Microscopic Interactions and Macroscopic Properties Final Report (1st
Microscopic Interactions and Macroscopic Properties Final Report (1st
Microscopic Interactions and Macroscopic Properties Final Report (1st
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</s