activated magnesium was not successful.<br />

To prepare the catalytic centres for a later reaction of olefins with the c.u.s., hydrocarbons<br />

(alkanes, alkenes) may be used as reducing agents. Here the extent of the reduction is<br />

usually lower. Cr reaches oxidation numbers near three [14]. The great variety of organic re-<br />

351


18 Topospecific Chemistry at Surfaces<br />

action products makes an interpretation difficult. Since for some polymerization processes<br />

the addition of a cocatalyst (AlR 3 , AlR n Cl 3–n , SnR 4 ) is useful or even necessary, the reductive<br />

potentials of these metal alkyls were tested as well. Here we observed a well-defined<br />

two step reaction from Cr(VI) via Cr(IV) to Cr(II) [41].<br />

18.4.2 Elimination of lig<strong>and</strong>s<br />

The removal of the oxygens of surface Cr(VI) may be classified as a reductive elimination<br />

of lig<strong>and</strong>s. If from the beginning the metal is applied in a low valence state the formation of<br />

the c.u.s. requires just the (thermal) removal of the outer lig<strong>and</strong>s from the anchored surface<br />

compound:<br />

surface(Si-O) 2 ML n<br />

! surface(Si-O) 2 M_ n +nL.<br />

Fixation <strong>and</strong> lig<strong>and</strong> desorption usually are not separated processes, but are characterized<br />

by an overlap that can be studied by analysing the exhaust of a TPD experiment. In<br />

Fig. 18.2 this is shown for FeCl 2 . The reaction can be carried out in an inert, i. e. oxygenfree,<br />

gas stream (Ar) or under high vacuum conditions. Cr <strong>and</strong> Mo as well may be also applied<br />

to the support in low valence state as Cr(III) [42–47] or Cr(II)/Mo(II) [23].<br />

Figure 18.2: Desorption of lig<strong>and</strong>s from the system FeCl 2 7aq/silica [24]. Release of H 2 O <strong>and</strong> HCl vs. T.<br />

352


18.5 Physical properties of the coordinatively unsaturated sites<br />

18.5 Physical properties of the coordinatively unsaturated sites<br />

18.5.1 Topologically different sites<br />

As stated earlier, the preparation of the c.u.s. on amorphous substrates does not yield a uniform<br />

surface compound. Depending on the individual position of the metal on the surface,<br />

usually two discrete, topologically different ensembles of species are formed simultaneously.<br />

Their relative amounts depend on a variety of parameters. The temperatures of activation<br />

<strong>and</strong> reduction have the most serious influence. Of practical importance are further the nature<br />

of the support, its thermal history, the concentration of the metal, the metal compound used<br />

for impregnation, <strong>and</strong> (if applied) the reducing agent.<br />

18.5.2 Optical <strong>and</strong> magnetic properties<br />

For measuring the relative amount of the different species, physical methods play the most<br />

important role since they do not influence their ratio. The impressive colours of the catalysts<br />

recommend the use of optical (reflectance) spectroscopy. In the chromium case, for example,<br />

the characteristic absorptions (Fig. 18.3) allow a deconvolution <strong>and</strong> – with certain assump-<br />

Figure 18.3: UV-VIS-NIR reflectance spectra of surface chromium on silica. — T act = 800 8C <strong>and</strong> T red<br />

350 8C, green product; - - - T act = 500 8C <strong>and</strong> T red = 500 8C, blue product.<br />

353


18 Topospecific Chemistry at Surfaces<br />

tions with respect to the relative e values – an estimation of the relative amounts of the A<br />

<strong>and</strong> B species [13, 37, 38]. If we vary the activation temperature at otherwise constant preparation<br />

conditions the intensities of the characteristic b<strong>and</strong>s for Cr(II)A <strong>and</strong> B indicate that<br />

the A type is increasing up to 600 8C with a corresponding decrease of the B type<br />

(Fig. 18.4). This coincides roughly with the release of water from vicinal SiOH groups<br />

(680 8C, DTG). Heating the reduced sample (vacuum, 600 8C) diminishes the intensity of<br />

the Cr(II)A b<strong>and</strong>s but does not cause the appearance of new absorptions to be attributed to a<br />

Cr(II)C – provoking some doubts about this species, see below.<br />

Figure 18.4: Areas F rel of the main absorption b<strong>and</strong>s of Cr(II)A at 6.8610 3 cm –1 <strong>and</strong> 11.9610 3 cm –1<br />

(black), Cr(II)B at 8.6610 3 cm –1 <strong>and</strong> 13.5610 3 cm –1 (open) in deconvoluted UV-VIS-NIR reflectance<br />

spectra vs. T act .<br />

The results with other metals are less advanced. The study of the iron system suggested<br />

the use of Mößbauer spectroscopy. Unfortunately, the spectra showed a very modest<br />

resolution because of the experimental conditions (c (Fe) < 8%, airtight cuvette, no enrichment<br />

of Fe-57 isotope). The spectra cannot be interpreted conclusively [24].<br />

354


18.6 Chemical properties of the coordinatively unsaturated sites<br />

18.6 Chemical properties of the coordinatively unsaturated sites<br />

The different grade of saturation defining the two species A <strong>and</strong> B is of course reflected by<br />

their chemical reactivity.<br />

The most striking observation offers the reaction with oxygen: while Cr(II)B dominated<br />

samples show here a dark red glowing chemoluminescence, Cr(II)A dominated samples<br />

react under emission of a bright blazing red light [48, 49]. Both phenomena parallel the<br />

features shown by the photoluminescence of the activated samples, mentioned above, especially<br />

the critical dependence on the activation temperature (Fig. 18.1). The reaction pathway<br />

is well understood, including the intermediate formation <strong>and</strong> decomposition of peroxo species<br />

[50]. Surface Cr(III) does not react with oxygen. As to other metals, the reduced surface<br />

species of vanadium <strong>and</strong> molybdenum show a green or a blue chemoluminescence respectively.<br />

These reactions are not yet explored in detail. An intermediate peroxide formation is<br />

observed in the vanadium case [30, 40].<br />

Addition of donor lig<strong>and</strong>s results in the formation of well-defined complexes. Varying<br />

the preparation conditions of the catalysts <strong>and</strong> p, T of the lig<strong>and</strong> addition, the resulting UV-<br />

VIS-NIR spectra allow a discrimination of the different b<strong>and</strong> systems together with the determination<br />

of the correct parameters for the deconvolution [37, 38]. This requires some instinct.<br />

In the chromium case, together with the 2 b<strong>and</strong>s for Cr(III) <strong>and</strong> 263 b<strong>and</strong>s for the<br />

two Cr(II) species there appear the new b<strong>and</strong>s for Cr(II)A.L n with n = 1, 2, 3 <strong>and</strong> Cr(II)B.L m<br />

with m = 1, 2. As shown by the stoichiometric range, the A species offers one more vacant<br />

coordination site. The complex formation occurs with practically every n-donor (including<br />

even R 3 SiOSiR 3 ) <strong>and</strong> p-donor (including C 6 H 6 ). The extent of the reaction is a function of<br />

the A/B ratio, the donor properties of L, the concentration (partial pressure) of L, <strong>and</strong> the<br />

temperature [33, 37, 38]. The formation of complexes with CO <strong>and</strong> NO was studied thoroughly<br />

in the IR elsewhere [39, 51–53]. There are indications for the occurrence of neighbouring<br />

Cr centres that accommodate CO as a bridge [54].<br />

It should be mentioned that organic p-systems, like the double bond in olefins, benzene,<br />

or the cyclopentadienyl rings are added as well. The formation of 1:1 or 1 : 2 complexes<br />

is obviously the first step of all interactions of olefins with the catalytic centre [55].<br />

Not all of the additions are completely reversible. The temperatures necessary to volatilize<br />

the lig<strong>and</strong> are often too high to keep the system redox inert. Some of the lig<strong>and</strong>s, e. g.<br />

NO, are irreversibly added, others as N 2 O are decomposed at moderate temperatures under<br />

vacuum conditions (here to N 2 <strong>and</strong> surface Cr(IV)=O) [38]. Surface Cr(II)(H 2 O) 2 undergoes<br />

a redox reaction at 600 to 800 8C to form surface Cr(III) <strong>and</strong> hydrogen [38, 56].<br />

18.6.1 Survey of catalytic reactions<br />

The early work with the c.u.s. already included the study of catalytic reactions [57]. On the<br />

one h<strong>and</strong>, these were redox cycles like<br />

CO + 1/2 O 2 ! CO 2 or SO 2 + 1/2 O 2 ! SO 3<br />

355


18 Topospecific Chemistry at Surfaces<br />

that include both, the oxidised <strong>and</strong> the reduced form of the metal. On the other h<strong>and</strong>, there<br />

are reactions that concern the c.u.s. only, like the polymerization / oligomerisation, isomerization<br />

<strong>and</strong> metathesis of olefins as well as the cyclisation of acetylenes <strong>and</strong> the Fischer-<br />

Tropsch-like processes.<br />

18.6.2 Olefin polymerization<br />

These reactions include the Phillips process, i. e. the polymerization of ethylene at<br />

T&105 8C <strong>and</strong> pressures around 30 bar [4, 5]. Of course, this procedure attracted our special<br />

attention. In spite of its worldwide use, which makes more than 1/4 of the world production<br />

of HDPE, <strong>and</strong> after almost four decades of research the mechanism of the reaction is<br />

still a matter of controversial discussions [1, 2]. The varying results are partially due to different<br />

experimental starting points (e. g. different supports), occasionally also to the use of<br />

inadequate measurements, <strong>and</strong> the neglect of involved parameters.<br />

If we consider the case of ethylene polymerization, the main experimental obstacles are<br />

the necessary application of pressure <strong>and</strong> the insolubility of the high molecular product. A<br />

REM picture (Fig. 18.5) shows the growing polyethylene worms [58, 59]. We tried to avoid<br />

these drawbacks studying the polymerization of higher (liquid) 1-olefins, preferentially 1-noctene<br />

[60, 61]. With n C 6 4 the polymers are easily soluble in alkane solvents. This allows<br />

to apply high concentration of the monomer at normal pressure, to analyse the reaction mixture<br />

at any wanted time, <strong>and</strong> to use classical procedures for the study of the products [61].<br />

According to these experiments the polymerization of olefins is caused by a member<br />

(all members?) of the A species, while the B type centres cause some isomerization of the<br />

monomer. With catalysts of the same preparative history the specific polymerization activity<br />

is not dependent on the Cr concentration up to 1% [61, 62]. The reaction rate (as measured<br />

by the decrease of the monomer) is decreasing with n C from ethylene to 1-hexene [22].<br />

Figure 18.5: REM picture of growing polyethylene on surface Cr(II) [58, 59] with a rock of inert<br />

Cr 2 O 3 . The bar corresponds to 1 mm.<br />

356


18.6 Chemical properties of the coordinatively unsaturated sites<br />

Higher olefins polymerize practically with the same rate as C 6 . The comparison was carried<br />

out at deep temperatures to allow the gas phase olefins to be present in solution in the same<br />

starting concentration as the liquid members. This offers the chance of copolymerizing different<br />

olefins with n 6 6 without discrimination [60]. As to the products, the 1-olefins form<br />

comb-like polymer molecules with a practically unbranched backbone. The broad distribution<br />

of the molecular weight (MW) around 10 5 shows at low polymerization temperatures a<br />

three-modal profile that collapses at higher reaction temperatures to a mono-modal distribution<br />

with a lower MW value for its centre [61]. The chains contain one double bond per molecule<br />

[61].<br />

The use of the new system at ambient pressure <strong>and</strong> at moderate temperatures allows<br />

to follow the reaction kinetics by analysing the concentration of educts <strong>and</strong> products as a<br />

function of time. Batch experiments showed that there is a marked induction period of slow<br />

reaction progress (Fig. 18.6) [22, 60, 63]. The systems exhibit the characteristics of an autocatalytic<br />

reaction. Addition of new monomer up to the original starting concentration during<br />

the conversion of the monomer does not result in a new induction period but shortly after<br />

completion of the reaction the system answers to another addition of monomer with a new<br />

induction period [14, 64]. In steady state experiments the polymerization was proceeding<br />

over days with constant reaction rate, i. e. without decrease of the catalyst’s activity. The re-<br />

Figure 18.6: Reaction of surface Cr(II), T act = 800 8C <strong>and</strong> T red = 350 8C, with 1-octene at room temperature<br />

[22]. Decrease of monomer <strong>and</strong> increase of polymer show a marked induction period; the formation<br />

of isomers (triangles) is marginal.<br />

357


18 Topospecific Chemistry at Surfaces<br />

action was shown to be of first order in monomer, solving an old controversy in the literature<br />

[63]. The ethylene system does not allow clear answers since the solid reaction product<br />

causes difficulties in the data interpretation, not to speak of problems with collecting the<br />

samples at high pressure <strong>and</strong> high temperature conditions.<br />

A st<strong>and</strong>ard reaction with 1-octene (polytest) was used for a quick characterisation of<br />

the single catalysts. The time t for 50% consumption of the monomer measures the general<br />

activity in 1/t. The time for 10% consumption gives an information on the induction period.<br />

The GC analysis after the completion of the reaction shows an additional formation of low<br />

MW products, for example isomers [14].<br />

Of course it was tempting to connect the observed catalytic properties with the variation<br />

of certain parameters. Primarily the old question of oxidation number <strong>and</strong> coordination<br />

number of the active centres deserved new studies. By variation of the reducing agent – now<br />

including methane, ethene, ethylene, <strong>and</strong> isobutene – we learned that an oxidation number<br />

of 2.0 is not favourable. Surprisingly catalysts with oxidation numbers of 3.0 +0.5 were by<br />

far superior to the fully reduced samples [14]. This result agrees to our earlier observation<br />

that surface Cr(III) species, prepared from surface Cr(II)(H 2 O) 2 by heating in argon, are<br />

good catalysts for ethylene polymerization [56]. Even sheer heating of a Cr(II) catalyst to<br />

800 8C in vacuo increases the activity considerably while the oxidation number rises simultaneously<br />

[23]. Lunsford [43] published results, which show a clear coincidence of the activities<br />

of a special surface Cr(III) <strong>and</strong> the heated Cr(II) catalysts. According to Garrone et al.<br />

[34] the latter should contain the species Cr(II)C besides residual Cr(II)A <strong>and</strong> the original<br />

share of Cr(II)B, which is not involved in the polymerization reaction. We are inclined to<br />

postulate that Cr(II)C does not exist at all but rather is a Cr(III) species. This is supported<br />

by the fact that there are no Cr(II)C b<strong>and</strong>s in the UV-VIS-NIR spectra. The absorptions may<br />

be hidden in the Cr(III) b<strong>and</strong>s. If so, the transition Cr(II) to Cr(III) must produce a reduction<br />

product, too. Hydrogen from neighbouring SiOH is the only c<strong>and</strong>idate, because of the lack<br />

of H 2 O molecules that are not present anymore:<br />

(surface SiO) 2 Cr(II) + surface SiOH ! (surface SiO) 3 Cr(III) + 1/2 H 2<br />

Heating of Cr(II) impregnated catalysts [23] to 800 8C in Ar produces contacts with<br />

the high oxidation number 2.80! that polymerize olefins without induction period. If the<br />

same starting product is activated in oxygen <strong>and</strong> reduced in CO (O.N. 2.14) it reacts with induction<br />

period, as usual. So we might conclude that the appearance of an induction period is<br />

caused by a slow transition of a precursor surface Cr(II) to surface Cr(III). However, the active<br />

centres in the first mentioned sample are obviously different from active species that<br />

are produced from Cr(II) during the induction period. Only the latter are transformed back<br />

to their precursors by lack of monomers [14].<br />

This behaviour dem<strong>and</strong>s a reconsidering of the Cr(II)A concept. Up to now the A species<br />

was regarded to be well off from reactive surface OH groups. The marked <strong>and</strong> steady increase<br />

of the A/B ratio up to an activation temperature of 600 8C, roughly correlated with the<br />

disappearance of vicinal SiOH at 680 8C [11], seems to support this assumption. Now this<br />

model may still be applied for some subspecies of Cr(II)A, but obviously not for that substantial<br />

part of the A centres which may undergo a redox reaction with neighbouring SiOH.<br />

There is another experimental access to this problem. If the support was treated in the<br />

beginning with D 2 O (H/D exchange 50 %), the final catalyst forms partially deuterated poly-<br />

358


18.6 Chemical properties of the coordinatively unsaturated sites<br />

1-octene <strong>and</strong> polyethylene [65]. Thermal deactivation (see below) yields hydrocarbons that<br />

contain one or two deuterium atoms [66]. Therefore the active centre had to have some<br />

SiOD surface groups in reach. Therefore the addition of lig<strong>and</strong>s should be rather considered<br />

as a kind of lig<strong>and</strong> replacement.<br />

This new model is indeed a very old one (Hogan, 1956) [5] <strong>and</strong> includes a simple formulation<br />

of the starting step of the polymerization cycle yielding an active species with<br />

Cr(IV),<br />

(surface SiO) 2 Cr(II) + surface SiOH + C 2 H 4 ! (surface SiO) 3 Cr(C 2 H 5 ).<br />

The consumption of the hydrogen by formation of an ethyl group is, however, in competition<br />

with the use of this hydrogen for the formation of an active surface Cr(III) from a<br />

Cr(II) precursor as formulated above. Further arguments against the ethylated Cr species<br />

come from the IR spectroscopy of centres in polymerization. In careful studies the Torino<br />

group [34] showed that there is no n CH of a CH 3 (end)group in the IR spectra. Correspondingly,<br />

the authors support a chromacycle model for the growing chain. If both, C 2 H 4 <strong>and</strong><br />

CO, are present at the virginal Cr(II) centre a surface chroma-cyclohexanone-1 is formed, as<br />

is proved by IR [34]. <strong>Final</strong>ly, according to the NMR spectra of polymers, formed with deuterated<br />

catalysts, the incorporation of the D atoms does not result in a formation of –CH 2 D<br />

groups [65].<br />

To make the picture even more complicated the steady increase of A-type centres with<br />

T act going from 300 8C to 600 8C is not paralleled by some other data which are ascribed to<br />

the Cr(II)A species or its Cr(VI) precursor. As shown in Fig. 18.1, the structured chemoluminescence,<br />

photoluminescence, polymerization activity, <strong>and</strong> content of CrC s-bonds in the<br />

working catalyst 1 rise steeply in the range of 450 8C to 700 8C with a turning point near<br />

600 8C. At this temperature the Cr(II)A b<strong>and</strong>s in the VIS spectra are already fully developed<br />

(Fig. 18.4).<br />

What is Cr(II)A? Originally defined by its reactivity vs. CO (in contrast to the more inert<br />

Cr(II)B [35]), the highly unsaturated character was quantified by measuring DH of the catalyst’s<br />

reaction with O 2 [55, 67–69]. The temperature-programmed desorption of innocent lig<strong>and</strong>s<br />

follows the same basic concept (population profile) [13, 38]. Another approach is the<br />

deconvolution of the spectra of real catalysts, first used by Blümel [37]. However, the rough<br />

coincidence of changes towards A properties by certain treatments is too inexact a measure. A<br />

comparison of Fig. 18.1 with Fig. 18.4 shows that further structuring is needed. Especially we<br />

have to become acquainted to the idea that at least part of the Cr(II)A is positioned in the<br />

neighbourhood of OH groups. This part is converted to Cr(II)C/Cr(III) by heating.<br />

Under topological aspects an impregnation of the support with dimeric Cr(II)acetate<br />

<strong>and</strong> its Mo <strong>and</strong> Cr/Mo analogues were of interest. There are clues that a pairwise fixation is<br />

obtained <strong>and</strong> pertained during the following preparation steps [23]. However, the differences<br />

in catalytic behaviour vs. st<strong>and</strong>ard catalysts are not very pronounced.<br />

1 Such bonds were shown to be present <strong>and</strong> determined quantitatively by quenching the polymerization<br />

by oxygen <strong>and</strong> analysing the aldehydes formed [70]. They represent up to 10% of the Cr(II) [71, 72].<br />

The details are not yet fully understood. With 1-octene as monomer the s-content is rising linearly<br />

with the progress of the reaction [73]. A well-expressed maximum of the s-bond formation is shown<br />

at a Cr concentration of 1%, maybe due to a topologically special situation [72].<br />

359


18 Topospecific Chemistry at Surfaces<br />

As expected all metals have their individual catalytic profile. The chromium system is<br />

by far the most suitable for the use as polymerization catalyst. Vanadium(III) without cocatalyst<br />

requires UV irradiation to yield at least some polyethylene besides oligomers [19].<br />

With addition of Al(R,Cl) 3 polyethylene is formed after a long induction period in reasonable<br />

yields but with very high MW of about 6610 6 [21]. Higher 1-olefins form dimers in<br />

good yields even if the double bond is in an internal position [19]. Alkanes (solvent!) may<br />

block the free coordination sites at surface V(III) <strong>and</strong> cause very long induction periods [19,<br />

20]. The reaction of reduced surface molybdenum with 1-olefins produces metathesis products<br />

besides oligomers [27]. With copper the catalytic reaction (with ethylene) was shown<br />

to be restricted to the Cu(II)A species. After a prior complex formation the polymerization<br />

produces oligomers with even <strong>and</strong> odd C numbers. Additionally benzene, toluene, <strong>and</strong> ethylbenzene<br />

are found [16, 74, 75].<br />

Since there is no obvious connection to topological facts, the effect of Lewis acids<br />

(Al(R,Cl) 3 , SnCl 4 ) as cocatalysts is mentioned here only marginally. The induction period is<br />

diminished <strong>and</strong> the reaction is accelerated while the spectrum of products is shifted towards<br />

isomerization [22, 23]. Furthermore the Lewis acids may attack the SiOSi bonds of the support<br />

<strong>and</strong> so – by loosening its rigidity – make the popping of the catalyst grains easier (polyethylene<br />

process!). At high concentrations of the cocatalyst, a splitting of SiOM bonds may<br />

occur as well [21]. <strong>Final</strong>ly Al(R,X) 3 acts as reducing agent with surface Cr(VI) [41].<br />

18.6.3 Other catalytic reactions<br />

As already mentioned the catalysts show a certain isomerization activity that may become<br />

the main reaction under certain experimental conditions. This is the only activity of the<br />

Cr(II)B species vs. olefins. The mechanism looks like an acid catalysis, using the protons<br />

from to chromium coordinated OH groups. Addition of Lewis acids supports this reaction.<br />

As mentioned above, Mo containing catalysts can react with olefins under metathesis<br />

[23, 27]. This shows that the carbene mechanism is at least not excluded for this type of surface<br />

compounds.<br />

Using mixtures of olefins <strong>and</strong> hydrogen as educts the olefins undergo hydrogenation<br />

under mild conditions <strong>and</strong> with quantitative yields [61]. With Ru catalysts this is the typical<br />

reaction [29].<br />

<strong>Final</strong>ly it should be mentioned that Cr(II) surface compounds may act as catalysts in a<br />

Fischer-Tropsch system. From H 2 <strong>and</strong> CO or CO 2 they produce methane with some ethane<br />

<strong>and</strong> propane. The reaction pathway may follow the mechanism given 1976 by Henrici-Olivé<br />

<strong>and</strong> Olivé [76]. Important steps are here the intermediate formation of a formaldehyde complex<br />

follow C 1 carbene complex [77].<br />

360


18.7 Deactivation<br />

18.7 Deactivation<br />

If the reaction of Cr(II) catalysts with olefins is carried out at higher temperatures, oligomer/polymer<br />

chains with odd carbon numbers appear, often with the same MS intensity as<br />

their even numbered neighbours [70]. Furthermore CH 2 units are transferred [66, 78], H 2 ,<br />

CH 4 ,C 2 H 6 ,C 3 H 8 , <strong>and</strong> n-C 4 H 10 are formed [66, 79, 80], <strong>and</strong> finally the catalyst is deactivated.<br />

All the Cr(II)A is converted into a new deep green species that was believed for a<br />

long time to be a bis-Z 3 -allyl-Cr(IV) complex [15, 59, 80]. These processes start slowly at<br />

120 8C [78], while at 300 8C the deactivation reaction is completed within seconds. The reactions<br />

require intermediates that may undergo a C–C cleavage, as it is commonly observed in<br />

carbene complex reactions. Such intermediates, recently proposed for the Ziegler-Natta<br />

polymerization [81, 82], can indeed also be formulated for the Phillips process [83]. Since<br />

the carbene complex moiety can also be formulated as a hydrido metallacycle (no CH 3 !),<br />

this model seems to us to be the most promising hypothesis, although some typical carbene<br />

reactions have not been observed yet [84].<br />

The new Cr surface compounds can be split from the support by Brønsted acids, e. g.<br />

HCl, <strong>and</strong> washed out with organic solvents (methanol, acetone, etc.). The active centres of<br />

the heterogeneous catalyst – or better their deactivation products – form a deep blue solution<br />

this way [59, 70, 83, 85].<br />

As expected, the material is not uniform but consists of an ensemble of homologue individuals,<br />

separated by CH 2 . The minimal C number is 8. The higher homologues are resulting<br />

from the different extent of the polymerization prior to deactivation. Recently we succeeded<br />

in separating the single fractions <strong>and</strong> in isolating crystals from the C 8 fraction. X-ray<br />

analysis showed surprisingly that the product is actually a cyclopentadienyl-Cr(III) complex<br />

of the formula (CpmCrCl 2 ) 2 , where Cpm is a 1,2,3-trimethyl-cyclopentadienyl lig<strong>and</strong> C 8 H 11<br />

(Fig. 18.7).<br />

In solution of donor solvents the dimer is split into monomers,<br />

(CpmCrCl 2 ) 2 +2D ! 2 CpmCrCl 2 .D .<br />

The surprising formation of a substituted Cp ring system may proceed via a ring slippage<br />

reaction from a corresponding metallacycle [86–89].<br />

The product belongs to a well-known group of complexes found by E.O. Fischer in<br />

1963 [90, 91]. The corresponding Cp derivative was also prepared by the reaction of surface<br />

Cr(II) with cyclopentadiene [55]. Unfortunately the conversion of ethylene into a Cp product<br />

is not a straightforward process: The hope coming to a new underst<strong>and</strong>ing of the structure<br />

of the active centres <strong>and</strong> for the appearance of odd C numbers in products from even olefins<br />

was disappointed. In our context we have to notify another example for the difference between<br />

Cr(II)A <strong>and</strong> B; though all Cr(II)A sites are able to form at least C 8 from C 2 , the B<br />

species are not converted. The dehydrogenation of four ethylene molecules to a C 8 H 11 moiety<br />

is obviously the source of the hydrogen needed to form the H 2 <strong>and</strong> the alkanes mentioned<br />

above.<br />

361


18 Topospecific Chemistry at Surfaces<br />

Figure 18.7: Structure of [CpmCrCl 2 ] 2 , determined by X-ray analysis [84].<br />

18.8 Summary <strong>and</strong> outlook<br />

As reported in the previous paragraphs, there is still no commonly accepted conclusive <strong>and</strong><br />

coherent picture of the surface chemistry of metal impregnated silica <strong>and</strong> related systems.<br />

Nevertheless, there are some unequivocal results. It is obvious, that there are different surface<br />

compounds, well differentiated by their oxidation state <strong>and</strong> by their topological environment.<br />

The latter does not offer a chaotic multitude. It leads to the occurrence of just a few<br />

(usually two) different types for a given metal unit. They can be characterised physically by<br />

their spectra <strong>and</strong> distinguished chemically by their different degree of unsaturation. The<br />

question of the real chemical environment of the species/subspecies, however, cannot be answered<br />

definitively yet. For instance the preparation procedure suggests to exclude SiOH in<br />

the vicinity of Cr(II)A, but the experiments with deuterated silica show the incorporation of<br />

neighbouring SiOD in the polymerization of 1-olefins at just these Cr(II)A centres. Our st<strong>and</strong>ard<br />

polymerization system (surface Cr(II) + 1-octene) provides an easy access to screening<br />

experiments. However, the inevitable large variety of reacting centres at the surface did not<br />

allow to relate unambiguously any response to any parameter. Although we were able to isolate<br />

the deactivation product of an active centre <strong>and</strong> to get its structure via X-ray analysis,<br />

the step from active to inactive species is still a matter of speculation.<br />

Prospects? Obviously, the multidimensional network of interactions needs further investigations.<br />

Whenever it is possible, additional methods should be incorporated. For example,<br />

the advanced mathematical concepts of building-up structures, like silica from given angles,<br />

coordination numbers, <strong>and</strong> bond lengths, may allow to explain the A <strong>and</strong> B-type topology.<br />

On the other h<strong>and</strong>, the advanced microscopic methods (STM, AFM, <strong>and</strong> LFM) could<br />

provide literally new insights in the fine structure of hydrated SiO 2 surfaces. A first step in<br />

this direction was successful [92]. The oxygen-covered 111 face of a quasi-cristobalite surface<br />

layer (Fig. 18.8) [91], may also be used for modelling silica surfaces, correspondingly.<br />

If we succeed to incorporate metal ions to these surfaces we might be able to simulate active<br />

centres <strong>and</strong> to determine their behaviour.<br />

362


References<br />

Figure 18.8: Lateral force microscopy (LFM) of a 111 face of quasi-cristobalite on monocrystalline silicon<br />

[91]. The bar corresponds to 10 Å.<br />

„Woran arbeiten Sie?“ wurde Herr K. gefragt. Herr K. antwortete: „Ich habe viel<br />

Mühe, ich bereite meinen nächsten Irrtum vor.“ Bertolt Brecht [93]<br />

References<br />

1. M. P. McDaniel: Advances in Catalysis, 33. 47 (1985)<br />

2. C. E. Marsden: Rubber <strong>and</strong> Composites, Processing <strong>and</strong> Applications, 21, 193 (1994)<br />

3. H. L. Krauss, F. Schwarzbach: Chem. Ber., 94, 1205 (1961)<br />

R. B. Johannesen, H. L. Krauss: ibid., 97, 2094 (1964)<br />

4. U. S. Pat. 2825 721, 2960 126 (1958)<br />

Brit. Pat. 790196, 80464 (1958)<br />

5. A. Clark, J. P. Hogan, R. L. Banks, W. S. Lanning: Ind. Eng. Chem., 48, 1152 (1956)<br />

V. A. Topchiev, B. A. Krentsch, A. J. Perelman, K. G. Miesserov: J. Polym. Sci., 34, 129 (1959)<br />

J. P. Hogan, J. Polym. Sci., 8, 2673 (1970)<br />

6. H. L. Krauss, E. Weber, N. Mövik: Z. Anorg. Allg. Chem., 338, 121 (1965)<br />

7. H. L. Krauss, H. Stach: Inorg. Nucl. Lett., 4, 393 (1968)<br />

8. H. L. Krauss, H. Stach: Z. Anorg. Allg. Chem., 366, 280 (1969)<br />

9. L. D. Evans, S. V. King: Nature, 212, 1353 (1966)<br />

A. Hasmy, M. Foret, J. Pelous, R. Julien: J. Phys. IV, 3 (C8), 365 (1993)<br />

V. Khavryuthenko, E. Sheka: React. Kinet. Catal. Lett., 50, 389 (1993)<br />

A. Nakano, L. Bi, R. K. Kalia, P. Vashishta: Phys. Rev. Lett., 71, 85 (1993)<br />

T. P. M. Beelen, W. H. Dokter, H. F. van G<strong>and</strong>eren, R. A. van<br />

Santen: Adv. Colloid Interface Sci. 50, 23 (1994)<br />

K. Uzelac, A. Hasmy, R. Julien: Phys. Rev. Lett. 74, 422 (1995)<br />

363


18 Topospecific Chemistry at Surfaces<br />

10. J. P. Peri, A. L. Hensley jr.: J. Phys. Chem., 72, 2926 (1968)<br />

11. H. L. Krauss, D. Naumann: Z. Anorg. Allg. Chem., 430, 23 (1977)<br />

12. A. Sebald, H. L. Krauss: unpublished.<br />

13. H. L. Krauss, E. Amberger, N. Arfsten, P. Blümel, W. Hammon, R. Höpfl, W. Riederer: Stoichiometric<br />

Reactions with Reduced Phillips Catalysts, in: Y. Imamoglu (ed.): Olefin Metathesis <strong>and</strong><br />

Polymerization Catalyszts, Kluwer Academic Publishers, Dordrecht, 359 (1990)<br />

14. B. Siebenhaar: PhD thesis, Universität Bayreuth (1991)<br />

15. K. Hagen: PhD thesis, Universität Bayreuth (1982)<br />

16. H. L. Krauss, G. Guldner, M. Hornscheidt, N. Larsen, R. Merkel, P. Morys, S. Schmerbeck,<br />

P. Zahn: Non-Chromium Surface Compounds, in: Y. Imamoglu, (ed.): Olefin Metathesis <strong>and</strong> Poymerization<br />

Catalysts, Kluwer Academic Publishers, Dordrecht, 385 (1990)<br />

17. N. Larsen: PhD thesis, Technische Hochschule München (1970)<br />

18. B. Horvath, J. Geyer, H. L. Krauss: Z. Anorg. Allg. Chem., 426, 141 (1976)<br />

19. G. Guldner, PhD thesis, Universität Bayreuth (1986)<br />

20. H. L. Krauss, P. Zahn: unpublished<br />

21. M. Pohl: PhD thesis, Universität Bayreuth (1993)<br />

22. E. Amberger: PhD thesis, Universität Bayreuth (1993)<br />

23. Th. Bohley: PhD thesis, Universität Bayreuth (1993)<br />

24. R. Merkel: PhD thesis, Universität Bayreuth (1991)<br />

25. H. L. Krauss, R. Gerritzen: Z. Anorg. Chem., 464, 99 (1980)<br />

26. M. Hornscheidt: PhD thesis, Freie Universität Berlin (1979)<br />

27. S. Schmerbeck: PhD thesis, Universität Bayreuth (1987)<br />

28. P. Morys, S. Schmerbeck: Z. Naturforsch., 42 b, 756 (1987)<br />

29. J. Köhler: PhD thesis, Universität Bayreuth (1992)<br />

30. R. Möseler, B. Horvath, D. Lindenau, E. G. Horvath, H. L. Krauss: Z. Naturforsch., 31 b, 892<br />

(1976)<br />

31. B. Rebenstorf: Acta Chim. Sc<strong>and</strong>., A31, 208, 877 (1977)<br />

32. M. Schraml-Marth, A. Wokaun, M. Pohl, H. L. Krauss: J. Chem. Soc., Faraday Trans., 87, 2635<br />

(1991)<br />

33. H. L. Krauss, D. Naumann: Z. Anorg. Allg. Chem., 446, 23 (1978)<br />

34. E. Garrone, G. Ghiotti, A. Zecchina: Coordination Chemistry of Silica Supported Cr Ions: Types<br />

of Cr(II) Ions on Reduced Samples, in: Y. Imamoglu (ed.): Olefin Metathesis <strong>and</strong> Polymerization<br />

Catalysts, Kluwer Academic Publishers, Dordrecht, 393 (1990)<br />

35. B. Rebenstorf: PhD thesis, Freie Universität Berlin (1975)<br />

36. H. L. Krauss, B. Rebenstorf, U. Westphal: Z. Anorg. Allg. Chem., 415, 57 (1975)<br />

37. B. Blümel: PhD thesis, Freie Universität Berlin (1979)<br />

38. R. Höpfl: PhD thesis, Universität Bayreuth (1981)<br />

39. E. Garrone, G. Ghiotti, C. Morterra, A. Zecchina: Z. Naturforsch., 42 b, 728 (1987)<br />

40. J. Geyer: PhD thesis, Freie Universität Berlin (1981)<br />

B. Horvath, J. Geyer, H. L. Krauss: Z. Anorg. Allg. Chem., 426, 141 (1976)<br />

41. H. L. Krauss, B. Hanke: Z. Anorg. Allg. Chem., 521, 111 (1985)<br />

42. V. A. Zakharov, Yu. I. Yermakov: Catal. Rev. Sci. Eng., 19, 67 (1979)<br />

43. D. D. Beck, J. H. Lunsford: J. Cat., 68, 121 (1981)<br />

44. D. L. Myers, J. H. Lunsford: J. Cat., 92, 260 (1985)<br />

45. D. L. Myers, J. H. Lunsford: J. Cat., 99, 140 (1986)<br />

46. J. H. Lunsford, S.-L. Fu, D. L. Myers: J. Cat., 111, 231 (1988)<br />

47. B. Rebenstorf: J. Cat., 84, 240 (1983)<br />

48. H. L. Krauss, H. Stach: Z. Anorg. Allg. Chem., 366, 34 (1969)<br />

49. P. Morys, R. Gerritzen, H. L. Krauss: Z. Naturforsch., 31 b, 774 (1976)<br />

50. P. Morys, U. Goerges, H. L. Krauss: Z. Naturforsch., 39 b, 458 (1984)<br />

51. G. Ghiotti, E. Garrone, A. Zecchina: J. Mol. Cat., 65, 73 (1991)<br />

52. G. Spoto, S. Bordiga, E. Garrone, G. Ghiotti, A. Zecchina, G. Petrini, G, Leofanti: J. Mol. Cat.,<br />

74, 175 (1992)<br />

364


References<br />

53. A. Zecchina, G. Spoto, G. Ghiotti, E. Garrone: J. Mol. Cat., 86, 423 (1994)<br />

54. B. Rebenstorf, R. Larsson: Z. Anorg. Allg. Chem., 478, 119 (1981); J. Mol. Cat., 11,<br />

247 (1981)<br />

55. H. L. Krauss, U. Westphal: Z. Naturforsch., 33 b, 1278 (1978)<br />

56. D. Naumann: PhD thesis, Freie Universität Berlin (1979)<br />

57. H. Stach: PhD thesis, Technische Universität München (1968)<br />

58. W. Riederer: PhD thesis, Universität Bayreuth (1989)<br />

59. H. A. Schmidt: PhD thesis, Universität Bayreuth (1992)<br />

60. K. Weiss, H. L. Krauss: J. Cat., 88, 424 (1984)<br />

61. G. Langstein: PhD thesis, Universität Bayreuth (1986)<br />

62. H. L. Krauss, B. Bojer: unpublished<br />

63. H. L. Krauss, G. Langstein: J. Mol. Cat., 65, 101 (1991)<br />

64. H. L. Krauss, B. Frank, B. Hanke, E. Hums, G. Langstein. D. Naumann, B. Siebenhaar, K. Weiss:<br />

Catalytic Reactions with Reduced Phillips Catalysts, in: Y. Imamoglu (ed.): Olefin Metathesis <strong>and</strong><br />

Polymerization Catalysts, Kluwer Academic Publishers, Dordrecht, 375 (1990)<br />

65. Q. Xing: PhD thesis, Universität Bayreuth (1993)<br />

66. E. Hums: PhD thesis, Universität Bayreuth (1981)<br />

67. H. L. Krauss, B. Rebenstorf, U. Westphal: Z. Anorg. Allg. Chem., 414, 97 (1975)<br />

68. H. L. Krauss, U. Westphal: Z. Anorg. Allg. Chem., 430, 218 (1977)<br />

69. W. Hammon: PhD thesis, Universität Bayreuth (1986)<br />

70. H. Schmidt: PhD thesis, Technische Universität München (1973)<br />

71. G. Ghiotti, E. Garrone, A. Zecchina: J. Mol. Cat., 46, 61 (1988)<br />

72. P. Wolff: PhD thesis, Universität Bayreuth (1991)<br />

73. H. L. Krauss, B. Putz, B. Bojer: unpublished<br />

74. R. Gerritzen, H. L. Krauss: Z. Anorg. Allg. Chem., 464, 99 (1980)<br />

75. R. Gerritzen: PhD thesis, Freie Universität Berlin (1981)<br />

76. G. Henrici- Olivé, S. Olivé: Angew. Chem., 88, 144 (1976); J. Mol. Cat., 4, 379 (1978)<br />

G. Henrici- Olivé, S. Olivé: ibid., 24, 7 (1984)<br />

77. E. Hums, H. L. Krauss: Z. Anorg. Allg. Chem., 527, 154 (1986)<br />

78. H. L. Krauss, G. Zeitler-Zürner: unpublished<br />

79. H. L. Krauss, E. Hums: Z. Naturforsch., 38 b, 1412 (1983)<br />

80. H. L. Krauss, K. Hagen, E. Hums: J. Mol. Cat., 28, 233 (1985)<br />

81. M. L. H. Green: Pure Appl. Chem., 50, 27 (1978)<br />

M. L. H. Green, A. Mahtab: J. C. S. Dalton, 1979, 262<br />

82. K. J. Irvin et al., J. C. S. Chem. Comm., 1978, 604<br />

83. H. L. Krauss, E. Hums: Z. Naturforsch., 34 b, 1628 (1979)<br />

H. L. Krauss, E. Hums: ibid., 35 b, 848 (1980)<br />

84. K. Weiss, K. Hoffmann: J. Mol. Cat., 28, 99 (1985)<br />

85. Q. Xing, H. L. Krauss, W. Milius: J. Mol. Cat., 90, 75 (1994)<br />

86. J. R. Bleeke et al.: Organometallics, 6, 1576 (1987)<br />

J. R. Bleeke et al.: J. A. C. S., 111, 4118 (1989)<br />

87. R. Ferede, T. Allison: Organometallics, 2, 463 (1983)<br />

88. W. R. Roper et al.: J. C. S. Chem. Comm., 1982, 811<br />

89. R. P. Hughes et al.: J. C. S. Chem. Comm., 1986, 1694<br />

90. E. O. Fischer, K. Ulm, P. Kuzel: Z. Anorg. Allg. Chem., 319, 253 (1963)<br />

91. F. H. Köhler, R. de Cao, K. Ackermann, J. Sedlmair: Z. Naturforsch., 38 b, 1406 (1983)<br />

92. Th. Schimmel, V. Popp, J. Küppers, H. L. Krauss: unpublished<br />

93. B. Brecht: in: Geschichten, Suhrkamp Verlag, Frankfurt/M. (1962)<br />

365


III<br />

Biopolymers<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed<br />

Chemistry of Biopolymers<br />

Stefan Limmer, Günther Ott, <strong>and</strong> Mathias Sprinzl<br />

19.1 Introduction<br />

In our laboratory several aspects of the protein biosynthesis are analyzed by different biochemical,<br />

molecular biological, <strong>and</strong> biophysical methods. In particular, research interest focused<br />

on tRNAs <strong>and</strong> the bacterial elongation factor Tu (EF-Tu). Both molecules specifically<br />

interact with each other in the course of the protein biosynthesis cycle.<br />

As indicated by its name, tRNA transports an amino acid to the ribosome where it is<br />

attached to the growing polypeptide chain. In this way, the tRNA forms the link between the<br />

genetic information being coded in the DNA <strong>and</strong> its RNA transcript (messenger RNA).<br />

For the exact binding of the aminoacylated tRNA at the A-site of the ribosome a complex<br />

between aminoacyl-tRNA <strong>and</strong> the active form of EF-Tu is required. Activation of EF-Tu<br />

is achieved by binding one molecule of GTP. By means of GTPase activity, GTP is cleaved<br />

into GDP <strong>and</strong> inorganic phosphate. This cleavage is accompanied by a dramatic conformational<br />

change of the protein structure [1, 2] whereby EF-Tu turns into its inactive GDP state.<br />

This transition can also be monitored via 1 H NMR studies [3]. In the GDP state, the capability<br />

of binding aminoacyl-tRNA is reduced by several orders of magnitude.<br />

By NMR studies information on structure <strong>and</strong> structure-function relationships of both,<br />

EF-Tu <strong>and</strong> tRNA, can be gained. In the case of tRNA, not only the whole molecule was analyzed,<br />

but also truncated structures that represented well-defined parts of the intact molecule<br />

<strong>and</strong> which still fulfilled certain biological functions, e. g. the acceptor stem of alanine-specific<br />

tRNA from Escherichia coli (E. coli) is correctly recognized by its cognate aminoacyltRNA<br />

synthetase <strong>and</strong> subsequently charged with alanine.<br />

NMR spectroscopy has proved to be a powerful method for the elucidation of the<br />

structure of relatively small macromolecules (molecular weights of up to ca. 10,000 Da). In<br />

particular, this applies to RNA molecules for which X-ray crystallographic analyses are relatively<br />

rare, mainly due to difficulties with their crystallization. Under certain conditions a total<br />

structure elucidation of these (small) molecules in their native environment (aqueous solution)<br />

at atomic resolution is feasible by means of multidimensional NMR methods [4–10].<br />

Moreover, even distinctly larger molecules, like whole tRNAs, can be at least partially characterized<br />

with reference to their structure, in particular secondary structure [11–14]. As detailed<br />

below, in certain cases NMR spectroscopy can also provide useful information for the<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

369


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

characterization of the interaction of relatively large molecular complexes, if e. g. certain<br />

isotopically labelled molecules act as NMR active probes. By such probes, the sites of a specific<br />

intermolecular interaction can be localized.<br />

In this article we present NMR studies with isotopically labelled aminoacyl-tRNA <strong>and</strong><br />

its complexes with EF-Tu 7GTP. Furthermore, we studied by combination of chemical RNAsynthesis<br />

<strong>and</strong> NMR spectroscopy, the structure of an RNA helix containing a single-str<strong>and</strong>ed<br />

dangling end <strong>and</strong> a specifically located non-Watson-Crick G-U base pair. By specific fluorescence-labelling<br />

of tRNA the dissociation constants for aminoacyl-tRNA, EF-Tu, <strong>and</strong> GTP<br />

were determined. <strong>Final</strong>ly, the structure <strong>and</strong> function of EF-Tu was studied by overexpression<br />

of thermostable EF-Tu originating from thermophilic bacterium Thermus thermophilus in<br />

E. coli, crystallisation, X-ray structure determination <strong>and</strong> site-directed mutagenesis.<br />

19.2 Site-specific NMR spectroscopy of chemically synthesized<br />

RNA duplexes<br />

One of the four arms of the tRNA secondary structure is the acceptor arm. It consists of seven<br />

base pairs <strong>and</strong> a single-str<strong>and</strong>ed 3'-terminus. At the 3'-terminal nucleotide (invariably<br />

A76), the amino acid is specifically attached. The aminoacylation of the tRNA occurs by esterification<br />

of the amino acid to the 2' or 3'-OH group of the ribose of adenosine A76 [15].<br />

This reaction is catalyzed by aminoacyl-tRNA synthetases.<br />

Earlier biochemical studies have demonstrated that in many cases the main identification<br />

features or recognition elements necessary for the correct recognition of tRNAs by their<br />

cognate aminoacyl-tRNA synthetases are localized in the acceptor stem [16–19]. Frequently,<br />

these recognition sites are referred to as identity elements since they make this particular<br />

tRNA unique <strong>and</strong> distinguishable from all other tRNAs for the corresponding aminoacyltRNA<br />

synthetase. For the tRNA Ala from E. coli it could be shown by the groups of Schimmel<br />

[20, 21] <strong>and</strong> McClain [19, 22–24] that a wobble base pair G3-U70 at position 3 of the<br />

acceptor arm represents the major identity element.<br />

Subsequently, it was demonstrated that even a truncated tRNA, consisting only of the<br />

acceptor arm with the single-str<strong>and</strong>ed 3'-end, is recognized as a substrate by the alanyltRNA<br />

synthetase (ARS) <strong>and</strong> correctly aminoacylated with alanine [21, 25], provided that the<br />

G3-U70 base pair is present at the correct location.<br />

This leads to the question whether this G-U pair causes local irregularities of the helical<br />

geometry, which are recognized by the ARS [24], or if certain functional groups<br />

(2-amino group of G3) are important in this recognition process [26, 27], or whether possibly<br />

both factors are affecting the ARS interaction with tRNA Ala . In addition, the influence<br />

of the ubiquitous terminal CCA sequence on the stability of the acceptor arm was studied.<br />

Earlier studies [28–30] indicated an ordered arrangement of the single-str<strong>and</strong> nucleotides.<br />

By varying both the number <strong>and</strong> the sequence of the single-str<strong>and</strong> nucleotides, their impact<br />

upon structure <strong>and</strong> stability of the duplex were analyzed. The duplex variants with altered<br />

370


19.2 Site-specific NMR spectroscopy of chemically synthesized RNA duplexes<br />

stem sequence were also investigated. NMR studies were complemented by recording the<br />

UV melting profiles [31] which permitted the determination of thermodynamic parameters.<br />

Sufficient amounts of cations are necessary for the formation of correct secondary<br />

<strong>and</strong> tertiary structures as well as for the stabilization of nucleic acid helices. In particular,<br />

divalent cations, especially Mg 2+ , play an important role in folding RNA. Consequently, the<br />

question arises whether short helical duplexes are also capable of creating specific binding<br />

sites for divalent cations. We addressed this problem by using paramagnetic manganese ions<br />

which give rise to line broadening of the proton resonances which are located in the vicinity<br />

of the bound ion [32, 33]. The dependence of the specific ion binding capability on the sequence<br />

context was also analyzed [33].<br />

All ribooligonucleotides described in the following were chemically synthesized using<br />

the H phosphonate method [34, 35]. It was demonstrated that this method allows the preparation<br />

of the relatively large amounts (> 10 mg) of RNA being necessary for 2D NMR studies<br />

at reasonable costs <strong>and</strong> with the required purity. In addition, as compared to phosphoramidite<br />

chemistry, the H phosphonate chemistry offers the advantage of recovery <strong>and</strong> reusing<br />

of synthons. This is particularly important for expensive modified or isotopically labelled<br />

nucleotides.<br />

19.2.1 Stability of tRNA-derived acceptor stem duplexes<br />

By checking the imino resonance region of 1 H NMR spectra of nucleic acids information<br />

about the secondary structure can be obtained quickly. Sufficiently narrow resonance lines<br />

will be observed only if stable hydrogen bonds between the paired bases are formed. In that<br />

case the exchange of the protons of the endocyclic base nitrogens with the surrounding<br />

water is drastically slowed down as compared to the non-paired state. Fast exchange of the<br />

imino protons with the water protons gives rise to line broadening which increases with<br />

growing exchange rate. Above a certain limit (several hundreds of Hz, corresponding to exchange<br />

rates above 10 3 s –1 ) these signals become too broad to be detected any more at the<br />

typical signal-to-noise ratios [36, 37].<br />

Usually, a separate imino signal is observed for each regular, stable Watson-Crick<br />

base pair. This resonance may be, however, strongly broadened due to the enhanced base<br />

pair opening rate for terminal base pairs ( fraying), <strong>and</strong> is often not detectable at moderately<br />

high temperatures. The spectral region of the imino resonances at 10–15 ppm is well separated<br />

from that of the other proton signals as, e. g., aromatic <strong>and</strong> ribose protons. Thus, in certain<br />

cases, an assignment of all imino protons to the individual base pairs of intact tRNAs<br />

was feasible [11–14].<br />

The unusual G-U pair, in contrast to classical Watson-Crick base pairs, gives rise to<br />

two well-separated imino lines since in this case two imino protons are involved in base pair<br />

hydrogen bonding. The G imino resonance is easily detectable since it is shifted strongly upfield<br />

to about 10–11 ppm. Moreover, the two G-U imino protons display an extraordinarily<br />

strong nuclear Overhauser effect (NOE) due to their close spatial proximity.<br />

371


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

Figure 19.1: Sequence of the acceptor stem duplex of E. coli tRNA Ala with the numbering of base pairs<br />

used in the text.<br />

In our studies the influence of the single-str<strong>and</strong>ed terminus (ACCA) on the stability of<br />

the duplex was analyzed. The starting sequence used in these investigations was that of the<br />

acceptor stem of tRNA Ala from E. coli (Fig. 19.1).<br />

Both, 1 H NMR studies <strong>and</strong> thermodynamic analyses, have been performed with the<br />

original (wild-type) sequence <strong>and</strong> with sequence variants having gradually truncated singlestr<strong>and</strong>ed<br />

regions, up to the duplex without unpaired nucleotides [32]. In the following the<br />

corresponding sequences are referred to as 18mer, 17mer, etc.<br />

A comparison of the imino region of the 1 H NMR spectra of the five different duplex<br />

variants reveals that after attachment of the first unpaired nucleotide (A73) onto the C72 of<br />

the first base pair (G1-C72) the imino resonance of G1 is substantially shifted as compared<br />

to the 14mer. This is also valid for the imino lines of base pairs 2 (G2-C71), <strong>and</strong> – though<br />

clearly less pronounced – for the G imino resonance of base pair 3 [32]. This upfield shift is<br />

increased upon attachment of further nucleotides of the single-str<strong>and</strong> terminus with even the<br />

fourth nucleotide (A76) causing a measurable upfield shift of the imino resonances of base<br />

pairs 1 <strong>and</strong> 2 (Fig. 19.2).<br />

Figure 19.2: Upfield shift of the imino resonances for base pairs 1 (y), 2 (o), 3G (_), <strong>and</strong> 3U (V)<br />

upon stepwise attachment of 3' nucleotides onto the base pair 1 of the 14mer at 4 8C. For comparison,<br />

the upfield shifts are given for the variants with an IU base pair at position 3 (solid symbols).<br />

372


19.2 Site-specific NMR spectroscopy of chemically synthesized RNA duplexes<br />

The upfield shifts are reflected also in a quite similar dependence of melting temperature<br />

T m <strong>and</strong> Gibbs free energy DG of duplex formation (Tab. 19.1) on the number of singlestr<strong>and</strong><br />

nucleotides.<br />

Both, melting temperature <strong>and</strong> Gibbs free energy, systematically increase with rising<br />

number of single-str<strong>and</strong> nucleotides at the 3'-terminus. The largest stabilization increment is<br />

thereby contributed already by the attachment of the first 3'-nucleotide (A73) though further<br />

single-str<strong>and</strong> nucleotides, too, affect the total duplex stability to a measurable degree. The<br />

remarkable imino line narrowing of the first base pair upon attachment of the 3'-A73 can be<br />

explained by the stabilizing effect of this nucleotide for both, the individual base pair <strong>and</strong><br />

the duplex, as a whole. This nucleotide is by no means as mobile <strong>and</strong> disordered as it is suggested<br />

by the frequently used term dangling nucleotide. Rather it is stacked almost regularly,<br />

i. e. well ordered above the base pair that terminates the duplex stem.<br />

Table 19.1: Thermodynamic parameters for duplex formation <strong>and</strong> activation enthalpies for imino proton<br />

exchange of base pair 2. T m is the melting temperature, H a (2) the activation enthalpy for the exchange of<br />

the G-N1H of base pair 2 with water, <strong>and</strong> G 0 the Gibbs free energy for duplex formation at 37 8C. Typical<br />

experimental errors are +0.5 8C for T m , +2% for DG 0 , <strong>and</strong> +5% for H a (2) . Single-str<strong>and</strong> concentrations<br />

in the melting curve analyses were always 3 mM.<br />

Duplex<br />

(2)<br />

DH a T m<br />

DG8<br />

[kJ/mol] [8C] [kJ/mol]<br />

14mer 223 35.9 –33.4<br />

without dangling end<br />

15mer 266 43.3 –41.0<br />

( 3' A)<br />

16mer 256 44.6 –41.5<br />

( 3' CA)<br />

17mer 295 45.6 –43.3<br />

18mer 301 45.9 –43.9<br />

( 3' ACCA)<br />

18mer/5mM 338 50.8 –48.9<br />

MgCl 2<br />

By overlap of the geometrically more extended purine base over the preceding pyrimidine<br />

base of the same str<strong>and</strong> (C72) with the purine base (G1) of the opposite str<strong>and</strong>, both<br />

nucleotides of the terminal base pair of the stem are locked together more strongly than<br />

without this non-paired nucleotide. In this way the overall duplex stability is enhanced, <strong>and</strong><br />

at the same time the opening rate of the terminal base pair <strong>and</strong> consequently the imino proton<br />

exchange rate of the G1 are considerably reduced.<br />

The upfield shift, too, is directly related to the ordered stacking of the single-str<strong>and</strong><br />

nucleotide upon the terminal base pair. The ring current effects in the conjugated ring systems<br />

of the bases produce strong shielding of the protons in the case of well ordered stacked<br />

nucleotides [38].<br />

373


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

A possible explanation of the incremental upfield shift is based on the assumption that<br />

with rising numbers of single-str<strong>and</strong> nucleotides the stacking order of the preceding<br />

(5'-neighbouring) nucleotides are successively enhanced. This enhanced stacking order, which<br />

can be visualized by an increase of the residence probability in the ideally stacked state, also<br />

affects the terminal base pairs of the stem <strong>and</strong> the stability of the duplex as a whole.<br />

Another parameter that reflects the stability of individual base pairs <strong>and</strong>, to a certain<br />

degree, of the whole duplex is the activation enthalpy of the imino proton exchange. It can<br />

be derived from the temperature dependence of the imino signal line widths, under the assumption<br />

of an Arrhenius-like thermal activation of this exchange process. Such an analysis<br />

has been carried out for several imino resonances that were separable <strong>and</strong> which could be<br />

detected over a sufficiently broad temperature range. Of particular interest in this context<br />

was the imino proton exchange of base pair 2 which is, on the one h<strong>and</strong>, not too far from<br />

the single-str<strong>and</strong> terminus to be totally unaffected by it. On the other h<strong>and</strong>, however, it is<br />

even without single-str<strong>and</strong> nucleotides stable enough. Moreover, it is directly adjacent to the<br />

G3-U70 wobble pair.<br />

The determined activation enthalpies (Tab. 19.1) display the same dependence on the<br />

number of single-str<strong>and</strong> nucleotides as melting temperature, Gibbs free energies, <strong>and</strong> chemical<br />

shift changes. It has to be concluded that the activation enthalpies do not reflect the base<br />

pair opening behaviour <strong>and</strong> correspondingly the stability of one individual base pair, but<br />

rather of several (3–4) neighbouring base pairs at a time. Hence it indicates the cooperative<br />

character of this process for which the sequence context plays an important role.<br />

In all duplexes with G-U <strong>and</strong> I-U pairs (I = inosine) the activation enthalpy for the exchange<br />

of the G <strong>and</strong> I imino proton were distinctly greater than that of the U imino proton.<br />

This could be attributed to an asymmetric opening mechanism of this base pair or different<br />

accessibilities of the concerned imino protons for the phosphate ions in the buffer that catalyze<br />

the proton exchange [37].<br />

Beyond the duplexes with single-str<strong>and</strong>ed ends of differing length other variants with<br />

altered sequences in the single-str<strong>and</strong> <strong>and</strong> in the stem have been studied [32]. Most of these<br />

variations resulted in reductions of duplex stability as compared to the original sequence. As<br />

expected, replacement of the G3-U70 with a G3-C70 base pair yielded an increased duplex<br />

stability. Surprisingly, a duplex with an I3-U70 is substantially less stable than the duplex containing<br />

G3-U70, because I differs from G only by a lacking 2-amino group. This seems to indicate<br />

an involvement of the NH 2 group in the stabilization of the duplex structure, possibly<br />

via the formation of specific hydrogen bonds with an immobilized water molecule that has<br />

been found in an X-ray structural analysis of an RNA duplex containing a G-U pair [39].<br />

19.2.2 Manganese ion binding sites at RNA duplexes<br />

In an X-ray analysis of the yeast tRNA Phe , a specific magnesium-binding site in the vicinity<br />

of a G4-U69 wobble base pair was found. Thus it seemed attractive to study if a G-U pair is<br />

a specific binding site for magnesium ions. Addition of the diamagnetic Mg 2+ cation has<br />

only little, though measurable, effects on the chemical shifts of the imino resonances<br />

(< 0.05 ppm). Since it is known from a number of biological <strong>and</strong> biochemical studies that<br />

374


19.2 Site-specific NMR spectroscopy of chemically synthesized RNA duplexes<br />

Mg 2+ ions can be replaced by manganese ions, Mn 2+ has been employed as a probe for identification<br />

of metal binding sites in RNA duplexes. The paramagnetic manganese causes line<br />

broadening of resonances originating from protons in the vicinity of the bound Mn 2+ . The<br />

observed line width increase varies, under certain conditions, approximately as r –6 with r<br />

being the distance between paramagnetic ion <strong>and</strong> the concerned proton [40].<br />

Indeed, specific imino line broadenings were found upon addition of manganese ions<br />

for the tRNA Ala acceptor stem duplexes with <strong>and</strong> without ACCA terminus (Fig. 19.3).<br />

Figure 19.3: Effect of the addition of Mn 2+ on the imino proton resonance signals at 4 8C of the duplex<br />

without a single-str<strong>and</strong>ed end (14mer) (a) <strong>and</strong> the duplex with an ACCA 3'-terminus at base pair 1<br />

(18mer) (b). In each case, the lower trace belongs to the solution without Mn 2+ <strong>and</strong> the upper trace<br />

gives the spectrum after addition of either 7.5 mM (a) or5mM (b) MnCl 2 . The solutions contain 100<br />

mM NaCl <strong>and</strong> either 7.5 mM (a) or 5 mM (b) MgCl 2 .<br />

It should be noted that the molar ratio of manganese ions to RNA was 1:150 to<br />

1:200. From the appearance of the observed spectra it must be concluded that a fast exchange<br />

on the NMR time scale occurs with rates for the exchange of ions among different<br />

RNA molecules of more than 10 3 s –1 . Thus, the binding is not a strong one. Nevertheless,<br />

the particularly distinct broadenings of the imino resonance of base pair 2 <strong>and</strong> the U imino<br />

line of base pair 3 (G3-U70) indicate that the manganese ion is preferably found between<br />

base pairs 2 <strong>and</strong> 3. In conjunction with the information obtained from two-dimensional nuclear<br />

Overhauser spectroscopy (NOESY) experiments (Section 19.2.3.) the most probable<br />

binding site of the manganese ion was located in the major groove between the bases G2<br />

<strong>and</strong> G3 of the short str<strong>and</strong>.<br />

These manganese ion-binding studies have been extended by analyzing several other<br />

RNA duplexes that were derived from tRNA acceptor stems [33]. These included sequence<br />

variants of the E. coli tRNA Ala (with G3-C70 <strong>and</strong> I3-U70 base pairs), E. coli tRNA Gly ,<br />

tRNA His acceptor duplexes, as well as yeast tRNA Phe , <strong>and</strong> tRNA Asp acceptor stems.<br />

375


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

As mentioned before, tRNA Phe from yeast also has a G4-U69 wobble pair in its acceptor<br />

arm in the vicinity of which a magnesium binding site was determined in the native<br />

tRNA. With the chemically synthesized acceptor duplex line broadening effects upon manganese<br />

ion addition were observed (Fig. 19.4) which are most strongly pronounced for the<br />

imino proton of the third <strong>and</strong> the U imino proton of the fourth (G4-U69) base pair. The<br />

ranking of the specific broadenings corresponds fully to the pattern that also has been observed<br />

with the tRNA Ala duplex if the G-U pair is chosen as a reference point, i. e. largest<br />

line width increase for the G imino proton of the 5'-neighbour of the guanosine of the G-U<br />

pair, followed by the U imino proton of the G-U pair <strong>and</strong> so on [33].<br />

1<br />

5<br />

6<br />

2<br />

4(U)<br />

4(G)<br />

A<br />

C<br />

C<br />

A<br />

1 G-C<br />

2 C-G<br />

3 G-C<br />

4 G-U<br />

5 A-U<br />

6 U-A<br />

7 U-A<br />

14<br />

13 12 11<br />

ppm<br />

Figure 19.4: Imino region of the H NMR spectra of the acceptor arm duplex derived from yeast<br />

tRNA Phe before (lower trace) <strong>and</strong> after (upper trace) addition of 7.5 mM MnCl 2 (RNA concentration<br />

1.65 mM, 7.5 mM MgCl 2 ) at 277 K.<br />

It could be concluded that G-U pairs indeed have a potential for specific binding of<br />

divalent cations. However, similar broadening effects for the tRNA Ala acceptor stem variant<br />

with a G3-C70 instead of the G3-U70 base pair were also observed. There the largest line<br />

width increase is likewise observed for the G imino proton of base pair 2, followed by a<br />

slightly smaller broadening for the G3 imino resonance [33]. Analysis of several further duplexes<br />

led to the conclusion that stretches of four consecutive purines also create sites of<br />

preferred binding of manganese ions.<br />

376


19.2 Site-specific NMR spectroscopy of chemically synthesized RNA duplexes<br />

19.2.3 Structural determination of short RNA duplexes by 2D NMR spectroscopy<br />

Though one-dimensional 1 H NMR studies of the tRNA Ala acceptor arm gave hints to the special<br />

role of the G-U base pair, detailed information on the possible modifications of the local<br />

helical geometry in the vicinity of the G-U pair cannot be obtained by these experiments.<br />

Elucidation of the fine structure of biological macromolecules at atomic resolution is<br />

possible by multi-dimensional NMR [4, 10, 41–43].<br />

Two-dimensional NOESY allows detection of spatial proximity between protons that<br />

are separated by less than 4.5 Å. By the correlated spectroscopy method (COSY) scalar couplings<br />

between protons, which are separated by at most three (in some exceptions four) chemical<br />

bonds, are used. In the case of ribooligonucleotides such couplings consequently can<br />

only be observed between the protons of the ribose rings <strong>and</strong> the C5 <strong>and</strong> C6 protons of the<br />

pyrimidine bases of a given nucleotide. The NOESY method, however, permits the detection<br />

of contacts between the protons of different nucleotides.<br />

For molecules with a regular secondary structure, like RNA duplexes, there are numerous<br />

characteristic interproton distances below 5 Å between adjacent nucleotides [41, 42]<br />

which give rise to typical cross peak patterns in the NOESY spectra. For a regular A-RNA,<br />

the distance between the aromatic proton (C6H or C8H) of a given nucleotide <strong>and</strong> the 2'-ribose<br />

proton of the 3'-neighboured nucleotide, as well as the one between the aromatic proton<br />

<strong>and</strong> the 3'-ribose proton of the same residue is particularly small (ca. 2.0–2.5 Å). The cross<br />

peaks in the NOESY spectrum caused by the interaction between the corresponding protons<br />

are therefore especially strong.<br />

For mixing times larger than 200 ms contacts between protons which are more distant<br />

than 4.5 Å are detectable. This is due to a process that is referred to as spin diffusion in<br />

which NOE contacts are mediated by an additional proton which acts as a kind of relay station.<br />

In this way, NOE contacts between the 1'-ribose protons <strong>and</strong> the aromatic protons<br />

(C6H/C8H) of both, the same <strong>and</strong> the 3'-neighboured nucleotide, can be observed. Thus,<br />

complete assignment paths between 1' <strong>and</strong> aromatic protons of all the residues within one<br />

str<strong>and</strong> can be drawn as long as the helical geometry is not perturbed. There are also contacts<br />

between protons of nucleotides in the two different str<strong>and</strong>s of the duplex, e. g. between C2H<br />

of an adenosine <strong>and</strong> the 1'-ribose proton of the 3'-neighboured nucleotide of the opposite<br />

str<strong>and</strong> [4].<br />

Making use of these techniques as well as of 13 C <strong>and</strong> 1 H hetero-COSY experiments an<br />

assignment of the aromatic protons of C5H, C6H, C8H, C2H, <strong>and</strong> all 1', 2', 3', <strong>and</strong> part of<br />

4'-ribose protons of the tRNA Ala acceptor stem was achieved (Fig. 19.5) [44].<br />

In the spectra at low mixing time (80 ms) as well as in the frame of the signal-to-noise<br />

ratio, which was attainable with the existing equipment <strong>and</strong> the available sample concentrations,<br />

only those cross peaks were found that are particularly strong for A-RNA helices, as<br />

2'-ribose-aromatic interresidue cross peaks, beyond the contacts which always give rise to intense<br />

cross peaks independent of helix type (1'-2'-ribose NOEs, 5–6 pyrimidine contacts).<br />

From these results, a regular helical structure was deduced.<br />

At longer mixing times many further contacts become detectable due to spin diffusion<br />

that cannot be used for a quantitative evaluation in terms of proton distances. However, these<br />

contacts allow a qualitative comparison of equivalent interactions – as e. g. 5–5, 6–6, <strong>and</strong><br />

5–6 contacts – between different adjacent nucleotides.<br />

377


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

Figure 19.5: Schematic representation of the observed internucleotide NOESY contacts in the E. coli<br />

tRNA Ala derived acceptor arm duplex (18mer/GU) for mixing time of 300 ms at 303 K.<br />

In Fig. 19.5 all the NOE contacts between the protons of adjacent nucleotides detected<br />

for 300 ms mixing time NOESY spectra are compiled.<br />

The number of the internucleotide contacts, visualized in Fig. 19.5 by connecting<br />

lines, gives an approximate impression of the stacking between the individual bases. There<br />

is a particularly efficient stacking between nucleotides G2 through U6 in the purine-rich<br />

(short) str<strong>and</strong> as well as between G68 <strong>and</strong> C69, C71 <strong>and</strong> C72, <strong>and</strong>, remarkably, between<br />

A73, C74 <strong>and</strong> C75 of the longer str<strong>and</strong>. Obviously there is poorer stacking between U70 <strong>and</strong><br />

C71 than between most of the other adjacent nucleotides.<br />

The small number of the observed internucleotide NOEs between C72 <strong>and</strong> A73 suggests<br />

a weak stacking. However, there are cross peaks of the C2 proton of A73 with both,<br />

the aromatic C8H <strong>and</strong> the 1'-ribose proton of G1 of the opposite str<strong>and</strong>. The former cross<br />

peaks indicate an overlap of the purine base plane of A73 with that of G1, which corroborates<br />

the before-mentioned interlocking effect of the purine base in position 73 <strong>and</strong> provides<br />

a structural explanation for the stabilizing effect of 3'-dangling purine nucleotides. At the<br />

same time, the C8 proton of A73 is moved away from the aromatic protons of C72 making<br />

the lack of the aromatic-aromatic cross peaks conceivable. Some NOE contacts have been<br />

378


19.2 Site-specific NMR spectroscopy of chemically synthesized RNA duplexes<br />

observed between C75 <strong>and</strong> the terminal A76, suggesting at least a partial stacking of A76<br />

upon C75. Thus it can be concluded that the stacking order within the single-str<strong>and</strong> terminus<br />

is comparable to the one in the interior of the duplex.<br />

19.2.4 NMR derived model of the tRNA Ala acceptor arm<br />

As mentioned above, the cross peak intensities from NOESY spectra taken at long mixing<br />

times cannot be related in a simple <strong>and</strong> direct way to distances between two protons due to spin<br />

diffusion effects that mask the actual proton distances. A possibility to extract such information<br />

is provided by relaxation matrix analysis that accounts for all dipolar interactions of a given<br />

proton <strong>and</strong> hence takes spin diffusion effects explicitly into consideration. Several computational<br />

procedures have been developed which iteratively back-calculate an experimental<br />

NOESY spectrum, starting from a certain molecular model that is altered in many cycles of the<br />

iteration process to fit best the experimental NOESY data. In each cycle, the calculated structures<br />

are refined by restrained molecular dynamics <strong>and</strong> free energy minimization [42, 43].<br />

We have performed an approximate calculation of the duplex structure of the tRNA Ala acceptor<br />

arm on the basis of the 300 ms NOESY spectrum using the iterative relaxation matrix<br />

analysis procedure IRMA [45] of the FELIX software package (Biosym Technologies, San<br />

Diego). The structures were refined employing the DISCOVER module of the INSIGHT II molecular<br />

graphics software package (Biosym Technologies). As a starting structure, a canonical<br />

A-RNA duplex was used. If uracil (U70) is substituted for the C70 of the regular Watson-Crick<br />

duplex with a G3-C70 base pair, there is a distinct base plane overlap not only between the purines<br />

(G4 through G2) but also between the pyrimidines (C69 to C71) (Fig. 19.6 a).<br />

The arrangement calculated for a helix with G3-U70 base pair is given in Fig. 19.6 b.<br />

The most striking difference between the Structures 19.6a <strong>and</strong> 19.6b is that in the latter a<br />

distinct destacking is found between the bases U70 <strong>and</strong> C71. Moreover, a regular G-U hy-<br />

Figure 19.6: Arrangement of the three consecutive base pairs G2-C71, G3-U70, <strong>and</strong> G4-C69; a) in the<br />

starting structure (as created with the BIOPOLYMER module of INSIGHT II) obtained from a regular<br />

Watson-Crick A-RNA duplex with a G3-C70 base pair after subsequent replacement of the C70 residue<br />

by a uridine. No further adjustment of the U70 base position has been made <strong>and</strong> hence no correct hydrogen<br />

bonding pattern is found; b) in the final (average) structure derived from relaxation matrix analysis<br />

of the 300 ms NOESY spectrum using the IRMA procedure [45] <strong>and</strong> restrained molecular dynamics<br />

calculation. Note that G3 <strong>and</strong> U70 now are forming a regular base pair. The bases G3 <strong>and</strong><br />

U70 are indicated by thick lines.<br />

379


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

Figure 19.7: Model of the tRNA Ala acceptor arm structure as determined on the basis of the 300 ms<br />

NOESY spectrum using the IRMA procedure [45] <strong>and</strong> restrained molecular dynamics. Three structures<br />

resulting from different calculations are superimposed. The continuation of the helix geometry into the<br />

single-str<strong>and</strong>ed terminus is clearly visible. The black sphere marks the most probable location of the<br />

bound manganese ion in the vicinity of G3-U70 base pair.<br />

drogen bond geometry for the paired bases is found though this has not been implemented<br />

as a constraint into the calculation. Figure 19.7 displays a superposition of three duplex<br />

structures, obtained for different independent calculations each starting from the same regular<br />

A-type helix. The rmsd value between each two structures amounts to ca. 1.35 Å.<br />

19.2.5 Chemical shifts <strong>and</strong> scalar coupling as an indicator of RNA structure in the<br />

vicinity of a G-U pair<br />

If the chemical shift differences of the equivalent 1'-ribose <strong>and</strong> aromatic protons of the wildtype<br />

G3-U70 sequence <strong>and</strong> the G3-C70 variant sequence are plotted versus the nucleotide<br />

position, only small values (< 0.05 ppm) are found except at the wobble pair site (3–70) or<br />

in its immediate vicinity (Fig. 19.8). The difference of the chemical shift values of the C5<br />

380


19.2 Site-specific NMR spectroscopy of chemically synthesized RNA duplexes<br />

Figure 19.8: Differences Dd in ppm between the chemical shifts in the 18mer/GU <strong>and</strong> the 18mer/GC<br />

duplex for aromatic <strong>and</strong> H1' proton resonances at T = 303 K, given separately for both str<strong>and</strong>s.<br />

protons of U70 <strong>and</strong> C70 amounts to 0.22 ppm. The one for the difference of the C5 protons<br />

of C71 is similarly large. If the intrinsic chemical shift differences between cytidine <strong>and</strong> uridine<br />

are taken into account, the correct shift difference of position 70 is increased to ca.<br />

0.4 ppm.<br />

Since the upfield shifts of the resonance signals (in particular of the pyrimidine C5<br />

protons) of nucleotides upon transition from the totally structurally disordered coil to the<br />

well ordered duplex state are caused by the ring current effects of the stacked bases the<br />

downfield shift of U70 <strong>and</strong> C71 proton signals in the G3-U70 duplex can be interpreted in<br />

terms of a reduction of stacking interaction between U70 <strong>and</strong> C71, as compared to the regular<br />

G3-C70 duplex.<br />

The results presented above demonstrate that there is a deviation from the regular A-<br />

helical geometry in the wild-type tRNA Ala acceptor arm that is mainly characterized by a<br />

displacement of the U70 base. Thereby the base plane overlap between C71 <strong>and</strong> U70 is dis-<br />

381


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

tinctly diminished in comparison to a regular Watson-Crick duplex with a G3-C70 base pair<br />

in place of the G3-U70 wobble pair. Probably, the discontinuity in the helix geometry introduced<br />

by the wobble pair is recognized by the ARS. Exactly this has been suggested by<br />

McClain et al. [24] from in vivo studies using suppressor tRNAs in E. coli.<br />

19.2.6 Structure of aminoacyl-tRNA <strong>and</strong> transacylation of the aminoacyl residue<br />

Esterification of the amino acid with the tRNA occurs either at the 2' or 3'-hydroxyl group<br />

of the terminal adenosine residue A76, according to the specificity of the corresponding<br />

aminoacyl-tRNA synthetase [15]. After dissociation of the aminoacyl-tRNA from the<br />

synthetase, the amino acid residue can transacylate between the 2' <strong>and</strong> 3' positions. The rate<br />

of this process, which correlates with the deacylation rate of the amino acid from the tRNA,<br />

depends on temperature <strong>and</strong> pH value as well as on the specific type of the attached amino<br />

acid [46]. Methionine that is found at the N-terminus of polypeptides often is formylated in<br />

procaryotes. This formylation takes place after the esterification of the corresponding<br />

tRNA fmet with methionine.<br />

A way to study the effect of formylation on Met-tRNA structure was opened by specifically<br />

labelling the amino acid with 13 C. In the 13 C NMR spectrum of the yeast<br />

[e- 13 CH 3 ]Met-tRNA i Met a single 13 C resonance is observed which corresponds to the methyl<br />

group. This indicates a rapid (on the NMR time scale) exchange of the amino acid residue<br />

between the 2' <strong>and</strong> 3' acylation positions (Fig. 19.9 a). A quite similar behaviour is found if<br />

E. coli tRNA fMet is aminoacylated with [e- 13 CH 3 ]Met (Fig. 19.9 b).<br />

By contrast, formylation of the methionine gives rise to two separate resonances in the<br />

spectrum of [e- 13 CH 3 ]fMet-tRNA fMet with roughly equal intensity that can be assigned to<br />

the 2' <strong>and</strong> 3' isomers of the aminoacyl-tRNA, respectively (Fig. 19.9c). Thus, formylation<br />

leads to a drastic reduction of the transacylation rate permitting the detection of signals for<br />

2' <strong>and</strong> 3' acylation isomers. From the frequency splitting of the two lines (0.35 ppm) it can<br />

be concluded that the transacylation rate in this case has to be less than ca. 150 s –1 .<br />

By analyzing the fMet-adenosine monophosphate, additional information on the transacylation<br />

process was obtained. This molecule corresponds to the aminoacylated 3'-terminal<br />

A76 of the fMet-tRNA fMet . Owing to its low molecular weight <strong>and</strong> the small number of protons<br />

the resolution of the proton NMR spectrum is naturally much better than for the whole<br />

tRNA. Indeed, the chemical shifts for all protons of the 2' <strong>and</strong> 3' isomerss can be detected.<br />

From the smallest measured chemical shift difference between the lines corresponding to 2'<br />

<strong>and</strong> 3' isomers (0.0073 ppm for the formyl proton) an upper limit for the transacylation rate<br />

in fMet-AMP of ca. 11.5 s –1 can be deduced. In contrast to fMet-tRNA fMet , the ratio of the<br />

3' <strong>and</strong> 2' isomers populations is here about 1.9 : 1. However, in native tRNA the isomer ratio<br />

<strong>and</strong> the transacylation rate could be different.<br />

A similar experiment as with Met-tRNA was performed with 13 C Val-tRNA Val . In this<br />

case, the amino acid was 13 C labelled at the carbonyl position of the valine. Here, likewise,<br />

two separate carbonyl resonances have been observed that corresponded to the 2' <strong>and</strong> 3' acylation<br />

positions [47]. The intensity ratio of 3' <strong>and</strong> 2' isomers amounts to ca. 7 : 3. The slow<br />

transacylation correlates well with the very slow deacylation rate of Val-tRNA Val .<br />

382


19.3 Structure of elongation factor Tu<br />

Figure 19.9: 125.7 MHz proton-decoupled 13 C NMR spectra a) of yeast [e-CH 3 - 13 C]Met-tRNA Met<br />

i ,<br />

b) E. coli [e-CH 3 - 13 C]Met-tRNA fMet , <strong>and</strong> c) E. coli [e-CH 3 - 13 C]fMet-tRNA fMet .2' <strong>and</strong> 3' isomers can<br />

be distinguished only after formylation of the a-amino group due to the decrease in the transacylation<br />

rate.<br />

19.3 Structure of elongation factor Tu<br />

Bacterial elongation factor Tu (EF-Tu) is a GTPase which promotes the binding of aminoacyl-tRNA<br />

to the codon-programmed ribosomes. During its functional cycle it consecutively<br />

binds several lig<strong>and</strong>s: GDP, elongation factor Ts, GTP, aminoacyl-tRNA, <strong>and</strong> ribosomes<br />

[48]. It functions as a timing device, determining the relation between the rate of protein<br />

biosynthesis <strong>and</strong> its error frequency [49].<br />

The cycle of EF-Tu during protein biosynthesis, emerging from the studies on<br />

T. thermophilus system [2] conducted in our laboratory, is shown in Fig. 19.10. EF-Tu 7 GDP<br />

binds EF-Ts <strong>and</strong> a tetrameric complex (EF-Tu 7 EF-Ts) 2 is formed. The dissociation constant<br />

for this interaction is about 10 –8 M, in the case of T. thermophilus elongation factors. The<br />

functional reason for the formation of an (EF-Tu 7 EF-Ts) 2 tetramer, also observed with other<br />

elongation factors, is not yet understood. EF-Ts functions as nucleotide exchange factor for<br />

EF-Tu since it promotes dissociation of GDP. Aminoacyl-tRNA in the presence of GTP is required<br />

for dissociation of T. thermophilus (EF-Tu 7EF-Ts) 2 <strong>and</strong> aminoacyl-tRNA7EF-<br />

Tu 7 GTP ternary complex is formed. The high concentration of GTP (10 –3 M) <strong>and</strong> aminoacyl-tRNA<br />

(10 –4 M) in the bacterial cell is the main factor which drives the cycle of EF-Tu<br />

383


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

Figure 19.10: Functional cycle of EF-Tu. The cycle is driven by GDP to GTP exchange which is catalyzed<br />

by EF-Ts <strong>and</strong> GTPase, stimulated by programmed ribosome with A-site bound aminoacyl-tRNA. EF-<br />

Tu 7EF-Ts complex occurs in vitro as (EF-Tu7EF-Ts) 2 tetramer. This fact is not indicated in the figure.<br />

toward the formation of aminoacyl-tRNA7EF-Tu 7 GTP ternary complex. The aminoacyltRNA<br />

is transported to ribosomal A-site where its codon-dependent binding to mRNA-programmed<br />

ribosomes occurs. Correct reading of the codon by the aminoacyl-tRNA induces a<br />

GTPase which is required for the dissociation of EF-Tu 7GDP from ribosomes [50].<br />

19.3.1 Sequence of Thermus thermophilus EF-Tu<br />

The sequence of T. thermophilus EF-Tu was determined by gene analysis <strong>and</strong> DNA sequencing<br />

[51]. The regulatory GTPases, like other proteins, are composed of structural modules, which we<br />

will call domains. GTP is bound to a nucleotide binding domain containing conserved sequences<br />

which are present in all regulatory GTPases (Fig. 19.11). These sequences GlyXXXXGly-<br />

LysThr 25 , (Arg)XXThr 62 ,AspXXGly 84 ,AsnLysXAsp 139 , <strong>and</strong> SerAla 175 fulfill a specific function<br />

in nucleotide binding, GTP hydrolysis, <strong>and</strong> regulation of the conformational switch of EF-Tu.<br />

The L 2 region of the nucleotide binding domain (Fig. 19.11), connecting the C-terminus<br />

of the helix A with the N-terminus of the antiparallel b sheet b, was named the effector<br />

loop since it changes its structure between GDP <strong>and</strong> GTP forms of the G protein [52]. Most<br />

mutations inhibiting the interaction with the GTPase activating protein (GAP) are located<br />

within this L 2 region [53].<br />

In most GTP/GDP binding elongation factors, the C-terminal part of L 2 contains an<br />

ArgXXThr 62 consensus sequence. The arginine59 of this consensus sequence in T. thermophilus<br />

EF-Tu is a trypsin hypersensitive site. The threonine62 is one of the residues coordinating<br />

the magnesium ion <strong>and</strong> the g-phosphate of the EF-Tu bound GTP. There is less sequence<br />

homology in the N-terminus of L 2 than in the C-terminus of this loop (Fig. 19.11).<br />

The N-terminal part of L 2 in EF-Tu probably interacts with aminoacyl-tRNA <strong>and</strong> ribosome<br />

[2]. A remarkably high homology between different bacterial EF-Tus exists in the region<br />

around helix B (amino acid residues 83–106), where the catalytic site for GTPase is located.<br />

The GTPase-dependent conformational changes of EF-Tu are propagated from this highly<br />

conserved region [1].<br />

384


19.3 Structure of elongation factor Tu<br />

Figure 19.11: Sequence homology in eubacterial EF-Tu. The bold capitals above the line show the<br />

homologous sequences occurring in all regulatory GTPases including G-proteins <strong>and</strong> proteins of ras family.<br />

The capitals above the line give the amino acid residues conserved in all bacterial EF-Tus, derived<br />

from 36 known sequences. The numbers indicate the position of these residues in Thermus thermophilus<br />

EF-Tu. The secondary structure elements of T. thermophilus EF-Tu are shown below the line. Open boxes<br />

(a–r) are b structures <strong>and</strong> the hatched areas are a-helices. The alternative black area of the helix B occurs<br />

in the EF-Tu7GDP structure <strong>and</strong> indicates the change in the position of this element upon regulatory<br />

switch from GDP to GTP bound conformation [1].<br />

385


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

There are many conserved aminoacyl residues in domains II <strong>and</strong> III. The function of<br />

most of these residues is not yet known. Arginine300 [54] of domain II apparently plays an<br />

important role in stabilization of interdomain interactions <strong>and</strong> threonine394 was suggested<br />

as a phosphorylation site regulating aminoacyl-tRNA binding to EF-Tu 7GTP [55]. The<br />

amino acid residues on the interface between the domains I <strong>and</strong> II are often conserved <strong>and</strong><br />

are important targets for binding antibiotics.<br />

19.3.2 Crystallization, X-ray analysis, <strong>and</strong> the tertiary structure<br />

The crystallization of native elongation factor Tu isolated from bacterial cells, i. e. in complex<br />

with GDP, proved to be difficult. Useful crystals for X-ray structure analyses could be<br />

obtained only from a partially trypsinized E. coli EF-Tu 7 GDP missing a part of the L 2 region<br />

(amino acid residues 45–58 E. coli numbering). Through laborious efforts of several<br />

crystallographic groups, the structure of the partially proteolyzed E. coli EF-Tu 7 GDP has<br />

been determined up to 2.5 Å resolution [56].<br />

The structure of EF-Tu in the active GTP form, namely that of EF-Tu7GppNHp from<br />

T. thermophilus, was determined up to 1.7 Å resolution [1]. The EF-Tu structures in GDP<br />

<strong>and</strong> GTP bound forms can now be compared (Fig. 19.12). However, we still have to keep in<br />

Figure 19.12: Structures of E. coli EF-Tu7GDP [56] <strong>and</strong> T. thermophilus EF-Tu7GppNHp [1]. The nucleotide<br />

binding domains were placed at approximately the same location in order to demonstrate the<br />

movements of the helix B <strong>and</strong> the domains II/III (dark lines). The L 2 region connecting the helix A with<br />

the antiparallel b sheet b, c is missing in the EF-Tu7GDP structure (arrows).<br />

386


19.3 Structure of elongation factor Tu<br />

mind that the structure of EF-Tu7GDP is not derived from the native protein but from its<br />

proteolytic derivative.<br />

EF-Tu is composed of three domains (Fig. 19.12). Domain I, which harbours the nucleotide<br />

binding site, contains five b structures <strong>and</strong> six a helices (compare with Fig. 19.11).<br />

Domains II <strong>and</strong> III are composed exclusively of antiparallel b sheets forming two b barrels.<br />

The most striking difference between EF-Tu 7 GDP <strong>and</strong> EF-Tu 7 GppNHp is the relative location<br />

of the three domains. In EF-Tu 7 GDP, domain II is separated from domain I, whereas<br />

the EF-Tu7GppNHp structure is compact with an interface formed between all three domains.<br />

Large intramolecular movement must occur during transition from the GDP to the<br />

GTP bound form. Since the interface between domains II <strong>and</strong> III does not change during<br />

this dramatic conformational change, a movement of domain I takes place involving distances<br />

which are more than one third of the molecular diameter. The GDP form of EF-Tu is<br />

stabilized only by interactions between domains I <strong>and</strong> III. These interactions must be disrupted<br />

by going from GDP to the GTP form, a process which does not take place spontaneously<br />

since EF-Ts, aminoacyl-tRNA, <strong>and</strong> GTP are required to promote it (Fig. 19.10).<br />

19.3.3 Nucleotide binding <strong>and</strong> GTPase reaction<br />

The nucleotide interacts with consensus sequences located in domain I (Fig. 19.11). An important<br />

interaction of guanine nucleotides with magnesium ion is coordinated by the b <strong>and</strong><br />

g-phosphates of GTP <strong>and</strong> four other lig<strong>and</strong>s located in the phosphate loop (Thr25), effector<br />

loop (Asp51, Thr62), <strong>and</strong> GTPase loop (Asp81).<br />

An intriguing mechanistic question concerns the mode by which the signal induced by<br />

interaction of a GTPase with GAP is transmitted to the GTPase centre to trigger the hydrolysis<br />

of GTP to GDP. In the case of EF-Tu, the rate of intrinsic GTPase activity is stimulated<br />

100,000-fold by the binding of aminoacyl-tRNA <strong>and</strong> ribosomes [50]. Thus cooperative interaction<br />

of programmed ribosomes <strong>and</strong> the aminoacyl-tRNA7EF-Tu 7GTP complex is essential<br />

for the EF-Tu-GTPase stimulation. Evidence is available which suggests that the L 2 region<br />

of EF-Tu participates in this process. Cleavage of the L 2 region at arginine59 with trypsin<br />

does not affect the binding of GTP, GDP, or the interaction with aminoacyl-tRNA [57]<br />

but abolishes the ability of ribosomes to stimulate the GTPase [58]. Thus it seems likely that<br />

a cut in the L 2 region intercepts the transfer of information from the GTPase site of the protein<br />

to the ribosomal interaction site.<br />

The central feature of the GTPase mechanism in EF-Tu is the catalytic triad composed<br />

of aspartic acid87, histidine85, <strong>and</strong> a catalytic water molecule placed in the vicinity of the g-<br />

phosphate of GTP [1] (Fig. 19.13). This isolated water molecule was found in all structures of<br />

GTPases crystallized with slowly hydrolyzable GTP analogues. The intrinsic GTPase of EF-<br />

Tu is (in the absence of ribosomes <strong>and</strong> aminoacyl-tRNA) probably catalyzed by different mechanisms<br />

than the 100,000-fold stimulated GTPase of the same protein taking place during<br />

the decoding process. Thus the GTPase in EF-Tu is regulated. The action of His85, which acts<br />

as a nucleophile for water activation, is blocked by a lipophilic gate formed by Ile61 <strong>and</strong><br />

Val20 as well as by the involvement of Asp87 in a salt bridge with Arg59. Binding of aminoacyl-tRNA<br />

<strong>and</strong> the ribosomal A-site provides for movement of the effector loop L 2 <strong>and</strong> the<br />

387


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

Figure 19.13: Schematic presentation of the model for GTPase mechanism in EF-Tu.<br />

loosening of these interactions. This leads to the activation of the catalytic triad. The GTPase<br />

is facilitated in addition by interactions involving Mg 2+ <strong>and</strong> several other amino acid residues<br />

(Lys24, Thr62, Gly84) with the pentacoordinated phosphate transition state. This mechanism<br />

is supported by experiments with EF-Tu mutated at His85 <strong>and</strong> Thr62. Their mutations to Ile85<br />

<strong>and</strong> Ala82, respectively, completely abolish the GTPase [54]. It is remarkable that the two residues<br />

of the L 2 region, arginine59 <strong>and</strong> isoleucine61, which seem to prevent the optimal formation<br />

of this catalytic triad, are conserved in all elongation factors Tu. This provides supporting<br />

evidence for their involvement in the GTPase trigger mechanism.<br />

19.3.4 Mechanism of GTP induced conformational change of EF-Tu<br />

The EF-Tu 7 GDP <strong>and</strong> EF-Tu7GTP structures in Fig. 19.12 provide a picture of the dramatic<br />

change caused by the presence of g-phosphate group of GTP in complex with EF-Tu. This<br />

conformational change is propagated from domain I by a flip of the glycine84 peptide bond<br />

<strong>and</strong> results in a large movement of domains II/III. In the GDP structure the NH group of<br />

this glycine residue is located on the N-terminus of helix B of domain I <strong>and</strong> stabilizes this<br />

helix by interaction with the side-chain of aspartate87. The peptide bond between proline83<br />

<strong>and</strong> glycine84 moves by approximately 1508 in the transition between the inactive GDP <strong>and</strong><br />

active GTP forms of EF-Tu. A new interaction between the NH group of glycine84 <strong>and</strong> the<br />

g-phosphate of GTP is formed. Consequently, the first turn of the helix B melts. In addition<br />

helix B changes its location in the structure:<br />

a) it is translated along its axis by one turn,<br />

b) its axis is tilted by about 308, <strong>and</strong> finally<br />

c) a slight turn around its axis takes place [1].<br />

These movements alter the intramolecular interactions of several amino acids in helix<br />

B. For example, asparagine91 interacts, in the GDP bound form of the protein, with cysteine82<br />

of domain I, whereas in the GTP bound form it interacts with arginine300 of domain<br />

II – a residue playing a pivotal role in stabilization of interdomain interaction [59] –<br />

<strong>and</strong> isoleucine63 of the GTPase regulating region of domain I. The centrally located helix B<br />

388


19.3 Structure of elongation factor Tu<br />

(Fig. 19.12) thus functions like a cylinder of a lock <strong>and</strong> is hidden in the open state EF-<br />

Tu 7 GDP. In the locked state EF-Tu 7 GTP it enters the complementary site of the neighbouring<br />

domain II. The moveable parts of this system are the glycines84 <strong>and</strong> 95. Mutation of<br />

glycine83 of E. coli EF-Tu (equivalent to glycine84 in T. thermophilus EF-Tu) interferes with<br />

the conformational change of EF-Tu <strong>and</strong> prevents the interaction of EF-Tu 7GTP with aminoacyl-tRNA<br />

[60]. Helix B with its GTP/GDP-dependent open/closed forms is well suited to<br />

interact with an effector protein. Since in the EF-Tu 7 GTP the domains II/III interact with<br />

the helix B, it is most likely that these domains represent here an integrated effector.<br />

19.3.5 Aminoacyl-tRNA in complex with EF-Tu7GTP<br />

The dissociation constants of aminoacyl-tRNA EF-Tu 7 GTP ternary complexes were measured<br />

at equilibrium by several independent methods [61–63]. Site-specific fluorescence labelling<br />

of tRNA on the cytidin74 with AEDANS allowed an equilibrium measurement of<br />

the complex formation between aminoacyl-tRNA <strong>and</strong> EF-Tu 7 GTP [62]. This method was<br />

used to determine the equilibrium dissociation constants K D for different tRNA isoacceptors<br />

[64], modified tRNAs [47, 65–68], modified EF-Tu, <strong>and</strong> EF-Tu mutants [54, 69, 70]. These<br />

are summarized in Tab. 19.2. The K D values for different aminoacyl-tRNAs <strong>and</strong> EF-Tu species<br />

are in the range of 5610 –10 –10 –8 M [64]. The missing aminoacyl residue (tRNA7EF-<br />

Tu 7 GTP interaction) lowers the affinity by four orders of magnitude [63]. About 10 base<br />

pairs of the aminoacyl arm are recognized by EF-Tu 7GTP, as shown by investigation of aminoacyl-RNA<br />

minihelices [68]. Whereas the positional isomers of aminoacyl-tRNA (2' or 3')<br />

do not have a strong effect on complex formation, the attachment of the aminoacyl residue<br />

by an amide bond to the 3'-terminal ribose strongly inhibits the ternary complex formation<br />

(Tab. 19.2a). An aminoacylated-tRNA-like structure containing a pseudoknot in the acceptor<br />

Tables 19.2: Apparent equilibrium dissociation constants K D for aminoacyl-tRNA EF-Tu7GTP ternary<br />

complex as determined by a fluorescence-spectrometric method according to [62].<br />

Table 19.2a: Variation of aminoacyl-tRNA using T. thermophilus EF-Tu7GTP. Asp-tRNA Asp originates<br />

from yeast <strong>and</strong> is fully modified. A minihelix consists of two RNA oligonucleotides derived from aminoacyl<br />

arm of tRNA. The number of base pair counted from aminoacyl end is indicated in parentheses. His-<br />

TYMV RNA is an aminoacylated tRNA-like structure originating from turnip yellow mosaic virus. AE-<br />

DANS indicates that the fluorescence reporter group was used as an indicator for the measurements of the<br />

interaction with EF-Tu7GTP whereas 2' or 3' dA, 3'NH, <strong>and</strong> oxi/red denote different non-isomerisable<br />

aminoacyl-tRNAs [28].<br />

Aminoacyl-tRNA K D [10 –10 M] Fold decrease Reference<br />

Asp-tRNA Asp (modified) 16.1 [68]<br />

Asp-tRNA Asp (unmodified) 26.5 1.65 [68]<br />

Asp-Minihelix Asp (12 bp) 62.0 3.85 [68]<br />

Asp-Minihelix Asp (13 bp) 92.1 5.72 [68]<br />

Asp-Minihelix Asp (8 bp) 1900 118 [68]<br />

389


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

Table 19.2a (continued)<br />

Aminoacyl-tRNA K D [10 –10 M] Fold decrease Reference<br />

Asp-Minihelix Asp (7 bp) > 2680 166 [68]<br />

Minihelix Asp > 51200 3180 [68]<br />

His-tRNA His (Yeast) 45.1 2.80 unpublished<br />

His-TYMV RNA (3'-end) 44.7 2.78 unpublished<br />

Ala-tRNA Ala (E. coli) 18.9 unpublished<br />

Ala-Minihelix Ala (12 bp) 110 5.82 unpublished<br />

Ala-Microhelix Ala (7 bp) > 4040 214 unpublished<br />

(AEDANS)Tyr-tRNA Tyr 17.0 [72]<br />

Tyr-tRNA Tyr 17.4 1.02 unpublished<br />

Tyr-tRNA Tyr 2'dA 42.0 2.47 unpublished<br />

Tyr-tRNA Tyr 3'dA 120 7.06 unpublished<br />

Phe-tRNA Phe 22.8 1.34 unpublished<br />

Phe-tRNA Phe 3'dA 189 11.1 unpublished<br />

Phe-tRNA Phe 3'NH 1230 72.4 unpublished<br />

Phe-tRNA Phe oxi/red 1830 108 unpublished<br />

Table 19.2b: Variations of aminoacyl-tRNA using E. coli EF-Tu7GTP. When not indicated tRNAs originate<br />

from E. coli.<br />

Aminoacyl-tRNA K D [10 –10 M] Fold decrease Reference<br />

Tyr-tRNA Tyr (AEDANS) 2.4 [62]<br />

Gln-tRNA Gln 1.9 0.79 [64]<br />

Tyr-tRNA Tyr 5.6 2.33 [64]<br />

Phe-tRNA Phe 6.0 2.50 [62]<br />

AcPhe-tRNA Phe 3800 1580 [63]<br />

tRNA Phe 26100 10875 [63]<br />

tRNA (unfractionated) 28200 11750 [63]<br />

Ser-tRNA Ser (GCU) 7.0 2.92 [64]<br />

Ser-tRNA Ser (UGA) 7.2 3.0 [64]<br />

Ser-tRNA Sec 500 208 [65]<br />

Met-tRNA Met 7.3 3.04 [64]<br />

Met<br />

Met-tRNA f 173 72.1 [63]<br />

Met<br />

fMet-tRNA f 1360 567 [63]<br />

Met-tRNA Met i (Yeast) 6000 2500 [66]<br />

Met-tRNA Met i (Yeast, ox) 110 45.8 [66]<br />

Thr-tRNA Thr 7.7 3.21 [64]<br />

Trp-tRNA Trp 13.5 5.63 [64]<br />

Glu-tRNA Glu 17.0 7.08 [64]<br />

Leu-tRNA Leu (UAA) 22.3 9.29 [64]<br />

Leu-tRNA Leu (GAG) 29.7 12.4 [64]<br />

Asp-tRNA Asp 31.2 13.0 [64]<br />

Leu-tRNA Leu (CAA, E. coli) 33.8 14.1 [64]<br />

Val-tRNA Val (GAC) 36.2 15.1 [64]<br />

Val-tRNA Val (UAC) 47.1 19.6 [64]<br />

Leu-tRNA Leu (CAG) 64.1 26.7 [64]<br />

390


19.3 Structure of elongation factor Tu<br />

arm does not disturb the interaction with EF-Tu [71]. Different aminoacyl-tRNA species interact<br />

with EF-Tu with different affinities. These sequence <strong>and</strong> aminoacyl-dependent variations<br />

are up to 64-fold (Tab. 19.2b). Peptidyl-tRNAs <strong>and</strong> initiator Met-tRNAs interact only<br />

weakly with EF-Tu 7 GTP (Tab. 19.2b).<br />

19.3.6 1 H NMR of yeast Phe-tRNA Phe EF-Tu 7GTP complex<br />

For codon-specific binding of aminoacyl-tRNA to the ribosome the tRNA must form a complex<br />

with the elongation factor Tu. The complex formation manifests itself also in the imino<br />

resonance region of the 1 H NMR spectrum of the tRNA. Since the molecular mass of tRNA<br />

(25,000 Da) is only about one third of the molecular mass of the aminoacyl-tRNA-EF-<br />

Tu 7 GTP complex (70,000 Da) the complex formation should become apparent by an increase<br />

of the line widths of the imino resonances. This was indeed observed with a complex<br />

of yeast Phe-tRNA Phe <strong>and</strong> EF-Tu 7 GTP from T. thermophilus (Fig. 19.14).<br />

Figure 19.14: The imino proton resonances in the NMR spectrum of free Phe-tRNA Phe <strong>and</strong> in ternary<br />

complex with EF-Tu 7GTP <strong>and</strong> EF-Tu7GDP [47]. a) 270 µM tRNA Phe , b)270 µM Phe-tRNA Phe EF-<br />

Tu 7GTP ternary complex, c) the same complex as in (b) after GTP hydrolysis <strong>and</strong> deacylation of PhetRNA<br />

Phe . The dashed lines in (b) <strong>and</strong> (c) correspond to the 1 H NMR imino proton spectrum of free<br />

tRNA Phe , as shown in (a). The numbers denote the imino proton resonances of the base pairs in the<br />

tRNA Phe acceptor stem as assigned by [74]. The numbers in circles indicate the signals of the first three<br />

imino protons in the acceptor stem, which disappear from the NMR spectrum after ternary complex formation.<br />

391


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

The spectra recorded immediately after complex formation, display resonance signals<br />

which are broadened by a factor of about 2.6 in comparison to free tRNA. In addition, it<br />

should be kept in mind that during the acquisition of the spectrum (about 1 hour), part of<br />

the GTP has already been cleaved to GDP [75]. In the GDP bound form, however, the affinity<br />

of EF-Tu for aminoacyl-tRNA is considerably lower than in the GTP bound form<br />

(Tab. 19.2c).<br />

Table 19.2c: Variation of EF-Tu using Tyr-tRNA Tyr (AEDANS).<br />

EF-Tu (from T. thermophilus) K D [10 –10 M] Fold decrease Reference<br />

EF-Tu . GTP 17.0 [72]<br />

EF-Tu . GDP 98000 5760 [72]<br />

EF-Tu . GTP (Thr62Ala) 618 36.4 unpublished<br />

EF-Tu . GTP (Thr62Ser) 31.7 1.86 unpublished<br />

EF-Tu . GTP (His67Ala) 60.9 3.58 unpublished<br />

EF-Tu . GTP (His85Gln) 260 15.3 [54]<br />

EF-Tu . GTP (His85Leu) 54.6 3.21 [54]<br />

EF-Tu . GTP (Glu55Leu) 16.3 0.96 unpublished<br />

EF-Tu . GTP (Glu56Ala) 49.2 2.89 unpublished<br />

EF-Tu . GTP (Arg59Thr) 36.4 2.14 [47]<br />

EF-Tu7GTP 181–190 24.0 1.41 unpublished<br />

wild type (6His-tag) 82.3 4.84 unpublished<br />

EF-Tu-Domain I >2120 125 unpublished<br />

EF-Tu f 35 2.06 [72]<br />

EF-Tu . GMPPCP 82.0 4.82 unpublished<br />

EF-Tu . GMPPNP 98.9 5.82 unpublished<br />

EF-Tu . f EF-Ts 50.6 2.98 unpublished<br />

EF-Tu (from E. coli) K D [10 –10 M] Fold decrease Reference<br />

EF-Tu . GTP 2.4 [62]<br />

EF-TuA R 37 15.4 [73]<br />

EF-TuB o 56 23.3 [73]<br />

Trypsin cleaved EF-Tu 54 22.5 [73]<br />

TPCK-treated EF-Tu 2400 1000 [73]<br />

The imino spectrum of the Phe-tRNA Phe -EF-Tu 7 GTP complex also reveals several<br />

distinct line shifts in comparison to the spectrum of the free tRNA. Moreover, some of the<br />

imino resonances, namely the ones of base pairs 1 through 3 of the acceptor stem, seem to<br />

be missing. However, it cannot be excluded that these lines overlap with other lines, due to<br />

larger shift changes. The assumption of vanishing imino resonances is also supported by the<br />

work of Heerschap et al. [76], who observed a disappearance of the imino resonances of the<br />

first acceptor stem base pairs upon complex formation of tRNA Phe with E. coli EF-Tu.<br />

As the ternary complex dissociates in the course of the GTP hydrolysis <strong>and</strong> the PhetRNA<br />

Phe deacylates, the line width of the imino resonances decreases again <strong>and</strong> the chemical<br />

shifts adopt the positions they had before in the free tRNA. However, even after com-<br />

392


19.3 Structure of elongation factor Tu<br />

plete GTP hydrolysis <strong>and</strong> total deacylation the line widths do not fully recover the initial values<br />

of the spectra of free tRNA Phe , indicating a complex between tRNA Phe <strong>and</strong> EF-Tu7GDP<br />

(Fig. 19.14). Since the above-described NMR experiments have been undertaken at complex<br />

concentrations of about 0.3 mM, the still noticeable association of deacylated tRNA <strong>and</strong> EF-<br />

Tu 7 GDP represents no contradiction to the numerous biochemical studies which typically<br />

use concentrations below ca. 1 µM.<br />

19.3.7 13 C NMR studies of the Val-tRNA Val EF-Tu7GTP ternary complex<br />

E. coli Val-tRNA Val , 13 C labelled at the carbonyl group, was used for 13 C NMR analysis of<br />

the ternary complex formed by Val-tRNA Val <strong>and</strong> EF-Tu 7 GTP [47]. Labelling the carbonyl<br />

group appeared particularly favourable since it represents the carbon atom which functions<br />

as a linker between the ribose of the tRNA terminal adenosine76 <strong>and</strong> the amino acid residue.<br />

This linkage is the most important prerequisite for ternary complex formation (Tab. 19.2b).<br />

Free Val-tRNA provides two peaks in the 13 C NMR spectrum around 170 ppm corresponding<br />

to 2' <strong>and</strong> 3' isomers, with a slight preference for 3' derivative (Fig. 19.15 a). Complex formation<br />

of Val-tRNA Val with EF-Tu 7 GTP leads to a drastic upfield shift to a region around<br />

64 ppm (Fig. 19.15 b). Such an extremely large upfield shift cannot be explained by conformational<br />

changes of the protein or isomerizations of Val-tRNA. Rather it requires the assumption<br />

of substantial alterations in the electronic environment of the carbonyl group. A<br />

change of the hybridization state from sp 2 (carbonyl) to sp 3 (orthoester) upon complex formation<br />

of Val-tRNA Val with EF-Tu 7 GTP is a conceivable explanation for the observed resonance<br />

shift.<br />

Figure 19.15: Part of the proton-decoupled 125.7 MHz 13 C NMR spectrum of a) E. coli [1-C- 13 C] ValtRNA<br />

Val . The resonances at 169.8 ppm <strong>and</strong> 169.5 ppm represent the 3' <strong>and</strong> 2' isomers of Val-tRNA Val ,<br />

respectively; b) [1-C- 13 C] Val-tRNA Val EF-Tu7GDP/GTP. The resonances at 63.7 ppm <strong>and</strong> 63.5 ppm<br />

were suggested to belong to the GDP <strong>and</strong> GTP forms of the protein, respectively [75].<br />

393


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

Such an orthoester can be formed either with the oxygen atoms of both, the 2' <strong>and</strong> the<br />

3'-hydroxyl groups of the A76 ribose (Fig. 19.16 a) or with a nucleophilic functional group<br />

of the protein (Fig. 19.16 b). The structure in Fig. 19.16 a seems to be reasonable as during<br />

the (relatively slow) transacylation reaction of the amino acid residue this intermediate state<br />

has to be passed [77]. The structure shown in Fig. 19.16 b would explain the observed chemical<br />

shift as well. In this case a nucleophile leading to an easily cleavable transient bond<br />

should be provided by the protein. A deprotonated carboxylate located in the interface between<br />

domains I <strong>and</strong> II of EF-Tu [1] is the most obvious c<strong>and</strong>idate for such a function.<br />

A<br />

B<br />

Figure 19.16: Orthoester acid intermediate structure of the aminoacyl residue in the Val-tRNA Val EF-<br />

Tu 7GTP ternary complex. The function of nucleophile is fulfilled a) either by the 2'-OH group or b) by<br />

a functional group of the EF-Tu.<br />

Evidently, EF-Tu stabilizes the orthoester intermediate possibly via electrostatic or hydrogen<br />

bond interactions. This could also be the reason for the strong upfield shift of this resonance,<br />

which is extraordinary as compared to the available model molecules. Moreover, a<br />

strong conformational constraint in the vicinity of the sp 3 hybridized carbonyl group could<br />

contribute to this strong upfield shift.<br />

The relative intensities of the two lines at 63.7 ppm <strong>and</strong> 63.5 ppm are varying with time<br />

after complex formation. The 63.7 ppm line increases at the expense of the 63.5 ppm signal<br />

[47]. The disappearance of the 63.5 ppm resonance proceeds with the time constant of the<br />

GTP hydrolysis [3]. Thus, the two signals could be associated with two slightly differing conformations<br />

of the orthoester in the GTP <strong>and</strong> GDP forms of EF-Tu. This assumption is corroborated<br />

by the observation of only one single resonance at 63.7 ppm for a complex of [1- 13 C]ValtRNA<br />

Val with EF-Tu which has bound the slowly hydrolyzable GTP analogue GMPPNP [47].<br />

Positional isomers (2' or 3') of aminoacyl-tRNA do not have a dramatic effect on the<br />

affinity of EF-Tu 7 GTP. However, the replacement of the ester bond (–O–CO–) for the<br />

amide bond (–NH–CO–) attachment of aminoacyl residue to tRNA strongly influences the<br />

affinity (Tab. 19.2 a).<br />

394


19.3 Structure of elongation factor Tu<br />

19.3.8 Role of EF-Tu in complex with aminoacyl-tRNA<br />

Domains II/III function as an effector promoting the binding of aminoacyl-tRNA. In the inactive<br />

GDP conformation these domains are tilted away from the domain I, forming a large<br />

cavity. This conformational change has considerable functional consequences, since EF-<br />

Tu 7 GDP interacts with aminoacyl-tRNA 5000 times less efficiently than EF-Tu7GTP. Thus<br />

it is conceivable to suggest that the binding site of the main determinant of the aminoacyltRNA<br />

binding, namely the aminoacyl-residue, is placed in the binding site which is formed<br />

by the EF-Tu 7GDP to EF-Tu7GTP conformational change. Indeed it was demonstrated that<br />

the reactive e-bromoacetyl-lysyl-tRNA Lys labels specifically a histidine67 residue located<br />

near the interface of domain I <strong>and</strong> II in EF-Tu 7 GTP [78]. The X-ray structure determination<br />

of aminoacyl-tRNA7EF-Tu 7 GppNHp [79] <strong>and</strong> aminoacyl-AMP EF-Tu 7 GppNHp ternary<br />

complexes confirm the affinity labelling results <strong>and</strong>, in addition to histidine67, locate glutamate271<br />

of domain II in the vicinity of aminoacyl-tRNA binding site. Aminoacyl residues<br />

50–60 in the effector loop of EF-Tu Cys82 in domain I [64], arginine300 [54], <strong>and</strong> threonine393<br />

[55] of domain III are other aminoacyl residues from the domain interfaces which<br />

are involved in aminoacyl-tRNA binding. Mutations in EF-Tu which affect GTPase activity<br />

(Thr62, His85) do not influence aminoacyl-tRNA binding (Tab. 19.2 c).<br />

The aminoacyl domain of tRNA composed of the CCA end, seven base pairs of the<br />

aminoacyl stem <strong>and</strong> five base pairs of the T stem, is sufficient to promote efficient binding<br />

to the EF-Tu7GTP [68]. About 10 base pairs of the aminoacyl domain are bound to domain<br />

III of EF-Tu [79]. This binding is probably governed by ionic interactions of the RNA-phosphate<br />

backbone. The sequence of nucleotides, however, also plays some role in this process.<br />

EF-Tu 7GTP serves for aminoacyl-tRNA not only as a vehicle transporting it to the ribosome<br />

but in addition as a matrix setting the correct conformation of the L shaped molecule [80].<br />

The precise distance between the anticodon loop <strong>and</strong> aminoacyl-residue is evidently required<br />

for aminoacyl-tRNA functioning during translation [50].<br />

19.3.9 EF-Tu interaction with EF-Ts<br />

Elongation factor Ts forms stable complexes with EF-Tu <strong>and</strong> fulfills the role of the nucleotide<br />

exchange protein (NEP). It is not known how this protein binds to EF-Tu <strong>and</strong> how the<br />

acceleration of the GDP dissociation is achieved. However, it is clear that EF-Tu 7 EF-Ts interaction<br />

changes the structure of nucleotide binding domain. Such effects were identified<br />

especially in the L 2 region of EF-Tu, which alters its protease accessibility upon EF-Tu 7 EF-<br />

Ts complex formation [81]. Replacement of the lysine in the NKXD-guanine binding region<br />

of E. coli EF-Tu for glutamate causes a dissociation of nucleotide <strong>and</strong> formation of stable<br />

EF-Tu 7EF-Ts complexes. This effect can be reversed by a second mutation in EF-Tu which<br />

prevents EF-Tu 7 EF-Ts interaction. Such mutants were found by replacement of some amino<br />

acids in the C-terminal region of domain I [82] (residues 154–199). The binding of EF-Ts<br />

may dislocate the guanine binding SA consensus element (Fig. 19.2) leading to dissociation<br />

of the bound nucleotide. The interaction of transducin-b with rhodopsin takes place presum-<br />

395


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

ably also in the vicinity of the region of the GTP-binding domains [83]. In addition to domain<br />

I of EF-Tu, EF-Ts apparently interacts also with domains II or III. This is evident from<br />

investigation of EF-Tu species of domains I/II [84] <strong>and</strong> II/III [72].<br />

19.3.10 Site-directed mutagenesis of EF-Tu<br />

An elegant way to achieve a site-directed replacement of one or more amino acids in the<br />

protein is provided by genetically induced mutations into particular genes followed by expression<br />

of the mutant protein. This method was extensively used in the investigations of<br />

T. thermophilus EF-Tu. A compilation of mutants is presented in Tab. 19.3. In many cases<br />

the replacement of invariant <strong>and</strong> functionally important amino acids leads to the loss of the<br />

Table 19.3: Derivatives of EF-Tu, prepared by site-directed mutagenesis of the wild type T. thermophilus<br />

tufA gene variants, were overexpressed in E. coli, purified, <strong>and</strong> fully characterized.<br />

EF-Tu Mutant <strong>Properties</strong> References<br />

A) Single Mutants<br />

Glu55Leu Ribosome-induced GTPase affected unpublished<br />

Glu56Ala Ribosome-induced GTPase affected unpublished<br />

Arg59Thr Poly(Phe) synth. affected, binding of aa-tRNA affected [47]<br />

Thr62Ser Nucleotide binding is affected unpublished<br />

Thr62Ala<br />

Nucleotide binding is affected,<br />

aa-tRNA binding is affected,<br />

Ribosome binding is defective<br />

unpublished<br />

His67Ala aa-tRNA binding is affected unpublished<br />

His85Gln Ribosome <strong>and</strong> aa-tRNA binding is affected [54]<br />

His85Gly Protein degradation; cell growth affected [54]<br />

His85Leu Slower GTPase; ribosome binding is affected [54]<br />

Asp81Ala Protein degradation [54]<br />

Cys82Ala As wild type unpublished<br />

Arg300Ile Protein degradation [54]<br />

EF-Tu(NHis6) aa-tRNA binding affected unpublished<br />

EF-Tu (CHis6) Ribosome induced GTPase affected unpublished<br />

B) Double Mutants<br />

His85Gly/Gly233Asp Protein degradation; normal cell growth unpublished<br />

Cys82Ala/Thr394Cys aa-tRNA binding defective unpublished<br />

C) Deletion Mutants<br />

181–190 Thermostability decreases [85]<br />

212–405 (Dom. I) Thermostability decreases [85]<br />

313–405 (Dom. I/II) Higher intrinsic GTPase; thermostable [85]<br />

1–316 (Dom. III) No functional [85]<br />

1–208 (Dom. II/III) Thermostable; EF-Ts interaction [72]<br />

212–405/His85Gly Thermostability decreases; no ribosome interaction [85]<br />

396


19.4 Summary <strong>and</strong> conclusions<br />

specific function of the protein. Thus the GTPase was affected by replacement of His85 or<br />

Thr62, the binding of aminoacyl-tRNA was impaired when Arg59 or Thr294 were mutated,<br />

<strong>and</strong> the stability of protein was decreased by mutations at Asp81 or Arg300. Remarkably,<br />

the deletion of domain III led to the stimulation of intrinsic GTPase. Some mutations were<br />

introduced in order to increase the stability of EF-Tu (Arg59) against proteolysis to improve<br />

the quality of crystals (His67) or to allow crystallisation of EF-Tu with affinity reagents <strong>and</strong><br />

transition state analogues (Cys82). Many of these investigations are still in progress.<br />

19.4 Summary <strong>and</strong> conclusions<br />

The size or properties of the molecule often do not allow the use of direct physical or biochemical<br />

methods to study the structure <strong>and</strong> function of biological macromolecules. Even in<br />

the case of the tertiary structure determination by NMR or X-ray crystallography the structure<br />

dynamics, mechanism of catalytic reactions, <strong>and</strong> their regulation remain to be studied<br />

by site-specific methods. In the present work we concentrated our effort on the study of<br />

RNA structure, structure <strong>and</strong> mechanism of a regulated GTPase (represented by an elongation<br />

factor Tu) <strong>and</strong> on a protein-RNA complex (formed by aminoacyl-tRNA, EF-Tu, <strong>and</strong><br />

GTP). Fluorescence spectroscopy (after specific labelling of RNA), electron microscopy<br />

(after attachment of a gold cluster to tRNA), NMR spectroscopy (using specifically introduced<br />

stable isotopes), ESR spectroscopy (with nitroxyl radicals bound to substrates of enzymatic<br />

reactions), affinity labelling of proteins (with reactive substrates), <strong>and</strong> site-directed<br />

mutagenesis in combination with the X-ray structure analysis of proteins were the utilized<br />

methods. The results of this investigation provided insight into the regulation <strong>and</strong> function<br />

of GTPases <strong>and</strong> allowed the discovery of some rules governing the RNA-protein interaction.<br />

Despite the large amount of new information which was gained from this research, most rewarding<br />

is the fact that we are now better prepared to ask new relevant questions for our future<br />

work.<br />

Acknowledgement<br />

The research laboratory in Bayreuth was supported by the Deutsche Forschungsgemeinschaft,<br />

Sonderforschungsbereich 213, D4, <strong>and</strong> D5. We thank R. Hilgenfeld <strong>and</strong> J. Nyborg for<br />

cooperation on the study of EF-Tu 7 GppNHp structure, M. Daniel for help with preparation<br />

of the manuscript, <strong>and</strong> all co-workers who participated in the work. Their names can be<br />

found as co-authors in the referred work from our laboratory.<br />

397


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

References<br />

1. H. Berchtold, L. Reshetnikova, C.O.A. Reiser, N.K. Schirmer, M. Sprinzl, R. Hilgenfeld: Nature,<br />

365, 126 (1993)<br />

2. M. Sprinzl: TIBS, 19, 245 (1994)<br />

3. S. Limmer, C.O.A. Reiser, N.K. Schirmer, N.W. Grillenbeck, M. Sprinzl: Biochem., 31, 2970<br />

(1992)<br />

4. G. Varani, I. Tinoco Jr.: Q. Rev. Biophys., 24, 479 (1991)<br />

5. S.A. White, M. Nilges, A. Huang, A.T. Brünger, P.B. Moore: Biochem., 31, 1610 (1992)<br />

6. C. Cheong, G. Varani, I.J. Tinoco: Nature, 346, 680 (1990)<br />

7. H.A. Heus, A. Pardi: Science, 253, 191 (1991)<br />

8. M. Chastain, I.J. Tinoco: Biochem., 31, 12 733 (1992)<br />

9. A.A. Szewczak, P.B. Moore: J. Mol. Biol., 247, 81 (1995)<br />

10. F.H.T. Allain, G. Varani: J. Mol. Biol., 250, 333 (1995)<br />

11. A. Heerschap, C.A.G. Haasnoot, C.W. Hilbers: Nucleic Acids Res., 10, 6981 (1982)<br />

12. A. Heerschap, C.A.G. Haasnoot, C.W. Hilbers: Nucleic Acids Res., 11, 4483 (1983)<br />

13. A. Heerschap, C.A.G. Haasnoot, C.W. Hilbers: Nucleic Acids Res., 11, 4501 (1983)<br />

14. D.R. Hare, N.S. Ribeiro, D.E. Wemmer, B.R. Reid: Biochem., 24, 4300 (1985)<br />

15. M. Sprinzl, F. Cramer: Proc. Natl. Acad. Sci. USA, 73, 3049 (1975)<br />

16. P. Schimmel: Interpretation of experiments that delineate transfer RNA recognition in vivo <strong>and</strong> in<br />

vitro, in: F. Eckstein, D.M. Lilley (eds.): Nucleic acids <strong>and</strong> molecular biology, Springer, Berlin,<br />

p. 274 (1990)<br />

17. L.H. Schulman: Progr. Nucl. Acids Res. Mol. Biol., 41, 23 (1991)<br />

18. R. Giegé, J.D. Puglisi, C. Florentz: Progr. Nucl. Acids Res. Mol. Biol., 45, 129 (1993)<br />

19. W.H, McClain: J. Mol. Biol., 234, 257 (1993)<br />

20. Y.-M. Hou, P. Schimmel: Nature, 333, 140 (1988)<br />

21. C. Francklyn, K. Musier-Forsyth, P. Schimmel: Eur. J. Biochem., 206, 315 (1992)<br />

22. W.H. McClain, K. Foss: Science, 240, 793 (1988)<br />

23. W.H. McClain, K. Foss: Science, 241, 1804 (1988)<br />

24. H.W. McClain, Y.-M. Chen, K. Foss, J. Schneider: Science, 242, 1681 (1988)<br />

25. C. Francklyn, P. Schimmel: Nature, 337, 478 (1989)<br />

26. K. Musier-Forsyth, N. Usman, S. Scaringe, J. Doudna, R. Green, P. Schimmel: Science, 253, 784<br />

(1991)<br />

27. K. Musier-Forsyth, P. Schimmel: Nature, 357, 513 (1992)<br />

28. M. Sprinzl, F. Cramer: Progr. Nucl. Acids Res. Mol. Biol., 22, 2 (1979)<br />

29. K. Nagamatsu, Y. Miyazawa: J. Biochem., 94, 1967 (1983)<br />

30. H. Paulsen, W. Wintermeyer: Eur. J. Biochem., 138, 117 (1984)<br />

31. J.D. Puglisi, I.J. Tinoco: Methods Enzymol., 180, 304 (1989)<br />

32. S. Limmer, H.-P. Hofmann, G. Ott, M. Sprinzl: Proc. Natl. Acad. Sci. USA, 90, 6199 (1993)<br />

33. G. Ott, L. Arnold, S. Limmer: Nucleic Acids Res., 21, 5859–5864 (1993)<br />

34. L. Arnold, J. Smrt, J. Zajicek, G. Ott, M. Schiesswohl, M. Sprinzl: Collect. Czech. Chem. Commun.,<br />

56, 1948 (1991)<br />

35. G. Ott, L. Arnold, J. Smrt, M. Sobkowski, S. Limmer, H.-P. Hofmann, M. Sprinzl: Nucleosides &<br />

Nucleotides, 13, 1069 (1994)<br />

36. C.W. Hilbers: Hydrogen-bonded proton exchange <strong>and</strong> its effect on NMR spectra of nucleic acids,<br />

in: R.G. Schulman (ed.): Biological applications of magnetic resonance, Academic Press, New<br />

York, p. 1 (1979)<br />

37. J.L. Leroy, D. Broseta, N. Bolo, M. Guéron: NMR studies of base-pair kinetics of nucleic acids, in:<br />

P.H. van Knippenberg, C.W. Hilbers (eds.): Structure <strong>and</strong> dynamics of RNA, Plenum Press, New<br />

York, p. 31 (1986)<br />

398


References<br />

38. D.B. Arter, P.G. Schmidt: Nucleic Acids Res., 3, 1437 (1976)<br />

39. S.R. Holbrook, C. Cheong, I. Tinoco, S.-H. Kim: Nature, 353, 579 (1991)<br />

40. O. Jardetzky, G.C.K. Roberts: in: NMR in molecular biology, Academic Press, Orl<strong>and</strong>o, (1981)<br />

41. K. Wüthrich: in: NMR of proteins <strong>and</strong> nucleic acids, Wiley, New York, (1986)<br />

42. S.S. Wijmenga, M.M.W. Mooren, C.W. Hilbers: NMR of nucleic acids; from spectrum to structure,<br />

in: G.C.K. Roberts (ed.): NMR of Macromolecules, IRL Press, Oxford, p. 217, (1993)<br />

43. W.R. Croasmun, R.M.K. Carlson: in: Two-dimensional NMR spectroscopy, VCH, Weinheim,<br />

(1994)<br />

44. S. Limmer, B. Reif, G. Ott, L. Arnold, M. Sprinzl: FEBS Lett. 385, 15 (1996)<br />

45. R. Boelens, T.M.G. Koning, G.A. van der Marel, J.H. van Boom, R. Kaptein: J. Magn. Resonance,<br />

82, 290 (1989)<br />

46. M. Taiji, S. Yokoyama, T. Miyazawa: Biochem., 22, 3220 (1983)<br />

47. C. Förster, S. Limmer, W. Zeidler, M. Sprinzl: Proc. Natl. Acad. Sci. USA, 91, 4254 (1994)<br />

48. Y. Kaziro: Biochim. Biophys. Acta, 505, 95 (1978)<br />

49. R.C. Thompson: TIBS, 13, 91 (1988)<br />

50. M.V. Rodnina, R. Fricke, L. Kuhn, W. Wintermeyer: EMBO J., 14, 2613 (1995)<br />

51. L. Seidler, M. Peter, F. Meissner, M. Sprinzl: Nucleic Acids Res., 15, 9263 (1987)<br />

52. M.V. Milburn, L. Tong, A.M. de Vos, A. Brünger, Z. Yanaizumi, S. Nishimura, S.-H. Kim:<br />

Science, 247, 939 (1990)<br />

53. M.S. Marshall: TIBS, 18, 250 (1993)<br />

54. W. Zeidler, C. Egle, S. Ribeiro, A. Wagner, V. Katunin, R. Kreutzer, M. Rodnina, W. Wintermeyer,<br />

M. Sprinzl: Eur. J. Biochem., 229, 596 (1995)<br />

55. C. Alex<strong>and</strong>er, N. Bilgin, C. Lindschau, J.R. Mesters, B. Kraal, R. Hilgenfeld, V.A. Erdmann, C.<br />

Lippmann: J. Biol. Chem., 270, 14541 (1995)<br />

56. M. Kjeldgaard, J. Nyborg: J. Mol. Biol., 223, 721 (1992)<br />

57. A. Wittinghofer, R. Frank, R. Leberman: Eur. J. Biochem., 108, 423 (1980)<br />

58. M.E. Peter, N.K. Schirmer, C.O.A. Reiser, M. Sprinzl: Biochem., 29, 2876 (1990)<br />

59. W. Zeidler, R. Kreutzer, M. Sprinzl: FEBS Lett., 319, 185 (1993)<br />

60. K.-W. Hwang, F. Jurnak, D.L. Miller: A mutation that hinders the GTP induced aminoacyl-tRNA<br />

binding of elongation factor Tu, in: L. Bosch, B. Kraal, A. Parmeggiani (eds.): Guanine-nucleotide<br />

binding proteins, Plenum Press, New York, , p. 77 (1989)<br />

61. A. Louie, F. Jurnak: Biochem., 24, 6433–6439 (1985)<br />

62. G. Ott, H.G. Faulhammer, M. Sprinzl: Eur. J. Biochem., 184, 345 (1989)<br />

63. F. Janiak, V.A. Dell, J.K. Abrahamson, B.S: Watson, D.L. Miller, D.L., A.E. Johnson: Biochem.,<br />

29, 4268 (1990)<br />

64. G. Ott, M. Schiesswohl, S. Kiesewetter, C. Förster, L. Arnold, V.A. Erdmann, M. Sprinzl: Biochim.<br />

Biophys. Acta, 1050, 222 (1990)<br />

65. C. Förster, G. Ott, K. Forchhammer, M. Sprinzl: Nucleic Acids Res., 18, 487 (1990)<br />

66. S. Kiesewetter, G. Ott, M. Sprinzl: Nucleic Acids Res., 18, 4677 (1990)<br />

67. B. Blechschmidt, V. Shirokov, M. Sprinzl: Eur. J. Biochem., 219, 65 (1994)<br />

68. J. Rudinger, B. Blechschmidt, S. Ribeiro, M. Sprinzl: Biochem., 33, 5682 (1994)<br />

69. J.P. Abrahams, J.J.C. Acampo, G. Ott, M. Sprinzl, J.M. de Graaf, A. Talens, B. Kraal: Biophys.<br />

Acta, 1050, 226 (1990)<br />

70. K. Bench, U. Pieper, G. Ott, N. Schirmer, M. Sprinzl, A. Pingoud: Biochimie, 73, 1045 (1991)<br />

71. R.L. Joshi, H. Faulhammer, F. Chapeville, M. Sprinzl, A.-L. Haenni: Nucleic Acids Res., 12, 7467<br />

(1984)<br />

72. M.E. Peter, C.O.A. Reiser, N.K. Schirmer, T. Kiefhaber, G. Ott, N.W. Grillenbeck, M. Sprinzl: Nucleic<br />

Acids Res., 18, 6889 (1990)<br />

73. G. Ott, J. Jonak, I.P. Abrahams, M. Sprinzl: Nucleic Acids Res., 18, 437 (1990)<br />

74. A. Heerschap, J.-R. Mellema, H.G.J.M. Janssen, J.A.L.I. Walters, C.A.G. Haasnoot, C.W. Hilbers:<br />

Eur. J. Biochem., 149, 649 (1985)<br />

75. C. Förster, S. Limmer, S. Ribeiro, R. Hilgenfeld, M. Sprinzl: Biochimie, 75, 1159 (1993)<br />

76. A. Heerschap, A.L.I. Walters, J.R. Mellema, C.W. Hilbers: Biochem., 25, 2707 (1986)<br />

399


19 Site-Directed Spectroscopy <strong>and</strong> Site-Directed Chemistry of Biopolymers<br />

77. S. Chládek, M. Sprinzl: Angew. Chem. Int. Ed., 24, 371 (1985)<br />

78. L.K. Duffy, L. Gerber, A.E. Johnson, D.L. Miller: Biochem., 20, 4663 (1981)<br />

79. P. Nissen, M. Kjeldgaard, S. Thirup, G. Polekhina, L. Reshetnikova, B.F.C. Clark, J. Nyborg: The<br />

three-dimensional structure of the ternary complex. 16th International Workshop, May 27–June 1,<br />

1995, University of Wisconsin-Madison, Abstract, 51 (1995)<br />

80. H.J. Adkins, D.L. Miller, A.E. Johnson: Biochem., 22, 1208 (1983)<br />

81. N.K. Schirmer, C.O.A. Reiser, M. Sprinzl: Eur. J. Biochem., 200, 295 (1991)<br />

82. Y.-W. Hwang, M. Carter, D.L. Miller: J. Biol. Chem., 267, 22198 (1992)<br />

83. J.P. Noel, H.E. Hamm, P.B. Sigler: Nature, 366, 654 (1993)<br />

84. S. Nock, A. Lechler, S. Ribeiro, R. Kreutzer: Biotechniques, 18, 608 (1995)<br />

85. S. Nock, N. Grillenbeck, M.R. Ahmadian, S. Ribeiro, R. Kreutzer M. Sprinzl: Eur. J. Biochem.,<br />

132 (1995)<br />

400


20 Spectroscopic Probes of Surfactant Systems<br />

<strong>and</strong> Biopolymers<br />

Alex<strong>and</strong>er Wokaun<br />

20.1 Introduction<br />

Characterization of surfactant systems is a challenging task because these systems, which often<br />

represent stable <strong>and</strong> homogeneous phases on the macroscopic scale, are typically consisting<br />

of microscopic or mesoscopic aggregates, such as micelles. Tremendous advances in the<br />

underst<strong>and</strong>ing of these systems has been made through direct visualization by freeze-fracture<br />

electron microscopy <strong>and</strong> by the application of scattering techniques rendering a wealth of information<br />

on aggregate shapes <strong>and</strong> sizes.<br />

In order to investigate the microscopic properties of the individual aggregates or compartments<br />

constituting the surfactant systems toposelective methods of spectroscopy have<br />

been applied in this study. Focusing on the microscopic dynamics, three complementary<br />

methods will be discussed. First, the self-diffusion coefficients of individual components<br />

have been distinguished by FT-NMR pulsed gradient spin echo experiments. This technique<br />

has been applied to characterize the diffusion of micelles with solubilized hydrocarbons <strong>and</strong><br />

residual mobilities in cubic phases (ringing gels) that are formed by aggregation of swollen<br />

micelles.<br />

An alternative method for tracing the motion of micellar aggregates is their selective<br />

labelling by photochromic dye probes. Diffusion coefficients are determined by monitoring<br />

the time dependence of laser induced gratings using the forced Rayleigh scattering (FRS)<br />

technique. In this way, the influence of alcohol content on the diffusion of a system forming<br />

multilamellar vesicles has been investigated. In a further study, the influence of solubilized<br />

water on the diffusion of reverse micelles in the AOT/octanol/water system <strong>and</strong> diffusion in<br />

the corresponding bicontinuous gel phase have been studied.<br />

A fascinating aspect of surfactant systems resides in the fact that within the same chemical<br />

system one-dimensional aggregates (rod-like micelles), two-dimensional bilayers, <strong>and</strong><br />

space-filling continuous structures may be formed. This aspect of dimensionality has been<br />

probed by studying the time-dependence of fluorescence, as induced by lipophilic quencher<br />

molecules diffusing within dimensionally restricted regions of space.<br />

The second part of this study was devoted to the spectroscopic probing of biopolymers.<br />

Transitions between right <strong>and</strong> left-helical conformations of oligonucleotide duplexes<br />

were monitored by Raman <strong>and</strong> 2D NMR techniques.<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

401


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

The conditions for the applicability of surface enhanced Raman spectroscopy for studies<br />

of oligo <strong>and</strong> polynucleotides has been assessed. In particular, surface enhanced resonance<br />

Raman spectroscopy (SERRS) has been used to study the binding of chromophores to<br />

DNA str<strong>and</strong>s at probe concentrations as low as 10 –6 M.<br />

20.2 Diffusion in surfactant systems<br />

Three different spectroscopic techniques have been used to determine molecular <strong>and</strong> micellar<br />

diffusion coefficients in surfactant systems. Each of these methods is probing a different<br />

aspect of molecular mobilities. First, the NMR pulsed gradient spin echo (PGSE) method is<br />

used to determine composition-induced changes in the diffusion coefficients of micelles <strong>and</strong><br />

inverse micelles in ternary surfactant systems. In particular, differences between the diffusion<br />

coefficients in micellar <strong>and</strong> cubic phases are monitored. Second, the diffusion of absorbing<br />

dye probes in light-induced gratings is monitored by the method of FRS, providing<br />

information on the structure of multilamellar vesicles. Third, the dynamics of fluorescent<br />

probes due to diffusing quenchers is investigated, rendering information on the dimensionality<br />

of diffusion pathways <strong>and</strong> hence on the structure of lyotropic mesophases.<br />

20.2.1 Structural characteristics of micellar solutions, cubic phases,<br />

<strong>and</strong> multilamellar vesicles from NMR self-diffusion measurements<br />

Ternary systems consisting of surfactant, hydrocarbon <strong>and</strong> water exhibit a rich variety of<br />

phases. With increasing fraction of the hydrophobic components, frequently a sequence of<br />

structural changes of the type<br />

micelles ) hexagonal phase ) lamellar phase ) inverse hexagonal phase )<br />

inverse micelles<br />

is observed [1]. Additional to the mentioned phases, the occurrence of cubic phases has recently<br />

been discovered [2, 3] which include both bicontinuous structures [4] <strong>and</strong> cubic or<br />

disordered packages of micellar aggregates. Of particular interest are the so-called ringing<br />

gel phases [5], which exhibit a metallic sound upon mechanical excitation, <strong>and</strong> have been<br />

extensively investigated within the Sonderforschungsbereich 213.<br />

In our experiments, we have focused attention on a homologous series of ternary systems<br />

consisting of alkyl-dimethylaminoxide surfactants, C n H 2n+1 (CH 3 ) 2 N + O – (C n DMAO), a linear<br />

hydrocarbon (C m H 2m+2 ) or a cyclic hydrocarbon (C m H 2m ), <strong>and</strong> water as the solvent.<br />

A systematic investigation of the concentration dependence in the system C 14 DMAO/C 6 H 12 /<br />

402


20.2 Diffusion in surfactant systems<br />

H 2 O [6] showed that water constituted the continuous phase in both, the micellar solution <strong>and</strong><br />

the ringing gel phase. In the latter, the diffusion coefficients of hydrocarbon <strong>and</strong> surfactant<br />

were lower by one <strong>and</strong> two orders of magnitude, respectively, as compared to the micellar solution.<br />

Thus, we are dealing with an aggregated network where the residual mobilities are caused<br />

by exchange of surfactant <strong>and</strong> solute molecules between neighbouring micelles in the gel.<br />

In order to investigate whether the mentioned results are generic properties of the<br />

mentioned class of ringing gels, a series of micellar solutions <strong>and</strong> of gel samples was prepared<br />

where the water weight fraction was fixed to 70 % <strong>and</strong> 57.50 %, respectively, varying<br />

the surfactant <strong>and</strong> hydrocarbon chain lengths. Self-diffusion coefficients have been determined<br />

by the stimulated echo experiment described by Tanner [7]. Details of the experiment<br />

as well as the exact composition of the samples with the corresponding phase diagrams of<br />

the ternary surfactant systems have been given recently [8].<br />

The influence of the surfactant chain length on the self-diffusion of the solvent (here<br />

D 2 O) is shown in Fig. 20.1. For both, the micellar phase (rhombs) <strong>and</strong> the cubic phase<br />

(squares), the solvent diffusion coefficient is seen to rise with the alkyl chain length of the<br />

surfactant. In this series, the hydrodynamic radius of the micelles is known to increase with<br />

the surfactant length. Hence, at constant weight fraction of the micelles, the system<br />

C 16 DMAO/C 6 H 12 /D 2 O contains fewer but larger micelles, as compared to C 12 DMAO/<br />

C 6 H 12 /D 2 O. As these micelles may be considered as obstacles for the diffusion of water the<br />

observed trend is then easily interpreted in terms of a st<strong>and</strong>ard model [9].<br />

Figure 20.1: Diffusion coefficient of water in ternary C n DMAO/C 6 H 12 /D 2 O systems, as a function of<br />

the surfactant chain length n. Rhombs: micellar phase; squares: ringing gel phase. Details of the compositions<br />

are given by Panitz et al. [8].<br />

The self-diffusion coefficient of the solubilized hydrocarbon was determined by Fourier<br />

transformation of the second half-echo [8]. In the micellar phase an interesting trend is<br />

observed (Fig. 20.2): the diffusion coefficient first increases with the hydrocarbon chain<br />

length up to m = 10, then a pronounced decrease for the larger dodecane <strong>and</strong> tetradecane solutes<br />

follows.<br />

Light scattering experiments by Oetter <strong>and</strong> Hoffmann [5] have shown that the hydrodynamic<br />

radius of the micelles decreases (together with the total capacity for solubilization)<br />

403


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

Figure 20.2: Influence of the hydrocarbon chain length m on the micellar self-diffusion coefficient D in<br />

ternary C 14 DMAO/C m H 2m+2 /D 2 O surfactant systems. For m = 6, cyclohexane C 6 H 12 has been chosen as<br />

the solute.<br />

as the chain length of the hydrocarbon solute increases from C 6 H 12 to C 10 H 22 . This result<br />

agrees with the increase in the diffusion coefficient for the same series in Fig. 20.2.<br />

In order to account for the subsequent decrease in the diffusion coefficient for longer<br />

hydrocarbon chains we have to recall that the C 14 DMAO surfactant by itself forms rod-like<br />

micelles in the binary solution in water. Due to the addition of a critical amount of hydrocarbon<br />

these rods are transformed into spherical micelles. As mentioned above, the solubilization<br />

capacity is rapidly decreasing with the size of the solute. Thus, for very long hydrocarbons<br />

the solubility limit is reached prior to induction of the rod ) sphere transition. In fact,<br />

for the C 14 DMAO/C 14 H 30 /D 2 O system the amount of solubilized hydrocarbon in the investigated<br />

sample was only slightly above the critical concentration required for the mentioned<br />

transition. In such a situation small dynamic deviations from an average spherical shape of<br />

the micelles are expected to occur. Thus a lower diffusion coefficient is observed for this<br />

system – <strong>and</strong> to a lesser extent also for the homologous system containing C 12 H 26 – which<br />

is attributed to shape fluctuations of the micelles [8].<br />

The NMR investigations have been complemented by studying the solvent diffusion in<br />

multilamellar vesicles in a related system. When the ionic surfactant C 14 H 29 N + (CH 3 ) 3 Br –<br />

(CTAB) <strong>and</strong> the cosurfactant n-C 6 H 13 OH are added to the above-mentioned C 14 DMAO/<br />

H 2 O system vesicles <strong>and</strong> lamellar phases are formed [10]. The aim of this study was to<br />

complement the rheological measurements in these systems, where a decrease of both, the<br />

storage modulus G' <strong>and</strong> the yield stress, with increasing salinity of the water had been<br />

found.<br />

Results are shown in Fig. 20.3 for three systems containing NaCl (c = 0, 10, 100 mM).<br />

The measured signal may be attributed to extravesicular water. The diffusion coefficient of<br />

the solvent D 2 O increases by about 50 % upon the addition of NaCl (c = 10 mM), in agreement<br />

with a lowering of the yield stress observed in the rheological measurements [10].<br />

Using the equation D H2 O = D o (1 – F v ), where D o is the diffusion coefficient of free<br />

water <strong>and</strong> F v is the volume fraction of the micelles, an increase of D H2 O appears to be accompanied<br />

by a decrease in F v . Assuming a constant total surface area, a decrease of F v<br />

must be due to deformation of the vesicles from the maximum volume spheres.<br />

404


20.2 Diffusion in surfactant systems<br />

Figure 20.3: Diffusion coefficient of the solvent D 2 O in a system consisting of C 14 DMAO (c = 90 mM),<br />

CTAB (c = 10 mM), <strong>and</strong> n-C 6 H 13 OH (c = 220 mM), which forms multilamellar vesicles. The dependence<br />

on diffusion time D is investigated. Circles: no salt added; rhombs: NaCl (c = 10 mM); squares:<br />

NaCl (c = 100 mM).<br />

Increasing the NaCl concentration to 100 mM gives rise to a further strong increase in<br />

the diffusion coefficient. Interestingly, the apparent coefficient depends now on the diffusion<br />

time set by the experiment, i. e. the time t between the two gradient pulses. Such a dependence<br />

is st<strong>and</strong>ardly interpreted as being due to restricted diffusion. Both, the absolute increase <strong>and</strong><br />

the diffusion time dependence, hint to a further deformation of the multilamellar vesicles into<br />

planar lamellar stacks or aggregates, which are partially ordered within the NMR tube.<br />

20.2.2 Probing of mobilities in multilamellar vesicles by forced<br />

Rayleigh scattering<br />

Recently, increasing use has been made of laser-induced gratings for visualizing static <strong>and</strong><br />

dynamic properties of gases <strong>and</strong> condensed phases [11]. In FRS two coherent laser beams<br />

are made to interfere in the volume of the sample. In our case that was a surfactant solution<br />

doped with a chromophoric dye probe. Excitation of the absorber molecules in the regions<br />

of high light intensities of the interference pattern gives rise to a spatial modulation of both,<br />

the absorption <strong>and</strong> of the refractive index of the sample. The lattice constant L of this grating<br />

is determined by the wavelength l <strong>and</strong> by the angle y between the two beams (Fig. 20.4),<br />

L ˆ l=2 sin…y=2† :<br />

…1†<br />

The hologram induced in this manner may be probed by a second laser of wavelength<br />

l', which is diffracted by the grating (Fig. 20.4). For thick holograms it is essential that the<br />

grating is probed at an angle j for which the Bragg condition (Eq. 2) for diffraction on the<br />

planes within the sample is fulfilled (Fig. 20.4),<br />

2 L sinj ˆ m l 0 …2†<br />

405


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

ϕ<br />

ka<br />

kb<br />

q<br />

Λ<br />

Figure 20.4: FRS experimental setup. Beam (1) is the writing laser gated by a Kerr cell (5) between<br />

crossed polarizers (4, 7). The probe laser (10) is diffracted by the grating in the sample (11); the diffracted<br />

order is isolated by an iris diaphragm (13), <strong>and</strong> recorded by a pin diode (15). The Bragg condition<br />

for reading out the grating in a thick hologram is illustrated on the right-h<strong>and</strong> side.<br />

In order to monitor the diffusion of the probe molecules the sample is illuminated for<br />

a short time interval gating the laser with a Kerr cell (Fig. 20.4). The length of the excitation<br />

pulse must be carefully matched to the intrinsic relaxation time of the photochromic probe.<br />

After terminatio of excitation the grating is washed out by diffusion of both, excited <strong>and</strong> unexcited<br />

dye molecules. Hence, the diffraction efficiency of the grating is diminished <strong>and</strong> the<br />

intensity I (t) of the diffracted probe laser beam decreases according to the equation<br />

I …t† ˆ…A e t=s ‡ B† 2 ‡ C 2 :<br />

…3†<br />

The decay constant s is influenced by both, the intrinsic relaxation time s dye of the<br />

probe (i. e. the time constant of relaxation of the photoisomer produced) <strong>and</strong> the diffusion<br />

constant D, according to the equation<br />

s 1 ˆ s 1<br />

dye ‡ Dq2 ;<br />

…4†<br />

where q =2p/L. Hence, if the grating constant L is varied via the angle of intersection y<br />

(Fig. 20.4), a plot of s –1 vs. q 2 will yield the desired diffusion coefficient as the slope.<br />

This procedure is illustrated in Fig. 20.5, using as an example the diffusion of the hydrophobic<br />

dye methyl red (2-carbohydroxy-4'-dimethylamino-azobenzene) in simple linear chain<br />

aliphatic alcohols. This photochromic probe molecule is transformed from the trans into the<br />

cis form by irradiation with the Ar + laser wavelength of 514.5 nm. The lifetime s dye varies between<br />

about 20 ms <strong>and</strong> 1 s, depending on the medium <strong>and</strong> its viscosity. The dependence of<br />

the diffusion coefficient on the viscosity is clearly seen from the different slopes in Fig. 20.5.<br />

The power of the method will be illustrated by two applications. First, we investigate<br />

the influence of the alcohol content on the quaternary system for which the formation of<br />

multilamellar vesicles has been discovered by Hoffmann et al. [10]. Second, diffusion in the<br />

water-in-oil microemulsion <strong>and</strong> in the bicontinuous cubic gel phase of the AOT/octanol/<br />

water system is investigated.<br />

The multilamellar vesicle system was investigated without the addition of salt. The<br />

surfactant mixture was held constant at 90 mM C 14 DMAO <strong>and</strong> 10 mM C 14 H 29 N + (CH 3 ) 3 Br – ,<br />

406


20.2 Diffusion in surfactant systems<br />

Figure 20.5: Evaluation of diffusion coefficients from angle-dependent FRS measurements. The grating<br />

decay rate s –1 is plotted against q 2 (Eq. 4). The diffusion of methyl red in simple linear chain alcohols<br />

was investigated.<br />

with the hexanol (C 6 OH) concentration being varied. Methyl red was added in a concentration<br />

of 7610 –5 M. Polarization microscopy clearly shows [12] that the dye molecules are attached<br />

to the vesicles <strong>and</strong> hence are probing diffusion processes within the vesicles.<br />

The results presented in Fig. 20.6 comprise diffusion measurements both, in the micellar<br />

L 1 phase (c (C 6 OH) = 0, 30, <strong>and</strong> 60 mM) <strong>and</strong> in the L a phase (c (C 6 OH) = 150 mM). The<br />

diffusion coefficient is largest for c (C 6 OH) = 0 M; it drops rapidly upon the addition of<br />

30 mM of hexanol as rod-like micelles are formed. Interestingly, the diffusion coefficient increases<br />

again if the alcohol concentration is doubled. This behaviour has been successfully<br />

interpreted [12] by noting that increasing the hexanol concentration from 30 to 60 mM<br />

causes a shortening of the rod-like micelles, which according to the theory of Doi <strong>and</strong> Edwards<br />

[13] gives rise to an increase in the diffusion coefficient.<br />

Figure 20.6: Influence of the hexanol concentration on the diffusion coefficient of methyl red in the<br />

C 14 DMAO/CTAB/C 6 H 13 OH system (see text).<br />

407


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

In the L a phase (c (C 6 OH) > 150 mM in Fig. 20.6), the diffusion coefficient of methyl<br />

red is found to be low <strong>and</strong> approximately constant. Freeze-fracture microscopy has revealed<br />

that the system consists of densely packed multilamellar vesicles (Hoffmann 1995). Their<br />

average size (1 µm) is smaller than the spacing of the light-induced grating (15–25 µm),<br />

but some large vesicles with sizes up to 30 µm are present as well. Hence, the observed decay<br />

of the FRS signal is due to the exchange of dye molecules between surfactant bilayers<br />

on the one h<strong>and</strong> <strong>and</strong> due to diffusion within giant vesicles on the other h<strong>and</strong>. Variation of<br />

the alcohol content within the L a phase changes the number of the vesicles but does not alter<br />

the qualitative nature of the diffusional processes, thus underst<strong>and</strong>ing the constancy of the<br />

measured diffusion coefficient.<br />

The section is concluded by reporting results for the AOT/octanol/water system. The<br />

water-soluble dye congo red – a structurally related photochromic azo dye [12] – has been<br />

used as a probe. Compositions corresponding to two cross sections through the ternary phase<br />

diagram [14] are investigated. Details are given by Hahn et al. [12].<br />

First, a series of four samples within the region of water-in-oil (w/o) microemulsion<br />

phase L 2 was investigated (Fig. 20.7 a). The diffusion constant of the probe (residing within<br />

the water phase) is seen to decrease strongly with the water content, to stay constant above a<br />

water concentration of 50 wt%. This behaviour is interpreted in terms of the structural models<br />

of Scriven [15] <strong>and</strong> of Fontell [1]. The microemulsion is thought to undergo continuous<br />

a)<br />

b)<br />

Figure 20.7: Diffusion coefficients in the AOT/octanol/water system determined by FRS. a) variation of<br />

the water content within the L 2 phase; b) transition from the L 2 into the I 2 phase induced by decreasing<br />

the octanol weight fraction.<br />

408


20.2 Diffusion in surfactant systems<br />

structural changes with increasing water content, i. e. swelling of inverse rod-like micelles<br />

<strong>and</strong> formation of long narrow water channels within which the diffusion of the dye probe<br />

would be strongly impeded. As mentioned above, the spacing of the light-induced grating<br />

amounts to 15–25 µm. Over such distances, diffusion within the water channels may be excluded.<br />

Thus, the observed decay is ascribed to motion of the aggregates as a whole, which<br />

is slowed down as the channels are swelling due to the addition of water.<br />

The transition from the L 2 into the cubic I 2 phase was investigated by varying the octanol<br />

content (Fig. 20.7b). The sample containing 20 wt% octanol corresponds to an L 2<br />

phase, whereas the three samples with 10 <strong>and</strong> 14 wt% of octanol, respectively, are cubic gels<br />

(I 2 phase). The diffusion coefficient is distinctly higher in the gels, however, the increase<br />

with respect to the w/o microemulsion appears to be a gradual one, rather than a stepwise<br />

change. The highest value recorded (5610 –11 m 2 s –1 ) is still four times smaller than the diffusion<br />

coefficient of congo red in water. The observed behaviour would not be consistent<br />

with the model of a gel consisting of closely packed, aggregated micelles. Rather, it suggests<br />

a bicontinuous structure of the gel, in which water channels are opening up with decreasing<br />

octanol content. This result is in agreement with the characterization of the gel by Gradzielski<br />

et al. [16], who found that the I 2 phase is characterized by higher long-range order.<br />

The mentioned description is also consistent with the structural models developed by Fontell<br />

[1] for related systems.<br />

20.2.3 Dimensionality of diffusion in lyotropic mesophases from fluorescence<br />

quenching<br />

Fluorescence quenching has been successfully used in surfactant research to determine aggregation<br />

numbers in micellar systems [17, 18]. Fluorescence quenching in a homogeneous<br />

phase leads to the well-known monoexponential decays but not if quencher diffusion is limited<br />

to restricted regions in space (e. g. the lipophilic regions in a surfactant system). It has<br />

early been realized [19] that the precise shape of the time dependence could be used to determine<br />

the dimensionality of the space available for diffusion. The probability for diffusive<br />

encounter between a fluorescent molecule <strong>and</strong> a given quencher decreases exponentially<br />

with time for three-dimensional diffusion. Whereas the integrated probability of encounter<br />

approaches unity in one or two dimensions the time dependence is distinctly different for<br />

each case.<br />

The aim of the present study was to demonstrate this concept within a single surfactant<br />

system forming both, one-dimensional aggregates (rod-like micelles) <strong>and</strong> a two-dimensional<br />

a-lamellar phase. The C 14 DMAO/C 7 H 15 OH/H 2 O system was chosen for this purpose because<br />

it forms in addition to the above-mentioned phases two distinct L 3 mesophases of different<br />

viscosity [20]. The structure of these cubic phases has been the subject of intense investigations.<br />

Freeze fracture electron microscopy demonstrated their bicontinuous character [21].<br />

In our study, the C 14 DMAO concentration was held constant at 100 mM, while the<br />

heptanol concentration was varied between 0 <strong>and</strong> 200 mM in order to realize compositions<br />

that correspond to the mentioned phases. With view to comparability of earlier studies in<br />

rod-like micelles [22], pyrene (c =10 –5 M) was chosen as the fluorescent probe <strong>and</strong> benzo-<br />

409


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

phenone as the quencher diffusing exclusively within the lipophilic regions of the surfactant<br />

system. Details of the experiment <strong>and</strong> of the data reduction procedure have been described<br />

in the thesis work of Wiesner [23] <strong>and</strong> Meyer [24].<br />

Experimental results for a micellar phase consisting of rod-like micelles <strong>and</strong> of an a-<br />

lamellar phase in the same system are compared in Fig. 20.8. Fluorescence decay curves are<br />

presented as logarithmic plots. The deviations from exponentiality in the micellar system<br />

(Fig. 20.8a) are evident as all traces are curved even for long times except for the system<br />

containing no quencher (top trace). Evidently, the decay is faster in the a-lamellar phase<br />

(Fig. 20.8b) when comparing runs at the same quencher concentration. This result meets the<br />

intuitive expectation that the approach of quenchers from all sides within a lamellar plane<br />

should give rise to a faster decay as compared to one-dimensional r<strong>and</strong>om walk within a cylindrical<br />

or rod-like micelle.<br />

a) b)<br />

Figure 20.8: Fluorescence decay of pyrene in the C 14 DMAO/C 7 H 15 OH/H 2 O system. a) c (C 7 H 15 OH) = 0 M,<br />

rod-like micelles; benzophenone quencher (top to bottom: c = 0, 0.1, 0.3, 0.5, 0.7 mM).<br />

b) c (C 7 H 15 OH) = 62 mM, a-lamellar phase; benzophenone quencher (top to bottom: c = 0, 0.1, 0.3, 0.5,<br />

0.7 mM).<br />

Corresponding representative results for the L 3 phase (Fig. 20.9) show again a distinctly<br />

non-exponential decay. In view of the poorer signal-to-noise ratio, obtained with this<br />

sample, an extensive data set was acquired with this sample.<br />

For the analysis the analytic result of Almgren et al. [22], of Alsins <strong>and</strong> Almgren [25]<br />

for 1D diffusion, <strong>and</strong> the model of Owen [19] for 2D diffusion have been used. Their equations<br />

were implemented in a fitting routine which performed, for a given sample, a global<br />

analysis of the entire data set recorded at different quencher concentrations.<br />

Prior to deriving values of the diffusion coefficients, it was checked whether the quality<br />

of the non-exponential decays recorded was sufficient to warrant a distinction of one <strong>and</strong><br />

two-dimensional diffusion. In fact it was found that the use of the inappropriate model (e. g.<br />

2D diffusion for the rod-like micelles) resulted in considerably higher mean square deviations,<br />

thus demonstrating the significance of the fits. Technically, the intrinsic fluorescence<br />

decay time of pyrene was derived from the curve without added quencher, the quenching<br />

rate constant k q was adjusted as a global parameter <strong>and</strong> an average diffusion constant was<br />

then derived using the known quencher concentrations [24].<br />

410


20.2 Diffusion in surfactant systems<br />

Figure 20.9: Fluorescence decay of pyrene in the L 3 phase of the C 14 DMAO/H 2 O system (c (C 7 H 15 OH)<br />

= 190 mM). Temperature was held constant at 25 8C. Concentration of benzophenone quencher (top to<br />

bottom: c = 0, 0.3, 0.6, 0.67, 0.75, 0.83, 1.0, 1.33 mM).<br />

The selection of the appropriate model for the L 3 phase deserves some comment. According<br />

to the accepted model [21,26], it consists of a bicontinuous structure featuring inner<br />

<strong>and</strong> outer water channels separated by surfactant bilayers. An idealized structure resembling<br />

a Gaussian minimal surface with positive <strong>and</strong> negative curvature at each point has been proposed<br />

[27]. The real structure probably corresponds to a multiply branched, irregular analogue<br />

of the Gaussian surface. The decisive feature for the present analysis is that the surfactant<br />

molecules are arranged in a (positively <strong>and</strong> negatively curved) surface in space, such<br />

that the application of the 2D model of diffusion is appropriate, see below.<br />

Results for the diffusion coefficients derived from the fits are compiled in Tab. 20.1.<br />

Surveying the results we see that the diffusion coefficient of the benzophenone quencher is<br />

2–3 times higher in the two investigated lamellar phases than in the micellar phases. At first<br />

sight, this result seems to contradict the higher viscosity of the a-lamellar phase. However,<br />

we have to consider that we are dealing with the motion of the comparatively large benzophenone<br />

molecule, containing two phenyl rings, within the lipophilic region consisting of<br />

the surfactant alkyl chains. A simple model suggests that a preferred orientation of the plane<br />

of the benzophenone molecule is perpendicular to the rod axis <strong>and</strong> that diffusional motion<br />

parallel to the rod axis is severely impeded. In contrast, within a surfactant bilayer there are<br />

many more ways for the translational motion of benzophenone within the lipid region, so<br />

that the higher diffusion coefficient obtained for the latter phase may be understood.<br />

Most interesting is the result that the diffusion coefficient in the L 3 phase, 6.7610 –11 m 2<br />

s –1 , is significantly smaller than the one in the a-lamellar phase, although we have argued that<br />

diffusion in the L 3 phase should be essentially two-dimensional. Resolution of this apparent contradiction<br />

comes from a recent theoretical analysis by Anderson <strong>and</strong> Wennerström [4] who<br />

showed that the diffusion coefficient of a r<strong>and</strong>om walker on a Gaussian minimum surface should<br />

be only about 2/3 of its value on a plane. In view of this result, which strictly applies only for an<br />

idealized regular structure, the results in Table 20.1 are fully supporting the presently accepted bicontinuous<br />

model of the L 3 phase.<br />

411


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

Table 20.1: Quencher diffusion coefficients in the C 14 DMAO/C 7 H 15 OH/H 2 O system.<br />

C 7 H 15 OH concentration/mM Diffusion coefficient D/10 –11 m 2 s –1 Phase <strong>and</strong> model<br />

0 4.9 ± 2.2 rod-like micelles<br />

1D diffusion<br />

20 3.0 ± 0.9 rod-like micelles<br />

1D diffusion<br />

62 12.4 ± 5.5 lamellar phase<br />

2D diffusion<br />

170 11.6 ± 2.4 lamellar phase<br />

2D diffusion<br />

190 6.7 ± 3.4 L 3 phase<br />

modified 2D diffusion<br />

20.2.4 Summary of results<br />

In Section 20.2 the use of three spectroscopic methods has been exemplified for obtaining<br />

information on diffusion coefficients in surfactant systems. Regarding the micellar phases<br />

both, NMR self-diffusion measurements <strong>and</strong> FRS, have been successfully used to monitor<br />

changes in the micellar diffusion coefficients resulting from size <strong>and</strong> shape changes. This information<br />

complements the results from static <strong>and</strong> dynamic light scattering experiments in a<br />

most valuable manner.<br />

Two structurally different types of cubic phases have been clearly distinguished by the<br />

diffusion measurements. In the C n DMAO/hydrocarbon/water ringing gels which consist of<br />

aggregated micelles, residual mobilities resulting from exchange processes between the micelles<br />

have been characterized by NMR. On the other h<strong>and</strong>, the increase in long-range order<br />

<strong>and</strong> the opening of water channels in the bicontinuous AOT/octanol/water gels was evidenced<br />

by the increased diffusion coefficient of the water soluble congo red dye probe, monitored<br />

by FRS.<br />

For the quaternary C 14 DMAO/CTAB/hexanol/water system NMR was used to investigate<br />

salt-induced changes in the volume fraction occupied by the multilamellar vesicles, as<br />

well as deformation <strong>and</strong> shape alterations. In the same system, FRS yielded information on<br />

changes in diffusion coefficients upon the transition from rod-like micelles to the vesicular<br />

phase, induced by increasing the alcohol concentration.<br />

For the C 14 DMAO/heptanol/water system monitoring the time dependence of fluorescence<br />

due to diffusing quenchers resulted in a clear distinction of one <strong>and</strong> two-dimensional<br />

diffusion. For the L 3 phase an experimental confirmation of a recent theoretical model for<br />

diffusion on a Gaussian minimal surface has been achieved.<br />

412


20.3 Vibrational spectroscopy <strong>and</strong> conformational analysis of oligonucleotides<br />

20.3 Vibrational spectroscopy <strong>and</strong> conformational analysis<br />

of oligonucleotides<br />

Three projects have been pursued, with the aim of providing spectroscopic tools for the research<br />

of Section 19. First, conformational changes in oligodeoxyribonucleotide duplexes<br />

were monitored by Raman spectroscopy <strong>and</strong> by 2D NMR techniques. The second aim was<br />

to assess the potential of surface enhanced Raman spectroscopy (SERS) for the characterization<br />

of oligonucleotides. Third, the interaction of intercalating <strong>and</strong> groove binding dye<br />

probes with DNA str<strong>and</strong>s has been investigated by resonance Raman spectroscopy.<br />

20.3.1 Spectroscopic characterization of right <strong>and</strong> left-helical forms<br />

of a hexadecanucleotide duplex<br />

Changes between the right <strong>and</strong> left-helical forms of DNA are important for underst<strong>and</strong>ing<br />

control mechanisms of replication [28]. In particular, Jovin et al. [29] have characterized the<br />

structural parameters of Z DNA <strong>and</strong> have indicated Raman b<strong>and</strong>s that might be used as indicators<br />

for the B , Z transition.<br />

Our first goal was to assess the sensitivity of normal Raman spectroscopy for monitoring<br />

the right-to-left conformational transition. Defined substrates for this investigation, i. e.<br />

the hexadecanucleotides d (CG) 8 <strong>and</strong> d(m 5 CG) 8 , have been synthesized by Weber et al. [30].<br />

Duplexes were obtained in the right-helical B conformation from the synthesis, transforming<br />

them into the left-helical Z conformation by exposure to high salt concentration<br />

(c (NaCl) & 3M).<br />

Distinct differences between the Raman spectra of the two forms are evident in<br />

Fig. 20.10. In particular, the phosphodiester vibration of the backbone at a frequency of<br />

824 cm –1 is characteristic for the C 2' -endo pucker of the ribose rings in the B conformation.<br />

In the spectrum of the Z conformation, this b<strong>and</strong> is downshifted to 786 cm –1 <strong>and</strong> overlaps a<br />

cytosine ring breathing vibration.<br />

Further pronounced changes concern vibrations within the guanine ring system, which<br />

is oriented more towards the exterior surface of the double helix in the Z conformation because<br />

of the C 3' -endo-syn conformation at the ribose rings. The ring breathing vibration is<br />

downshifted from 680 to 612 cm –1 upon the B ) Z transition. In addition, several ring<br />

stretching vibrations of guanine in the 1300–1360 cm –1 region are affected (Fig. 20.10).<br />

2D NMR spectroscopy is an established tool for elucidating the three-dimensional<br />

structure of biopolymers in solution. We have used 2D NOE spectroscopy as an alternative<br />

technique for monitoring the B , Z conformational transition. The procedure is exemplary<br />

illustrated in Fig. 20.11 for the oligonucleotide d (m 5 CG) 8 . As mentioned above, the orientation<br />

of guanine relative to the ribose ring is distinctly different for the right <strong>and</strong> left-helical<br />

forms. Consequently, there are in the B conformation strong cross peaks between the guanine<br />

proton GH 8 <strong>and</strong> the ribose proton GH2'. In contrast, for the Z conformation, the H 8<br />

413


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

Figure 20.10: Raman spectra of the deoxyribonucleotide duplex d(CG) 8 .<br />

Top: B conformation, c = 39 mM; bottom: Z conformation, c = 24 mM.<br />

Figure 20.11: NOESY spectrum of d(m 5 CG) 8 in the Z conformation (medium: c (MgCl 2 ) = 10 mM).<br />

Experimental conditions have been given in Ref. [30].<br />

414


20.3 Vibrational spectroscopy <strong>and</strong> conformational analysis of oligonucleotides<br />

proton of the base <strong>and</strong> the GH1' proton of the sugar are spatially close. The spectrum demonstrates<br />

that from the presence of the respective cross peak the conformation present under<br />

the given conditions can be unambiguously inferred [30].<br />

20.3.2 SERS spectra of deoxyribonucleotides<br />

For SERS spectroscopy of biomolecules [31], the use of small sample volumes is m<strong>and</strong>atory.<br />

For this purpose, two spectroelectrochemical cells were constructed in which the liquid sample<br />

is held in a thin cylindrical volume between a silver rod electrode <strong>and</strong> the optical access<br />

window. The design of the cells (from glass or teflon) <strong>and</strong> the oxidation-reduction cycle, used<br />

for the required roughening of the silver, have been described in detail by Zimmermann [32].<br />

The power of the technique is illustrated by the SERS spectrum of a mononucleotide,<br />

5'-adenosine-monophosphate (5'-rAMP). The top spectrum in Fig. 20.12 (left) was recorded<br />

with a concentration of 1.5 mM 5'-rAMP in phosphate buffer, with the electrolyte concentration<br />

adjusted to 100 mM. In particular, attention is drawn to the excellent signal-to-noise<br />

ratio achieved at a millimolar concentration.<br />

The high salinity conditions that are often used to induce a transition into left-helical conformations<br />

in oligo or polynucloetides were critically considered for the envisaged application.<br />

Figure 20.12: SERS spectra of mononucleotides excited at 530.9 nm with the power of 29 mW <strong>and</strong> recorded<br />

in a spectroelectrochemical cell using a resolution of 7 cm –1 [33].<br />

Left diagram: 5'-rAMP (1.5 mM), top: NaCl (84 mM), KCl (16 mM), U Ag/AgCl = –0.5 V; bottom:<br />

NaCl (3.3 M), KCl (80 mM), U Ag/AgCl = –0.6 V.<br />

Right diagram: top: 5'-rCMP (2 mM), phosphate buffer (2 mM);<br />

bottom: 5'-rCMP (9 mM), phosphate buffer (9 mM), NaCl (0.1 M), U Ag/AgCl = –0.3 V.<br />

415


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

Therefore, the SERS spectrum of 5'-rAMP was also recorded in the presence of NaCl (3.3 M).<br />

The obtained high quality spectrum (Fig. 20.12, lower trace) demonstrates that SERS spectroscopy<br />

is feasible in the media that are typically used to prepare <strong>and</strong> study the Z conformation.<br />

Corresponding experiments with 5'-cytidine-monophosphate (5'-rCMP) showed an<br />

optimum adsorption potential of –0.3 V vs. Ag/AgCl (Fig. 20.11, right). Because the silver<br />

surface is positively charged at this potential, chloride <strong>and</strong> phosphate ions as well as impurities<br />

from the solution are adsorbed much more strongly to the electrode surface as compared<br />

to the potential of –0.5 V where the 5'-rAMP adsorption had been found to be an optimum<br />

(left). Therefore, the sensitivity in the 5'-rCMP, recorded at a concentration of 2 mM, was<br />

considerably lower as compared to the 5'-rAMP spectrum [33].<br />

An attempt to record the SERS signals of the oligonucleotide d(CG) 8 resulted in an<br />

uninformative spectrum (Fig. 20.13, top). In spite of the comparatively high total concentration<br />

(2 mg/ml & 3 mM), the signal-to-noise ratio is poor <strong>and</strong> only a few vibrations that are<br />

characteristic for d(CG) 8 are detected. In the bottom trace, the normal Raman spectrum recorded<br />

at a 10 times higher concentration is reproduced from Fig. 20.10 for comparison.<br />

Figure 20.13: SERS spectrum of d(CG) 8 (2 mg/mL), excited at 530.9 nm with the power 29 mW. Conditions:<br />

phosphate buffer (0.17 M), NaCl (0.33 M), U Ag/AgCl = –0.1 V. In the bottom trace, the normal<br />

Raman spectrum of d(CG) 8 is reproduced from Fig. 20.10 for comparison.<br />

Several reasons may be given for the unsatisfactory sensitivity of the SERS spectrum<br />

of d(CG) 8 . A comparatively more positive potential (–0.1 V vs. Ag/AgCl) had to be applied<br />

in order to promote adsorption of the negatively charged exterior of the double helix<br />

on the silver electrode <strong>and</strong> to avoid destabilization of the duplex [31]. At this potential,<br />

however, the binding of d(CG) 8 is severely competed by adsorption of phosphate ions<br />

from the buffer <strong>and</strong> chloride ions from the electrolyte, as shown by corresponding test experiments<br />

with 5'-rCMP. From these experiments we conclude that only at very low buffer<br />

<strong>and</strong> electrolyte concentrations favourable conditions might be found for recording sensitive<br />

SERS spectra.<br />

416


20.3 Vibrational spectroscopy <strong>and</strong> conformational analysis of oligonucleotides<br />

20.3.3 Studies of chromophore-DNA interaction by vibrational spectroscopy<br />

The binding of dye molecules onto DNA str<strong>and</strong>s is an important process in biochemistry<br />

[28]. Depending on its nature, this interaction may result in mutagenic, carcinogenic, <strong>and</strong> cytotoxic<br />

properties of the chromophore so that specific dyes have been suggested for the use<br />

as antitumor pharmaceuticals. Furthermore DNA binding dyes have been extensively used<br />

as stains in biochemical fluorescence microscopy work.<br />

In the present project we have investigated whether surface enhanced resonance Raman<br />

spectroscopy (SERRS) could be used to study various types of DNA-chromophore interactions.<br />

In this technique, a silver colloid is added to the solution containing the molecule<br />

to be studied [34]. First, we had to test whether the complex, between dye <strong>and</strong> DNA, would<br />

remain unaltered upon adsorption to the silver colloid surface which is a necessary condition<br />

for a valid application of the method. A second system specific question concerns the useful<br />

information that can be extracted from frequency shifts <strong>and</strong> relative intensity alterations in<br />

the vibrational spectrum of the DNA-bound chromophore [35–37].<br />

As a first example, the complex of acridine orange (AO) with calf thymus DNA has<br />

been studied [38]. The stability of the complex upon adsorption on the silver colloid was<br />

confirmed by absorption spectroscopy (Fig. 20.14, right). In solution the absorption maximum<br />

at 504 nm (trace b) of AO in the complex is red shifted as compared to 495 nm (trace<br />

a) of the spectrum of the free AO molecule. The analogous red shift is maintained when the<br />

complex is added into a solution containing the silver colloid (trace d).<br />

Highly sensitive SERS spectra have been recorded at the remarkably low AO concentration<br />

of 2.5610 –6 M (Fig. 20.14, left). The spectra of the free (trace a) <strong>and</strong> of the DNAbound<br />

dye (trace b) appear to be quite similar <strong>and</strong> only small changes in relative b<strong>and</strong> intensities<br />

are detected in the difference spectra (trace c). A careful analysis has been based on a<br />

Figure 20.14: Interaction of AO with calf thymus DNA. Left diagram: SERRS spectrum of the free<br />

AO (a) <strong>and</strong> of the AO-DNA (b) complex, both excited at 476.2 nm with a power of 25 mW. The top<br />

trace (c) represents the difference between the normalized spectra (b)-(a). Right diagram: absorbance<br />

spectra of AO (2.5610 –6 M); aqueous solution (a), complex with calf thymus DNA (b), free AO adsorbed<br />

on silver colloid (c), complex with calf thymus DNA adsorbed on silver colloid (d).<br />

417


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

complete assignment of the rich vibrational spectrum [38]. The observed alterations are<br />

compatible with a geometry in which the DNA double helix is adsorbed parallel to the silver<br />

surface [31] with the plane of the intercalated dye molecule approximately perpendicular to<br />

the helix axis.<br />

As a second example, the groove binding dye Hoechst 33 258 [39] was investigated.<br />

The SERRS spectrum has been recorded at a concentration of 10 –6 M. It was analyzed in<br />

detail <strong>and</strong> assigned using normal coordinate analysis [32]. The influence of protonation on<br />

the binding to the silver surface was studied <strong>and</strong> an adsorption geometry with the two-benzimidazole<br />

rings approximately parallel to the silver surface was inferred.<br />

Besides intercalation <strong>and</strong> groove binding the ionic interaction of charged reagents<br />

with the exterior surface of the double helix is a third mode of molecular interaction with<br />

DNA str<strong>and</strong>s. In this context, platinum(II) complexes have been intensely investigated with<br />

respect to their cytostatic properties. In the classic representative dichlorodiamin-platinum(II)<br />

or cisplatinum the binding to DNA is initiated by dissociation of a chloride lig<strong>and</strong>. In<br />

this context, we have investigated a series of novel dichlorobis(cycloalkylamin)Pt(II) complexes<br />

[40], in which the size of the cycloalkyl ring (C n H 2n–1 -) of the amine lig<strong>and</strong>s was varied<br />

from n = 3 to n = 8 [41]. The aim of the study was to detect influences of the lig<strong>and</strong><br />

size on the Pt-Cl bond strength in the complexes. But analysis showed that it is fairly constant<br />

throughout the investigated series. The pronounced differences in pharmaceutical activity<br />

seem to be mainly caused by the increasing lipophilicity of the higher membered rings.<br />

20.3.4 Summary of results<br />

Vibrational spectroscopy <strong>and</strong>, in particular, Raman scattering have been used to elucidate selected<br />

interactions of DNAs. The B , Z conformational transition of oligo-DNA duplexes<br />

<strong>and</strong> the interaction of the intercalating dye AO with calf thymus DNA have been explicitly<br />

analyzed in the preceding sections.<br />

Surface enhancement of the Raman signals due to adsorption of the analyte onto<br />

either silver electrodes or colloids provides large gains in sensitivity in suitable cases. Limitations<br />

are implicit in the requirement of adsorption on the silver surface. In the case of negatively<br />

charge DNA double helices this leads to unfavourable competition with the binding<br />

of phosphate ions from the buffer at positive surface potentials <strong>and</strong> to desorption at negative<br />

surface potentials. Therefore, for intact double helices it seems to be difficult to take full advantage<br />

of the enhancement, demonstrated for the smaller building blocks.<br />

Very intense Raman signals may be obtained if the enhancement, due to addition of<br />

silver colloids, is acting in combination with resonant enhancement of the Raman scattering<br />

cross section by close-lying electronic transitions. This favourable situation is found for<br />

many important dye molecules that have been shown to interact with DNA in a specific<br />

manner. The present results on intercalating dyes demonstrate the high potential of SERRS<br />

for studying chromophore-DNA interactions.<br />

418


20.4 Related projects carried out within the framework of the Collaborative Research Centre<br />

20.4 Related projects carried out within the framework<br />

of the Collaborative Research Centre<br />

The projects discussed in Section 20.2 have been centred around the spectroscopic probing<br />

of the structure <strong>and</strong> dynamics of surfactant systems <strong>and</strong> lyotropic mesophases, in particular<br />

hydrocarbon gels. Three related investigations not yet discussed in detail due to limitations<br />

in space shall be briefly mentioned. These are the study of the conformation of surfactant<br />

molecules by surface enhanced Raman spectroscopy, the development of a lipophilic dye<br />

probe for surfactant systems, <strong>and</strong> an outlook onto a different class of reaction gels created<br />

by hydrolysis of metal alkoxide precursors.<br />

The conformation of several surface-active molecules, polymers (alkyl-trimethylammonium<br />

surfactants, polyvinyl alcohol), <strong>and</strong> of model compounds such as choline has been<br />

studied by SERS [42, 43]. In particular, information on the geometry of their binding to the<br />

surface of silver colloids was obtained. Special methods for the synthesis of monodisperse<br />

colloids have been developed [44].<br />

The microscopic polarity in the interior of micellar aggregates is of intrinsic interest<br />

in surfactant research. A well-known approach to this information is the use of solvatochromic<br />

dye probes. For these chromophores, the wavelengths of absorption <strong>and</strong> emission (<strong>and</strong><br />

hence the Stokes shift) depend on the environment in which they are dissolved. A large variety<br />

of dye molecules exhibiting this property has been tested with surfactant systems. However,<br />

inspection reveals that most of these dyes contain ionic groups <strong>and</strong>, therefore, are<br />

either water-soluble or tend to bind to the hydrophilic exterior of micellar aggregates.<br />

Within the present project, we have synthesized <strong>and</strong> tested [23] the lipophilic dye<br />

probe 4-dioctylamino-anthra-thiadiazol-1,2-dion (TDA), for the structure see Fig. 20.15. As<br />

a measure for the solvatochromic properties of the compound, the Stokes shift of various<br />

O<br />

O<br />

N<br />

Oc<br />

N<br />

Oc<br />

N<br />

S<br />

Figure 20.15: Stokes shift as a function of solvent orientational polarization df for TDA. The structure<br />

of the molecule is shown on the right-h<strong>and</strong> side. Solvents: hexane (1), toluene (2), CCl 4 (3), dioxane<br />

(4), dibutylether (5), chlorobenzene (6), diethylether (7), ethylacetate (8), 1-chlorobutane (9), tetrahydrofuran<br />

(10), CH 2 Cl 2 (11), dimethylsulfoxide (12), dimethylformamide (13), acetonitrile (14).<br />

419


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

simple solvents is presented in Fig. 20.14. It shows a remarkable range of variation from<br />

about 500 cm –1 to about 2000 cm –1 . Solvent polarity is characterized in Fig. 20.15 by the<br />

orientational polarization df, which includes the effects of both, static <strong>and</strong> dynamic polarization<br />

of the solvent [45]. Linearity in the plot of the Stokes shift vs. df, as expected according<br />

to Lippert [46], shows that TDA is a promising c<strong>and</strong>idate for polarity probing in surfactant<br />

systems.<br />

The fluorescence lifetime of TDA is highest (10–12 ns) in hydrocarbon solvents <strong>and</strong><br />

decreases to 0.6 ns in polar solvents, such as acetonitrile. A plot of the decay rate against<br />

the dimensionless parameter E T N [47], which characterizes solvent polarity, has been given by<br />

Wiesner [23]. In contrast to the pronounced polarity dependence the lifetime is found to depend<br />

only weakly on the viscosity of the environment.<br />

Chromophore doping has been used not only as a spectroscopic probe but also in order<br />

to get non-linear optical properties for Langmuir-Blodgett films. The aim of the latter<br />

project was to monitor changes in molecular order <strong>and</strong> orientation when a spreaded film of<br />

arachidic acid (AA) is transferred from the water surface to a solid substrate in the Langmiur<br />

trough. For this aim three dyes of the phenylethenyl-pyridinium <strong>and</strong> of the benzaldehyde-hydrazone-type<br />

were dissolved in the AA film. Then surface second harmonic generation<br />

(SHG) was measured before <strong>and</strong> after transferring the films to a glass slide [48].<br />

In the context of this study both, SHG [49] <strong>and</strong> enhanced Raman scattering [50], from<br />

multilayer samples have been studied. These substrates, which consist of a silver isl<strong>and</strong> film<br />

separated from a silver mirror by a dielectric spacer layer, excel by remarkable reflectivity<br />

<strong>and</strong> absorption properties that have been accounted for by appropriate modelling [51].<br />

Polymers doped with triazene chromophores were used for sensitizing non-absorbing<br />

polymers, such as PMMA, for excimer laser ablation at 308 nm [52]. This project, which<br />

provided a strong link with the polymer-related activities described in Chapter III, resulted<br />

in the development of a new class of photosensitive polymers containing the triazene<br />

(–N=N–N–) functional group in the main chain [53].<br />

This Section is concluded by briefly mentioning structural investigations on a different<br />

type of gels, which were generated by network formation during hydrolysis of metal alkoxide<br />

precursors. Mixed oxides are materials of high technological interest <strong>and</strong> the sol-gel technique<br />

[54] is one of the most promising routes for preparing homogeneous materials at low<br />

process temperatures. In the present context, the use of these mixed oxide gels as catalysts<br />

or catalyst supports is of particular interest.<br />

Depending on the drying method employed, mixed oxides derived from the sol-gel<br />

process are designated either as xerogels (solvent evaporation) or as aerogels (supercritical<br />

drying). The properties of TiO 2 /SiO 2 -xerogels to be used as catalyst supports were studied<br />

by Schraml-Marth et al. [55], using vibrational spectroscopy. Walther et al. [56] have characterized<br />

the connectivity of vanadia <strong>and</strong> silica in V 2 O 5 /SiO 2 -xerogel catalysts by 29 Si<br />

MASNMR.<br />

Mixed oxide aerogels composed of titania, vanadia, <strong>and</strong> niobia have been shown to be<br />

highly active catalysts for the selective catalytic reduction (SCR) of nitric oxides with ammonia.<br />

These gels were characterized by FTIR, Raman, photoelectron spectroscopy, <strong>and</strong> secondary<br />

ion mass spectrometry [57–59].<br />

420


20.5 Remarks <strong>and</strong> acknowledgements<br />

20.5 Remarks <strong>and</strong> acknowledgements<br />

Spectroscopic tools are one of the links for our investigations of two seemingly different<br />

classes of materials, namely the surfactant systems <strong>and</strong> the biological polymers. Specific<br />

conclusions for each of the individual projects are presented in Sections 20.2.4 <strong>and</strong> 20.3.4,<br />

respectively. We now want to focus on some features common to both activities.<br />

In the studies of biopolymers, emphasis has been given to the development of methods<br />

that are sufficiently sensitive to detect small toposelective changes in large molecules often<br />

only available in small quantities. Vibrational <strong>and</strong> NMR spectroscopy have been successfully<br />

used to characterize conformational changes in oligonucleotides <strong>and</strong> chromphore-DNA interactions.<br />

Particular emphasis was devoted to the use of surface enhanced <strong>and</strong> resonant Raman<br />

scattering techniques.<br />

Structural characterization also formed the basis for the investigations of surfactants<br />

(SERS, dye probes) <strong>and</strong> of gel systems (NMR <strong>and</strong> vibrational spectroscopies). Here the<br />

main challenge was the monitoring of the dynamics of molecules <strong>and</strong> aggregates constituting<br />

the various investigated lyotropic mesophases. NMR self-diffusion measurements of micelles<br />

<strong>and</strong> solvents, diffusion of dye probes in laser induced gratings by FRS, <strong>and</strong> time-dependent<br />

fluorescence techniques were combined to unravel the mobilities of different entities.<br />

As in these mesoscopic systems a variety of motional processes is occurring simultaneously<br />

on vastly different time scales, the use of different spectroscopic techniques, delivering<br />

complementary information, was m<strong>and</strong>atory for a consistent interpretation of the<br />

fascinating dynamics of these systems.<br />

The results reported in this chapter have been obtained by the dedicated effort of past<br />

<strong>and</strong> present members of the group. The corresponding publications, mentioning the names<br />

of all persons involved, have been quoted in the respective sections. With regard to the<br />

NMR self-diffusion measurements the author is particularly indebted to J.-C. Panitz, K.L.<br />

Walther, Th. Schaller, <strong>and</strong> A. Sebald. FRS experiments have been carried out by G. Rehder<br />

<strong>and</strong> Ch. Hahn. Fluorescence decay by diffusing quenchers was studied by J. Wiesner,<br />

G. Meyer, <strong>and</strong> M. Klenke.<br />

The investigations of biopolymers <strong>and</strong> of DNA-chromophor interactions have been advanced<br />

by F. <strong>and</strong> B. Zimmermann. We are grateful to Th. Lippert for his valuable contributions.<br />

Members of the team involved in the other projects have been mentioned in<br />

Section 20.4.<br />

This spectrum of activities has only been made possible thanks to the fruitful interaction<br />

with colleagues in joint projects within the Sonderforschungsbereich. In particular, the<br />

author would like to express his gratitude to H. Hoffmann <strong>and</strong> M. Sprinzl for numerous discussions.<br />

Many of the reported investigations have been stimulated by their suggestions. We<br />

are indebted to C.D. Eisenbach for his support in developing the dye probes <strong>and</strong> to O. Nuyken<br />

for a fruitful collaboration on the subject of photopolymers.<br />

421


20 Spectroscopic Probes of Surfactant Systems <strong>and</strong> Biopolymers<br />

References<br />

1. K. Fontell: Coll. Polym. Sci., 268, 264 (1990)<br />

K. Fontell: J. Colloid Interface Sci., 43, 156 (1973)<br />

2. K. Fontell: Adv. Coll. Interface Sci., 41, 127 (1992)<br />

3. C. La Mesa, A. Khan, K. Fontell, B. Lindman: J. Colloid Interface Sci., 103, 373 (1985)<br />

4. D.A. Anderson, H. Wennerström: J. Phys. Chem., 94, 8683 (1990)<br />

5. G. Oetter, H. Hoffmann: Colloids Surf., 38, 225 (1989)<br />

6. K.L. Walther, M. Gradzielski, H. Hoffmann, A. Wokaun: J. Colloid Interface Sci., 153, 272 (1992)<br />

7. J.E. Tanner: J. Chem. Phys., 52, 2523 (1970)<br />

8. J.-C. Panitz, M. Gradzielski, H. Hoffmann, A. Wokaun: J. Phys. Chem., 98, 6812 (1994)<br />

9. P.O. Eriksson, G. Lindblom, E.E. Burnell, G.J. T. Tiddy: J. Chem. Soc. Faraday Trans. 1, 84, 3129<br />

(1988)<br />

10. H. Hoffmann, C. Thunig, P. Schmiedel, U. Munkert: to be published (1995)<br />

11. G. Hall, B. Whitaker: J. Chem. Soc. Faraday Trans., 90, 1 (1994)<br />

12. Ch. Hahn, A. Wokaun: Langmuir, 13, 391 (1997)<br />

13. M. Doi, S. Edwards: J. Chem. Soc. Faraday Trans., 74, 560 (1978)<br />

M. Doi, S. Edwards: J. Chem. Soc. Faraday Trans., 74, 918 (1978)<br />

14. P. Mariani, V. Luzzati, H. Delacroix: J. Mol. Biol., 204, 165 (1988)<br />

15. L. Scriven: in: K. Mittal (ed.): Micellization, Solubilization, <strong>and</strong> Microemulsions, Vol. 2, p. 877<br />

(1977)<br />

16. M. Gradzielski, H. Hoffmann, J.-C. Panitz., A. Wokaun: J. Colloid Interface Sci., 169, 103 (1995)<br />

17. P.P. Infelta, M. Grätzel, J.K. Thomas: J. Phys. Chem., 78, 190 (1974)<br />

18. P. Lianos, J. Lang, J. Sturm, R. Zana: J. Phys. Chem., 88, 819 (1984)<br />

19. C.S. Owen: J. Chem. Phys., 62, 3204 (1974)<br />

20. C.A. Miller, M. Gradzielski, H. Hoffmann, U. Krämer, C. Thunig: Prog. Colloid Sci., 84, 243<br />

(1991)<br />

21. C. Thunig, G. Platz, H. Hoffmann: Proc. Workshop on Structure <strong>and</strong> Conformation of Amphiphilic<br />

Membranes, Jülich, (1991)<br />

22. M. Almgren, J. Alsins, E. Mukhtar, J. van Stam: J. Phys. Chem., 92, 4479 (1988)<br />

23. J. Wiesner: PhD thesis, Universität Bayreuth (1992)<br />

24. M. Meyer: PhD thesis, Universität Bayreuth (1995)<br />

25. J. Alsins, M. Almgren: J. Phys. Chem., 94, 3062 (1990)<br />

26. G. Porte, J. Marignan, P. Bassereau, R. May: J. Phys. Paris, 49, 511 (1988).<br />

27. J.M. Seddon, J.L. Hogan, N.A. Warrender, E. Pebay-Peyroula: Progr. Colloid Polym. Sci., 81, 189<br />

(1990)<br />

28. W. Saenger: Principles of Nucleic Acid Structure, Springer, Berlin, (1989).<br />

29. T.M. Jovin, L.P. McIntosh, D.J. Arndt-Jovin, D.A. Zarling, M. Robert-Nicoud, J.H. van de S<strong>and</strong>e,<br />

K.F. Jorgenson, F. Eckstein: J. Biomol. Struct. Dyn., 1, 21 (1985)<br />

30. J. Weber, A. Wokaun: Mol. Phys., 74, 293 (1991)<br />

31. E. Koglin, J.-M. Sequaris: Topics in Current Chemistry, 134, 1 (1986)<br />

32. F. Zimmermann, B. Zimmermann, J.-C. Panitz, A. Wokaun: J. Raman Spectrosc., 26, 435 (1995)<br />

33. F. Zimmermann: PhD thesis, Universität Bayreuth (1994)<br />

34. T.M. Cotton, S.G. Schultz, R.P. Van Duyne: J. Am. Chem. Soc., 104, 6528 (1982)<br />

35. G. Smulevich, A. Feis: J. Phys. Chem., 90, 6388 (1986).<br />

36. T.F. Barton, R.P. Cooney, W.A. Denny: J. Raman Spectrosc., 23, 341 (1992)<br />

37. J. Aubard, M.A. Schwaller, J. Pantigny, J.P. Marsault, G.J. Lévy: J. Raman Spectrosc., 23, 373<br />

(1992)<br />

38. F. Zimmermann, B. Hossenfelder, J.-C. Panitz, A. Wokaun: J. Phys. Chem., 98, 12796 (1994)<br />

39. C. Zimmer, U. Wähnert: Prog. Biophys. Mol. Biol., 41, 31 (1986)<br />

422


References<br />

40. H.E. Howard-Lock, C.J.L. Lock, G. Turner, M. Zvagulis: Can. J. Chem., 59, 2737 (1981)<br />

41. J. Kritzenberger, F. Zimmermann, A. Wokaun: Inorg. Chim. Acta, 210, 47 (1993)<br />

42. J. Wiesner, A. Wokaun, H. Hoffmann: Prog. Colloid Polym. Sci., 76, 271 (1988)<br />

43. F. Zimmermann, A. Wokaun: Progr. Colloid Polym. Sci., 81, 242 (1990)<br />

44. P. Barnickel, A. Wokaun, W. Sager, F. Eicke: J. Colloid Interface Sci., 148, 80 (1992)<br />

45. W. Liptay: Z. Naturforsch., 18A, 1441 (1965)<br />

46. E. Lippert: Z. Elektrochem., 61, 962 (1957)<br />

47. Ch. Reichardt, E. Harbusch-Görnert: Liebigs Ann. Chem., 1983, 721 (1983)<br />

48. M. Klenke, H.G. Bingler, A. Wokaun: to be published (1998)<br />

49. H.G. Bingler, H. Brunner, M. Klenke, A. Leitner, F.R. Aussenegg, A. Wokaun: J. Chem. Phys., 99,<br />

7499 (1994)<br />

50. H.G. Bingler, H. Brunner, A. Leitner, F.R. Aussenegg, A. Wokaun: Mol. Phys., 85, 587 (1995)<br />

51. A. Leitner, Z. Zhao, H. Brunner, F.R. Aussenegg, A. Wokaun: Appl. Opt., 32, 102 (1993)<br />

52. Th. Lippert, A. Wokaun, J. Stebani, O. Nuyken, J. Ihlemann: Angew. Makromol. Chem., 213, 127<br />

(1993)<br />

53. Th. Lippert, J. Stebani, J. Ihlemann, O. Nuyken, A. Wokaun: J. Phys. Chem., 97, 12296 (1993)<br />

54. C.J. Brinker, G.W. Scherer: Sol-Gel Science, Academic Press, San Diego, (1990)<br />

55. M. Schraml-Marth, B.E. H<strong>and</strong>y, A. Wokaun, A. Baiker: J. Non-Cryst. Solids, 143, 93 (1992)<br />

56. K.L. Walther, A. Wokaun, A. Baiker: Mol. Phys., 71, 769 (1990)<br />

57. U. Scharf, M. Schneider, A. Baiker, A. Wokaun: J. Catal., 149, 344 (1994)<br />

58. M. Schneider, M. Maciejewski, S. Tschudin, A. Wokaun, A. Baiker: J. Catal., 149, 326 (1994)<br />

59. M. Schneider, U. Scharf, A. Wokaun, A. Baiker: J. Catal., 150, 284 (1994)<br />

423


21 Energy Transport by Lattice Solitons in a-Helical<br />

Proteins<br />

Franz-Georg Mertens, Dieter Hochstrasser, <strong>and</strong> Helmut Büttner<br />

21.1 Introduction<br />

For about 40 years the question of energy transport in muscle proteins has gained considerable<br />

attention. The biological energy quantum of 0.42 eV is given by the hydrolysis of adenosine<br />

triphosphate (ATP). It is assumed that this energy is transported practically without<br />

loss <strong>and</strong> is eventually used for the contraction of muscle fibers. The soliton concept can<br />

give an elegant answer to this question because here the dispersion of the energy can be prevented<br />

by non-linear effects.<br />

The fibers of striated muscles in vertebrates contain many myofibrils consisting of sarcomers.<br />

Each sarcomer consists of parallel-running thick <strong>and</strong> thin filaments. The basic mechanism<br />

for the muscle contraction consists in a sliding of the thin filaments relative to the<br />

thick ones [1]. This sliding can be described by phenomenological models consisting of a sequence<br />

of molecular processes [2, 3].<br />

The thick filaments consist of myosin molecules which resemble rods with a diameter<br />

of about 40 Å <strong>and</strong> a length of about 1600 Å. The myosin consists of two polypeptide chains<br />

forming a double a-helical structure. At one end there are globular heads where the ATP hydrolysis<br />

takes place.<br />

In each helix there are three chains of peptide groups coupled by hydrogen bonds.<br />

Since 1973 Davydov [4] developed a quantum theory for solitons on these hydrogen bonded<br />

chains. The idea is that the energy from the ATP hydrolysis leads to an excitation of the<br />

amide-I vibration in the first peptide groups at one end of the chain. This vibration has an<br />

energy of about 0.205 eV <strong>and</strong> an electric dipole moment of 0.30 debye which is directed approximately<br />

along the helix axis. By the dipole-dipole interaction the next peptide groups on<br />

the chain can be excited, <strong>and</strong> so on. However, in this way the energy would not be transported<br />

but only dispersed. The essential point in Davydov’s assumption is that the vibrational<br />

excitations are coupled non-linearly to a deformation of the hydrogen bonds. If the coupling<br />

exceeds a certain threshold, solitons can be formed which are solutions of a non-linear<br />

Schrödinger equation. Further assumptions must be made in order to explain how the solitons<br />

eventually produce the relative sliding between thick <strong>and</strong> thin filaments.<br />

Davydov’s model has been refined or modified by several authors, which we do not discuss<br />

here because the stability of the solitons seems questionable. Both, thermal [5] <strong>and</strong><br />

424 Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


21.1 Introduction<br />

quantum fluctuations [6], reduce the lifetime such that it may not be large enough for the biological<br />

energy transport. (However, for one of the modified models the stability against thermal<br />

fluctuations seems to be better [7].) Moreover, some doubts have appeared whether the<br />

non-linear Schrödinger equation can be derived properly from Davydov’s Hamiltonian [8].<br />

In 1984 Yomosa [9] proposed an alternative model which is much simpler than Davydov’s<br />

model because only the hydrogen bonds are involved in the energy transport. The peptide<br />

groups are rigid, i. e. they are represented by single masses which are coupled by<br />

strongly non-linear hydrogen bonds. Thus the energy is transported by lattice solitons. This<br />

has several advantages:<br />

a) The conditions for the occurrence of solitary waves on a one-dimensional lattice are<br />

very weak [5]. This means that every realistic interaction potential can be used, e. g. Lennard-Jones<br />

or Morse potentials. Solitons in the mathematical sense exist only for a completely<br />

integrable model, the Toda lattice, but for simplicity we will generally use the term soliton.<br />

b) Lattice solitons are non-topological, i. e. there is no energy gap. Thus the whole energy<br />

of a soliton can be converted into the mechanical work for the muscle contraction [9]. By<br />

contrast, Davydov solitons show an energy gap <strong>and</strong> only the kinetic part of the energy can<br />

be converted.<br />

c) Using lattice solitons, a molecular interpretation can be given [9] for the phenomenological<br />

rowing boat model [3] which describes the relative sliding between thick <strong>and</strong> thin filaments.<br />

d) Molecular dynamics simulations by Perez <strong>and</strong> Theodorakopoulos [11] have shown that<br />

even at room temperature lattice solitons are very stable against thermal fluctuations. Moreover,<br />

the lattice solitons are more stable than Davydov solitons if collisions between the two<br />

types of solitons are considered [11].<br />

e) Quantum mechanics is necessary only for the initial condition. The idea is that the energy<br />

quantum of 0.42 eV, released by the ATP hydrolysis, produces impulsively a pulse-like<br />

compression of the hydrogen-bonded lattice. According to the inverse scattering theory [12]<br />

an arbitrary initial pulse develops into a finite number of solitons plus a radiative background<br />

(phonons). This result holds for integrable classical systems. However, at least for<br />

one integrable quantum lattice model soliton-like excitations have been found. For the quantum<br />

Toda lattice in the strong-coupling regime there is a branch of excitations with a dispersion<br />

curve (energy vs. momentum) which is practically identical to that of the classical solitons,<br />

though the ground state shows large quantum fluctuations [13]. In fact, a quantum<br />

Toda lattice with parameters appropriate for the a-helix [9] can be shown to be in the<br />

strong-coupling, i. e. semiclassical, regime [14]. Reference [13] used the Bethe ansatz. This<br />

method has been complemented recently by a semiclassical quantization of the periodic<br />

Toda chain with practically identical results in the limit of an infinite chain [15].<br />

f) The reasoning of point (a) also holds for quantum lattice models with realistic interaction<br />

potentials. Such a model can always be replaced in a good approximation by a corresponding<br />

classical model with a renormalized potential [16], which fulfills the weak conditions<br />

for the existence of solitary waves. In the case of the Toda lattice the form of the potential<br />

remains unchanged <strong>and</strong> only the parameters are renormalized.<br />

425


21 Energy Transport by Lattice Solitons in a-Helical Proteins<br />

For the above reasons lattice solitons are very good c<strong>and</strong>idates for the energy transport.<br />

In this paper we generalize Yomosa’s model [9] in two ways:<br />

1) The peptide groups are no longer rigid. For simplicity we consider only one internal degree<br />

of freedom (Section 21.2) which leads already to interesting new features for the solitons<br />

(Section 21.4). The generalization to more degrees of freedom is straightforward<br />

(Section 21.7).<br />

2) Contrary to Yomosa, we do not work in the continuum limit. In fact, discreteness effects<br />

turn out to be decisive. These effects are first taken into account by applying the quasicontinuum<br />

approach (QCA) of Collins [17]. We use a rederivation of the approach in Fourier<br />

space [18] which offers several advantages (Section 21.3). However, at least for a part of the<br />

relevant energy range, the QCA is not sufficient (Section 21.5). Therefore we apply an iterative<br />

method [18] where the accuracy of taking into account the discreteness effects can be<br />

increased as much as necessary (Section 21.6). <strong>Final</strong>ly, we discuss the energy loss of the<br />

solitons due to the emission of optical phonons (Section 21.6). The results of the various<br />

sections are always checked by computer simulations.<br />

21.2 The model<br />

Following Yomosa [9], Perez, <strong>and</strong> Theodorakopoulos [11] we consider a one-dimensional<br />

lattice of hydrogen-bonded peptide groups. The non-linear interactions between neighbouring<br />

peptide groups are described by a suitable potential V, e. g. a Toda potential or a Lennard-Jones<br />

potential, with parameters from the literature (Section 21.5).<br />

In contrast to Refs. [9, 11] we do not neglect the internal vibrations of the peptide<br />

groups. In a first step we describe each group by two masses M 1 <strong>and</strong> M 2 coupled by a linear<br />

interaction with eigenfrequency O 0 . The resulting Lagrangian is<br />

L ˆ X <br />

1<br />

2 M 1 A _ 2 n ‡ 1 2 M 2 B _ 2 n VA … n‡1 B n †<br />

n<br />

<br />

1<br />

2 mO2 0 … B n A n † 2 ; …1†<br />

where m is the reduced mass; A n <strong>and</strong> B n are the displacements from the equilibrium positions<br />

of M 1 <strong>and</strong> M 2 of the n th peptide.<br />

The neglection of anharmonic terms for the internal vibrations is justified by the fact<br />

that the covalent bonds within the peptide groups are considerably stronger than the hydrogen<br />

bonds between the groups. Thus the relative displacements for the internal motion,<br />

D n ˆ … A n B n †=a ; …2†<br />

are expected to be much smaller than the relative displacements,<br />

426


21.2 The model<br />

j n ˆ … A n‡1 B n †=a ; …3†<br />

of the hydrogen bonds.<br />

D n <strong>and</strong> j n are defined in units of the equilibrium distance a of neighbouring hydrogen<br />

bonds, times in units of O 0 –1 , <strong>and</strong> energies in units of mO 0 2 a 2 . After this scaling we write the<br />

non-linear interaction in the form aV (j 2 n ), where the harmonic part of V has the form 1 2 j2 n.<br />

The dimensionless parameter a measures the strength of the non-linear interaction compared<br />

to the linear one. In this notation the equations of motion are<br />

j n ‡ a dV 1<br />

…<br />

dj n 1 ‡ m mD n ‡ D n‡1 † ˆ 0 …4†<br />

D n ‡ D n<br />

<br />

a<br />

1 ‡ m m dV ‡ dV <br />

ˆ 0<br />

dj n j n 1<br />

…5†<br />

with the mass ratio m = M 1 /M 2 .<br />

As D n appears only linearly, it can be eliminated which leads to a fourth order equation<br />

in time<br />

d 2<br />

dt 2<br />

<br />

j n ‡ j n ‡ a dV <br />

dj n<br />

<br />

‡ c 2 m a 2 dV<br />

dj n<br />

dV<br />

dj n‡1<br />

<br />

dV<br />

ˆ 0<br />

dj n 1<br />

…6†<br />

with<br />

c 2 m ˆ<br />

m<br />

…1 ‡ m† 2 ˆ m<br />

M ;<br />

…7†<br />

where M = M 1 + M 2 is the total mass.<br />

Linearization (dV/dj n % j n ) yields the dispersion curves of acoustic <strong>and</strong> optical phonons<br />

o 2 ‡ a<br />

…q† ˆ1<br />

2<br />

The sound velocity is<br />

( <br />

16a<br />

1 1<br />

q ) 1=2<br />

…1 ‡ a† 2 c2 m sin2 : …8†<br />

2<br />

c s ˆ<br />

r<br />

a<br />

c m :<br />

1 ‡ a<br />

…9†<br />

427


21 Energy Transport by Lattice Solitons in a-Helical Proteins<br />

21.3 Quasicontinuum approximation<br />

For the diatomic chain with non-linear nearest-neighbour interactions a st<strong>and</strong>ard decoupling<br />

technique in the continuum limit [19] can be used <strong>and</strong> yields acoustic pulse-type solitary<br />

waves <strong>and</strong> optical envelope-type solitary excitations [20]. These calculations are rather involved,<br />

even more for our model which has not only alternating masses but also alternating<br />

interactions. We prefer not to apply this method because there are two general problems:<br />

a) Only polynomial interaction potentials can be used; usually the realistic potentials are<br />

exp<strong>and</strong>ed up to the fourth order, assuming that the anharmonicities are small.<br />

b) In the st<strong>and</strong>ard continuum approximation the difference operator is exp<strong>and</strong>ed up to a<br />

certain order which can lead to an ill-posed Cauchy problem [21]. This difficulty occurs because<br />

the dispersion due to the discreteness of the system is not taken into account consistently.<br />

For the monoatomic chain both problems have been overcome by the QCA of Collins<br />

[17] for solitary waves <strong>and</strong> periodic modes. Here the difference operator is inverted instead<br />

of exp<strong>and</strong>ed. Rosenau [21] developed a still more general approximation scheme which reduces<br />

to the result of Collins in the case of solitary waves.<br />

For the diatomic chain Collins [22] considered only pulse-like solitary waves. We are<br />

not interested here in the optical solitons because for the energy transport the pulse solitons<br />

are much better c<strong>and</strong>idates:<br />

1) They are supersonic, in contrast to the optical ones.<br />

2) The mechanism needed for the conversion of the energy into a shortening of muscle fibers<br />

seems to be simple only for pulse solitons [9].<br />

Since the QCA overcomes the above-mentioned problems we now apply it to our diatomic<br />

model with alternating interactions. However, we use a rederivation of the QCA in<br />

Fourier space [18]. This formulation is much simpler than the original one [17] <strong>and</strong> can<br />

easily be generalized from the monoatomic chain to our model.<br />

We are interested in solitary waves<br />

j n …t† ˆj…n ct† ˆj…z† …10†<br />

with velocity c, satisfying decaying boundary conditions. Because of these conditions the<br />

Fourier transform exists,<br />

~j…q† ˆ R1 dz e iqz j…z† ;<br />

1<br />

…11†<br />

<strong>and</strong> analogously ~ D (q) for D (z). Moreover, we define the force<br />

428


21.3 Quasicontinuum approximation<br />

F…z† ˆa<br />

dV<br />

dj…z†<br />

…12†<br />

<strong>and</strong> its Fourier transform ~ F (q).<br />

The equations of motion now read<br />

c 2 q 2 ~j ‡ ~F<br />

1<br />

1 ‡ m eiq ‡ m ~ D ˆ 0 ; …13†<br />

<br />

c 2 q 2 ‡ 1 ~D<br />

1<br />

1 ‡ m e iq ‡ m ~F ˆ 0 : …14†<br />

The elimination of ~ D yields<br />

c 2 q 2 1 c 2 q 2 <br />

~j…q† ˆ 4c 2 q <br />

m sin2 c 2 q 2 F…q† ~ ; …15†<br />

2<br />

which can also be obtained directly from Eq. 8.<br />

According to Ref. [18] the QCA now consists in the following procedure. We write<br />

Eq. 15 as<br />

A…q† ~j…q† ˆ ~F…q†<br />

…16†<br />

<strong>and</strong> exp<strong>and</strong> A (q) in a Taylor series for |q|


21 Energy Transport by Lattice Solitons in a-Helical Proteins<br />

with<br />

V ef f ˆ aV…j†<br />

1<br />

2 a 0 j 2 ; …21†<br />

where we have set V (0) = 0.<br />

Because of c 2 m ^ 1/4, a 2 in Eq. 19 is always positive <strong>and</strong> can be interpreted as an effective<br />

mass. Equation 20 has pulse-like, solitary solutions if there is a range of j for which<br />

V eff (j) ^ 0, where the equality must hold at the boundaries of this range. For the considered<br />

intermolecular potentials we have V eff (j) ^0 for a negative range j 1 ^j^0, with a 0 >0<br />

<strong>and</strong> a < a 0 (because V % j 2 /2 for small j). These conditions lead to<br />

c s c c m<br />

…22†<br />

which means we get supersonic, compressional solitary waves with amplitude j 1 . The meaning<br />

of the upper limit c m will be discussed in the next Section.<br />

The shape j (z) of the pulses is obtained by the integration of Eq. 20:<br />

z…j† ˆ<br />

Z j<br />

dj 0<br />

p : …23†<br />

2V ef f …j 0 †=a 2<br />

j 1<br />

For the Toda or Lennard-Jones potential this integral can only be calculated numerically.<br />

However, if we use an expansion of these potentials up to the fourth order<br />

V…j† ˆ1<br />

2 j2 2bj 3 ‡ 2gj 4 ; …24†<br />

where 9 b 2


21.4 Velocity range for the quasicontinuum approach<br />

~D…q† ˆm ‡ e iq<br />

1 ‡ m<br />

c 2 q 2<br />

~j…q† : …28†<br />

4c 2 m sin2 …q=2† c 2 q2 In order to be consistent with the approximations leading to Eq. 17, we exp<strong>and</strong> Eq. 28<br />

to order q 2 <strong>and</strong> perform an inverse Fourier transformation, which yields<br />

<br />

D…z† ˆa 0 j…z†<br />

1<br />

1 ‡ m j0 …z†‡<br />

<br />

1<br />

21‡ … m†<br />

a 0 c 2 <br />

m<br />

12c 2 j 00 …z† : …29†<br />

21.4 Velocity range for the quasicontinuum approach<br />

There are restrictions for the velocity of the solitons, both in principle <strong>and</strong> for technical reasons.<br />

In Eq. 17 the function<br />

A…q† ˆ<br />

c 2 q 2 …1 c 2 q 2 †<br />

4c 2 m sin2 …q=2† c 2 q 2 …30†<br />

has been exp<strong>and</strong>ed in a Taylor series with the implication that the Fourier transform ~j (q) of<br />

the solitary wave is negligible for |q|6 q c . Here q c is the radius of convergence, defined by<br />

the first non-trivial solution of<br />

2c m sin q ˆ cq :<br />

2<br />

…31†<br />

In Fig. 21.1 this condition is visualized as the intersection of the straight line cq with<br />

the dispersion curve o m (q) =2c m |sin (q/2)|. Going back to the original units, o m can be<br />

identified as the dispersion of a monoatomic chain of masses M = M 1 + M 2 with a linear interaction<br />

with coupling constant mO 2 0. For this reason we see now that the upper limit c m for<br />

the soliton velocity in Eq. 22 results from the linear interaction in our diatomic model with<br />

alternating linear <strong>and</strong> non-linear interactions. For the usual diatomic model there is no upper<br />

limit for the velocity [22].<br />

In practice, i. e. for solitons with a finite width, the condition (Eq. 22) must be tightened<br />

to<br />

c s


21 Energy Transport by Lattice Solitons in a-Helical Proteins<br />

Figure 21.1: Phonon dispersion curves. o + (q): optical, o – (q): acoustic, o m (q): monoatomic (from<br />

Eq. 31). The intersections with the straight line o = cq are discussed in Sections 21.4 <strong>and</strong> 21.6. Parameters:<br />

a = 0.6, m =2,c/c s = 1.3.<br />

range of soliton velocities satisfying both conditions in Eq. 32 we must dem<strong>and</strong> a P 1. In<br />

fact, this is fulfilled for our model since the hydrogen bonds between the peptide groups are<br />

considerably weaker than the covalent bonds within these groups (a is the ratio of the coupling<br />

strengths, see Section 21.2). In Section 21.5 we will give estimates for a.<br />

So far we have discussed only the limitations which result from the finite radius of<br />

convergence for the Taylor series of A (q). Moreover, we must test whether the truncation<br />

(Eq. 17) behind the second term of the series is justified. For a given mass ratio m, i. e. for<br />

given c m , we choose a velocity c satisfying Eq. 32 <strong>and</strong> estimate the q range for which the<br />

truncated expansion agrees well with A (q). Then we choose a potential V <strong>and</strong> calculate the<br />

solitary wave j (z) which belongs to c. Its Fourier transform ~j (q) must be negligible outside<br />

the above-mentioned q range. The results are given in the next Section.<br />

21.5 Solitary waves for realistic parameter values<br />

The lattice constant a for the H-bonded peptide chains in the a-helix is about 4.5 Å [9]. As a representative<br />

value for the eigenfrequencies of a peptide group we choose O 0 = 3.11610 14 s –1<br />

from the amide-I vibration. The total mass M = M 1 + M 2 corresponds to the mass of a peptide<br />

group plus an average residue in a muscle protein, which gives together about 100 proton<br />

masses [9, 10]. The mass ratio m = M 1 /M 2 is kept as a free parameter which is varied between 1<br />

<strong>and</strong> 10 (our results are invariant under the transformation m 7 ! 1/m).<br />

We model the hydrogen bonds between neighbouring peptide groups by a suitable<br />

non-linear interaction, e. g. a Toda potential with parameters fitted to an ab initio self-consistent-field<br />

molecular-orbital calculation for an H bond in a formamide dimer [9]. In our dimensionless<br />

units this corresponds to<br />

432


21.5 Solitary waves for realistic parameter values<br />

aV T …j† ˆ a <br />

b 2 exp… bj†‡bj 1<br />

…33†<br />

with b = 18 <strong>and</strong> a = 0.00123/c m 2 . Note that in Section 21.2 the interaction was introduced in<br />

the form aV, where V = j 2 /2 for j?0. With these parameters the sound velocity is<br />

c s % 4900 m/s for 1 ^m^10.<br />

As a second example we take a Lennard-Jones potential with parameters fitted to the<br />

equilibrium distance a <strong>and</strong> the bond energy [11, 23], which gives<br />

V LJ …j† ˆ 1 <br />

<br />

1 2<br />

72 …1 ‡ j† 12 …1 ‡ j† 6 ‡ 1<br />

<strong>and</strong> a = 0.000811/c m 2 . The corresponding sound velocity is about 4000 m/s for 1^m^10.<br />

For fixed mass ratio m, i. e. for fixed c m , we choose a velocity c in the range of Eq. 32<br />

calculating the corresponding solitary wave j (z) <strong>and</strong> its Fourier transform ~j (q). A first test<br />

shows that ~j (q c )/~j (0) is indeed negligible for a large range of velocities (about 30% above<br />

c s ), as expected from the discussion of the radius of convergence q c in Section 21.4.<br />

However, this large velocity range is considerably reduced by the second test described<br />

in Section 21.4. For the potentials <strong>and</strong> parameter values used here the QCA is valid for velocities<br />

c which do not exceed the sound velocity by more than about 5 to 10%. Eventually<br />

we perform a final test by comparing them with the results of a computer simulation. If we<br />

take Eq. 23 <strong>and</strong> Eq. 29 as initial conditions <strong>and</strong> integrate the difference-differential Eqs. 4<br />

<strong>and</strong> 5 numerically then the time evolution of a single pulse (Fig. 21.2) shows that the QCA<br />

solution is a good approximation. Only very small oscillations (phonons) appear immediately<br />

after the start. After a while these phonons are left behind <strong>and</strong> the pulse travels without<br />

changing its shape (Section 21.6). Figure 21.3 shows the scattering of two solitons.<br />

…34†<br />

Figure 21.2: Computer simulation of time-evolution of a single pulse for a chain of 200 unit cells with<br />

m = 1 <strong>and</strong> the Toda potential (Eq. 33). As initial condition a QCA solution with c/c s = 1.05 is used.<br />

After these tests we turn to the essential question whether the solitons of the QCA can<br />

transport enough energy. Figure 21.4 shows the energy E of a single soliton as a function of<br />

c/c s , for Toda <strong>and</strong> Lennard-Jones potentials. E must be compared with the biological energy<br />

quantum of 0.42 eV (Section 21.1). Yomosa [9] assumed that these quanta set the initial<br />

conditions for the lattice solitons. Naturally, an arbitrary initial condition generally produces<br />

several solitons plus an oscillatory background. Knowing that nature usually works fairly ef-<br />

433


21 Energy Transport by Lattice Solitons in a-Helical Proteins<br />

Figure 21.3: Collision of two QCA solitons, same parameters as in Fig. 21.2.<br />

Figure 21.4: Soliton energy vs. velocity. QCA results for Toda (solid line) <strong>and</strong> for Lennard-Jones interaction<br />

(dashed line). Iterative solutions for Toda (x) <strong>and</strong> for Lennard-Jones interaction (y).<br />

fectively, <strong>and</strong> in order to be on the safe side, let us assume that the energy of one quantum<br />

essentially goes into one or two solitons. Then we see from Fig. 21.4 that we need a velocity<br />

of at least 1.2c s for which the QCA clearly is no longer valid. Naturally, this velocity is only<br />

a rough estimate that depends on the energy unit mO 2 0 a 2 , i. e. on our choice of O 0 . But in<br />

any case we need a new method in order to h<strong>and</strong>le the case of larger velocities. This is treated<br />

in the next Section.<br />

21.6 Iterative method <strong>and</strong> stability<br />

With increasing velocity <strong>and</strong> energy the solitary waves become narrower <strong>and</strong> their width can<br />

be in the order of the lattice constant. In this case the QCA does not hold. And it would not<br />

help to take more terms of the Taylor expansion of A (q) in Eq. 16 into account. The resulting<br />

higher order differential equations cannot be integrated like Eq. 17. Moreover, even the<br />

infinite Taylor series cannot fully represent A (q) because of the finite radius of convergence.<br />

Fortunately, an iterative method [18] has been developed for the monoatomic chain.<br />

Here the accuracy of taking into account the discreteness effects is increased systematically.<br />

In the case of the Toda lattice the iteration converges to the exact one-soliton solution.<br />

434


21.6 Iterative method <strong>and</strong> stability<br />

This method can also be applied to the pulse-type solitary waves of our diatomic<br />

model. Our basic Eq. 15 can be written in the form<br />

~j…q† ˆA 1 …q† ~F …q† ;<br />

…35†<br />

which already suggests an iteration because ~F (q) is the Fourier transform of F (q), (Eq. 12).<br />

However, instead of Eq. 35 we use a slightly different form which will turn out later to be<br />

more convenient. We first split the linear part of the force (Eq. 12) from the non-linear part,<br />

denoted by G,<br />

<br />

F …j…z†† ˆ aj…z†‡G……j…z†† : …36†<br />

Then we insert Eq. 36 into Eq. 15 <strong>and</strong> isolate ~j<br />

<br />

a c 2 q 2 o 2 m<br />

~j…q† ˆ<br />

…q†<br />

<br />

~G…q†<br />

c 2 q 2 o 2 ; …37†<br />

‡ …q† c2 q 2 o 2 …q†<br />

where o + (q) <strong>and</strong> o m (q) are the dispersion curves from Eq. 8 <strong>and</strong> Eq. 31, respectively.<br />

Similar to the monoatomic case [18], an iteration for Eq. 37 would converge only to<br />

the trivial solution j (z) : 0. This can be prevented by keeping ~j (0) constant during the<br />

iteration [18]. This condition leads to the elimination of c from Eq. 37 <strong>and</strong> thus to a new<br />

iteration procedure<br />

with<br />

<br />

a c 2 i<br />

~j i‡1 …z† ˆ<br />

q2 o 2 m …q†<br />

<br />

~G<br />

o 2 ‡ …q† c<br />

2<br />

i q 2 o 2 i …q† …38†<br />

…q†<br />

c 2 i q2<br />

c 2 i ˆ c 2 s<br />

~j…0†‡ ~G i …0†<br />

~j…0†‡c 2 s =c2 m ~ G i …0† :<br />

…39†<br />

This implies c s


21 Energy Transport by Lattice Solitons in a-Helical Proteins<br />

c i changes during the iteration <strong>and</strong> converges towards a value which usually is considerably<br />

lower than the initial one. By the way, this behaviour is qualitatively similar to a computer<br />

simulation. When starting with an approximate solution as initial condition, an adaptation<br />

to the lattice by the emission of phonons is observed. The resulting solitary wave has a<br />

lower velocity (<strong>and</strong> energy) than the initial wave. For the potentials <strong>and</strong> parameters, used<br />

here, a sufficient accuracy is achieved after at most 8 iterations. As a test, these results were<br />

used as input for a computer simulation. Contrary to the QCA result for a velocity outside<br />

of the validity range (Fig. 21.5a), the shape of the pulse remains unchanged, no adaptation<br />

to the lattice is observed (Fig. 21.5b). The soliton character of the pulses is also seen clearly<br />

in scattering experiments, even in the highly discrete regime (Fig. 21.6). Figure 21.4 shows<br />

the soliton energies. They are considerably higher than the QCA results for the same velocities.<br />

Figure 21.5: Computer simulations for initial conditions with c/c s = 1.28 for the QCA (a) <strong>and</strong> the iterative<br />

solution (b), using the Toda potential <strong>and</strong> m =1.<br />

However, a complete convergence cannot be achieved, i. e. an exact solitary wave solution<br />

probably does not exist. In fact, contrary to the monoatomic case [18], our iteration procedure<br />

(Eq. 38) shows a pole, namely at the solution q 0 of<br />

c i q ˆ o ‡ …q† :<br />

…41†<br />

The existence of the pole means that Eq. 38 makes sense only if ~G (q) <strong>and</strong> ~j (q) are<br />

negligible for |q| 6 q 0 . In practice a cut-off is necessary in each step of the iteration.<br />

Contrary to the finite convergence radius q c which limits the validity range of the<br />

QCA (Section 21.4), the occurrence of the pole is not a technical problem but is connected<br />

with a physical effect. Peyrard et al. [24] showed by computer simulations <strong>and</strong> theoretical investigations<br />

that a solitary wave with velocity c cannot be stable if the line cq has an intersection<br />

with one of the phonon branches o (q). The solitary wave permanently looses energy<br />

by the emission of phonons with frequency spectrum centred around o(q 0 ) where q 0 is the<br />

intersection point. Due to the energy loss the velocity c decreases gradually <strong>and</strong> the width of<br />

the wave increases. Thus the discreteness effects which are responsible for the phonon emission<br />

become smaller <strong>and</strong> smaller. Eventually the solitary wave is practically stable because<br />

the energy loss is negligible.<br />

This scenario holds for our model, too. The condition (Eq. 41) corresponds to the intersection<br />

of c i q with the optical phonon branch (Fig. 21.1). In a computer simulation two<br />

very different time scales appear. Starting with any initial condition including the QCA re-<br />

436


21.7 Conclusion<br />

Figure 21.6: Computer simulation of the collision of two very narrow solitons (c/c s = 1.9) from the<br />

iterative solution for the Toda potential <strong>and</strong> m =1.<br />

sult the above-mentioned adaptation to the lattice only takes a rather short time (Fig. 21.5a).<br />

But this adaptation does not take place when the result from our iterative method is used as<br />

initial condition (Fig. 21.5 b). Contrary to the adaptation process, the emission of optical<br />

phonons occurs for a much longer time, which in principle goes on forever, while the solitary<br />

wave disappears asymptotically.<br />

However, this effect is extremely small when a-helix parameters are used. The situation<br />

can be illustrated by drawing Fig. 21.1, now using the a-helix parameters of<br />

Section 21.5. As a is very small there is a wide gap between the acoustic <strong>and</strong> optical phonon<br />

branches <strong>and</strong> the intersection point q 0 is far outside of the first Brillouin zone (e. g. q 0 % 5p<br />

for c % 2 c s ). Therefore the above condition for the convergence of the iteration, i. e. ~j (q 0 )<br />

<strong>and</strong> ~G (q 0 ) are negligible, is indeed very well fulfilled <strong>and</strong> the energy loss due to emission<br />

of optical phonons is negligible.<br />

21.7 Conclusion<br />

Lattice solitons remain good c<strong>and</strong>idates for the explanation of the energy transport in the<br />

a-helix if an internal vibration mode of the peptide groups is taken into account. Our theory<br />

can easily be generalized for a lattice with a basis of more than two atoms, e. g. three masses<br />

representing a peptide group N–C=O. The neglection of anharmonicities for the internal vibrations<br />

is the essential point which allows the elimination of coordinates in the equations of<br />

motion. This neglection is justified because the covalent bonds within the peptide groups are<br />

considerably stronger than the hydrogen bonds between the peptides.<br />

Discreteness effects are very important. They are partially incorporated by the quasicontinuum<br />

approximation. However, for the parameters of the a-helix, a systematic implementation<br />

of the discreteness is necessary, which is achieved by an iterative method. The<br />

convergence of this method is limited in principle by an instability of the solitons due to the<br />

emission of optical phonons. However, this effect turns out to be negligible.<br />

In our opinion mainly two questions remain:<br />

437


21 Energy Transport by Lattice Solitons in a-Helical Proteins<br />

a) Stability of solitons<br />

The coupling between the three hydrogen bonded chains of an a-helix possibly reduces or<br />

destroys the stability of the solitons on the chain. So far this problem has not yet been<br />

tackled analytically. Molecular dynamics simulations, which include transversal fluctuations,<br />

do not show stable solitons [25]. On the other h<strong>and</strong>, we have performed simulations for a<br />

Toda chain that branches into two Toda chains with different parameters [26]. Here we<br />

choose the parameters of the first branch so that the interactions become strong <strong>and</strong> nearly<br />

linear, representing the covalent bonds between the hydrogen-bonded chains. The parameters<br />

of the second branch are chosen similar to the original chain, namely representing the weak,<br />

very non-linear hydrogen bonds. The simulations show that a soliton, which is initiated on<br />

the original chain, prefers to go into the second branch. Very little energy is lost via the first<br />

branch, unless the energy of the soliton is very small. This is an interesting result, but naturally<br />

it does not prove the stability of solitons on the very complicated structure of three<br />

coupled chains.<br />

b) Detection of solitons<br />

In contrast to the topological solitons, it was not clear for a long time whether non-topological<br />

lattice solitons yield a clear-cut signature in the dynamic form factor S (q,o), which can<br />

be measured by inelastic neutron scattering. Recently combined Monte Carlo molecular<br />

dynamics simulations yielded only one peak in S(q,o) for the thermal equilibrium <strong>and</strong> the<br />

soliton <strong>and</strong> phonon contributions could not be distinguished [27]. However, if certain nonequilibrium<br />

configurations are used as initial condition for the molecular dynamics, additional<br />

small peaks on the high-frequency side of the main peak appear which can be identified<br />

with lattice solitons. Therefore, the question now is whether the ATP hydrolysis produces<br />

such initial conditions.<br />

References<br />

1. H.E. Huxley, J. Hanson: Nature, 173, 973 (1954)<br />

H.E. Huxley, J. Hanson, in: Structure <strong>and</strong> Function of Muscle, Academic, New York, p. 183<br />

(1960)<br />

A.F. Huxley, R. Niedergerke: Nature, 173, 971 (1954)<br />

A.F. Huxley: Proc. Biophys. 7, 255 (1957)<br />

2. A.F. Huxley, A. Simmons: Nature, 233, 533 (1971)<br />

3. J.M. Murray, A.Weber: Sci. Amer., 230, 1974 (1974)<br />

4. A.S. Davydov: J. Theor. Biol., 38, 559 (1973)<br />

A.S. Davydov: Stud. Biophys., 47, 221 (1974)<br />

A.S. Davydov: Phys. Scr., 20, 387 (1979)<br />

A.S. Davydov: Biology <strong>and</strong> Quantum Mechanics, Pergamon, New York, (1982)<br />

5. P.S. Lomdahl, W.C. Kerr: Phys. Rev. Lett., 55, 1235 (1985)<br />

A.F. Lawrence, J.C. McDaniel, D.B. Chang, B.M. Pierce, R.R. Birge: Phys. Rev. A, 33, 1188<br />

(1986)<br />

6. H. Bolterauer, M. Opper, Z. Phys. B, 82, 95 (1991)<br />

438


References<br />

7. L. Cruzeiro, J. Halding, P.L. Christiansen, O. Skovgaard, A.C. Scott: Phys. Rev. A, 37, 880 (1988)<br />

8. D.W. Brown, K. Lindenberg, B.J. West: Phys. Rev. A, 33, 4104 (1986)<br />

D.W. Brown, B.J. West, K. Lindenberg: Phys. Rev. A, 33, 4110 (1986)<br />

9. S. Yomosa: J. Phys. Soc. Japan, 53, 3692 (1984)<br />

S. Yomosa: Phys. Rev. A, 32, 1752 (1985)<br />

10. M.K. Ali, R.L. Somorjai: J. Phys. A, 12, 2291 (1979)<br />

J.R. Rolfe, S.A. Rice, J. Dancz: J. Chem. Phys., 70, 26 (1979)<br />

M.A. Collins: Chem. Phys. Lett., 77, 342 (1981)<br />

H. Bolterauer, in: J.T. Devreese, L.F. Lemmens, V.E. van Doren (eds): Recent Developments in<br />

Cond. Matter Physics,Vol. 4, Plenum, New York, p. 199 (1981)<br />

11. P. Perez, N. Theodorakopoulos: Phys. Lett., 117 A, 405 (1986)<br />

P. Perez, N. Theodorakopoulos: Phys. Lett., 124 A, 267 (1987)<br />

12. C.S. Gardner, J.M. Greene, M.D. Kruskal, R.M. Miura: Phys. Rev. Lett., 19, 1095 (1967)<br />

13. F.G. Mertens, M. Hader, in: S. Takeno (ed.): Dynamical Problems in Soliton Systems, Springer<br />

Series in Synergetics, Vol. 30, Springer, Berlin, p. 89 (1985)<br />

14. D. Hochstrasser, F.G. Mertens, H. Büttner: Phys. Rev. A, 40, 2602 (1989), Appendix A<br />

15. F. Göhmann, W. Pesch, F.G. Mertens: J. Phys. A, 26, 7589 (1993)<br />

16. F. Göhmann, F.G. Mertens: J. Phys. A, 25, 649 (1992)<br />

17. M.A. Collins: Chem. Phys. Lett., 77, 342 (1981)<br />

M.A. Collins, S.A. Rice: J. Chem. Phys., 77, 2607 (1982)<br />

18. D. Hochstrasser, F.G. Mertens, H. Büttner, Physica D, 35, 259 (1989)<br />

19. H. Büttner, H. Bilz, in: A.R. Bishop, T. Schneider (eds.): Solitons in Cond. Matter Physics,<br />

Springer, Berlin, p. 162 (1978)<br />

20. St. Pnevmatikos, M. Remoissenet, N. Flytzanis: J. Phys. C, 16, L305 (1983)<br />

St. Pnevmatikos, N. Flytzanis, M. Remoissenet: Phys. Rev. B, 33, 2308 (1986)<br />

21. P. Rosenau: Phys. Lett. A, 118, 222 (1986)<br />

Phys. Rev. B, 36, 5868 (1987)<br />

22. M.A. Collins: Phys. Rev. A, 31, 1754 (1985)<br />

23. M. Levitt: J. Mol. Biol., 168, 595 (1983)<br />

24. M. Peyrard, St. Pneuvmatikos, N. Flytzanis: Physics, 19 D, 268 (1986)<br />

25. O.H. Olsen, P.S. Lomdahl, W.C. Kerr: Phys. Lett., A 136, 402 (1989)<br />

26. D. Hochstrasser: PhD thesis, Universität Bayreuth (1989)<br />

27. A. Neuper, F.G. Mertens, in: M. Peyrard (ed): Nonlinear Excitations in Biomolecules, Les Editions<br />

de Physique, Les Ulis <strong>and</strong> Springer, Berlin, (1995)<br />

439


IV<br />

Appendix<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1


22 Documentation of the Collaborative Research Centre 213<br />

Chairmen<br />

Prof. Dr. Markus Schwoerer<br />

Prof. Dr. Heinz Hoffmann<br />

22.1 List of Members<br />

Name Institute Membership<br />

Blumen, A. Experimentalphysik 1987–1991<br />

Büttner, H. Theoretische Physik 1984–1995<br />

Dormann, E. Experimentalphysik 1984–1990<br />

Eisenbach, C.D. Makromolekulare Chemie 1988–1995<br />

Faulhammer, H. Biochemie 1984–1989<br />

Fesser, K.** Theoretische Physik 1984–1995<br />

Friedrich, J. Experimentalphysik 1984–1989<br />

Friedrich, J. Experimentalphysik 1993–1995<br />

Haarer, D. Experimentalphysik 1984–1995<br />

Höcker, H. Makromolekulare Chemie 1984–1986<br />

Hoffmann, H. Physikalische Chemie 1984–1995<br />

Kalus, J. Experimentalphysik 1984–1992<br />

Kiefer, W. Experimentalphysik 1984–1986<br />

Kramer, L. Theoretische Physik 1984–1995<br />

Krauss, H.L. Anorganische Chemie 1984–1992<br />

Lattermann, G. Makromolekulare Chemie 1989–1995<br />

Laubereau, A. Experimentalphysik 1984–1995<br />

Mertens, F.G. Theoretische Physik 1987–1995<br />

Nuyken, O. Makromolekulare Chemie 1988–1992<br />

Pobell, F. Experimentalphysik 1987–1995<br />

Richter, W. Experimentalphysik 1990–1995<br />

Rösch, P. Struktur und Chemie der Biopolymere 1993–1995<br />

* “Promotion” supported by the Sonderforschungsbereich 213<br />

** “Habilitation” supported by the Sonderforschungsbereich 213<br />

Macromolecular Systems: <strong>Microscopic</strong> <strong>Interactions</strong> <strong>and</strong> <strong>Macroscopic</strong> <strong>Properties</strong><br />

Deutsche Forschungsgemeinschaft (DFG)<br />

Copyright © 2000 WILEY-VCH Verlag GmbH, Weinheim. ISBN: 978-3-527-27726-1<br />

443


22 Documentation of the Collaborative Research Centre 213<br />

Name Institute Membership<br />

Schmidt, M. Makromolekulare Chemie 1993–1995<br />

Schwoerer, M. Experimentalphysik 1984–1995<br />

Seilmeier, A. Experimentalphysik 1993–1995<br />

Spiess, H. W. Makromolekulare Chemie 1984–1986<br />

Sprinzl, M. Biochemie 1984–1995<br />

Strohriegl, P. Makromolekulare Chemie 1989–1995<br />

Wokaun, A. Physikalische Chemie 1987–1995<br />

Rentsch, S. Optik- und Quantenelektronik 1992–1995<br />

University of Jena<br />

22.2 Heads of Projects (Teilprojektleiter)<br />

22.2.1 Projektbereich A: Gemeinsame Einrichtungen<br />

A 1 FTIR-Gerät Krauss 1984–1992<br />

A 5 Rasterelektronenmikroskopie Haarer 1987–1995<br />

A 6 Transmissionselektronenmikroskopie Eisenbach 1989–1995<br />

22.2.2 Projektbereich B: Festkörper<br />

B 1 ENDOR in Diacetylenkristallen: Schwoerer 1984–1986<br />

die molekulare Elektronik der topochemischen Reaktion<br />

B 2 UV-Holographie und das Wachstum von Schwoerer 1984–1989<br />

Mikrostrukturen in Diacetylen-Einkristallen<br />

B 3 Photoleitung und Redox-Photochemie Molekularer Haarer 1984–1989<br />

und hochpolymerer Festkörper<br />

B 4 Disubstituierte Diacetylene mit besonderen Dormann 1984–1990<br />

di- und pyroelektrischen Eigenschaften<br />

B 5 Kommensurable und inkommensurable Büttner 1984–1986<br />

strukturelle Übergänge in Charge-Transfer-Kristallen<br />

B 6 Nichtlineare Anregungen in konjugierten Polymeren Fesser 1984–1995<br />

B 7 Untersuchung elementarer Anregungen mit Kalus 1984–1992<br />

hochausgelöster inelastischer Röntgenstreuung<br />

444


22.2 Heads of Projects (Teilprojektleiter)<br />

B 8 Elektron-Phonon-Wechselwirkung und elektronische Büttner 1990–1995<br />

Korrelation in quasi-eindimensionalen Charge-<br />

Transfer Systemen<br />

B 9 Lochbrennspektroskopie an excitonischen Friedrich 1987–1989<br />

Zuständen in Gläsern<br />

B 10 Spektroskopie und Kalorimetrie an polymeren Haarer/Pobell 1987–1995<br />

und nicht-kristallinen Substanzen<br />

B 11 Hierarchische Modelle zum Ladungs- und Blumen 1987–1991<br />

Massentransport in ungeordneten Medien<br />

B 12 Lochbrenn-Spektroskopie in Polymeren: Haarer 1990–1995<br />

Elektrische Feldeffekte und Druckeffekte<br />

B 13 Nichtlineare Optik in Makromolekülsystemen: Schwoerer 1990–1995<br />

Materialien und Bauelemente<br />

B 14 Lichtinduzierte spektrale Diffusion von Richter 1990–1995<br />

Farbstoffmolekülen in glasartigen Matrizen<br />

B 15 Lochbrennspektroskopie an excitonischen Friedrich 1992–1995<br />

Zuständen von Aggregatketten<br />

22.2.3 Projektbereich C: Funktionale Systeme – Mizellen, Oberflächen<br />

und Polymere<br />

C 1 Viskoelastische Tensidlösungen Hoffmann 1984–1995<br />

C 2 Lyotrope nematische und cholesterische Mesophasen Hoffmann 1984–1995<br />

C 3 Flüssigkristalline Haupt- und Seitenkettenpolymere Höcker 1984–1986<br />

C 4 Stabilität und Selektion in den strukturbildenden Kramer 1984–1995<br />

Instabilitäten von flüssigen Kristallen<br />

C 5 Struktur und Reaktivität von koordinativ Krauss 1984–1990<br />

ungesättigten Oberflächenverbindungen<br />

C 6 CARS-Spektroskopie an fluoreszierenden Materialien Kiefer 1984–1986<br />

C 7 Darstellung und Untersuchung von Oligomeren Höcker 1984–1986<br />

und Polymeren mit definierter Struktur und<br />

besonderen Eigenschaften<br />

C 8 NMR-Untersuchungen der molekularen Bewegung Spiess 1984–1986<br />

und der molekularen Ordnung in festen Polymeren<br />

C 9 Zeitaufgelöste Schwingungsspektroskopie Laubereau 1984–1995<br />

an Polymeren mit ultrakurzen Laserimpulsen<br />

C 10 Zeitaufgelöste toposelektive Spektroskopie von asso- Laubereau 1984–1986<br />

ziierten Flüssigkeiten mit ultrakurzen Laserimpulsen<br />

C 12 Spektroskopische Charakterisierung flüssigkristalliner Wokaun 1987–1995<br />

Tensidphasen und lichtempfindlicher Polymere<br />

C 13 Untersuchungen zur Konformation und Segment- Nuyken 1988–1992<br />

beweglichkeit von Polymeren<br />

C 14 Neue flüssigkristalline Modellsubstanzen und Polymere Lattermann 1988–1995<br />

445


22 Documentation of the Collaborative Research Centre 213<br />

C 15 Photoleitende Polymere Strohriegl 1988–1995<br />

C 16 Segmentierte Poyurethan-Elastomere mit und ohne Eisenbach 1988–1991<br />

Wasserstoffbrückenbindungen: Synthese von Modellsystemen,<br />

Struktur, Morphologie und Eigenschaften<br />

C 17 Spektrales Verhalten von Photochromen Eisenbach 1988–1995<br />

und Fluorophoren in Polymeren und deren<br />

Verwendung als Molekülsonden<br />

C 18 Brummgele Hoffmann 1990–1995<br />

C 19 Spin-Korrelation bei magnetischen Lipidschichten Mertens 1990–1995<br />

C 21 Toposelektive Reaktionen zur Darstellung und Schmidt 1993–1995<br />

Charakterisierung supramolekularer Strukturen<br />

C 22 Molekulare Verstärkung von Polymeren Eisenbach 1993–1995<br />

durch Polydiacetylen-Propfcopolymere<br />

C 23 Untersuchung des Energietransfers in Copolymeren Seilmeier 1993–1995<br />

durch ultraschnelle Thermometrie nach Anregung<br />

im mittleren Infrarot<br />

22.2.4 Projektbereich D: Biopolymere<br />

D 1 Optische Spektroskopie an DNA-Daunomycin Haarer 1984–1990<br />

Intercalationskomplexen<br />

D 2 Stark-Effekt an optisch anisotropen photochemischen Friedrich 1984–1986<br />

Löchern: Antennenpigmente in organischen Gläsern<br />

D 3 Elektronenspinresonanz-Spektroskopie mit Faulhammer 1984–1989<br />

Proteinen und Nukleoprotein-Komplexen<br />

D 4 Magnetische Kernresonanzspektroskopie spezifisch Sprinzl 1984–1995<br />

markierter Ribonukleinsäuren, Ribonukleoproteinkomplexe<br />

und Oligodesoxynukleotide<br />

D 5 Spektroskopische <strong>Report</strong>ergruppen zur Untersuchung Sprinzl 1984–1995<br />

der Struktur und Dynamik der Protein-Nukleinsäure-Wechselwirkung<br />

D 6 Strukturelle Untersuchungen an Oligonukleotiden Wokaun 1987–1995<br />

und DNA-Assoziationsverbindungen<br />

D 7 Gittersolitonen auf Polypeptid-Ketten in Proteinen Mertens 1987–1989<br />

D 8 Die Struktur von Makromolekülassoziaten in Lösung Rösch 1993–1995<br />

am Beispiel von Protein-Nukleotid Komplexen<br />

Ye 1 Untersuchung photoleitender Polymerthiophene auf Rentsch 1992–1995<br />

der Pikosekunden- und Subpikosekunden Zeitskala<br />

Yw 1 Untersuchung photoleitender Polythiophene Laubereau 1992–1995<br />

auf der Piko-Subpikosekunden-Zeitskala<br />

mit optischen und elektrischen Methoden<br />

446


22.3 Guests<br />

22.3 Guests<br />

(Guests with a duration of stay longer than 2 weeks)<br />

Name Acad. grad. Institute Year<br />

Allakhverdiev, K. Prof. Academy of Sciences, Baku 1991/95<br />

Aranson, I. Dr. Uni Jerusalem, Israel 1990/92<br />

1994<br />

Arnold, L. Dr. Insitute of Organic Chemistry. 1989/90<br />

Czech. Acad. of Sc., Prague 1991/93/94<br />

Bachvarov, I. Dr. Uni Sofia, Griechenl. 1993<br />

Bara, M. Dr. Université Paris 1985/86<br />

Bartusch, G. Dipl.-Chem. TU Dresden 1990/91<br />

Boesch, R. Dr. Universität Dijon 1992<br />

Brezesinski, G. Dr. Martin-Luther-Universität Halle Wittenberg 1989<br />

Buka, A. Prof. Inst. f. Physics, Budapest 1989/90/91<br />

1992/93/94<br />

Caceres, M. Dr. Centro Atomico, Bariloche 1990/92<br />

Cameron, J.H. Dr. Heriot-Watt University Edinburgh 1995<br />

Chen, S.H. Prof. Massachusetts inst. of Technology, USA 1987/88<br />

Chigrinov,V. Prof. Organic Intermediats & Dyes Inst., Moscow 1992/94<br />

Danielius, R. Dr. Vilna, Litauen 1989<br />

Delev,V. Dr. Baschkirische Uni, Ufa 1994<br />

Dobrov,V. Dr. State University Moscow 1994<br />

Eber, N. Dr. Academy of Sciences Budapest, Hungarian 1994/95<br />

Favorova, O. Dr. Soviet Academy of Sciences, Moscow 1989<br />

Fidy, J. Dr. Inst. of Biophysics Budapest 1994<br />

Foltynowicz Poznan 1991<br />

Gadonas Dr. Universität Vilnius 1989/90<br />

Gaididei, Y. Prof. Academy of Sciences Kiev, Ukraine 1994<br />

Gammel, J.T. Dr. National Laboratory, Los Alamos 1990<br />

Getautis,V. Dr. Uni Kaunas, Litauen 1994<br />

Golov, A. Dr. Academy of Sciences Chernogolovka, Moscow 1994<br />

Gouvèa, M.E. Dr. Universidade Federal de Minas Gerais, Brasilien 1993/94<br />

Grazulevicius, J. Dr. Technicol University Kaunas 1992/95<br />

Herenyi, L. Institut of Biophysics University of Budapest 1994<br />

Hiltrop, K. Dr. University Paderborn 1987<br />

Imae, T. Prof. Nagoya University Japan 1991<br />

Imrich, J. Dr. P.J. Safarik University Kosice 1994<br />

Ivanov, B. Dr. Institute for Metical Physics Kiev, Ukraine 1994<br />

Jonak, J. Dr. Institute of Molecular Biology 1988<br />

Czech. Acad. of Sc., Prague<br />

Joshi, R.L. Université Paris VII 1986<br />

447


22 Documentation of the Collaborative Research Centre 213<br />

Name Acad. grad. Institute Year<br />

Kasperczyk, J. Dr. Zawiercie, Polen 1989/90<br />

1991/92<br />

Katunin,V. Dr. Inst. of Nuclear Physics, Soviet Acad. of Sc., 1993/95<br />

St. Petersburg<br />

Kauffmann, H. Prof. Universität Wien 1991<br />

Kharlamov, B. Dr. Academy of Sciences, Moscow 1992/94<br />

Kikas, J. Dr. Inst. of Physics, Tartu 1994/95<br />

Kim, H. S. Prof. University of Seoul Südkorea 1991<br />

Kohler, B. Prof. University California 1984/85/<br />

88/90<br />

Korrovits,V. Dr. Academy of Sciences, Tartu 1988<br />

Kotomin, E. Dr. Universität Riga 1990/91<br />

Kozar, T. Dr. Academy of Sciences Kosice Slovakia 1993<br />

Krekhov, A.P. Dr. Baschkirische Uni,.Ufa 1993/94<br />

1995<br />

Kumar, A. Dr. Calcutta-University, Indien 1988/89/90<br />

Kurlat, D.H. Prof. University Buenos Aires Argentinien 1987<br />

Li, W. Dr. 1991/92<br />

Lin, T.L. Prof. Universität Taiwan 1991<br />

Mainz<br />

Matyjaszewsky, K. Prof. Carnegia-Mellon-Uni. Pittsburgh, USA 1991<br />

Mikhin, N. Dr. Academy of Sciences Kharkov, Ukrainian 1995<br />

Miller, C.A. Prof. Rice University Dep. Houston Texas, USA 1989/95<br />

Mistriotis, A. Dr. Universität Kreta 1991<br />

Möhle, L. Dr. Leuna-Werk, Merseburg 1990/91<br />

Muthukumar, M. Prof. University of Massachusetts 1994<br />

Nawrot, B. Dr. Technical University Poznan Pol<strong>and</strong> 1994/95<br />

Nesrullajev, A.N. Prof. Universität Baku 1989<br />

Nikogosyan, D.N. Prof. Inst. of Spectroscopy 1993<br />

Academy of Sciences, Moscow<br />

Nool<strong>and</strong>i, J. Prof. Xerox Research Centre of Canada 1991/92<br />

Ofeng<strong>and</strong>, J. Prof. Roche Inst. of Mol. Biology Nutley, New Jersey 1988<br />

Ollikainen, O. Academy of Sciences, Tartu 1991<br />

Personov, R. Prof. Inst. of Molecular Spectroscopy, Dep. Moscow 1987/91<br />

Planer-Kühner, G. MPI f. Polymerforschung Pushchino, Russian 1990/91<br />

Quèmarais, P. Prof. LEPES – CNRS, Grenoble 1994/95<br />

Ragunathan,V. Dr. Bangalore, Indien 1989<br />

Rebane, A. Dr. Academy of Sciences, Tartu 1987/88/92<br />

Renge, I. Dr. Academy of Sciences, Tartu 1991<br />

Reshetnikova, L. Dr. Inst. of Mol. Biology Acad. of Sciences, Moscow 1990/91<br />

Riecke, H. Prof. Uni Northwestern, USA 1991/94<br />

Rieckhoff, K. Dr. Simon Frazer University North Vancouver 1987/90<br />

448


22.3 Guests<br />

Name Acad. grad. Institute Year<br />

Rötger, A. Dr. Universitè Fourier, Grenoble 1994<br />

Ruckenstein, E. Prof. University Buffalo, USA 1985/86<br />

Rudinger, J. Dr. Centre National de la Recherche Scientifique, 1992/93<br />

Strasbourg<br />

Salaev, F. Prof. Academy of Sciences of Baku, Aserbaijan 1992<br />

Santa, I. Dr. Academy of Sciences Budapest, Ungarn 1990/91<br />

Sarbak, Z.M. Dr. Universität Poznan, Polen 1989<br />

Schott, M. Prof. Université Paris VII 1985<br />

Schneider, A. Martin-Luther-University 1991<br />

Inst. of Biochemistry, Halle-Wittgenstein<br />

Sédlak, E. University Kosice 1994<br />

Serra i Albet, A. Dr. University Barcelona 1987<br />

Shalaby, G.A. Faculty of Science 1995<br />

Minoufia University Shebin El-kom, Ägypten<br />

Shchipunov,Y.A. Prof. Academy of Sciences Vladivostok, Russia 1993<br />

Shirokov,V. Dr. Academy of Sciences 1991<br />

Sigler, P. B. Prof. Dep. of Molecular Biophysics <strong>and</strong> Biochemistry, 1995<br />

Yale University, New Haven<br />

Silber, M. Dr. Passadena, Caltech, USA 1991<br />

Smrt, J. Dr. Inst. of Organic Chemistry 1988/89<br />

Czech. Acad. of Sc., Prague 1990<br />

Sobkowski, M. Dr. Inst. of Bioorganic Chemistry, 1991<br />

Polish Acad. of Sc., Poznan<br />

Spirin, A. Prof. Inst. of Protein Research, Academy of Sciences, 1989<br />

Pushchino, Russia<br />

Stasko, A. Prof. Technical University, Bratislava, CSFR 1990/91<br />

Steinberg, S. V. Dr. Inst. of Molecular Biology 1991/92<br />

Soviet Acad. of Sc., Moscow 1993/94<br />

Stoylov, S.P. Prof. Academy of Sciences Sofia, Bulgarien 1993/94/95<br />

Suzuki, H. Dr. NTT Opto-Electronic Laboratories Tokai-Mure, 1989/92/93<br />

Naka-Cun1<br />

Svitova, T. Prof. Academy of Sciences, Chem. Dep., Moscow 1994<br />

Szkaradkiewicz, K. University of Poznan 1993/97<br />

Tamori, K. Academy of Sciences, Tokyo 1990<br />

Tezak, D. Dr. University Zagreb, Croatia 1988<br />

Thelakkat, M. Dr. NSS College, Manjeri, 1995<br />

University of Calicut, India<br />

Trommsdorff, P. Prof. University of Grenoble 1990<br />

Tunkin,V. Dr. University Moscow 1991/92<br />

Ueda, T. Dr. University of Tokyo 1989<br />

Vainer,Y. Dr. Academy of Sciences Moscow 1993/94<br />

Valiente-Martinez, M. Dr. Universidad de Alcala de Henares, Spanien 1992<br />

449


22 Documentation of the Collaborative Research Centre 213<br />

Name Acad. grad. Institute Year<br />

Vitukhnovski, A. Dr. Academy of Sciences Moscow 1990<br />

Vodopyanov, K Dr. Inst. of General Physics Moscow 1990/92<br />

Wada, Y. Prof. University of Tokyo 1988/93<br />

Wysin , G. Dr. Kansas State University 1990/91<br />

Xu, J. Chengdu University, China 1990<br />

22.4 Co-workers<br />

Name Acad.grad Institute Project Period<br />

Adam, D. Exp.Physik B3 1991/92/93/94<br />

Aechtner, Dr. Exp.Physik C9/10 1984/85/86/89<br />

Aggarwal, K. Dr. Makro.Chem. C13 1989<br />

Ahmadian, R. Dr.* Biochem. D5 1989/91/92<br />

Amberger, E. Dr.* Anorg.Chem. C5 1989/90/91<br />

Ambrosch, A. Makro.Chem.. C14 1990/91/92<br />

Angel, G. Dr.* Exp.Physik C9 1984–89<br />

Angel, M. Dr.* Phys.Chem. C1/C2 1984/85/86<br />

Angel, S. Dr.* Phys.Chem. C2 1987/88/89<br />

Angstl, R. Dr.* Exp.Physik B2 1985–89<br />

Bächer, R. Dr.* Phys.Chem. C2 1986/87/88<br />

Bachmann, S. Biochem. D4/D5 1987/88/89<br />

Bara, M. Exp.Physik B1 1985/86<br />

Baranowski, D. Theo.Physik B8 1990–94<br />

Barnickel, P. Dr.* Phys.Chem. D6 1989/91<br />

Barth, K. Dr.* Exp.Physik B14 1991/92/93/94<br />

Bauer, F. Exp.Physik B13 1994/95<br />

Bauer, H.-D. Dr.* Exp.Physik B2/B13 1985/86/88–91<br />

Beck, M. Biochemie D4/D5 1985–89<br />

Beginn, Ch. Dr.* Makro.Chem. C15 1993<br />

Beginn, U. Makro.Chem. C14 1993<br />

Beikmann, M. Biochemie D4/D5 1993/94/95<br />

Bettenhausen, J. Makro.Chem. C15 1993/94/95<br />

Bieger, P. Dr.* Biochem. D5 1984–89<br />

Bittl, Th. Biopolymere D8 1993<br />

Blank, J. Biochemie D5 1991/92<br />

450


22.4 Co-workers<br />

Name Acad.grad Institute Project Period<br />

Blechschmidt, B. Dr. Biochemie D5 1994/95<br />

Blumen, A. Prof. Exp.Physik B11 1987–91<br />

Bodenschatz, E. Dr.* Theor.Physik C4 1985–89<br />

Bojer, B. Anorg.Chem. A1/C5 1987/88/89<br />

Bratengeier, K. Exp.Physik C9 1984/85/86<br />

Breinl, W. Dr.* Exp.Physik D2 1986<br />

Bronold, F. Theo.Physik B6 1991/95<br />

Buchwald, E. Exp.Physik B13 1995<br />

Büttner, H. Prof. Theor.Physik B6/B8 1984–1995<br />

Burger, A. Dr.* Phys.Chem. C1 1987/88/89<br />

Burkhardt,V. Dr.* Makro.Chem. C13 1989/90/91/92<br />

Caceres, O. Dr. Theor.Physik C4 1990<br />

Carron, K. Dr. Phys.Chem. D6 1986/87/88<br />

Cuc, T. T. Dr. Makro. Chem. C3/C7 1984/85<br />

Dahinten, T. Exp.Physik C23 1993/94<br />

Decker, W. Theor.Physik C4 1991/92/93<br />

Deeg, M. Theor.Physik B8 1993/94/95<br />

Denninger, G. Dr.** Biochemie/Exp.Phys. D3/B4 1984–89<br />

Denzner, M. Anorg.Chem. C5 1988/89<br />

Diemer, E. Phys.Chem. C18 1988–92<br />

Dirnberger, K. Makro.Chem. C17 1993/94/95<br />

Döge, B. Phys.Chem. C1 1986/87/88/89<br />

Dörfler, S. Biochemie D4/D5 1994/95<br />

Dohlus, R. Dr.* Exp.Phys. C9 1986/87<br />

Domes, H. Dr.* Exp.Phys. B3 1984–89<br />

Dormann, E. Prof. Exp.Phys. B4 1984–1990<br />

Dreßel, U. Exp.Phys. B4 1988/89<br />

Ebert, G. Dr.* Phys.Chem. C1 1986/87/88/89<br />

Eisenbach, C.D. Prof. Makro.Chem. C16/C17 1988–1995<br />

Ernst, U. Anorg.Chem. A1/C5 1985–89<br />

Esquinazi, P. Dr. Exp.Phys. B10 1988/89<br />

Faulhammer, H. Dr. Biochemie D3 1984–1989<br />

Fehn, T. Exp.Phys. B13 1992/93<br />

Fehske, H. Dr.** Theor.Phys. B8 1990–1995<br />

Feile, M. Exp.Phys. B4 1988/89<br />

Feng, Q. Dr.* Theor.Phys. C4 1990/91<br />

Fesser, K. Prof.** Theor.Phys. B6 1984–1995<br />

Feyerherd, B. Makro.Chem. C14 1992<br />

Ficht, K. Dr.* Makro.Chem. C17 1990/91/92<br />

Fickenscher, M. Exp.Phys. C9 1985/86/89/90<br />

Fiebig, A. Exp.Phys. B4 1986–90<br />

Fischer, K. Dr.* Makro.Chem. C17 1986/87/88/89<br />

451


22 Documentation of the Collaborative Research Centre 213<br />

Name Acad.grad Institute Project Period<br />

Fischer, W. Dr.* Biochemie D4 1984/85<br />

Flöser, G. Dr.* Exp.Phys. D1 1984/85/86/88<br />

Förster, Ch. Dr.* Biochemie D4/D5 1990/91/92/93<br />

Franzke, D. Dr.* Phys.Chem. C12 1987/88/89/91<br />

Friedrich, J. Prof. Exp.Phys. B9/B15 1984–1989<br />

1993–1995<br />

Frosch, H. Theor.Phys. B5 1984/85<br />

Gafert, J. Exp.Phys. B15 1994/95<br />

Gagel, R. Dr.* Exp.Phys. C9/10 1986–91<br />

Gammel, T. Theor.Phys. B8 1990<br />

Gebhardt, H. Exp.Phys. B4 1988<br />

Gebhardt-Singh, E. Dr.* Bioochemie D5 1985/86<br />

Geissinger, P. Dr.* Exp.Phys. B12 1990–94<br />

Geist, F. Exp.Phys. B7 1991/92<br />

Giering, T. Exp.Phys. B12 1994/95<br />

Gläser, D. Exp.Phys. A6 1986/87/88/89<br />

Gollner, G. Makro.Chem. C14 1988/89<br />

Gotschy, B. Exp.Phys. B2 1987<br />

Gradl, G. Dr.* Exp.Phys. D2 1987/88/89<br />

Gradzielski, M. Dr.* Phys.Chem. C18 1988/89/90/91<br />

Graener, H. Dr.** Exp.Phys. C9 1986/87/88/89<br />

Grillenbeck, N. Biochemie D4/D5 1988–92<br />

Grottenmüller, R. Makro.Chem. C21 1993/94/95<br />

Grotz-Green, C. Anorg.Chem. C5 1988/89<br />

Gruner-Bauer, P. Dr.* Exp.Phys. B4 1986–91<br />

Güttler, W. Dr. Exp.Phys. A3 1984–89<br />

Haarer, D. Prof. Exp.Phys. A5/B3/B12 1984–1995<br />

Häfner, W. Dr. Exp.Phys. C6 1984–89<br />

Haegel, F.-H. Dr.* Phys.Chem. C1/D6 1984–88<br />

Häfner, W. Phys.Chem. D6 1986/87/88/89<br />

Härtl, P. Dr.* Biochemie D3 1985/86/87/88<br />

Hahn, Ch. Phys.Chem. D6 1995<br />

Halbritter, A. Biochemie D4/D5 1987/88/89<br />

Hammon, W. Dr.** Anorg.Chem. C5 1986<br />

Hecht, E. Phys.Chem. C18 1993/94/95<br />

Heindl, D. Exp.Phys. B4 1986/87<br />

Heitz, T. Makro.Chem. C3 1986<br />

Henglein, F. Phys.Chem. D6 1988/89<br />

Herold, M. Exp.Phys. B13 1992/93/94<br />

Herrmann, F. Dr.* Biochemie D8 1994/95<br />

Hertel, G. Dr.* Phys.Chem. C2 1986/87/88/89<br />

Hirschmann, R. Dr.* Exp.Phys. B9 1987/88/89<br />

452


22.4 Co-workers<br />

Name Acad.grad Institute Project Period<br />

Hochstraßer, D. Dr.* Theor.Phys. D7 1986/87/88/89<br />

Höcker, H. Prof. Makro.Chem. C3/C7 1984–1986<br />

Hoff, H. Dr.* Makro.Chem. C17 1988/90/91<br />

Hoffmann, H. Prof. Phys.Chem. C1/C2/C18 1984–1995<br />

Hoffmann, K. Dr.* Anorg.Chem. C5 1986<br />

Hofmann, J. Makro.Chem. C22 1993/94/95<br />

Hofmann, M. Biochemie D4/D5 1985/86<br />

Hofmann, M. Dr.* Exp.Phys. C9 1992/93/94/95<br />

Hofmann, W. Dr.* Exp.Phys. B7 1984–88<br />

Hoffmüller, P. Biochemie D4/D5 1995<br />

Hopfmüller, H. Exp.Phys. C6 1984/85/86<br />

Huber, G. Dr.* Phys.Chem. C1/C2 1984–89<br />

Hübner, J. Exp.Phys. B13 1993/94/95<br />

Hüser, B. Makro.Chem. C8 1984/85/86<br />

Hums, E. Dr. Anorg.Chem. C5 1984/85<br />

Illner, J.- Ch. Dr.* Phys.Chem. C1 1991–95<br />

Juarez, M. de L.T. Dr.* Theor.Phys. C4 1988/89<br />

Kaiser, M. Dr.* Theor.Phys. C4 1987/88/89<br />

Kalus, J. Prof. Exp.Phys. B7 1984–1992<br />

Kanischka, G. Makro.Chem. C3/C7 1984/85/86<br />

Kaul, H. Dr.* Exp.Phys. B2 1988/89<br />

Kiefer, W. Prof. Exp.Phys. C6 1984–1986<br />

Klenke, M. Phys.Chem. C12 1992/93/94/95<br />

Knöchel, F. Makro.Chem. C7 1986<br />

Köhler, G. Dr.* Exp.Phys. B11 1987–92<br />

Köhler, M. Exp.Phys. B15 1995<br />

Köhler, W. Dr.* Exp.Phys. D2 1984–88<br />

Kohles, N. Exp.Phys. C9 1984/85/86/87<br />

König, R. Dr. Exp.Phys. B10 1988–93<br />

Krämer, U. Dr.* Phys.Chem. C2 1987/88/89<br />

Kramer, L. Prof. Theor.Phys. C4 1984–1995<br />

Krapf, M. Exp.Phys. B7 1990/91<br />

Krauss, H.L. Prof. Anorg.Chem. A1/C5 1984–1992<br />

Kreutzer, R. Dr. Biochemie D4/D5 1990/91/92<br />

Kuang, W. Theor.Phys. C4 1986/87<br />

Lattermann, G. Dr. Makro.Chem. C14 1989–1995<br />

Laubereau, A. Prof. Exp.Phys. C9/10/Yw1 1984–1995<br />

Lenk, M. Biochemie D4/D5 1991–95<br />

Lenz, U. Dr.* Phys.Chem. C1 1988/89/90/92<br />

Lieberth, M. Exp.Phys. B14 1990/91<br />

Li, J. Exp.Phys. B10 1993/94<br />

Limmer, St. Dr. Biochemie D4 1990–94<br />

453


22 Documentation of the Collaborative Research Centre 213<br />

Name Acad.grad Institute Project Period<br />

Lindenberger, F. Exp.Phys. C20 1991/1992<br />

Löbl, M. Dr.* Phys.Chem C1/C2 1984/85<br />

Löw, A. Biochemie D3 1989<br />

Lorenz, W. Exp.Phys. B10/B13 1990/91/92<br />

Luding, St. Exp.Phys. B11 1990/91/92<br />

Maier, H. Exp.Phys. B10 1992/93/94/95<br />

Martins, J. M. N. Makro.Chem. C14 1994/95<br />

Materny, A. Exp.Phys. B2 1987/88<br />

Matuschek, D. Exp.Phys. B11 87/88/89/90<br />

Mertens, F.G. Prof. Theor.Phys. C19/D7 1987–1995<br />

Meyering-Vos, M. Dr. Biochemie D4/D5 1995<br />

Merkel, K.-R. Dr.* Anorg.Chem. A1/C5 1985/86<br />

Meyer, G. Dr.* Phys.Chem. C12 1988/89/92<br />

Meyer, H. Dr.* Exp.Phys. B3 1989/90/91/92<br />

Mirke, M. Makro.Chem. C21 1993/94/95<br />

Morys, P. Prof. Dr. Anorg.Chem. C5 1986/87/88/89<br />

Müller, K.-P. Dr.* Exp.Phys. B10 1986–91<br />

Müller, R.G. Dr.* Makro.Chem. C15 1988/89<br />

Müller-Horsche, E. Dr.* Exp.Phys. B3 1984/85<br />

Müller-Nawrath, R. Dr. Exp.Phys. B1 1984/85/86<br />

Munkert, U. Dr.* Phys.Chem. C2 1995<br />

Nattler, G. Exp.Phys. C9 1988/89<br />

Nawrot, B. Dr. Biochemie D4/D5 1994<br />

Nefzger, H. Dr.* Makro.Chem. A6 1986/87/88/89<br />

Nittke, A. Exp.Phys. B10 1993/94<br />

Nomayo, M. Phys.Chem. D6 1993/94<br />

Nuyken, O. Prof. Makro.Chem. C13 1988–1992<br />

Oetter, G. Dr.* Phys.Chem. C1 1986/87/88<br />

Ott, G. Dr. Biochemie D4 1988–1994<br />

Paap, H.-G. Theor.Phys. C4 1985/86<br />

Pecher, U. Theor.Phys. B8 1990/91/92/95<br />

Pesch, W. Prof. Dr. Theor.Phys. C4 1984–88<br />

Peter, M. Dr.* Biochemie D5 1986/89<br />

Platz, G. Prof. Dr. Phys.Chem. C1/C2 1984/85<br />

Pöhlmann, E. Exp.Phys. B4 1985/86<br />

Pöhlmann, T. Makro.Chem. C13 1990/91/92<br />

Pößnecker, G. Dr.* Phys.Chem. C2 1987/88/89/90<br />

Pobell, F. Prof. Exp.Phys. B10 1987–1995<br />

Pschierer, H. Exp.Phys. B15 1993/94<br />

Purucker, H.-G. Dr.* Exp.Phys. C9 1987/88/92<br />

Raible, Ch. Exp.Phys. B13 1991/92<br />

Rauscher, A. Dr.* Phys.Chem. C1 1987–92<br />

454


22.4 Co-workers<br />

Name Acad.grad Institute Project Period<br />

Rehage, H. Dr.** Phys.Chem. C1/C2 1984–89<br />

Rehberg, I. Dr. Theor.Phys. C4 1988/89<br />

Reichl, J. Dr.* DipL.Phys. B5 1984–86/88/89<br />

Reichstein, W. Exp.Phys. A5 1985–89<br />

Reiser, Ch. Dr.* Biochemie D5 1987–91/93<br />

Rentsch, S. Doz., Dr. Exp.Phys. Ye1 1992–1995<br />

Renner, H. Dr.* Phys.Chem. C2 1994/95<br />

Reul, St. Dr.* Exp.Phys. B10 1991/92<br />

Ribbe, A. Dr.* Makro.Chem. C17 1992<br />

Richter, W. Dr. Exp.Phys. B14 1990–1995<br />

Riederer, W. Dr.* Anorg.Chem. C5 1986/87/88/89<br />

Robisch, P. Phys.Chem. C2 1988/89<br />

Roder, S. Exp.Phys. B7 1986<br />

Rösch, P. Prof. Struktur und Chem. d. Biop. D8 1993–1995<br />

Röthlein, P. Exp.Phys. B7 1990<br />

Röttger, J. Makro.Chem. C3/C7 1984/85/86<br />

Rothemund, P. Phys.Chem. C1 1986/87<br />

Schellenberg, P. Dr.* Exp.Phys. B15 1994<br />

Schellhorn, M. Dr.* Makro.Chem. C14 1990/91/92/93<br />

Schießwohl, M. Biochemie D4 1990/91/92<br />

Schirmer, N. Dr.* Biochemie D5 1991/92<br />

Schmelzer, U. Dr. Exp.Phys. B7 1986/87/88/89<br />

Schmerbeck, S. Dr.* Anorg.Chem. C5 1984/85/86<br />

Schmid, W. Dr.* Exp.Phys B2/B13 1994<br />

Schmidt, H. Anorg.Chem. A1 1986/87/88/89<br />

Schmidt, St. Makro.Chem. C13/C14 1990/92<br />

Schmidt, M. Prof. Makro.Chem. C21 1993–1995<br />

Schmutzler, M. Exp.Phys C9 1985/86<br />

Schnabel, E. Phys.Chem. C1 1986/87<br />

Schneider, J.M. Makro.Chem. C15 1992<br />

Schnitzer, H.-J. Theor.Physik C19 1993/94/95<br />

Schnörer, H. Dr.* Exp.Phys. B11 1987–92<br />

Schörner, H. Makro.Chem. C15 1986/87/88/89<br />

Schultes, H. Dr.* Exp.Phys. B4 1984/85/86/87<br />

Schulz, S. Dr. Phys.Chem. C1 1990/91/92<br />

Schumann, Ch. Anorg.Chem. A1 1990/91/92/93<br />

Schw<strong>and</strong>ner, B. Dr.* Phys.Chem. C1 1984/85/86<br />

Schwarz-Schultz Dr. Biochemie D4 1985/86/87<br />

Schwoerer, M. Prof. Exp.Phys. B1/B2/B13 1984–1995<br />

Seidler, L. Dr.* Biochemie D5 1984/85/86<br />

Seilmeier, A. Prof. Exp.Phys. C23 1993–1995<br />

Sesselmann, R. Dr.* Exp.Phys. D1 1987/88<br />

455


22 Documentation of the Collaborative Research Centre 213<br />

Name Acad.grad Institute Project Period<br />

Siebenhaar, B. Dr.* Anorg.Chem. C5 1987/88/89<br />

Spiess, H.W. Prof. Makro.Chem. C8 1984–1986<br />

Sprinzl, M. Prof. Biochemie D4/D5 1984–1995<br />

Sühler, G. Phys.Chem. C1/C2 1993/94/95<br />

Staufer, G. Dr.* Makro.Chem. C14 1986/87/88/89<br />

Stöcklein, W. Exp.Phys. B1 1985<br />

Strohriegl, P. Dr.** Exp.Phys. B2/B4 1984–89<br />

Struller, B. Dr.* Phys.Chem. C2 1986/87/88/89<br />

Sum, U. Dr.* Theor.Phys. B6 1984–89<br />

Terskan-Reinhold, M. Makro.Chem. C17 1992/93/94<br />

Thaufelder, H. Dr.* Makro.Chem. C16/17 1986–91<br />

Thom, W. Theor.Physik C4 1987/88<br />

Thurn, H. Dr. Phys.Chem. C1 1987<br />

Thurn, R. Exp.Phys. C6 1984–89<br />

Trapper, U. Theor.Phys. B8 1995<br />

Treiber, M. Theor.Phys. C4 1992//93/94/95<br />

Troeger, P. Dr.* Exp.Phys. C10 1984/85<br />

Überla, H. Dr.* Theor.Phys. B5 1984/85/86<br />

Ulbricht, W. Dr. Phys.Chem. C1/C2 1984–89<br />

Völkel, A. Theor.Phys. C19 1990/91/92/93<br />

Vogtmann, T. Dr.* Exp.Phys. B13 1987/88/89/92<br />

Voit, J. Dr.** Theor.Phys. B8 1986/87/88/89<br />

Voit, S. Exp.Phys. B13 1990<br />

Vollstädt, K.-U. Exp.Phys. B13 1990/91/92<br />

Vornlocher, H.-P. Biochemie D4/D5 1992/93<br />

Voss, D. Exp.Phys. B3 1995<br />

Waas,V. Dr.* Theor.Phys. B8 1986/87/88/89<br />

Wagner, R. Dr.* Exp.Phys. B2 1987–91<br />

Walther, K.L. Dr.* Phys.Chem. C12 19898/89<br />

Wanka, G. Phys.Chem. C2 1988/89<br />

Weber, R. Dr.* Phys.Chem. C2 1984/85/86<br />

Weber, S. Dr.* Theor.Phys. B5/B8 1984–89<br />

Wefing, St. Makro.Chem. C8 1984/85/86<br />

Weidlich, K. Dr.* Exp.Phys. C9 1994/95<br />

Weiss K. Dr.** Anorg.Chem. A1 1986/87/88/89<br />

Weigel, J. Phys.Chem. C12 1993/94<br />

Weißhaar, M. Dr. Biochemie D4/D5 1987/88/89<br />

Wiesner, J. Dr.* Phys.Chem. C12 1986–90<br />

Wietasch, H. Makro.Chem. C14 1994/95<br />

Winter, H. Exp.Phys. B4 1987/88/89<br />

Wittenbeck, P. Phys.Chem. C12 1988/89<br />

Wokaun, A. Prof. Phys.Chem. C12/D6 1987–1995<br />

456


22.5 International Cooperation<br />

Name Acad.grad Institute Project Period<br />

Wolf, M. Dr.* Theor.Phys. B6 1989–94<br />

Wolff, P. Anorg.Chem. C5 1987/88/89<br />

Wolfrum, K. Exp.Phys. C9 1986/87/92<br />

Wunderlich, I. Dr.* Phys.Chem C1/C2 1984–89<br />

Yamaguchi, Y. Phys.Chem. C1 1993/94/95<br />

Ye, T. Dr.* Exp.Phys. C9 1987–91<br />

Zahn, M. Anorg.Chem. C5 1986/87/88/89<br />

Zimmermann, F. Dr.* Phys.Chem. D6 1988–93<br />

Zimmermann, H. Exp.Phys. B7 1984/85<br />

Zimmermann, W. Dr* Exp.Phys. C4 1984–89<br />

Zollfrank, J. Dr.* Exp.Phys. B9 1987/88<br />

Support of young scientists<br />

* “Promotion” supported by the Sonderforschungsbereich 213<br />

** “Habilitation” supported by the Sonderforschungsbereich 213<br />

<strong>and</strong> about 200 diploma theses<br />

22.5 International Cooperation<br />

A6<br />

Prof. Hashimoto, Kyoto University, Kyoto, Japan<br />

Dr. H.G. Schmeler, Miles Inc. Pitsburgh, PA, USA<br />

B3<br />

NTT, Tokyo, Japan<br />

Universität Wien, Österreich<br />

B6<br />

Los Alamos National Laboratory T-11, USA<br />

Prof. Y. Ono, Japan<br />

B8<br />

Dr. Bishop, LANL, USA<br />

Los Alamos National Laboratory, Theoretical Group, Los Alamos, NM, USA<br />

B10<br />

MIT, Boston, USA<br />

Prof. Yu Kogan, Dr. Burin, Kurchatov Institut, Moskau, Rußl<strong>and</strong><br />

Dr. R. Nava, Universidad Central de Venezuela, Caracas<br />

457


22 Documentation of the Collaborative Research Centre 213<br />

B12<br />

Prof. R. Silbey, MIT Cambridge, USA<br />

Prof. Skinner, Dept. of Chemistry, University Wiscousin, USA<br />

C1/C2/C18<br />

Prof. Dr. J.S. C<strong>and</strong>au, Universite Louis Pasteur, Strasbourg, Frankreich<br />

Dr. M.E. Cates, Cavendish, Laboratory Cambridge, Großbritannien<br />

Prof. Dr. C.A. Miller, Rice-University, Houston USA<br />

Prof. Dr. D. Langevin, Universite Paris VI, Paris, Frankreich<br />

Dr. K. Esumi, Nagoya University Tokyo, Japan<br />

Dr. Kell Mortensen, Risö Laboratorium, Dänemark<br />

Doz. Dr. P. Stern, Akad. d. Wissenschaften Tschechien<br />

Dr. D. Roux, CNRS Pessac, Frankreich<br />

C4<br />

Weizmann Institut, Dept. of Physics, Israel<br />

Group de Physique de Solides, Orsay, Frankreich<br />

Dept. of Physics, University California, Santa Barbara, USA<br />

Dept. of Physics, Hebrew-University, Jerusalem, Israel<br />

Dept. of Physics, CALTECH, Passadena, USA<br />

Dept. of Physics, Northwestern University, Evanston, USA<br />

Centro Atomico, Bariloche, Argentinien<br />

Prof. G. Ahlers, Santa Barbara, USA<br />

Prof. A. Chuvyrov, Ufa, Russl<strong>and</strong><br />

Prof. V. Steinberg, Rehovot, Israel<br />

Dpto. Estructura y Constiuyentos de la Materia, Facultät de Fisica, Universität Bacelona, Spanien<br />

C10<br />

Laser Research Center,Vilnius, Litauen<br />

C12<br />

ETH Zürich<br />

Prof. F.R. Aussenegg/Univ. Doz. Dr. A. Leitner, Institut für Experimentalphysik, Graz, Österreich<br />

Prof. A. Baiker, Chemieingenieurwesen und Industrielle Chemie, ETH Zürich, Schweiz<br />

PD. Dr. A. Leitner, Institut für Experimentalphysik, Universität Graz, Österreich<br />

C13<br />

Slowakische Techn. Hochschule, CSFR<br />

C14<br />

Dr. A.M. Levelut, Universite Orsay, Frankreich<br />

Dr. B. Gallot, Laboratoire des Materiaux Organiques, CNRS, BP 24,Vernaison, Frankreich<br />

C16/17/22<br />

Prof. Blackwell, Case Western Reserve University, Clevel<strong>and</strong>, Ohio, USA<br />

Prof. Mac Knight, University of Massachusetts, Amherst, USA<br />

Dr. J. Nool<strong>and</strong>i, Xerox Research Centre Canada, Missisauga, Ontario, Canada<br />

Prof. Newkomo, Univ. of South Florida, Tampa, Florida<br />

458


22.6 Funding<br />

C19<br />

Center of Nonlinear Studies, Los Alamos, National Laboratory, Los Alamos USA<br />

Servece National des Champes Intenses, Grenoble, Frankreich<br />

C21<br />

N. Pogodina, Institute of Physics, Universite St. Petersburg, Russia<br />

D4/D5<br />

Prof. Dr. A. Redfield, Br<strong>and</strong>eis University, USA<br />

Prof. Dr. O. Uhlenbeck, University of Boulder, USA<br />

Prof. Dr. A. Spirin, Academy of Science, Russia<br />

Prof. Dr. M. Boublik, Roche Institute of Molecular Biology, Nutley, New Jersey, USA<br />

Prof. Dr. B. F. Clark, Aarhus, Denmark<br />

Prof. Dr. J. Nyborg, Aarhus, Denmark<br />

Prof. Dr. O. Lavrik, Russia<br />

Prof. Dr. L. Bosch, Leiden University, The Netherl<strong>and</strong>s<br />

Prof. Dr. P. Sigler,Yale University, USA<br />

Prof. Dr. J. Heinfeld, USA<br />

Dr. M. Makinen, USA<br />

Dr. L. Arnold, Czech Republic<br />

Prof. Dr. L. Spremulli, USA<br />

Prof. Dr. R. Giegé, France<br />

Prof. Dr. A. Barciszewski, Pol<strong>and</strong><br />

D6<br />

Prof. Dr. H. Eicke, Institut f. Phys. Chemie, Universität Basel, Schweiz<br />

D8<br />

Prof. K. Kirschner, Schweiz<br />

Prof. A. Yaniv, Israel<br />

Prof. M. Breitenbach, Österreich<br />

Dr. M. Auer, Österreich<br />

22.6 Funding<br />

The Collaborative Research Centre 213 was supported by grants of the Deutsche Forschungsgemeinschaft<br />

totalling DM 28814200 in the period 1984–1995.<br />

459

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!